4,712 Matching Annotations
  1. Sep 2023
    1. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Review: In this manuscript the authors generated macrophages derived from the THP-1 cell line or human peripheral blood mononuclear cells stimulated with MCSF and infected them with alphaviruses some containing GFP expression cassettes. In Figure 1, they demonstrate that CHIKV infected these cells more robustly than RRV, SINV or the related ONNV. The authors generated an extensive array of CHIKV/ONNV chimeras to identify the viral proteins that dictate release from infected macrophages and narrowed it down to the envelop proteins E1 and E2. Fine mapping identified a couple of single mutations that affected macrophage infection outcomes. The authors then shifted their approach to identifying env protein interactors using a myc-tag pulldown methods followed by mass spectrometry. The assay identified a number of proteins including those involved in vesicular transport and interferon pathways. siRNA knockdown experiments were performed to identify interactors and many of them were shown to improve virus output.

      Critique: Overall, the manuscript is well written but in its current state it is more like two different stories because the effects of envelop proteins and list of interactors are not brought together in on one story. A possible fix to this problem would be inclusion of ONNV and CHIKV containing env mutations that do and do not restore viral release from macrophages into the pulldown/association experiments shown in Figure 6. The other major issue is the lack of protein data for the viral mutants relative to WT ONNV and CHIKV and assessment of viral RNA in the supernatants to determine whether the block is release or an earlier event since viral RNA levels in the cell seems to be the same or at least normalized. Lastly, knockdown experiments indicate an effect of things like OAS3 or other innate immune modulators. There are no controls to demonstrate that these are specific to CHIKV infection or if knockdown would assist growth of ONNV as well.

      Other points to consider:

      1. The title does not fit the manuscript findings and should be modified.
      2. It is unclear why the authors show results for SINV and RRV in Figure 1. Either these should be removed or the viruses should be carried throughout the experiments described in the Figure. Better yet would be to add additional alphaviruses to this analysis to determine if there are additional viruses that act similarly to CHIKV.
      3. Is the data presented in Figure 1A significant?
      4. The justification for inclusion of Figure 4A is lacking. It is unclear what this panel is supposed to be demonstrating.
      5. There is little justification for the candiates assessed in
      6. Extended data Figure 3 is very difficult to read due to the small font size.
      7. Just to be clear, the blots shown in Figure 6D are different from those depicted in Extended data Figure 4b, because some of them look very similar.

      Significance

      The study provides a fresh look at Alphavirus replication in macrophages. There are a number of issues that should be worked out that would enhance impact and interpretation of this study.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary: The authors utilize: 1) chimeric arthritogenic alphaviruses; evolution selection analyses with virus sequences isolated from human patients; and 3) mass spectrometry and proteomics to interrogate determinants of chikungunya virus (CHIKV) permissiveness in primary human macrophages and the human macrophage cell line, THP-1. The authors find that the vaccine strain, CHIKV 181/clone 25 replicates the most efficiently in primary monocyte-derived macrophages compared to other arthritogenic alphaviruses. Using o'nyong o'nyong (ONNV) as a comparison, the authors generate several chimeric viruses with CHIKV structural proteins and ONNV non-structural proteins (and vice versa) and perform a series of E1 and E2 domain swap experiments. They determine that both CHIKV structural proteins, E2 and E1, are necessary to confer efficient virus production over ONNV in the absence of a difference in viral RNA production. The authors also identify a specific residue in E1 that appears to be important for efficient virus production in THP-1 macrophage cell lines. Finally, using mass spectrometry, the authors identify two host proteins, SPCS3 and eIF3k, that bind to CHIKV E1 structural protein and appear to act as antiviral host factors.

      Major comments: The authors elegantly demonstrate that CHIKV structural proteins confer an advantage over ONNV structural proteins in a step in the replication cycle downstream of virus RNA synthesis, possibly virion assembly. This point would be strengthened determining the particle-to-PFU ratio of the parental viruses and the chimeras. Presumably, the ratio would increase in the chimeras containing CHIKV structural proteins. Additionally, the authors should consider performing virion assembly blocking assays with a small molecule inhibitor to determine if this abrogates the virus production advantage of CHIKV structural proteins within the ONNV backbone. Finally, the authors should perform competition experiments with the chimeric viruses and ONNV in macrophages to determine if the chimeras can outcompete the parental ONNV strain. Based on their data, the chimeric viruses should outcompete. These experiments would likely take 3-4 weeks to complete.

      The authors use both primary macrophages and macrophage cell lines as their in vitro model system and make one of their major points (listed in the title) that the determinants they identified in the CHIKV structural proteins convert macrophages into dissemination vessels; however, they do not show: 1) an in vivo model that the CHIKV-ONNV chimeras disseminate more efficiently than the parental ONNV; and 2) that these chimeras generate virus more efficiently specifically in macrophages. It would be useful to show that ONNV and CHIKV have equivalent virion production in other cell lines and that the advantage conferred by CHIKV structural proteins in the ONNV backbone is specific to macrophages. The authors should also change their title to reflect that dissemination is not directly being addressed in their study; the implications of their in vitro experimentation in a mammalian host would be more appropriate for the discussion.

      OPTIONAL: The authors use CHIKV-ONNV chimeras but it would be interesting to test other chimeras to determine if CHIKV structural proteins confer the same advantage in the backbone of other arthritogenic alphaviruses. The study would also be strengthened by using a pathogenic strain of CHIKV instead of the vaccine strain, as this is significantly attenuated in vivo. In Figure 4, the authors identify residues in the CHIKV structural proteins that appear to be under positive selection in human subjects and generate point mutants in these residues with the corresponding ONNV residues. They find that one mutation, V1029I located in E1, completely abolishes virion production in THP-1 macrophage cell lines. However, in their previous chimeric experiments, they find that neither CHIKV E1 or E2 was sufficient to increase virus production in the ONNV backbone. The authors should address this discrepancy, otherwise they should consider moving the data in their point mutation experiments to a supplementary figure. While worthy of reporting, especially given the patient data, these experiments do not buttress the points made in the previous figures.

      The authors conclude their manuscript with an assessment of several host proteins, namely SPCS3 and eIF3k, that were identified by mass spectrometry and whose knockdown results in increased virion production. The authors speculate about the role of these proteins but do not provide any mechanistic detail on how they might be playing a role. It is unclear that the putative antiviral role of these proteins involves steps downstream of virus replication, especially given that the authors speculate translation might be affected by eIF3k which, if the case, RNA synthesis should also be expected to be affected.

      Overall, while the initial chimeric virus and domain swap approach is strong, the manuscript would benefit with a more thorough examination of virion assembly steps and a mechanistic link to virion production. Otherwise, the authors should revise the structure of their manuscript by de-emphasizing points about virion assembly and leave room for other mechanistic explanations of their chimeric data that more clearly link the host antiviral factor/E1 binding studies.

      Minor comments: In Figure 3e, the line under "with CHIKV E1" should be moved over to include the E2-II+E1 virus.

      Figure 5a, 5b, and 6a should be replaced with higher resolution images.

      Significance

      Strengths of the study include the initial chimeric virus and domain swap approach to determine factors that allow for the productive replication of chikungunya virus in macrophages compared to other arthritogenic alphaviruses. This approach yielded useful insights and could be adapted to other viruses. The study is limited, however, by the lack of mechanistic detail linking the antiviral host factors identified which bind to the E1 structural protein, and the advantage conferred by CHIKV structural proteins in the ONNV backbone. The study would be greatly improved by structural studies of the chimeric viruses that directly demonstrate more efficient virion production and that knockdown of the identified factors specifically affects virion production. This point could be addressed either through additional experimentation or tempering of the authors' conclusions about the mechanism by which CHIKV structural proteins provide an advantage over those of ONNV.

      The study advances knowledge in the field on what might advantage different pathogenic alphaviruses and explain differences in disease pathology. Additionally, the authors devise a simple and clever strategy that could be applied across different alphaviruses and would be useful to test in vivo in future studies. This study would be useful to a virology-specific audience.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary:

      In this work Yao et al. show CHIK is able to infect macrophages in contrast to other arthritogenic alphaviruses RRV, ONNV, and SINV. They use a series to chimeric viruses made with ONNV, the closest species to CHIK, and determine the E2-E1 proteins are important viral determinants which allow CHIK to replicate in machophages compared to ONNV. By comparing 397 CHIK sequences from infected patients, they identified 14 residues under pervasive and positive selection. Of these, 3 residues in E2 and 3 residues in E1 (amino acids) were different between CHIK and ONNV suggesting these residues contributed to the difference in macrophage tropism of CHIK compared to ONNV. The authors go on to determine what host factors the CHIK E2 protein is interacting with to presumably connect the viral and host determinants for CHIK infection in macrophages.

      Major concerns:

      1. The authors show one configuration of the E1-E2 heterodimer in Figure 4d. As shown, the E1 protein is exterior to the E2 protein and would suggest E1 is on the surface on the spike complex and virus surface. However, another configuration of the glycoproteins has E2 on the exterior of E1 and also on the exterior of the virus. The latter conformation is what has been observed in cryoEM studies of alphaviruses. The first configuation represents the E1-E2 between the three heterodimers which are important for spike assembly. The reason the orientation of the E2-E1 dimer is important is the authors speculate on the importance of the 6 CHIK residues not found in ONNV based on the structure, but the structural interpretation is, in my opinion, not correct.
      2. Validation of E1 interaction with SPSC3 and eIF3k needs to be stronger. Some concerns/questions are listed below. A myc tag was inserted between E3 and E2. How efficeintly does furin cleave E3 from E2 in this virus and how are viral titers of the myc-tagged virus compared to the non-tagged virus? I ask because is the IP looking at what is being pulled down by E2 or E3-myc-E2 that could be part of the spike polyprotein? The authors found E2 interacts with E3, E1 and a list of other host proteins. These results suggest several interactions including E2-host factor, E2-E1, E2-E3, E2-E1-host factor, E2-E3-E1, E2-E3-host factor. In figure 6d, and the subsequent conclusions, the authors suggest E1 is interacting with the host facor and do not see E2 alone and very low amounts of E3-E2-6K-E1. based on how the IP was performed I am not sure how an interaction between E1 and SPCS3 alone, without E2, would be detected. I would also like to see a reciprocal pull down using E1 and also E2 to see if these host factors are pulled down.
      3. If CHIK E1 is interacting with the host factors and that is antagonizing the antiviral response of SPSC3 (as one example), then what do pull downs using ONNV structural proteins look like? One would expect reduced interactions because the different amino acid causes a different E2-E1 dimer or attenuates the E1-host factor binding site.
      4. E1 and E2 are thought to interact during polyprotein translation and the initial dimer forms in the ER. If E1 is interacting wht SPSC3 in the ER, is E2 also present? Or is a population of E1 not interacting with E2 in order to inhibit SPSC3? I would love a model of how the authors see all these factors coming together for this new role of E1.

      Minor concerns:

      1. In Figure 1c, (-) RNA is shown but in the rest of the figures (+) RNA is shown. Show both or select one. I do find it interesting the (-) RNA levels are similar over time, even at 4 hours post transfection (early time). Related to this, ONNV has higher levels of (-) RNA but what is known about structural protein levels in ONNV and CHIK in macrophages? Are there comparable levels of CP and GP being produced?
      2. Figure 2e and figure 3 have ONNV has the first bar followed by CHIK. In figure 1 and 2b, CHIK is first and then ONNV. helps the reader to have the controls in the same order.
      3. Line 143-145 the authors discuss that when ONNV is the backbone and CHIK proteins are inserted the infection is more attenuated because of the E2 and E1 are from CHIK and ONNV, not the same virus (could also be E2-CP interactions are disrupted). However the chimeras made witht he CHIK backbone (in Figure 2) have a mismatch between E2 and E1 as well.
      4. When discussing the residues that were found in the FEL and MEME analysis, the authors start the amino acid numbering from CP and continue along the polyprotein. Usually when discussing amino acids in the structural proteins, each protein starts at amino acid 1. So V460 would be E2-V135. It would also be useful to know what the residues in ONNV were at these positions to see if amino acids changed in charge, size, bond forming potential, etc. Showing these residues in the E2-E1 conformation found in the virion would also allow one to find adjeacent residues that could explain differences in spike assembly and potentially where/how E1 is binding to a host protein.
      5. How effective is a non-attenuated CHIK strain in infecting macrophages? Could you make a SINV-La Reunion chimeric virus (which is BSL2) to see if a higher percentage of macrophages are infected and is this potentially contributing to the increased pathogenesis of La Reunion? Also how different is 181/25 with a pathogenic strain in the E2 and E1 resdiues? and compared to ONNV?
      6. When describing the last results section, "CHIK E1 binding proteins exhibit potent anit-CHIV activities" the authors use macrophages. In the rest of the text they consistently use THP-1 macrophages or human primary monocyte derived macrophages. The details of the cell type are extremely useful to the reader and having those in the last results section would be great.
      7. The paper is well-written. There is a slight disconnect as the authors go from discussing results in Figure 4 to Figure 5.

      Referees cross-commenting

      I agree with R#2 that having some Particle:PFU data would add some data to determine why such differences in titers/infectivity.

      I also see how this m/s could be split into two different m/s. One that focuses on the chimeric viruses and another that identifies the host factors important and goes in more depth with mechanism

      Significance

      Strengths:

      The authors have tackeled an intriguing question: why do some alphaviruses infect macrophages and others do not. They have used a chimeric approached to very systematically identify the viral determinants E2 and E1 as being important in macrophage infection. Using AP-MS they identify host factors that interact with E2 (possibly E2 and E1, see comments above) but if their findings that E1 has a role in attenuating a host antiviral factor, this would be fantastic.

      More and more examples of viral proteins having multiple roles during infection are in the literature. The idea that structural proteins also attenutate host antivirals is a developing field and vastly understudied. By fleshing out the results some more the authors might be onto something ery important in alphavirus virology.

      Limitations:

      The study has it is presented is limited in the validation of host factors and their interacting partners. I have many questions about the methodology, validation, and model from this last section.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      We thank the reviewers for their careful reading of the document and feedback which will help us to improve our manuscript. We will go through their comments one by one.

      Reviewer #1 (Evidence, reproducibility and clarity (Required)):

      This study would be much convincing if additional line of eukaryotic cells can be used to demonstrate the GEF-GAP synergy tis important for cell physiology. In addition, it would be best to demonstrate the spatiotemporal interaction of GEF-GAP using high-resolution live cell imaging.

      Response from the authors:

      The reviewer requests additional in vivo data to support our in vitro findings:

      (1) The reviewer requests in vivo data showing that GEF-GAP synergy is important for cell physiology. We believe that in order to show GEF-GAP synergy in vivo, Cdc42 cycling rates would need to be measured in vivo. For that single-molecule resolution is required – to track a single Cdc42 molecule and measure its GTPase cycling. We agree that such data would indeed be interesting, but are unaware of established techniques that would facilitate measurements of Cdc42 cycling rates in vivo.

      (2) The reviewer requests in vivo data showing the spatiotemporal interaction of GEF-GAP. Cdc24 and Rga2 are shown to interact (direct or mediated by another protein) (McCusker et al. 2007, Breitkreutz et al. 2010, Chollet et al. 2020). Cdc24 and Rga2 share 11 binding partners (https://thebiogrid.org/31724/table/saccharomyces-cerevisiae-s288c/cdc24.html, https://thebiogrid.org/32438/table/saccharomyces-cerevisiae-s288c/rga2.html) and have been found at the polarity spot (Gao et al. 2011). Live cell imaging of fluorescently tagged Cdc24 and Rga2 will show that they exhibit some interaction, but not specify the role of the interaction nor if the interaction is direct or mediated by one of the shared binding partners. In order to show a direct interaction between Cdc24 and Rga2, one could consider (A) super-resolution imaging or (B) FRET experiments: For both fluorescently tagged Cdc24 and Rga2 cell lines would need to be constructed.

      (A) Super-resolution imaging could show direct interaction between Cdc24 and Rga2, but even with the techniques available this would be on the limit. Further, it is usually done in fixed cells, and not in live cells (as requested from the reviewer).

      (B) To show a direct interaction of Cdc24 and Rga2 using FRET, suitable protein constructs would need to be engineered. We believe that the main obstacle in showing direct binding of Cdc24 and Rga2 using FRET is to design the fluorophore linker. The linker would need to be designed in such a way that it is flexible enough to give a FRET signal even if the two large proteins bind on the opposite sites of the fluorophore, but also is stiff/short enough to not show binding if both proteins are in close proximity through binding to a common binding partner.

      __We believe that an investigation of GEF GAP binding in vivo is beyond the scope of this study. Instead, we will further explore one possible mechanism underlying GEF GAP synergy - Cdc24 Rga2 binding - through conducting Size-Exclusion Chromatography Multi-Angle Light Scattering experiments with purified Cdc24 and Rga2 (alone and in combination). __

      Reviewer #1 (Significance (Required)):

      The revised study would provide first line evidence that GEF-GAP synergy to be general regulatory property in eukaryotic kingdom.

      Reviewer #2 (Evidence, reproducibility and clarity (Required)):

      The study entitled, "The GEF Cdc24 and GAP Rga2 synergistically regulate Cdc42 GTPase cycling" by Tschirpke et al., uses an in vitro GTPase assay to examine the GTPase cycle of Cdc42 in combination with its GEF and GAP effectors. The authors find that the Cdc24 GEF activity scales non-linearly with its concentration and the GAP Rga2 has substantially weaker effect on stimulating Cdc42 GTPase activity. Not surprisingly, the combined addition of Cdc24 and Rga2 lead to a substantial increase in Cdc42 GTPase activity.

      **Referees cross-commenting**

      In Zheng, Y., Cerione, R., and Bender, A. (1994) J. Biol. Chem. 269: 2369-2372 (Fig. 3C), the authors show that Cdc24 combined with the GAP Bem3 lead to a large synergy in boosting Cdc42 GTPase activity.

      Reviewer #2 (Significance (Required)):

      There is very little new information in this manuscript. Previous studies (Rapali et al. 2017) have shown that the scaffold protein Bem1 enhances the GEF activity of Cdc24. It is expected that the reconstitution of a GEF and GAP protein promote the GTPase cycle and indeed Zheng et al. (1994) showed that that Cdc24 combined with the GAP Bem3 lead to a large synergy in boosting Cdc42 GTPase activity. Hence the only potentially interesting finding in this work is that, in solution Cdc24 activity scales non-linearly with its concentration. However as this GEF and Cdc42 are associated with the membrane, the relevance of solution studies are less clear and furthermore the mechanistic basis for the non-linearity is not explored in detail. Given the limited new information from this work, the findings are, in their current form, too preliminary.

      Response from the authors:

      __We appreciate the reviewer recognizing our work on the non-linear concentration-dependence of Cdc24’s activity. We disagree that this is the only new finding in our study: __

      We explore the effect of Cdc24 and Rga2 on Cdc42’s entire GTPase cycle and show that Cdc24 and Rga2 synergistically upregulate Cdc42 cycling. So-far Cdc42 effectors were only characterized in isolation (with the exception of Cdc24-Bem1 (Rapali et al. 2017)) and through how they affect a specific GTPase cycle step. The regulation of single GTPase cycle steps through an effector yields mechanistic insight into this specific GTPase cycle step. However, it does not show how the effector affects overall GTPase cycling of Cdc42 – a process Cdc42 constantly undergoes in vivo. Our approach allows us to study synergistic effects between proteins affecting different GTPase cycle steps. Synergies are another regulatory layer of the polarity system, adding further complexity: Which polarity proteins exhibit synergy, to which extend? The assay employed here, which studies the entire GTPase cycle, enables studying the effect of any GTPase cycle regulator, alone and in combination with another regulator.

      The reviewer states that the GEF GAP synergy is to be expected, as it was already shown in Zheng et al. 1994. In Fig. 3C Zheng et al. shows the time course of the GTPase activity of Cdc42 in presence of Cdc24, Bem3, and Cdc24 plus Bem3. Fig. 3C is the only data in which the combined effect of a GEF (Cdc24) and a GAP (Bem3) is investigated. The data indicates synergy, but is neither discussed as such in the text of the publication, nor analyzed quantitatively. Further, only one concentration of each effector (GEF/GAP) is used and the study uses a Bem3 peptide containing codons 751-1128 (30%) of the full-length BEM3 gene. Zheng et al. 1994 gives an early indication of GEF GAP synergy, but does not claim, discuss, or further investigate the synergy as such. In contrast, we use full-length Rga2 (not Bem3) as GAP, conduct several concentration-dependent assays, and analyze them quantitatively. We thank the reviewer for pointing out the pioneering character of Zheng et al.‘s study and will mention it more prominently in our report. However, we disagree that Zheng et al. sufficiently studied the GEF GAP interaction. To our awareness no theoretical studies include a GEF GAP synergy term, which we would expect if GEF GAP synergy is well-established in the field.

      The reviewer criticizes the relevance of bulk in vitro studies (lacking membranes) of proteins that bind to membranes in vivo. We agree that the presence of a membrane can affect the protein’s property, and we can not exclude that membrane-binding could alter the magnitude of a GEF GAP synergy. However, we believe that membrane-binding does not impede the GEF GAP synergy altogether. If membrane binding would influence GTPase properties that strongly, other studies on Cdc42’s GTPase activity and GEF and GAP activity, that do not include a membrane, would be inconclusive as well (e.g. Zheng et al. 1993, Zheng et al. 1994, Zheng et al. 1995, Zhang et al. 1997, Zhang et al. 1998, Zhang et al. 1999, Zhang et al. 2000, Zhang et al. 2001, Smith et al. 2002, Rapali et al. 2017). Both studies mentioned by the reviewer (Zheng et al. 1994, Rapali et al. 2017) were also conducted without membranes present.

      We believe that an inclusion of membrane-binding into reconstituted Cdc42 systems will enhance our understanding of Cdc42 and recognize it as a next step, which could be enabled by the assay used in our study.

      Reviewer #3 (Evidence, reproducibility and clarity (Required)):

      This work reports a biochemical analysis of the effects of a recombinant yeast GEF (Cdc24) and GAP (Rga2) on Cdc42 GTPase cycling in vitro. The central conclusion is that the GEF and GAP act "synergistically", which occurs "due to proteins enhancing each other's effects". By this they appear to mean that the GEF enhances the GAP's activity and vice versa. I was not persuaded that this is correct, and was confused by many aspects of the approach and interpretation, as outlined below.

      1. GEF and GAP are expected to accelerate GTPase cycle synergistically even with no effect on each other's activity:

      The Cdc42 GTPase cycle is understood to occur via distinct steps (GDP release, GTP binding, and GTP hydrolysis): GDP release and GTP hydrolysis are intrinsically slow steps that are accelerated by GEFs (GDP release) and GAPs (GTP hydrolysis). This fundamental biochemistry was established in the 1990s using biochemical assays that measure each step independently. Here instead the authors use an assay that measures [GTP] decline in a mix with 5 uM starting GTP, 1 uM Cdc42, plus or minus some amount of GEF or GAP. They assume exponential decline of [GTP] with time, yielding a cycling "rate". If that is so, then one would expect that added GEF would accelerate only the first step, leaving a slow GTP hydrolysis step that limits the overall cycling rate, while added GAP would accelerate only the last step, leaving a slow GDP release step that limits the overall cycling rate. Adding both together would speed up both steps, and should therefore "synergistically" accelerate cycling. This would be expected based on previous work and does not imply that GEF or GAP are affecting each other's action (except trivially by providing substrate for the next reaction). If the authors wish to demonstrate that something more complex is indeed happening, they need to use assays that directly measure the sub-reaction of interest, as done by prior investigators.

      Response from the authors:

      The reviewer raises the point that we do not consider a simpler, rate-limiting model and that this rate-limiting model could explain our synergy between GAP and GEF in accelerating the GTPase cycle.

      We very much welcome this consideration of the reviewer! We will add a clarification to our manuscript to explain why a rate-limiting model/interpretation does not match our data.

      Intuitively, the rate-limiting model is appealing, as it permits interpretation of cycle rate increases in terms of individual biochemical steps. So, a consideration of this model is indeed relevant. However, as also noted by the reviewer in the next points, data from e.g., figure 3e are not compatible with a simple rate-limiting model with two steps (hydrolysis and nucleotide exchange). We will explain how the acceleration of the total rate by both GAP and GEF individually does not match the rate-limiting model, even if we assume maximal effects of adding GAPs and GEF to the cycle. For this purpose, we consider the rate-limiting model scenario where the sensitivity of the GTPase cycle to adding GAP/GEF is maximized, so the best case-scenario for the rate limiting step-model.

      In the rate-limiting step model, we assume that we have a GTPase cycle in which at least one of the three GTPase cycle steps is rate-limiting: (A) GTP binding, (B) GTP hydrolysis, and (C) GDP release.

      We assume that the addition of a GEF and GAP only accelerates GDP release and GTP hydrolysis respectively. Biochemically, all three steps in the GTPase cycle are expected to be relevant. However, here we will consider only the final two steps, as sensitivity to rate limitation by GAP/GEF is maximized when time spent in the GAP/GEF-independent step in the cycle (step A: GTP) is negligible (i.e. never rate-limiting). The two-step model thus consists of (1) a nucleotide exchange step (step C+A) which is dominated by GDP release (step C) and assumed to be accelerated exclusively by the GEF, and (2) a GTP hydrolysis step (step B) exclusively enhanced by the GAP.

      In the rate limiting step model GEF-GAP synergy can appear if one of the conditions applies:

      1. the addition of a GAP speeds up the GTP hydrolysis step so much that the hydrolysis step stops (or almost stops) being the rate-limiting step, or
      2. the addition of a GEF speeds up the GDP release step so much that the release step stops (or almost stops) being the rate-limiting step. In these conditions, the acceleration of the GTPase cycle, accomplished by adding only a GAP or adding only a GEF, is interdependent. Therefore, we consider the possible acceleration of the GTPase cycle by GAP and GEF individually, and compare these to our observations to determine whether the rate-limiting step model can explain our data.

      The GTPase cycle time Tc is thus composed of hydrolysis Th and nucleotide exchange time Te, and the rates r are connected through:

      1/rc=1/rh + 1/re

      If we compare the ratio of the rates with protein (GAP/GEF) added in the assay (index 1) with the basal rate without protein added (index 0), we obtain the cycle acceleration factor alpha:

      alpha=rc1/rc0=(1/rh0 + 1/re0)/(1/rh1 + 1/re1)=(re0 + rh0)/(re0*rh0/rh1 + rh0*re0/re1)

      Here, rc1 and rc0 are the total GTPase cycle rate with and without effector respectively, rh1 and rh0 are the GTP hydrolysis rate with and without effector respectively, and re1 and re0 are the nucleotide exchange rate with and without effectors respectively.

      There is indeed an interdependence created between how much the GAP and GEF can both accelerate the total cycle, if the GAP and GEF are assumed to only accelerate GTP hydrolysis and nucleotide exchange respectively. E.g., how much the total GTPase cycle rate rc is accelerated by an increase in GTP hydrolysis rate rh depends on and can be limited by the current nucleotide exchange rate re. However, this interdependence is too strict to match the data in Figure 3e, as we will explain in the next paragraphs:

      When we only add a GAP and the GAP accelerates only the GTP hydrolysis rate (re1=re0), then the maximal total GTPase cycle rate acceleration alphaGAP that the GAP can accomplish is when rh1>>rh0,re0:

      alphaGAP=rc1/rc0=(1/rh0 +1/re0)/(1/rh1+1/re0)=(re0+rh0)/(re0*rh0/rh1+rh0)

      ~(re0+rh0)/rh0=1+ re0/rh0

      We thus assume the GAP accelerates the cycle so much that the hydrolysis step is much faster than the exchange step, at which point the effect of adding more GAP would saturate. We note that we do not consider the GAP concentration regime where we see saturation, thus in reality the acceleration by the GAP is more restricted than predicted here.

      Analogously, if the GEF accelerates only the nucleotide exchange rate (rh1=rh0), then the maximum GTPase cycle rate ratio will be when re1>>re0,rh0 , yielding acceleration factor alphaGEF :

      alphaGEF= rc1/rc0=1+ rh0/re0

      Again, note we assume the GEF accelerates the cycle so much that the exchange step is much faster than the hydrolysis step, at which point the effect of adding more GEF would saturate. We note that we do not observe the GEF concentration regime where we see saturation, thus in reality the acceleration by the GEF is more restricted than predicted here.

      We see that the maximum gain in rates for GAP-only and GEF-only assays is limited by the same basal GTP hydrolysis and nucleotide exchange rates (rh0 and re0), leading to the following interdependence:

      alphaGAP=1+ 1/(alphaGEF -1)=alphaGEF/(AlphaGEF -1)

      In our GAP-only and GEF-only assays (Fig. 3e, Tab. 2), we see both a 2-fold and 100-fold increase in the total rate respectively. A 100-fold acceleration factor of the GEF would maximize the GAP acceleration factor to 1.01 (or alternatively, the 2-fold GAP acceleration would maximize the GEF acceleration to 2), which are both significantly lower than what we observe. So even though we made favorable assumptions for the rate-limiting model to maximize rate sensitivity to GAP/GEF, namely neglecting nucleotide binding and assuming GAP/GEF concentrations that saturate in their effects, we still cannot reproduce the acceleration factors in our GAP-only and GEF-only assays.

      Moreover, a rate-limiting step model would also imply saturation effects as stated in the next point of the reviewer. While we observe saturation in total rate acceleration for certain GAP concentrations, we use GEF and GAP concentrations in the combined protein assays for which no saturation effects were observed. Absence of saturation in both cycle steps simultaneously is also not reconcilable with the rate-limiting step model, as will be further discussed in the next point of the reviewer.

      In summary, this means that the rate-limiting model is not sufficient to explain our results: the GAP/GEF synergy we observe is not simply resulting from GEF and GAP independently lifting two different rate-limiting steps.

      Model-based interpretation of the GTPase assay is poorly supported:

      The assay employed measures overall GTP concentration with time. It is assumed (but not well documented-see below) that [GTP] declines exponentially, and that the rate constant for a particular condition can be fit by the sum of a series of terms that are linear or quadratic in the concentrations of Cdc42, GEF, and GAP. There is no theoretical derivation of this model from the elementary reactions, and the assumptions involved are not well articulated.

      As discussed in point 1 above, one would expect that a GEF or GAP alone could only accelerate the cycle to a certain point, where the other (slow) reaction becomes rate limiting. But that does not appear to be true for their phenomenological model, where slow steps (small terms in the sum) will always be overwhelmed by fast steps. This is not the traditional understanding of how GTPases operate.

      Response from the authors:

      The reviewer expresses the concern that because we do not derive our coarse-grained model from elementary reactions, we miss important effects that can occur when adding GAP and GEFs, particularly saturation.

      We understand the concern of the reviewer that if a rate-limiting step model is considered, saturation effects of GAP/GEF will limit the amount with which these effectors can speed up the total cycle. Our coarse-grained model indeed does not account for this saturation. However, as discussed in the previous point of the reviewer, we do not opt for the rate-limiting model interpretation, as the GAP and GEF effects are not compatible with the rate-limiting step model.

      Secondly, we agree that for high enough concentrations of GEF and GAPs, we would experience a saturation in the effect of adding the effectors. We are aware of this possibility, and we verify that we are not in saturation regimes with our added proteins by checking the plots of the individual protein titrations (see Figure 3a-d). If we enter the saturation regime, we expect a negative second derivative in the rate as function of protein concentration (the curve shallows off). We do not see this for any protein except for Rga2 at some point, as discussed in our main text of the manuscript. However, for this protein we only use the data in the linear regime for further analysis. In short, we understand the concern of the author but we empirically check that we are not in the saturation regime.

      Data that do not conform to expectation are not explained: Strangely, the data (as interpreted by the model assumptions) also appear inconsistent with the expectation of rate-limiting steps. GEF addition (alone) is said to accelerate cycling 100-fold, while GAP addition (alone) accelerates it 2-fold. But that would seem to imply that GDP release takes up >99% of the basal cycle (so accelerating that step alone reduces cycling time 100-fold), while GTP hydrolysis takes up >50% of the basal cycle (so accelerating that step alone reduces cycling time 2-fold). In the conventional understanding of GTPase cycles, these cannot both be be true (as the steps would then add to >100% of the basal cycle). There is no attempt to reconcile these findings with previous work.

      Response from the authors:

      The reviewer raises the point that our findings do not match the expectations of the rate-limiting model perspective.

      We fully agree with the reviewer that our data is not compatible with the rate-limiting step model. The 100-fold and 2-fold gain of the total cycle rates for GEF-only and GAP-only assays are one of our arguments against the rate-limiting model view, as described in the first point of the reviewer. Also, our lack of saturation as described in the previous point of the reviewer provides another argument against using expectations based on rate-limiting steps to interpret our findings.

      Lack of detailed timecourse data:

      The decline in [GTP] with time is stated to be exponential, allowing extraction of an overall cycling "rate". But this claim is supported only weakly (S3 Fig. 1 uses only 3 timepoints, is not plotted on semi-log axis, and does not report fit to exponential vs other models) and only for the Cdc42-alone scenario: no data at all are presented to support exponential decline in reactions with GEF or GAP. Most assays seem to measure only a single timepoint, so extraction of a "rate" is very heavily influenced by the unsupported assumption of exponential decline. And if the decline is not exponential, it becomes extremely difficult to interpret what a single timepoint means.

      Response from the authors:

      The reviewer requests additional timeseries data with GEF and GAP to support the assumption of an exponential decline of GTP in the assay and requests to plot it on a semi-log axis.

      We will add data for Cdc42 + Cdc24 and for Cdc42 + Rga2 with two to three time points, and plot it as requested on a semi-log axis.

      Other issues with interpretation of the data:

      (i) It is unclear why the authors chose to employ an assay that is much harder to interpret than the biochemical assays used by others. In biochemical studies, assays that report an output of multiple reactions are always harder to interpret than assays targeting a single reaction. As well-established assays are available for each individual step in GTPase cycles, any conclusions must be supported using such assays.

      Response from the authors:

      The reviewer wonders why an assay that investigates several GTPase steps at once was chosen over assays that investigate sub-steps of the GTPase cycle, given that these give more mechanistic insights.

      We agree that assays investigating GTPase cycle substeps can give more mechanistic insights into these specific steps. However, they do not allow to study how proteins affecting different steps act together. We were interested in investigating the overall GTPase cycle of Cdc42 and a possible interplay of GEFs and GAPs. Cdc42 GTPase cycling was found to be a requirement for polarity establishment (Wedlich-Soldner et al. 2004) and Cdc42 GTPase cycling is physiologically relevant. Ultimately, we hope that in vitro results provide stepping stones towards understanding the complex and less controlled in vivo environment. The in vivo environment often entails the output of many reactions combined, so there is every incentive to study aggregated effects of a full cycle which are not necessarily the sum of individual outputs.

      __We believe that both assay types – assays that investigate sub-steps and yield mechanistic details, and assays that investigate the entire cycle – are important and disagree that one assay type is superior to the other. Instead, we believe they complement each other. __

      (ii) The reported basal (and GEF/GAP-accelerated) rates are very slow, perhaps due to poor folding of recombinant proteins. This raises the possibility that much of the Cdc42 is inactive. If so, then accelerated GTP hydrolysis could come from increasing the active fraction of Cdc42, rather than catalyzing a specific step.

      Response from the authors:

      The reviewer wonders whether the reported rates are slow due to poor folding of recombinant Cdc42. We used S. cerevisae Cdc42, for which it has been shown that it has a significantly lower basal GTPase activity than Cdc42 of other organisms (see Zhang et al. 1999). Many other studies on Cdc42 were conducted with human Cdc42, which has a significantly higher basal GTPase activity (Zhang et al. 1999). We assessed the activity of several recombinantly expressed Cdc42 constructs previously (Tschirpke et al. 2023). We there observed that most constructs had a similar GTPase activity, only some purification batches and constructs had a significantly reduced GTPase activity (which might be linked to poor folding). The Cdc42 construct used here shows a similar activity as the active Cdc42 constructs in Tschirpke et al. 2023, and we therefore believe that it exhibits proper folding. If recombinant Cdc42 folds poorly, we would expect greater variations between Cdc42 constructs and purification batches (caused by different levels of folding/ a different fraction of active Cdc42) than what we observed previously (see Tschirpke et al. 2023).

      Tschirpke et al. 2023:

      Tschirpke et al. A guide to the in vitro reconstitution of Cdc42 activity and its regulation (2023) BioRxiv. (https://doi.org/10.1101/2023.04.24.538075) (in submission at Current Protocols)

      (iii) The GEF and GAP preparations include multiple partial degradation products and it is unclear whether the measured activities come from full-length proteins or more active fragments.

      Response from the authors:

      We agree with the reviewer that the Cdc24 and Rga2 preparations contain degradation products.

      It would be more ideal if the protein purifications were entirely pure, but this is experimentally very difficult to achieve for the used proteins (which are large and partially unstructured, making them prone to partial degradation). Further, it is not uncommon to use protein preparations where some degradation products were present (e.g. Zheng et al. 1993, Zheng et al. 1994). Other studies did not show their purified preparations.

      The vast majority of the Cdc24 preparation is the full-length protein. We therefore expect that the degradation fragments only contribute in a small extend to the overall protein behavior.

      The Rga2 preparation contains a higher amount of degradation product, but only larger size protein fragments (> 60kDa), suggesting that the fragments contain at least and more than 1/3 of the full-length protein (the protein fragments are thus the size or larger than of the GAP peptides used previously). The fragments could in principle have a higher or lower activity. We account for fragments of no/lower activity by comparing our cycling rates to those of BSA/Casein, which has no specific effect on Cdc42. The cycling rate Rga2 is almost an order of magnitude greater than that of BSA/Casein, suggesting that the effect of the full-length protein dominates. We could only imagine that a Rga2 fragment has a higher GAP activity if the fragment consists mainly of the GAP domain and if in Rga2 the activity of the GAP domain is downregulated. Nevertheless, we will do an additional experiment using a purified GAP domain peptide to assess that if a GAP domain by itself has a higher GAP activity than our Rga2 preparation. Using that data, we will discuss possible implication of the GAP fragments in our manuscript.

      (iv) Cdc42 cycling is also accelerated by BSA and casein, suggesting that there are poorly understood aspects of the assay and that GEF and GAP actions may (like BSA and casein) involve non-canonical effects on Cdc42. As GEF and GAP are expected to interact better with Cdc42 than BSA or casein, these effects could dominate the observed changes in GTP levels.

      Response from the authors:

      The reviewer raises the concern that the effects of the added effector proteins on the rates could be caused by non-canonical effects. We do not believe non-canonical effects play a relevant role in our assays. While BSA and casein accelerate the GTPase cycle in our assays, the GAP effect and GEF effect are orders of magnitude stronger.

      (v) Cdc42-alone cycling assays are said to be reproducible. However, assays with added GEF/GAP/BSA/Casein yield rates that vary almost an order of magnitude between replicates. This poor reproducibility further reduces confidence in the findings.

      Response from the authors:

      The reviewer is concerned about the variations in Cdc42 effector rates.

      __We disagree that the variations are concerning and believe to have accounted for them in our analysis: __The Cdc42 (Cdc42 alone) data is very reproducible (see Tschirpke et al. 2023). The GTPase assay is generally sensitive to small concentration changes and errors introduced through pipetting small volumes (as required for the assay). We believe that the small variation observed for Cdc42 alone is because Cdc42 has such a low basal rate and therefore the small concentration changes due to pipetting have a smaller effect. Once other effectors are added, especially highly GTPase stimulating ones as Cdc24, small concentration changes due to pipetting can lead to larger variations between assays (small variations in Cdc24 concentration lead to larger changes in remaining GTP due to Cdc24’s strong and non-linear effect on Cdc42). We conduct the assays multiple times to account for these variations. In our analysis we do not compare single rate numbers but the orders of magnitude of the rate, and report the variations present. Even given the present variations, the differences in effect sizes are still significant. We map and discuss assay variation in (Tschirpke et al. 2023), to which we refer to several times throughout the manuscript.

      Tschirpke et al. 2023:

      Tschirpke et al. A guide to the in vitro reconstitution of Cdc42 activity and its regulation (2023) BioRxiv. (https://doi.org/10.1101/2023.04.24.538075) (in submission at Current Protocols)

      (vi) It is unclear what timepoint was used for the different assays. 1.5 h at 30 degrees seems to be the standard here for the Cdc42-alone assays, but I assume that cannot be what was measured to assess GTP decline for GEF-containing assays as there would be very little GTP left at 1.5 h.

      Response from the authors:

      We used 60-100 min as incubation times for all assays. The assay data will be published on a data server, where all these numbers can be checked. We further added a clarification to the materials and methods section. In order to still have remaining GTP for the Cdc42 GEF mixtures after 60-100 min, we lowered the used protein concentrations.

      (vii) The graph reporting GEF activity is plotted only for [GEF]Response from the authors:

      The graphs show the full range of protein concentrations used.

      In order to calculate K1, K2, K3,Cdc24, K3,Rga2, K3,Cdc24,Rga2 from k1, k2, k3,Cdc24, k3,Rga2, k3,Cdc24,Rga2, …, a protein concentration has to be included in the term (as K1 = k1 [Cdc42], ….). In order to make K comparable, we chose to use 1uM for all protein concentrations. This was done to compare the cycling rate values of different proteins. 1uM was a choice, in the same fashion 0.2uM could have been chosen.

      __We will further discuss in the manuscript how the choices in protein concentration affect the effector strength on Cdc42. __

      (viii) S8 Data with casein seems very noisy and it is no longer at all clear that the quadratic fit for [Cdc24] is justified. Also, the symbol colors are very similar so it is hard to tell what data corresponds to what condition. The synergy between Cdc24 and Rga2 is also very noisy and the fits seem arbitrary.

      Response from the authors:

      The reviewer is concerned with (1) the noise in the S8 data, and (2) the Cdc42-Cdc24-Rga2 fits.

      (1) We acknowledge in the manuscript that the S8 data is noisy and should be viewed with caution. We do not put much emphasis on these data sets and their interpretation and show them only in the supplement.

      (2) We disagree that the Cdc42-Cdc24-Rga2 fits are arbitrary. The fits contain several data points per protein, and reproduce the rate values from Cdc42-Cdc24 and Cdc42-Rga2 assays well.

      The reviewer is concerned with the color scheme choice in the fits.

      __We will adapt the color scheme of the fits to make the colors more distinguishable. __

      (ix) It is disturbing that different Cdc42 constructs behave quite differently (S4). This suggests that protein behavior is influenced by the various added epitope tags and protease cleavage sites (they also leave the C-terminal CAAX box rather than removing the AAX as would happen in vivo). These features raise the concern that these findings may not be directly relevant to the situation with endogenous yeast Cdc42. Of course, it is also the case that relevant Cdc42 biochemistry occurs with prenylated Cdc42 on membranes.

      Response from the authors:

      The reviewer is concerned that the behavior of the Cdc42 constructs is influenced by their tags. In a previous manuscript (Tschirpke et al. 2023) we explored the effect of various N- and C-terminal tags on Cdc42, by comparing it to Cdc42 that is not tagged in that position. We found that most tags, including the tags present in the Cdc42 construct used here, do not affect Cdc42’s properties.

      Instead, we found a general, tag independent, heterogeneity in Cdc42 behavior (which can occur between purification batches and between constructs (but not between different assays)): in some batches GTPase activity depended quadratically on its concentration, others showed a linear relationship. Most batches exhibited a mixed behavior. The differences between the batches are generally small, and only visible in the activity to concentration plots and because of the assay’s high accuracy. We use a two-parameter fit (k1 [Cdc42] + k2 [Cdc42]2) to phenomenologically account for this heterogeneity, and to estimate the basal Cdc42 GTPase activity. We do not interpret this heterogeneity, as more research is needed. We believe that Cdc42 still has unexplored properties, of which this heterogeneous behavior can be one. We speculate in Tschirpke et al. 2023 that it is linked to Cdc42 dimerization mediated by its polybasic region, a relationship that is far from being fully understood yet. __We believe that it is of scientific interest to point out heterogeneous behaviors to encourage more research. __

      Tschirpke et al. 2023:

      Tschirpke et al. A guide to the in vitro reconstitution of Cdc42 activity and its regulation (2023) BioRxiv. (https://doi.org/10.1101/2023.04.24.538075) (in submission at Current Protocols)

      The reviewer is concerned that our findings are biologically not relevant, as our experiments (1) included Cdc42 that was not prenylated and (2) did not include membranes.

      (1) We here used recombinantly purified proteins, which do not contain posttranslational modifications, such as prenylations. So-far Cdc42’s prenyl group, which is responsible for binding it to membranes, has not been linked to its GTPase properties. We therefore believe that unprenylated Cdc42 is an equal choice to prenylated Cdc42 when studying Cdc42’s GTPase cycle. Further, the use of recombinantly purified proteins can be of advantage: when proteins are purified from their native host, the post-translationally modified protein is purified. However, many proteins contain a multitude of post-translational modifications (PTMs). Thus, the purified protein is a mixture of protein with different PTMs. For example, S. cerevisae Cdc42 undergoes ubiquitinylation (Swaney et al. 2013, Back, Gorman, Vogel, & Silva 2019), phosphorylation (Lanz et al. 2021), farnesylation and geranyl-geranylation (Caplin, Hettich, & Marshall 1994). We here used protein preparations that do not contain PTMs, and show how they behave. Natively purified proteins would be mixtures of various PTMs, and the observed protein behavior would be that of the mixture. If Cdc42’s PTMs affect it’s GTPase behavior, the observed behavior of natively purified Cdc42 would represent the average behavior of the mixture. It then would require additional work to disentangle which PTMs affect the GTPase cycling in which way. The use of recombinantly expressed Cdc42 does not require this work, and can set the baseline for how Cdc42 without PTMs behaves. If in the future a link between Cdc42’s GTPase behavior and PTMs are found, the work here could be used as a baseline for Cdc42’s behavior when it is without PTMs.

      (2) The concern about missing membranes was also raised by reviewer 2 (significance), and we like to refer to our response there.

      Reviewer #3 (Significance (Required)):

      The basic biochemistry of Cdc42 cycles was figured out about 30 years ago. However, those studies did not examine how combinations of Cdc42 regulators (as opposed to individual regulators) might interact to produce effects not expected from combining their individual actions. Recently, this combination approach did lead to interesting findings by Rapali et al. This approach is worthwhile and addresses a major question of interest to the broader field of GTPase biochemistry.

      One main limitation of this study is technical: the main assay is less informative (though perhaps easier) than traditional assays, and it is unclear whether the recombinant proteins employed retain their normal activities. Another limitation is the model-based interpretation of the assay that does not include the potential for rate-limiting steps.

      Response from the authors:

      We thank the reviewer for the detailed comments.

      One important point of confusion originated from our lack of discussion concerning a rate-limiting step model, which is an obvious starting point for modelling the GTPase cycle. We thank the reviewer for pointing this out, and we will include an explanation in our manuscript why we reject this model and instead opt for a coarse-grained model.

      Firstly, a rate-limiting model would generate saturation effects that we would observe when adding GEF and/or GAPs. In assays exploring GEF GAP synergy we use GEF and GAP concentrations for which no saturation effects were observed.

      Secondly, in our data we observed a two-fold increase of the total GTPase cycling rate when adding a GAP and a 100-fold rate increase when a GEF is added. These increases are not compatible with a model where either hydrolysis or nucleotide exchange limits the GTPase cycle. While a synergy could arise from the rate-limiting model perspective, the incompatibility of the rate-limiting model with the GAP-only and GEF-only assay data excludes this synergy explanation. Finally, through coarse-graining our model we avoid using single step parameters from literature which are incompatible in terms of proteins/buffers used. (For example; the mayor studies that kinetically characterized the individual GTPase steps of Cdc42 used human Cdc42 (Zhang et al. 1997, Zhang et al. 2000). Because human Cdc42 exhibits a higher basal GTPase activity (Zhang et al. 1999) we are skeptical how useful it is to transfer these parameters to S. cerevisae Cdc42.)

      At the same time, coarse-graining our model permits absorbing unidentified molecular details which is essential when we wish to incorporate BSA and casein rate contributions.

      The reviewer finds our assay, which investigates the GTPase cycle as a whole, less informative. Assays investigating single GTPase cycle sub-steps give more mechanistic insights into these steps. We opted for an assay that studies GTPase cycling as a whole instead, as we were interested in studying how proteins effecting different steps act together. We believe that both assay types are important as they complement each other.

      The reviewer is concerned about our use of recombinant proteins, and whether they retain their normal activities. We assessed Cdc42’s GTPase activity and the influence of added purification tags extensively (Tschirpke et al. 2023), and found that added tags do not affect Cdc42’s GTPase properties. We checked Cdc24’s GEF activity using the GTPase assay and found that it bound strongly to Bem1, as expected (Tschirpke et al. 2023). The Cdc24 concentrations needed to affect Cdc42’s GTPase activity were similar to those used previously (Rapali et al. 2017), suggesting that it is fully active. A similar comparison for Rga2 was not possible, as so-far only domains of Rga2 were used (Smith et al. 2002). We here used recombinantly purified proteins, which do not contain posttranslational modifications (PTMs). To our knowledge the PTMs of the herein used proteins are not linked to their GTPase/GEF/GAP properties. Thus, a lack of PTMs does not diminish our findings. Further, when proteins are purified from their native host, the post-translationally modified protein is purified. However, many proteins contain a multitude of post-translational modifications in vivo. Natively purified proteins would be mixtures of various PTMs, and the observed protein behavior would be that of the mixture. We here used protein preparations that do not contain PTMs, and show how they behave, setting the baseline for proteins without PTMs behaves. If in the future a link between GTPase behavior and PTMs are found, the work here could be used as a baseline for the proteins behavior when it is without PTMs.

      Reviewer #4 (Evidence, reproducibility and clarity (Required)):

      Summary

      The GTPase cdc42 is a key determinant of yeast polarization. Its activity is amplified at the site of polarization through a poorly defined positive feedback mechanism, and depends on numerous GAPs regulating GTP hydrolysis and the GEF cdc24 that regulates GDP release. These components have previously been evaluated for their quantitative effects on the individual steps in the GTPase cycle that they modulate, but potential interactions between the cdc24 GEF and any GAP could not be examined based on these assays. The authors validate and employ a bulk assay of the total GTPase cycle based on GTP consumption to study the activities of and potential interactions between cdc24 and the GAP Rga2. Fitting their data to a mathematical model, they come to three central conclusions: (1) the activating activity of cdc24 to activate cdc42 GTPase activity is nonlinear, showing a quadratic relationship, (2) Rga2 shows a much lower activating activity that is linear at low levels before saturating, and (3) there is a strongly synergistic interaction between the activating activities of cdc24 and Rga2. Some hypotheses for the mechanistic bases of these findings are hypothesized, but not further investigated. Their conclusions are well supported by the data which appears to be of sufficient rigor.

      Major comments

      The three main conclusions of the manuscript are well supported by the data and associated modeling.

      One unresolved issue is the discrepancy between the authors' conclusion that the non-linear activation by cdc24 is likely a result of oligomerization, whereas Mionnet et al 2008 reach the opposite conclusion. It seems that the authors wish to discount the Mionnet results because they used truncated constructs to test deficient oligomerization and an engineered construct to test induced oligomerization. If the authors are correct, then a relatively easy test would be to introduce the oligomerization deficient mutants defined by Mionnet into their fuill length construct and compare to wild type protein. While the authors' measured results don't depend on the offered mechanism and this experiment is therefore optional, their explanation is quite unsatisfying, especially since an experiment to resolve the difference is entirely feasible and not very strenuous.

      Response from the authors:

      __The reviewer suggests to conduct experiments with oligomerization deficient Cdc24 mutants to test our hypothesis that the non-linear concentration dependence of Cdc24’s activity is due to Cdc24 oligomerization. __

      We agree that this is an insightful experiment, and will conduct it. In order to observe the effect in our GTPase assays, we require a mutant that is oligomerizes substantially less than wild-type protein. Mionnet et al. constructed several Cdc24 mutants, but none were entirely oligomerization deficient. However, the DH5 (L339A/E340A) mutant showed a 10-fold reduction in oligomerization and the DH3 (F322A) mutant exhibited 2.5-fold reduction in oligomerization. We will therefore use the DH5 and DH3 mutant for two additional experiments.

      Minor comments

      The results in Fig S4 serve as assay validation, and this should be pointed out early in the Results section. I was initially concerned when the assay was described as based on consumption of GTP that a significantly diminished pool would alter the rate and thereby distort results, and being made aware of the S4 result would have alleviated that concern as I read further.

      Response from the authors:

      We believe that the reviewer refers to S3 (not S4). We appreciate this suggestion and now mention it earlier.

      On page 4 and Fig S4 the authors mention several cdc42 constructs, some of which show linear activity curves and others slightly non-linear curves. I was unable to find where these constructs or their differences are discussed. The authors should also tell us if the construct used for the remaining experiments was one of the two shown in S4, or a different one.

      Response from the authors:

      We added the requested information and explanations to the manuscript.

      It seems that in Fig 4 and Fig S8, some points are missing from the graphs. Were all concentrations for each condition not always assayed, or is some data omitted for some reason? For example, for the 0.125 microM Rga2 condition, only two points are shown vs 4 for some other conditions, and the two missing ones are expected to not be excluded by the >5% GTP remaining criterion.

      Response from the authors:

      The reviewer wonders whether Fig.4 and Fig. S8 miss data points. This is not the case, and __we added clarifying information to the manuscript. __

      In detail: Not all assays contain the same amount of data points/ concentrations for each protein. We first assessed Cdc42 alone using several Cdc42 concentration. We then examined the individual Cdc42 – effector mixtures, using a larger number of effector concentrations. We included a reduced number of effector concentrations in the assays containing two effectors and Cdc42. It would be ideal to include more concentrations, but this is not always feasible: The assay involves a multitude of pipetting steps and is sensitive to any pipetting errors. Further, assays can vary slights from each other, therefore all samples that ought to be compared need to be included in each assay.

      Each three-protein assay contains samples shown (Cdc42, Cdc42 + effector 1, Cdc42 + effector 2, Cdc42 + effector 1 + effector 2) and additional ‘buffer’ wells used for normalization. Each data point shown corresponds to the average of 3-4 replica samples per assay. We therefore did not include all concentrations in all conditions. As pointed out, Fig. 4a only shows two data points for the 0.125uM Rga2 axis (Rga2 + Cdc42 and Rga2 + Cdc24 + Cdc42). The rational was the following: We included three Cdc24 concentrations (for proper fitting for K3,Cdc24), three Rga2 concentrations (for proper fitting for K3,Rga2), and 5 mixtures of the used Cdc24 and Rga2 concentrations (for proper fitting for K3,Cdc24,Rga2).

      The Cdc42-Rga2-BSA and Cdc42-Rga2-Casein data is rather sparse and would benefit from additional data points. However, we only use those as control experiments and are cautious in their interpretation.

      In these graphs, a diamond symbol of slightly varying color is used for the different conditions. The different colors are hard to distinguish. Please use different shape symbols for the different conditions, and choose colors that are more distinct.

      Response from the authors:

      We will adapt the color scheme of the fits to make the colors more distinguishable.

      There are a few sentences that are of unclear meaning, for example on page 10, "It was suggested that each GAP plays a distinct role in Cdc42 regulation, of which the level of GAP activity could be a part of [Smith et al., 2002]." There are also typos and grammatical errors that should be fixed.

      Response from the authors:

      __We will further check the document for potentially unclear sentences and will try to clarify them, as well as further check for grammatical and spelling errors. __

      Reviewer #4 (Significance (Required)):

      Significance

      The most novel and important finding is the strong synergy observed between cdc24 and Rga2 in activating cdc42 GTPase activity. This is undoubtedly an important mechanism underlying positive feedback in polarization. The measured non-linear activity of cdc24 alone is also quite important given that availability of cdc24 is thought to be a critical in vivo stimulus for polarization. However, the unexplained discrepancy between this result and that of Mionnet leaves one to wonder which result is more reliable. Only Mionnet attempts to directly test whether oligomerization is important in cdc24 activity.

      The conclusions are of importance to a broad audience of cell biologists, though the lack of any mechanism for the synergy or the non-linearity of cdc24 activity somewhat diminishes significance.

      Note that my expertise and that of my co-reviewer is in the biology, and while we are able to follow the contributions of the modeling, we do not have the expertise to critically evaluate for potential errors or weaknesses in the modeling itself.

      The reviewer wonders whether our data or the data of Mionnet et al. on the link between Cdc24 oligomerization and its GEF activity is more reliable and suggests to conduct experiments with oligomerization deficient Cdc24 mutants.

      We thank the reviewer for this recommendation and we will do the suggested experiments to resolve the seemingly contradicting observations by us and Mionnet et al..

      The reviewer would find mechanistic insights into (2) the non-linear concentration dependence of Cdc24’s activity and (2) the Cdc24-Rga2 synergy useful.

      (1) We will conduct experiments with partially oligomerization deficient Cdc24 mutants, as suggested by the reviewer.

      (2) We speculate that Cdc24-Rga2 binding could lead to the synergy. ____We will add data on Cdc24 – Rga2 binding (in vitro: Size-Exclusion Chromatography Multi-Angle Light Scattering) to this study.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #4

      Evidence, reproducibility and clarity

      Summary

      The GTPase cdc42 is a key determinant of yeast polarization. Its activity is amplified at the site of polarization through a poorly defined positive feedback mechanism, and depends on numerous GAPs regulating GTP hydrolysis and the GEF cdc24 that regulates GDP release. These components have previously been evaluated for their quantitative effects on the individual steps in the GTPase cycle that they modulate, but potential interactions between the cdc24 GEF and any GAP could not be examined based on these assays. The authors validate and employ a bulk assay of the total GTPase cycle based on GTP consumption to study the activities of and potential interactions between cdc24 and the GAP Rga2. Fitting their data to a mathematical model, they come to three central conclusions: (1) the activating activity of cdc24 to activate cdc42 GTPase activity is nonlinear, showing a quadratic relationship, (2) Rga2 shows a much lower activating activity that is linear at low levels before saturating, and (3) there is a strongly synergistic interaction between the activating activities of cdc24 and Rga2. Some hypotheses for the mechanistic bases of these findings are hypothesized, but not further investigated. Their conclusions are well supported by the data which appears to be of sufficient rigor.

      Major comments

      The three main conclusions of the manuscript are well supported by the data and associated modeling.

      One unresolved issue is the discrepancy between the authors' conclusion that the non-linear activation by cdc24 is likely a result of oligomerization, whereas Mionnet et al 2008 reach the opposite conclusion. It seems that the authors wish to discount the Mionnet results because they used truncated constructs to test deficient oligomerization and an engineered construct to test induced oligomerization. If the authors are correct, then a relatively easy test would be to introduce the oligomerization deficient mutants defined by Mionnet into their fuill length construct and compare to wild type protein. While the authors' measured results don't depend on the offered mechanism and this experiment is therefore optional, their explanation is quite unsatisfying, especially since an experiment to resolve the difference is entirely feasible and not very strenuous.

      Minor comments

      The results in Fig S4 serve as assay validation, and this should be pointed out early in the Results section. I was initially concerned when the assay was described as based on consumption of GTP that a significantly diminished pool would alter the rate and thereby distort results, and being made aware of the S4 result would have alleviated that concern as I read further.

      On page 4 and Fig S4 the authors mention several cdc42 constructs, some of which show linear activity curves and others slightly non-linear curves. I was unable to find where these constructs or their differences are discussed. The authors should also tell us if the construct used for the remaining experiments was one of the two shown in S4, or a different one.

      It seems that in Fig 4 and Fig S8, some points are missing from the graphs. Were all concentrations for each condition not always assayed, or is some data omitted for some reason? For example, for the 0.125 microM Rga2 condition, only two points are shown vs 4 for some other conditions, and the two missing ones are expected to not be excluded by the >5% GTP remaining criterion.

      In these graphs, a diamond symbol of slightly varying color is used for the different conditions. The different colors are hard to distinguish. Please use different shape symbols for the different conditions, and choose colors that are more distinct.

      There are a few sentences that are of unclear meaning, for example on page 10, "It was suggested that each GAP plays a distinct role in Cdc42 regulation, of which the level of GAP activity could be a part of [Smith et al., 2002]." There are also typos and grammatical errors that should be fixed.

      Significance

      The most novel and important finding is the strong synergy observed between cdc24 and Rga2 in activating cdc42 GTPase activity. This is undoubtedly an important mechanism underlying positive feedback in polarization. The measured non-linear activity of cdc24 alone is also quite important given that availability of cdc24 is thought to be a critical in vivo stimulus for polarization. However, the unexplained discrepancy between this result and that of Mionnet leaves one to wonder which result is more reliable. Only Mionnet attempts to directly test whether oligomerization is important in cdc24 activity.

      The conclusions are of importance to a broad audience of cell biologists, though the lack of any mechanism for the synergy or the non-linearity of cdc24 activity somewhat diminishes significance.

      Note that my expertise and that of my co-reviewer is in the biology, and while we are able to follow the contributions of the modeling, we do not have the expertise to critically evaluate for potential errors or weaknesses in the modeling itself.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      This work reports a biochemical analysis of the effects of a recombinant yeast GEF (Cdc24) and GAP (Rga2) on Cdc42 GTPase cycling in vitro. The central conclusion is that the GEF and GAP act "synergistically", which occurs "due to proteins enhancing each other's effects". By this they appear to mean that the GEF enhances the GAP's activity and vice versa. I was not persuaded that this is correct, and was confused by many aspects of the approach and interpretation, as outlined below.

      1. GEF and GAP are expected to accelerate GTPase cycle synergistically even with no effect on each other's activity:

      The Cdc42 GTPase cycle is understood to occur via distinct steps (GDP release, GTP binding, and GTP hydrolysis): GDP release and GTP hydrolysis are intrinsically slow steps that are accelerated by GEFs (GDP release) and GAPs (GTP hydrolysis). This fundamental biochemistry was established in the 1990s using biochemical assays that measure each step independently. Here instead the authors use an assay that measures [GTP] decline in a mix with 5 uM starting GTP, 1 uM Cdc42, plus or minus some amount of GEF or GAP. They assume exponential decline of [GTP] with time, yielding a cycling "rate". If that is so, then one would expect that added GEF would accelerate only the first step, leaving a slow GTP hydrolysis step that limits the overall cycling rate, while added GAP would accelerate only the last step, leaving a slow GDP release step that limits the overall cycling rate. Adding both together would speed up both steps, and should therefore "synergistically" accelerate cycling. This would be expected based on previous work and does not imply that GEF or GAP are affecting each other's action (except trivially by providing substrate for the next reaction). If the authors wish to demonstrate that something more complex is indeed happening, they need to use assays that directly measure the sub-reaction of interest, as done by prior investigators. 2. Model-based interpretation of the GTPase assay is poorly supported:

      The assay employed measures overall GTP concentration with time. It is assumed (but not well documented-see below) that [GTP] declines exponentially, and that the rate constant for a particular condition can be fit by the sum of a series of terms that are linear or quadratic in the concentrations of Cdc42, GEF, and GAP. There is no theoretical derivation of this model from the elementary reactions, and the assumptions involved are not well articulated.

      As discussed in point 1 above, one would expect that a GEF or GAP alone could only accelerate the cycle to a certain point, where the other (slow) reaction becomes rate limiting. But that does not appear to be true for their phenomenological model, where slow steps (small terms in the sum) will always be overwhelmed by fast steps. This is not the traditional understanding of how GTPases operate. 3. Data that do not conform to expectation are not explained: Strangely, the data (as interpreted by the model assumptions) also appear inconsistent with the expectation of rate-limiting steps. GEF addition (alone) is said to accelerate cycling 100-fold, while GAP addition (alone) accelerates it 2-fold. But that would seem to imply that GDP release takes up >99% of the basal cycle (so accelerating that step alone reduces cycling time 100-fold), while GTP hydrolysis takes up >50% of the basal cycle (so accelerating that step alone reduces cycling time 2-fold). In the conventional understanding of GTPase cycles, these cannot both be be true (as the steps would then add to >100% of the basal cycle). There is no attempt to reconcile these findings with previous work. 4. Lack of detailed timecourse data:

      The decline in [GTP] with time is stated to be exponential, allowing extraction of an overall cycling "rate". But this claim is supported only weakly (S3 Fig. 1 uses only 3 timepoints, is not plotted on semi-log axis, and does not report fit to exponential vs other models) and only for the Cdc42-alone scenario: no data at all are presented to support exponential decline in reactions with GEF or GAP. Most assays seem to measure only a single timepoint, so extraction of a "rate" is very heavily influenced by the unsupported assumption of exponential decline. And if the decline is not exponential, it becomes extremely difficult to interpret what a single timepoint means. 5. Other issues with interpretation of the data:

      (i) It is unclear why the authors chose to employ an assay that is much harder to interpret than the biochemical assays used by others. In biochemical studies, assays that report an output of multiple reactions are always harder to interpret than assays targeting a single reaction. As well-established assays are available for each individual step in GTPase cycles, any conclusions must be supported using such assays.

      (ii) The reported basal (and GEF/GAP-accelerated) rates are very slow, perhaps due to poor folding of recombinant proteins. This raises the possibility that much of the Cdc42 is inactive. If so, then accelerated GTP hydrolysis could come from increasing the active fraction of Cdc42, rather than catalyzing a specific step.

      (iii) The GEF and GAP preparations include multiple partial degradation products and it is unclear whether the measured activities come from full-length proteins or more active fragments.

      (iv) Cdc42 cycling is also accelerated by BSA and casein, suggesting that there are poorly understood aspects of the assay and that GEF and GAP actions may (like BSA and casein) involve non-canonical effects on Cdc42. As GEF and GAP are expected to interact better with Cdc42 than BSA or casein, these effects could dominate the observed changes in GTP levels.

      (v) Cdc42-alone cycling assays are said to be reproducible. However, assays with added GEF/GAP/BSA/Casein yield rates that vary almost an order of magnitude between replicates. This poor reproducibility further reduces confidence in the findings.

      (vi) It is unclear what timepoint was used for the different assays. 1.5 h at 30 degrees seems to be the standard here for the Cdc42-alone assays, but I assume that cannot be what was measured to assess GTP decline for GEF-containing assays as there would be very little GTP left at 1.5 h.

      (vii) The graph reporting GEF activity is plotted only for [GEF]<0.2 uM, but the rates used in the subsequent experiments are reported for mixtures with 1 uM GEF. The full range of GEF data should be plotted.

      (viii) S8 Data with casein seems very noisy and it is no longer at all clear that the quadratic fit for [Cdc24] is justified. Also, the symbol colors are very similar so it is hard to tell what data corresponds to what condition. The synergy between Cdc24 and Rga2 is also very noisy and the fits seem arbitrary.

      (ix) It is disturbing that different Cdc42 constructs behave quite differently (S4). This suggests that protein behavior is influenced by the various added epitope tags and protease cleavage sites (they also leave the C-terminal CAAX box rather than removing the AAX as would happen in vivo). These features raise the concern that these findings may not be directly relevant to the situation with endogenous yeast Cdc42. Of course, it is also the case that relevant Cdc42 biochemistry occurs with prenylated Cdc42 on membranes.

      Significance

      The basic biochemistry of Cdc42 cycles was figured out about 30 years ago. However, those studies did not examine how combinations of Cdc42 regulators (as opposed to individual regulators) might interact to produce effects not expected from combining their individual actions. Recently, this combination approach did lead to interesting findings by Rapali et al. This approach is worthwhile and addresses a major question of interest to the broader field of GTPase biochemistry.

      One main limitation of this study is technical: the main assay is less informative (though perhaps easier) than traditional assays, and it is unclear whether the recombinant proteins employed retain their normal activities. Another limitation is the model-based interpretation of the assay that does not include the potential for rate-limiting steps.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      The study entitled, "The GEF Cdc24 and GAP Rga2 synergistically regulate Cdc42 GTPase cycling" by Tschirpke et al., uses an in vitro GTPase assay to examine the GTPase cycle of Cdc42 in combination with its GEF and GAP effectors. The authors find that the Cdc24 GEF activity scales non-linearly with its concentration and the GAP Rga2 has substantially weaker effect on stimulating Cdc42 GTPase activity. Not surprisingly, the combined addition of Cdc24 and Rga2 lead to a substantial increase in Cdc42 GTPase activity.

      Referees cross-commenting

      In Zheng, Y., Cerione, R., and Bender, A. (1994) J. Biol. Chem. 269: 2369-2372 (Fig. 3C), the authors show that Cdc24 combined with the GAP Bem3 lead to a large synergy in boosting Cdc42 GTPase activity.

      Significance

      There is very little new information in this manuscript. Previous studies (Rapali et al. 2017) have shown that the scaffold protein Bem1 enhances the GEF activity of Cdc24. It is expected that the reconstitution of a GEF and GAP protein promote the GTPase cycle and indeed Zheng et al. (1994) showed that that Cdc24 combined with the GAP Bem3 lead to a large synergy in boosting Cdc42 GTPase activity. Hence the only potentially interesting finding in this work is that, in solution Cdc24 activity scales non-linearly with its concentration. However as this GEF and Cdc42 are associated with the membrane, the relevance of solution studies are less clear and furthermore the mechanistic basis for the non-linearity is not explored in detail. Given the limited new information from this work, the findings are, in their current form, too preliminary.

    5. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      This study would be much convincing if additional line of eukaryotic cells can be used to demonstrate the GEF-GAP synergy tis important for cell physiology. In addition, it would be best to demonstrate the spatiotemporal interaction of GEF-GAP using high-resolution live cell imaging.

      Significance

      The revised study would provide first line evidence that GEF-GAP synergy to be general regulatory property in eukaryotic kingdom.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      We would like to thank Review Commons for their innovative approach to scientific peer-review and publishing. We thank all the Reviewers for their positive, highly complementary assessment of the manuscript and for highlighting the high quality and reproducibility of the work and the novelty and significance of the results: “The experiments are well-designed and perfectly executed and presented”; “I felt that this is a strong manuscript for peer-review as it serves diversified interests in modern cell biology.”; “The manuscript would be of interest to basic researchers working on epithelial development. Also potentially to basic researchers working on cancer, due to the mitotic errors described.”. We are grateful for the Reviewers’ comments and suggestions that have contributed to improving the revised manuscript. We have addressed all the Reviewers’ concerns, as detailed below in the point-by-point response to the Reviewers. Textual changes in the revised manuscript are marked in Blue.

      __Reviewer #1 (Evidence, reproducibility and clarity (Required)): __

      *The manuscript "Crosstalk between the plasma membrane and cell-cell adhesion maintains epithelial identity for correct polarised cell divisions" by Dr. Hosawi and colleagues reports the characterisation of the mitotic connection between plasma membrane dynamics and division orientation in polarised mammalian epithelial cells in culture. The authors start from the comparison of mitotic events of human mammary MCF10A cells grown at optimal density or at low density. They observed that only at optimal density MCF10A cells polarise by E-cadherin mediated cell-cell contacts, and display uniform membrane enrichment at the cortex, whereas cells grown at low density do not show cortical E-Cadherin enrichment, and distribute aberrantly the plasma membrane at one side and in cytoplasmic vesicles, generating daughter cells with unequal size. Consistently, further analyses revealed that low-density MCF10A cells undergo misoriented mitosis, with chromosome congression and misegregetion defects. Mechanistically, low density MCF10A cells fail to organise a symmetric mitotic spindle and center it in metaphase. This is due to an increased cortical actomyosin thickness coupled to abnormal astral microtubule stability. Building on previous data from the Elias lab, the authors uncover a role of the membrane-associated S100A11 protein in maintaining correct plasma membrane dynamics and E-cadherin localisation in mitosis. Further dissection of the molecular mechanism underlying this mitotic function od S10011A revealed that it enriches at the cortex only in optimal-density MCF10A cells, and promotes spindle orientation by association with LGN and E-cadherin, upstream of E-cadherin. This evidence depicts the plasma membrane and S100A11 proteins as a key mechanical sensors of cell-cell adhesion orchestrating the recruitment of E-cadherin and LGN-dependent force generators to ensure correct division orientation. *

      *Major points: *

      *- Important information is presented in Supplementary Figure S3. I suggest to move these panels in the main figures. Specifically, I would replace figure 4A with S3A showing the distribution of endogenous S100A11 in MCF10A cells, rather than the one of the GFP-tagged version which is over-expressed. *

      __Authors response: __We thank the Reviewer for this suggestion. We have now moved Figure S3A to Figure 4, to replace Figure 4A and show the localisation of endogenous S100A11 during mitosis and included new quantifications in new Figure 4B. We have moved Figure 3A to supplementary figures (new Figure S4A). We have amended the text of the results section and the Source Data file accordingly.

      *- The mechanisms of division orientation governed by S100A11 seems to impinge on the control of cortical F-actin and astral microtubule dynamics. This is illustrated in figure S3C, which in my opinion should be shown in the main figures with some more explanation / experiments. The authors mention the " tight actin F-actin bundles at the cell-cell contacts" that are lost in S100A11-depleted cells, and that interact with astral microtubules. However this is not fully clear in figure S3C. I think the authors should find a way to present better these evidence which is key in supporting their molecular model. *

      __Authors response: __As requested by the Reviewer we have now moved Figure S3C to the main manuscript, as new Figure 5. To clarify further the effect of S100A11 depletion on the tight actin bundle formation at the cell-cell contacts, we have now included a new illustration in new Figure 5C. Additionally, we have clarified further these findings in the results section (page 11). While we agree with the Reviewer that additional experiments, for example using live imaging of MCF-10A cells co-labelled for F-actin and tubulin, would help assess further the crosstalk between cortical actin and astral microtubules, based on our experience these live imaging experiments are challenging and can take up to several months to optimise and may not warrant successful outcome.

      *- I think the discussion would benefit from the addition of a graphical cartoon model illustrating the role of S100A11 in controlling plasma membrane dynamics in mitosis and spindle orientation. *

      __Authors response: __We thank the Reviewer for this suggestion. We have now added a graphical cartoon (new Figure 7), summarising the role of S100A11-mediated regulation of plasma membrane dynamics in polarised cell division orientation, progression and outcome. We hope this new illustration clarifies further the mechanisms described in this study.

      *- Finally, to understand the relevance of S100A11 in the context of 3D polarised mammary epithelia, it would be very interesting to analyse the effect of S100A11 knock-downn in mouse mammary epithelial acini grown in matrigel. This is not essential for the proposed studies, but would add biological relevance to the mechanisms characterised in 2D colture. *

      __Authors response: __We agree with the Reviewer that validating our findings in 3D cultures of mammary epithelial cells will be important to determine the influence of S100A11-mediated regulation of plasma membrane dynamics during mitosis on lumen formation and tissue morphogenesis. This is exactly the direction where the follow-up of these findings will go. While the first author who led this work has graduated and left our lab, we have recently recruited a new PhD student to address this important question, which will need a few years of investigation to provide important insights, similarly to what we did in our previous work (Fankhaenel et al., 2023 Nat Commun).

      *Minor comments: *

      *- It would be preferable to mention the known functions of S100A11 in the introduction rather than at the beginning of the paragraph at pg. 9. *

      __Authors response: __In response to the Reviewer’s suggestion, we have now moved the paragraph describing known functions of S100A11 to the introduction of the revised manuscript (see page 5).

      *- at pg 10, beginning of paragraph, I find it a weird phrasing that "LGN interacts with F-actin". As reported in the reference cited here, this is through Afadin, which binds simultaneously LGN and cortical F-actin. I would rephrase it. *

      __Authors response: __We thank the Reviewer for clarifying this point, which we have now rectified in the revised manuscript (see page 11).

      __Reviewer #1 (Significance (Required)): __

      *The description of cell adhesion as key factor instructing correct mitotic progression and execution of oriented division of vertebrate epithelial cells by controlling plasma membrane dynamics is novel and interesting for scientist in the spindle orientation/polarity field. The experiments are well-designed and perfectly executed and presented. I am in favour of publication of the manuscript, providing that a few points are addressed. *

      Authors response: We thank the Reviewer for their very positive evaluation of our work.

      __Reviewer #2 (Evidence, reproducibility and clarity (Required)): __

      *Establishment and maintenance of cell polarity are fundamental processes for physiology in multi-cellular organism given the fact that more than 380 million epithelial cell renewal for every second in human adults. However, the precise mechanisms linking plasma membrane polarity and cortical cytoskeleton dynamics of epithelial cells during mitotic exit and interphase remain ill-illustrated. Salah Elias and her colleagues experimentally manipulated the density of mammary epithelial cells in culture, which led to several mitotic defects. Specifically, they found that perturbation of cell-cell adhesion integrity impairs the dynamics of the plasma membrane during mitosis, affecting the shape and size of mitotic cells and resulting in defects in mitosis progression and generating daughter cells with aberrant cytoarchitecture. In these conditions, F-actin-astral microtubule crosstalk is impaired leading to mitotic spindle misassembly and misorientation, which in turn contributes to chromosome mis-segregation. Mechanistically, they identified the S100 Ca2+-binding protein A11 as a key membrane-associated regulator that forms a complex with E-cadherin and LGN to coordinate plasma membrane remodelling with E-cadherin-mediated cell adhesion and LGN-dependent mitotic spindle machinery. I felt that this is a strong manuscript for peer-review as it serves diversified interests in modern cell biology. *

      Authors response: We thank the Reviewer for their overall very positive feedback on our manuscript.

      __Reviewer #2 (Significance (Required)): __

      Several key cellular experiments should be repeated using a second line of epithelial cells such as RPE1.

      __Authors response: __We agree with the Reviewer it will be interesting to test our findings in other epithelial cells, including RPE1 cells, a widely used epithelial cell model to study the mechanisms controlling cell division. Nonetheless, we would like to emphasise that while our work demonstrates the importance of the interplay between plasma membrane dynamics and cell-cell adhesion for correct execution of polarised cell divisions in mammary epithelial cells, our aim is not to generalise the role of these S100A11-mediated mechanisms. An elegant study has shown that the mechanisms controlling plasma membrane remodelling and elongation during mitosis to ensure correct positioning of the mitotic spindle and symmetric division differ between HeLa cells and RPE1 cells (Kiyomitsu and Cheeseman, 2013 Cell). Additional experiments in a second cell line will require a thorough characterisation of the expression and localisation of S100A11 during the cell cycle, as well as the use of extensive and time-consuming knockdown and imaging experiments over several months and may lead to different observations requiring further mechanistic investigation, which is beyond the initial scope of this study. Additionally, the PhD student who led this study has graduated and left the lab and presently we don’t have capacity or resources to conduct these suggested experiments. Finally, to precisely address the Reviewer’s concern, we have now amended the revised manuscript to make our statements more specific to mammary epithelial cells throughout the text.

      __Reviewer #3 (Evidence, reproducibility and clarity (Required)): __

      *Summary: your understanding of the study and its conclusions. *

      *The scope of the study is to understand the links between cell-cell adhesion integrity, plasma membrane dynamics and mitotic spindle in mammalian epithelial tissues. To test this, the authors cultured mammary epithelial cells at optimal or low density as a way of perturbing cell-cell adhesion. The authors conclude that perturbing cell-cell adhesion alters plasma membrane dynamics, causing mitotic defects and that S100A11 coordinates this link via E-cadherin. Whilst this is an interesting manuscript, illustrating the differences of mitotic success in optimal density vs. low density cell cultures, I do not think that the conclusions are supported by the evidence presented for the reasons stated below. *

      *Major comments: major issues affecting the conclusions. *

      *- The manuscript clearly shows that culturing cells at a lower density results in a higher incidence of asymmetric division (figure 1) and mitosis defects (figure 2). Cells round more and faster and there is more actin at the cortex during rounding (figure 3). However, whilst differences in cell-cell adhesion are likely to play a role in mediating these effects, I don't think that it is possible to claim from the data presented that these defects are specifically due to cell-cell adhesion differences. This is because the morphology of cells at low density is also very different - cells appear more mesenchymal, with migratory front-rear polarity instead of apical-basal polarity. These cells will therefore have many differences between them, cell-adhesion being just one. The data is also not showing a 'loss' of cell-cell adhesion integrity but are rather illustrating the differences between cells that have formed cell-cell adhesions and those that have not. To really test the specific role of cell-cell adhesions, the authors would need to inhibit adhesions directly but without altering the cell density - for example via chemical or genetic perturbation within a confined environment. I suggest that the authors either need to do these experiments or to requalify what their data is telling us. *

      __Authors response: __We thank the Reviewer for their insightful discussion of the proposed mechanisms described in our manuscript. Several of the Reviewer’s comments pinpoint and exactly match the messages that we would like to convey to the scientific community. Therefore, to address the Reviewer’s comments, we have carefully requalified our statements in several places in the revised manuscript, to ensure they are more clear and more precise.

      We agree with the Reviewer’s comment that our experiments using sub-optimal density of mammary epithelial cells rather prevents the formation of cell-cell adhesions than disturbing them. The Reviewer is right, our experiments in low-density cultures suggest that perturbation of cell-cell adhesion formation impairs mammary epithelial identity, where cells lose their polarity and adopt a more mesenchymal phenotype, associated with plasma membrane remodelling defects. This affects the dynamics and progression of cell division. Nonetheless, our observations suggest an interplay between cell-cell adhesion and the plasma membrane to maintains correct cell shape during mitosis. To test this hypothesis, we explored the function of S100A11 which we have identified in the LGN interactome in mitotic mammary epithelial cells (Fankhaenel et al., 2023 Nat Commun), and which has been shown to interact with E-cadherin at adherens junctions in MDCK cells (Guo et al., 2014 Sci Signal). This, together with the fact that S100A11 controls plasma membrane repair (Jaiswal et al., 2014 Nat Commun), suggested S100A11 as an interesting candidate to investigate the interplay between cell-cell adhesion and membrane remodelling during mitosis. The data presented here suggest that we were right and the perturbation of our membrane-bound target, S100A11, indeed leads to the same mitotic phenotypes. S100A11 RNAi-mediated knockdown (48h) affects E-cadherin localisation at the plasma membrane and impairs cell-cell adhesion formation with effects on plasma membrane dynamics that phenocopy the defects observed in our low-density culture experiments. Remarkably, perturbation of cell-cell adhesions persisted in cell treated with si-S100A11 for 72h (see Figure S3). Of note, all our siRNA experiments have been carried out in cells cultured at optimal density to establish cell-cell adhesions. Thus, S100A11 knockdown allows genetic perturbation of E-cadherin-mediated cell-cell adhesion and recapitulates the plasma membrane and mitotic defects observed in sub-optimal cultures of mammary epithelial cells. Future experiments will be key to dissect these S100A11-mediated mechanisms to further understand how plasma membrane remodelling and cell-cell adhesions are coordinated during mitosis. Finally, as suggested by the Reviewer, we have now requalified our conclusions as appropriate in the revised manuscript.

      *- The current manuscript also demonstrates that cell adhesion is affected when S100A11 is knocked down (figure 4). It shows binding between and colocalization of S100A11 and E-cadherin, and shows that LGN cortical distribution is affected when S100A11 is knocked down (Figure 5). The results presented are suggestive of S100A11 being upstream of E-cadherin. However, I don't understand how the data shows "crosstalk between the plasma membrane, cell-cell adhesion, and the cell cortex during mitosis". For example, on P9: "We observed unequal distribution of CellMaskTM in a vast majority of S100A11-depleted cells (si-S100A11#1: ~79% versus si-Control: ~26%), indicating defects in plasma membrane remodelling (Figures 4B and 4C)." I don't agree that this demonstrates a defect in PM remodelling. Rather the cells in the representative images are less adherent and have adopted a more migratory cell state similar to that seen in figure 1 when seeded at low density. The fluidity of the much larger cells shown in knock down cells in panel F also appears higher, again suggesting an adhesion defect. *

      • *

      __Authors response: __The Reviewer has raised very important points here, which we would like to clarify.

      We agree with the Reviewer that our results in S100A11-depleted cells indicate impaired cell adhesions which generates cells displaying an invasive/migratory behaviour. However, our analysis of S100A11-depleted mitotic cells labelled with CellMaskTM reveals abnormal plasma membrane elongation generating two daughter cells displaying defective geometry as compared to control cells. These defects in the plasma membrane and cell shape were not noticeable upon E-cadherin knockdown (see previous Figures 5K and 5L; now new Figures 6K and 6L). Thus, our results strongly suggest that S100A11 acts as an upstream cue that coordinates plasma membrane dynamics with E-cadherin-mediated cell adhesions, and that additional mechanisms may be regulated by S100A11 to coordinate cell-cell adhesion with plasma membrane remodelling. How S100A11 ensures such a dynamic interplay between the plasma membrane and E-cadherin during mitosis remains a key question that we have not fully addressed in this initial study. An attractive mechanism could be mediated by the function of S100A11 in regulating the dynamic interaction between F-actin and the plasma membrane, as previously reported (Jaiswal et al., 2014 Nat Commun). Increasing evidence shows the importance of the crosstalk between the plasma membrane, the cortex and cell shape for correct execution of mitosis (Rizzelli et al., 2020 Open Biol). In our experiments, we show that impaired plasma membrane remodelling and cell shape are associated with defects in F-actin and astral microtubule organisation. Thus, our findings reinforce a model whereby S100A11 is a key membrane-associated protein that coordinates the crosstalk between the plasma membrane, cell-cell adhesion, and the cell cortex during mitosis. It will be key to characterise the interactome of S100A11 during mitosis to provide important mechanistic insights into this new role of S100A11; it is our intention to investigate this in the future.

      To address the points raised by the Reviewer, we have changed and clarified the statements they highlighted above, in the revised manuscript (pages 10 and 11).

      *- An earlier paper from the same lab this year identified Annexin A1 as directing mitotic spindle orientation via localising LGN at lateral cortex. During this earlier paper they also identified S100A11, which is a partner for Annexin A1. The authors could more clearly explain what S100A11 is in the current manuscript and how the current study builds on this earlier study. *

      __Authors response: __We thank the Reviewer for highlighting our previous work characterising the interactome of LGN in mitotic mammary epithelial cells (Fankhaenel et al., 2023 Nat Comms), and identifying Annexin A1 (ANXA1) as a polarity cue regulating the localisation and function of the evolutionarily conserved mitotic spindle orientation LGN complex. We also showed that ANXA1 direct partner S100A11 co-purifies with LGN and that perturbation of the ANXA1-S100A11 complex impairs the localisation of the LGN complex at the cell cortex during mitosis. Thus, as rightly pointed out by the Reviewer, this work builds on our previous work discussed above, but also on previous studies establishing S100A11 as a key regulator of plasma membrane repair by regulating the dynamic interplay between F-actin and the plasma membrane (Jaiswal et al., 2014 Nat Commun), and studies showing that S100A11 interacts with E-cadherin at adherens junctions (Guo et al., 2014 Sci Signal). To address the Reviewer’s point (also raised by Reviewer 1), we have now included a paragraph in the introduction (page 5) and results (page 10) of the revised manuscript describing these and other functions of S100A11 to provide a strong rational to our decision to investigate this protein.


      *- Based on the data presented, I suggest that the authors should requalify their data. I suggest that the conclusions that can be drawn from the data are that cellular state is important for regulating mitosis orientation and fidelity (i.e. adherent epithelia cells vs. less adherent more migratory cells). S100A11 is important for promoting cell-cell adhesions and might be upstream of the known role of E-cadherin in regulating spindle orientation. Whilst I suggest that more quantified experiments would need to be included in order to assess possible effects on plasma membrane remodelling, the manuscript could be generally improved by a clearer explanation of the open question that they are addressing and what specific advance this manuscript has made in relation to the current literature, including their own. I do not currently feel that the title of the manuscript is appropriate since I don't think that a crosstalk between the plasma membrane and cell-cell adhesion has been shown here. *

      __Authors response: __We would like to reiterate our agreement with the Reviewer’s suggestion about the conclusions drawn from our data. In the initial submission we proposed that perturbation of S100A11-mediated regulation of cell adhesion and plasma membrane impairs the identity of mammary epithelial cells, which affects their shape during mitosis leading to aberrant mitotic progression and outcome. While we have not checked the migratory behaviour of cells not forming cell-cell adhesions, we suggested that the cells adopted a mesenchymal phenotype. Furthermore, we discussed the implication of epithelial-to-mesenchymal transition on chromosome segregation fidelity and execution of mitosis, and how precisely they link with our study (see initial submission’s pages 14 and 19). As suggested by the Reviewer, we have now clarified further these observations in the results (pages 7 and 11) and discussion (pages 15 and 19) of the revised manuscript.

      We have quantified several aspects of the changes in plasma membrane dynamics and remodelling throughout, in the initial manuscript (Figure 1D-H; Figure 4C). To address the Reviewer’s point, we have now added quantifications of membrane blebbing (new Figure 1I).

      We would like to emphasise that the introduction of the initial manuscript has included the open questions that led to this study. These questions have been addressed further in the discussion, where we have also formulated new hypotheses and discussed what we think are the important outstanding questions for future investigations, in light of our findings. In this study we demonstrate that maintaining epithelial identity is essential for correct execution of polarised cell divisions. Our findings indicate that mammary epithelial cells grown at sub-optimal density lose their epithelial identity, which results in several mitotic defects. We propose a novel mechanism in which S100A11 acts as a molecular sensor of external cues coordinating the interplay between plasma membrane dynamics and cell-cell adhesion to maintain epithelial identity and integrity, thereby ensuring correct progression, orientation, and outcome of cell division. As suggested by the Reviewer, we have clarified further the advances made in this study, in the revised Results and Discussion sections.

      To address the Reviewer’s final point, we would like to suggest the following revised title “Interplay between the plasma membrane and cell-cell adhesion maintains epithelial identity for correct polarised cell divisions”, which we hope reflects better the results described in our studies.

      *Minor comments: important issues that can confidently be addressed. *

      - P3: I wouldn't describe the junctional proteins listed as polarity proteins.

      __Authors response: __We have now made this rectification in page 3, as suggested by the Reviewer.

      *- Figure 1 - can the membrane blebbing phenotype by quantified? At the moment this part is observational so can't really be used to determine the role of plasma membrane remodelling. *

      • *

      __Authors response: __We have now included quantifications of blebbing in the revised manuscript, as suggested by the Reviewer (new Figure 1I).

      *- Figure 3. I'm not sure what the 'subcortical actin cloud' measurement is. Figure 3G suggests it may be the distance from the cortex to the spindle pole but how does this relate to actin? *

      __Authors response: __The Reviewer is right, the subcortical actin cloud includes a pool of dynamic subcortical actin that extends from the cortex (excluding the stiff cortical actin) to the cytoplasm, interacting with the centrosomes and concentrating near the retraction fibres. The subcortical actin cloud has been shown to mediate cortical forces and to concentrate force-generating proteins at the retraction fibres acting on centrosome dynamics and pulling on astral microtubules to orient the mitotic spindle (for example, please see Kwon et al., 2015 Dev Cell). We have now included this clarification in the revised manuscript (page 10).

      *- Figure 4A. I can't see GFP-S100A11 accumulating at the cell surface. To me these images suggest that it is relatively ubiquitously expressed throughout the cytoplasm and surface, which is different to the later antibody stains, that show localisation at the cell surface. *

      __Authors response: __A similar point has been raised by Reviewer 1. Although our retroviral-mediated transduction allows to avoid excessive expression of GFP-S100A11, the ectopic S100A11 is expressed at higher levels as compared to its endogenous counterpart. Our live images show an accumulation of the protein at the cell surface, but relatively high levels are also visible in the cytoplasm (previous Figure 4A, new Figure S4A). By contrast immunolabelling for endogenous S100A11 shows an obvious accumulation of the protein at the plasma membrane. This difference could also be due to a dynamic behaviour of the protein translocation of GFP-S100A11 between the cell surface and cytoplasm that is captured in our live imaging. Similar slight differences between immunofluorescence and live imaging of cortical proteins involved in mitosis, such as Dynein, NuMA, LGN and CAPZB, have been reported in several studies (to cite a few: di Pietro et al., 2017 Curr Biol; Elias et al., 2014 Stem Cell Rep; Fankhaenel et al., 2023). To address this point, we have now moved the panel showing S100A11 immunofluorescence in Figure S3A to new Figure 4A (also see response to Reviewer 1 Major Point 1).

      *- Fig 4H doesn't show an active process of translocation of E-Cadherin to the cytoplasm. It shows representative images with slightly higher levels of E-Cadherin in the cytoplasm. This could be due to translocation or it could be to do with lack of E-Cadherin assembly. *

      • *

      __Authors response: __We thank the Reviewer for pointing this out. We have rectified this statement accordingly (page 11).

      *- 4I I don't understand where the line profile is derived from - where is apical and where is basal in the images? Could a diagram be included? *

      __Authors response: __We have now included an illustration of this quantification, in the revised manuscript (new Figure 4J).

      - The discussion could be shortened and more clearly written - perhaps with subheadings of the main findings.

      __Authors response: __We have clarified several ideas and statements, based on the specific points addressed above. While it is challenging to reduce the size of this section, given that the study addresses several mechanisms of mitosis, we have now shortened the discussion in the revised manuscript.

      *- Methods: Why is cholera toxin used in the cell culture medium? *

      • *

      __Authors response: __Cholera toxin is a key component of MCF-10A medium, which has been shown to stimulate cAMP activation promoting cell proliferation in culture. This culture protocol is a gold standard in the field (Debnath et al., 2023 Methods). Given that cholera toxin is a highly regulated chemical and takes several months to purchase, we have tried culturing MCF-10A without the toxin, but this negatively affected proliferation and passage of this cells. Therefore, we concluded that adding it to the culture medium is important.

      __Reviewer #3 (Significance (Required)): __

      *In general, this is an interesting paper about the fidelity of mitosis in cells in adherent monolayers vs. in more migratory, non-adherent states. There is existing literature on this topic (some cited in the manuscript, alongside reviews of the topic). *

      • *

      *The main conceptual advance, as far as I can see, is that S100A11 is important for promoting cell-cell adhesions and might be upstream of the known role of E-cadherin in regulating spindle orientation via LGN. The main limitation is that plating cells at different densities is not a direct 'perturbation' of cell-cell adhesion. This means that the phenotypes seen could be due to many factors, not just cell adhesion. Assessment of plasma membrane and cytoskeletal dynamics are also often observational and not conclusive. *

      • *

      *The manuscript would be of interest to basic researchers working on epithelial development. Also potentially to basic researchers working on cancer, due to the mitotic errors described. *

      • *

      *I have expertise in epithelial cell biology. *

      I estimate the authors would need between 3 and 6 months for revisions if they decide to do further experiments and between 1 and 3 months if they decide to re-qualify their claims.

      • *

      __Authors response: __We thanks the Reviewer for their overall positive feedback on our work and its broader importance for researchers in epithelial development and cancer biology.

      We would like to reiterate our agreement with the Reviewer’s assessment of the conceptual advances of our work. We show that S100A11 complexes with E-cadherin and LGN during mitosis to control cell-cell adhesion assembly and the mitotic spindle machinery, respectively, which in turn ensures faithful chromosome segregation. Our results also suggest that S100A11 lies upstream of E-cadherin in the regulation of the LGN-mediated mitotic spindle machinery. We also agree with the Reviewer that plating epithelial cells at low density does not directly affect cell-cell adhesion, because, in these culture conditions, cells are not dense enough to establish cell-cell contacts necessary to assemble stable adherens junctions. Rather, and as rightly pointed out by the Reviewer, cells grown at low density fail to maintain their epithelial identity and adopt a more mesenchymal and elongated behaviour, which is accompanied by dramatic changes in plasma membrane remodelling throughout mitosis. Interestingly, our results show that both S100A11 and E-cadherin do not localise at the plasma membrane in these sub-optimal culture conditions. This along with our results showing that depletion of S100A11 phenocopies the effect of low-density culture conditions on plasma membrane remodelling and E-cadherin mediated cell-cell adhesion assembly, allow us to propose a mechanism whereby the membrane-associated S100A11 protein acts as a molecular sensor of external cues bridging plasma membrane remodelling to E-cadherin-dependent cell adhesion to coordinate correct progression and outcome of mammary epithelial cell divisions.

      We are grateful for the Reviewer’s insightful discussion of our findings. As we discussed above in our responses to their specific points, we have requalified many of our statements to clarify further our main findings and conclusions throughout the revised manuscript. We have also added new quantifications in response to the Reviewer’s suggestions. We believe, that together, these revisions have advanced further the initial manuscript.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary: your understanding of the study and its conclusions.

      The scope of the study is to understand the links between cell-cell adhesion integrity, plasma membrane dynamics and mitotic spindle in mammalian epithelial tissues. To test this, the authors cultured mammary epithelial cells at optimal or low density as a way of perturbing cell-cell adhesion. The authors conclude that perturbing cell-cell adhesion alters plasma membrane dynamics, causing mitotic defects and that S100A11 coordinates this link via E-cadherin. Whilst this is an interesting manuscript, illustrating the differences of mitotic success in optimal density vs. low density cell cultures, I do not think that the conclusions are supported by the evidence presented for the reasons stated below.

      Major comments: major issues affecting the conclusions.

      The manuscript clearly shows that culturing cells at a lower density results in a higher incidence of asymmetric division (figure 1) and mitosis defects (figure 2). Cells round more and faster and there is more actin at the cortex during rounding (figure 3). However, whilst differences in cell-cell adhesion are likely to play a role in mediating these effects, I don't think that it is possible to claim from the data presented that these defects are specifically due to cell-cell adhesion differences. This is because the morphology of cells at low density is also very different - cells appear more mesenchymal, with migratory front-rear polarity instead of apical-basal polarity. These cells will therefore have many differences between them, cell-adhesion being just one. The data is also not showing a 'loss' of cell-cell adhesion integrity but are rather illustrating the differences between cells that have formed cell-cell adhesions and those that have not. To really test the specific role of cell-cell adhesions, the authors would need to inhibit adhesions directly but without altering the cell density - for example via chemical or genetic perturbation within a confined environment. I suggest that the authors either need to do these experiments or to requalify what their data is telling us. The current manuscript also demonstrates that cell adhesion is affected when S100A11 is knocked down (figure 4). It shows binding between and colocalization of S100A11 and E-cadherin, and shows that LGN cortical distribution is affected when S100A11 is knocked down (Figure 5). The results presented are suggestive of S100A11 being upstream of E-cadherin. However, I don't understand how the data shows "crosstalk between the plasma membrane, cell-cell adhesion, and the cell cortex during mitosis". For example, on P9: "We observed unequal distribution of CellMaskTM in a vast majority of S100A11-depleted cells (si-S100A11#1: ~79% versus si-Control: ~26%), indicating defects in plasma membrane remodelling (Figures 4B and 4C)." I don't agree that this demonstrates a defect in PM remodelling. Rather the cells in the representative images are less adherent and have adopted a more migratory cell state similar to that seen in figure 1 when seeded at low density. The fluidity of the much larger cells shown in knock down cells in panel F also appears higher, again suggesting an adhesion defect. An earlier paper from the same lab this year identified Annexin A1 as directing mitotic spindle orientation via localising LGN at lateral cortex. During this earlier paper they also identified S100A11, which is a partner for Annexin A1. The authors could more clearly explain what S100A11 is in the current manuscript and how the current study builds on this earlier study.

      Based on the data presented, I suggest that the authors should requalify their data. I suggest that the conclusions that can be drawn from the data are that cellular state is important for regulating mitosis orientation and fidelity (i.e. adherent epithelia cells vs. less adherent more migratory cells). S100A11 is important for promoting cell-cell adhesions and might be upstream of the known role of E-cadherin in regulating spindle orientation. Whilst I suggest that more quantified experiments would need to be included in order to assess possible effects on plasma membrane remodelling, the manuscript could be generally improved by a clearer explanation of the open question that they are addressing and what specific advance this manuscript has made in relation to the current literature, including their own. I do not currently feel that the title of the manuscript is appropriate since I don't think that a crosstalk between the plasma membrane and cell-cell adhesion has been shown here.

      Minor comments: important issues that can confidently be addressed.

      P3: I wouldn't describe the junctional proteins listed as polarity proteins. Figure 1 - can the membrane blebbing phenotype by quantified? At the moment this part is observational so can't really be used to determine the role of plasma membrane remodelling.

      Figure 3. I'm not sure what the 'subcortical actin cloud' measurement is. Figure 3G suggests it may be the distance from the cortex to the spindle pole but how does this relate to actin?

      Figure 4A. I can't see GFP-S100A11 accumulating at the cell surface. To me these images suggest that it is relatively ubiquitously expressed throughout the cytoplasm and surface, which is different to the later antibody stains, that show localisation at the cell surface.

      Fig 4H doesn't show an active process of translocation of E-Cadherin to the cytoplasm. It shows representative images with slightly higher levels of E-Cadherin in the cytoplasm. This could be due to translocation or it could be to do with lack of E-Cadherin assembly.

      4I I don't understand where the line profile is derived from - where is apical and where is basal in the images? Could a diagram be included?

      The discussion could be shortened and more clearly written - perhaps with subheadings of the main findings.

      Methods: Why is cholera toxin used in the cell culture medium?

      Significance

      In general, this is an interesting paper about the fidelity of mitosis in cells in adherent monolayers vs. in more migratory, non-adherent states. There is existing literature on this topic (some cited in the manuscript, alongside reviews of the topic).

      The main conceptual advance, as far as I can see, is that S100A11 is important for promoting cell-cell adhesions and might be upstream of the known role of E-cadherin in regulating spindle orientation via LGN. The main limitation is that plating cells at different densities is not a direct 'perturbation' of cell-cell adhesion. This means that the phenotypes seen could be due to many factors, not just cell adhesion. Assessment of plasma membrane and cytoskeletal dynamics are also often observational and not conclusive.

      The manuscript would be of interest to basic researchers working on epithelial development. Also potentially to basic researchers working on cancer, due to the mitotic errors described.

      I have expertise in epithelial cell biology.

      I estimate the authors would need between 3 and 6 months for revisions if they decide to do further experiments and between 1 and 3 months if they decide to re-qualify their claims.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Establishment and maintenance of cell polarity are fundamental processes for physiology in multi-cellular organism given the fact that more than 380 million epithelial cell renewal for every second in human adults. However, the precise mechanisms linking plasma membrane polarity and cortical cytoskeleton dynamics of epithelial cells during mitotic exit and interphase remain ill-illustrated. Salah Elias and her colleagues experimentally manipulated the density of mammary epithelial cells in culture, which led to several mitotic defects. Specifically, they found that perturbation of cell-cell adhesion integrity impairs the dynamics of the plasma membrane during mitosis, affecting the shape and size of mitotic cells and resulting in defects in mitosis progression and generating daughter cells with aberrant cytoarchitecture. In these conditions, F-actin-astral microtubule crosstalk is impaired leading to mitotic spindle misassembly and misorientation, which in turn contributes to chromosome mis-segregation. Mechanistically, they identified the S100 Ca2+-binding protein A11 as a key membrane-associated regulator that forms a complex with E-cadherin and LGN to coordinate plasma membrane remodelling with E-cadherin-mediated cell adhesion and LGN-dependent mitotic spindle machinery. I felt that this is a strong manuscript for peer-review as it serves diversified interests in modern cell biology.

      Significance

      Several key cellular experiments should be repeated using a second line of epithelial cells such as RPE1.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      The manuscript "Crosstalk between the plasma membrane and cell-cell adhesion maintains epithelial identity for correct polarised cell divisions" by Dr. Hosawi and colleagues reports the characterisation of the mitotic connection between plasma membrane dynamics and division orientation in polarised mammalian epithelial cells in culture. The authors start from the comparison of mitotic events of human mammary MCF10A cells grown at optimal density or at low density. They observed that only at optimal density MCF10A cells polarise by E-cadherin mediated cell-cell contacts, and display uniform membrane enrichment at the cortex, whereas cells grown at low density do not show cortical E-Cadherin enrichment, and distribute aberrantly the plasma membrane at one side and in cytoplasmic vesicles, generating daughter cells with unequal size. Consistently, further analyses revealed that low-density MCF10A cells undergo misoriented mitosis, with chromosome congression and misegregetion defects. Mechanistically, low density MCF10A cells fail to organise a symmetric mitotic spindle and center it in metaphase. This is due to an increased cortical actomyosin thickness coupled to abnormal astral microtubule stability. Building on previous data from the Elias lab, the authors uncover a role of the membrane-associated S100A11 protein in maintaining correct plasma membrane dynamics and E-cadherin localisation in mitosis. Further dissection of the molecular mechanism underlying this mitotic function od S10011A revealed that it enriches at the cortex only in optimal-density MCF10A cells, and promotes spindle orientation by association with LGN and E-cadherin, upstream of E-cadherin. This evidence depicts the plasma membrane and S100A11 proteins as a key mechanical sensors of cell-cell adhesion orchestrating the recruitment of E-cadherin and LGN-dependent force generators to ensure correct division orientation.

      Major points:

      • Important information is presented in Supplementary Figure S3. I suggest to move these panels in the main figures. Specifically, I would replace figure 4A with S3A showing the distribution of endogenous S100A11 in MCF10A cells, rather than the one of the GFP-tagged version which is over-expressed.
      • The mechanisms of division orientation governed by S100A11 seems to impinge on the control of cortical F-actin and astral microtubule dynamics. This is illustrated in figure S3C, which in my opinion should be shown in the main figures with some more explanation / experiments. The authors mention the " tight actin F-actin bundles at the cell-cell contacts" that are lost in S100A11-depleted cells, and that interact with astral microtubules. However this is not fully clear in figure S3C. I think the authors should find a way to present better these evidence which is key in supporting their molecular model.
      • I think the discussion would benefit from the addition of a graphical cartoon model illustrating the role of S100A11 in controlling plasma membrane dynamics in mitosis and spindle orientation.
      • Finally, to understand the relevance of S100A11 in the context of 3D polarised mammary epithelia, it would be very interesting to analyse the effect of S100A11 knock-downn in mouse mammary epithelial acini grown in matrigel. This is not essential for the proposed studies, but would add biological relevance to the mechanisms characterised in 2D colture.

      Minor comments:

      • It would be preferable to mention the known functions of S100A11 in the introduction rather than at the beginning of the paragraph at pg. 9.
      • at pg 10, beginning of paragraph, I find it a weird phrasing that "LGN interacts with F-actin". As reported in the reference cited here, this is through Afadin, which binds simultaneously LGN and cortical F-actin. I would rephrase it.

      Significance

      The description of cell adhesion as key factor instructing correct mitotic progression and execution of oriented division of vertebrate epithelial cells by controlling plasma membrane dynamics is novel and interesting for scientist in the spindle orientation/polarity field. The experiments are well-designed and perfectly executed and presented. I am in favour of publication of the manuscript, providing that a few points are addressed.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      The authors do not wish to provide a response at this time.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      This MS contains carefully carried out and well controlled experiments describing a new pFFAT in ELYS. There is a similarly convincing demonstration of functionally relevant colocalisation by proximity ligation assay (PLA), particularly that both ELYS and VAP are nuclear envelope proteins in interphase without interacting (neg control in Fig 4D).

      Major Issue: Functional significance

      A key conclusion is that experiments prove that "ELYS serves as the crucial initiation factor for post-mitotic NPC-assembly" (p5). However, evidence for this is lacking as this would require reconstitution of NPC assembly with a mutant form of ELYS carefully changing the FFAT motif (e.g. 1321A 1324E) and exclusion of other probable VAP targets in experiments with mutant VAP. VAPs are among the proteins with the highest number of documented interactors (see Huttlin 2015/7 etc, e.g. PMID 26186194), so knocking down VAP may have pleiotropic effects and quite indirect read-outs in many aspects of cell function. In addition, for this work specifically there are other NE proteins that are known interactors of VAP: Emerin (EMD) and LBR both interact with VAP (high-throughput data, VAPA and VAPB). EMD has a motif similar to the canonical phospho-FFAT: 98 SYFTTRT 104. LBR has no motif. These findings should not be overlooked in this work. For example, was the interaction with emerin (page 4) sensitive to mutating VAP or ELYS? Could the effect seen in Figure 5 result from interactions with proteins other them ELYS?

      Further experiments should be carried out to justify all statements in the current MS of functional significance. Instead of doing more experiments, an alternative for the authors would be to describe the current set of results more cautiously. However, that would require changing much of the impact of the current MS, from the title onwards.

      Moderate Issue: VAPA

      From the start of the Introduction and some elements of the Discussion, include VAPA in equal measure with VAPB. When describing interactions of ELYS with VAP note that Huttlin et al., reported interactions twice for each of VAPA and VAPB. When describing own results (James et al. 2019) and those of others (Saiz-Ros et al., 2019) that focused on VAPB, clarify if the authors' view is that VAPA would (or would not) have the same interaction.

      Is there any evidence that only VAPB is on NE? Note that some refs in the Introduction relate to VAPA: Mesmin (not VAPB); ACBD5: although article titles refer to VAPB, early work (10.1083/jcb.201607055) showed almost identical involvement of VAPA. Also, this redundancy likely explains "function of VAPB in mitosis is not essential," (in Discussion). The lack of effect of VAPA knock-down may indicate that in these cells VAPB is dominant, but does not exclude a role for VAPA when VAPB is reduced. That might be tested by depleting both. Even following that, there is MOSPD2 to consider

      Other aspects of the writing

      "two amino acid residues are crucial for the interaction (VAPB K87 and M89)." This is wrong. Many residues are critical, these are merely 2 of possibly >10 that were chosen by Kaiser et al (2005) to create their non-binder.. Others have used different mutations to block FFAT binding.

      "They may exhibit a certain binding preference to specific members of the VAP ... family...". I cannot think of any example. I note no citation is given.

      When listing many or all MSP proteins, the text should state that MOSPD2 is uniquely close to VAPA/B. CFAP65 is typically not mentioned in the VAP-like lists as it does not have any of the conserved sequence that binds FFAT. If however the authors wish to include all human MSP domain protein, they should also include Hydin.

      Slightly wrong to cite De Vos et al., 2012 about PTPIP51's FFAT as that paper makes no mention of the motif. Better pick Di Mattia (again)

      On VAPB (and also A) on INM: there are references to be cited esp. relating to intranuclear Scs2 in yeast (Brickner et al 2004, Ptak et al 2021)

      Citations for VAP at ER-mito contacts "De Vos et al., 2012; Gómez-Suaga et al., 2019; Stoica et al., 2014)". These all refer to the same bridging protein, PTPIP51. Reduce to one citation. Then mention other proteins at the same site VPS13A, mitoguardin(MIGA)-2 ...

      "The domain interacts with characteristic peptide sequences ..." add citation to this sentence

      "Several variants of such motifs have been described: (i)" ... "(ii)": (i) and (ii) are entirely unlinked. Delete these and also "Several variants of such motifs have been described." Which is repeated later

      "FFAT-like motifs come in different flavors and may even lack the two phenylalanine residues (Murphy and Levine, 2016)": while motifs can tolerate variation at both positions, this text is misleading as it implies much more variation than is known. The 1st F can only be conservatively substituted (Y).

      Minor aspects in Results:

      ORP1L peptide as positive control: cite Kaiser 2005

      Was phosphoproteomics done in such a way as to find peptides that have both S1314 and S1326?

      Figure 4D, row 2: Comment on intranuclear staining in Prophase (at approx 4 o'clock) of both ELYS & VAP that is PLA positive

      Referees cross-commenting

      I agree with this point from Reviewer #1. We all agree that the main issue can be resolved experimentally to determine the effect of subtle point mutations in ELYS. Both other reviewers have done a good job in finding issues with the experiments that can also be addressed.

      Significance

      This work documents an interaction between the protein ELYS, that is involved in the reformation of nuclear pore complexes after mitosis, and the ER membrane protein VAPB. The interactions was previously known through high-throughput studies, along with many 100's of others for VAP, but here it is studied in detail and with care, identifying how the motif is induced by phosphorylation of ELYS. The two proteins are co-localised using convincing proximity ligation assays. This biochemistry and cell biological localisation is well done.

      Functional experiments then show that VAP (in this case VAPB) knock-down affects mitosis and chromosome segregation. While the result is incontrovertible, it has many possible interpretations, mainly because VAP has hundreds of interactions, including with multiple proteins involved in mitosis beyond just ELYS. This means that there are major limitations on how the interaction and co-localisation should be interpreted, reducing the advance associated with the current manuscript to incremental, and the limiting the audience to specialized.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary:

      In this study, James et al. follow up on their prior discovery that the ER contact site protein VAPB localizes to the nuclear envelope and is a putative binding partner of the nucleoporin ELYS, which coordinates nuclear envelope reformation (NER) with nuclear pore complex (NPC) biogenesis at mitotic exit and is also a constituent of the nuclear facing Y-complexes of mature NPCs. Using a series of complementary biochemical approaches the authors 1) demonstrate that VAPB and ELYS directly interact, 2) map the binding sites on ELYS that are sufficient to bind VAPB, 3) show that mutations that disrupt VAPB-FFAT motif binding also abrogate binding to ELYS including of the full-length protein, 4) define mitotic phosphorylation sties on VAPB-bound ELYS, 5) demonstrate that phosphorylation of ELYS, specifically at the FFAT2 motif, is required for binding to VAPB, and 6) demonstrate that the phosphorylation of ELYS that regulates VAPB binding occurs in mitosis. Turning to cell biology, the authors find that VAPB, which is an established ER protein, has some preference for non-core regions during NER (like ELYS). In addition, PLA analysis suggests that the interaction of VAPB and ELYS is most robust during anaphase and is somewhat disrupted when the binding of VAPB to FFAT motifs is lost due to targeted mutation. Last, the authors demonstrate that depletion of VAPB leads to metaphase delay and lagging chromosomes.

      Major comments:

      The data supporting direct binding of peptides encoding the FFAT 1 and FFAT2 motif derived from ELYS to VAPB in a manner similar to other FFAT sequences is strong, as is the effect of phosphorylation of FFAT1 on the strength of this interaction.

      The evidence supporting the mitotic-specific nature of the ELYS-VAPB interaction is strong, and that this interaction is direct, is also strong, and was rigorously tested using a combination of endogenous expression, heterologous expression, and recombinant protein approaches. Moreover, the sensitivity of this interaction to established mutations in VAPB abrogating FFAT interactions reinforces the outlined underlying biochemical interaction mechanism. The essentiality of ELYS phosphorylation (and therefore the mechanism underlying the mitotic specificity of the interaction) is also strongly supported by the data using phosphatase treatment. Although it is an intuitive model, whether the cell biological evidence support the simplest view that the ELYS-VAPB complex bridges the nuclear envelope to chromatin during NER in late anaphase / at mitotic exit is far less solid and, at a minimum, alternative models should be considered/discussed. For example, how a delay in metaphase in the siVAPB condition is consistent with a role in NER, which occurs exclusively post-metaphase, is unclear. Is it not possible that the VAPB-ELYS complex is regulated by phosphorylation during mitotic progression such that VAPB and/or ELYS can only exert their biological effects when released from the complex? In other words, might ELYS be licensed to act in NER only when it is released from VAPB, which could prevent premature NER/NPC biogenesis? Subtleties of when during mitosis the phosphorylation occurs is challenging, and it could be that the anaphase A to anaphase B transition, when many mitotic entry phosphorylation events begin to be reversed, could be relevant here. Along these lines, in Fig. 5 how the VAPB knock-down does or does not recapitulate the phenotype of an ELYS knock-down in this cell type (and the effect of the combination, to address epistasis) is needed for context, as is whether VAPB knock-down affects ELYS distribution in mitosis. ELYS knock-down would also be very beneficial for the PLA analysis to establish the "floor" of measurable signal. Last, it is also possible that VAPB has other roles in mitosis that should be acknowledged - for example although it is in yeast, it is relevant that a VAPB orthologue Scs2 is required for normal nuclear envelope expansion in mitosis by regulating SUMOylation (Ptak, Saik et al., JCB, 2021 and Saik et al., JCB, 2023) - this work should be referenced as well. Of course, the ideal experiment would be one in which an ELYS knock-down is complemented with a resistant form that encodes the S to A mutations in the FFAT2 region to assess its localization and to see if it can complement the knock-out function of ELYS in post-mitotic NPC assembly or, as suggested by a sequestration model, it can drive the same metaphase delay seen upon VAPB knock-down. This is technically challenging for sure, particularly given the size of the ELYS gene, but it would address the cell biological function of this interaction in the most direct manner. Several other observations that could warrant further comment or study include 1) is there a VAPB signal at the metaphase poles as suggested by Fig. 4A and, if so, could this represent aa distinct mitotic function?; 2) Does the HA-VAPB KD/MD mutant localize differently in mitosis compared to the WT - it appears that it might be less enriched in non-core regions (Fig. 4E)?; 3) does VAPB alter post-mitotic NPC biogenesis/number?

      Minor comments:

      I would suggest avoiding the use of "novel" when describing newly assembled NPCs or post-mitotic nuclear envelope reformation, as its other meaning of "non-standard" makes this wording confusing. It is unclear whether when the authors state that ELYS localizes "to the nuclear side of the nuclear envelope" they are referring to the nuclear aspect of the NPC and/or a separate pool at the INM - please edit to clarify. More descriptive y-axes for the plots in Fig. 4F and 5F and related legends would be useful; although the details are in the methods section, it would be nice not to have to hunt them down. Also, please clarify the meaning of blue and orange points in Fig. 4F.

      Significance

      General assessment: The biochemical analysis is rigorous and compelling and establishes the mitotic-specific interaction of VAPB and ELYS including detailed information about the binding interface and its regulation by phosphorylation. The new insight provided into the function of this VAPB-ELYS interaction is somewhat less well developed as the current manuscript, in its current form, does not yet mechanistically define the function of the VAPB-ELYS interaction in mitosis.

      Advance: Conceptually, to the best of my knowledge, the idea that VAPB contributes to mitosis in mammalian cells is novel and is therefore impactful and will motivate further work. As the authors connect VAPB biochemically to ELYS, an established factor that promotes the coordination of NER and NPC biogenesis, this interaction is likely to be mechanistically important, although the specific details by which this interaction facilitate normal mitotic progression is not yet clear.

      Audience: This work will be of interest to a broad swath of cell biologists including those interested in NPCs, the nuclear envelope, the ER, membrane remodeling, and chromosome segregation.

      My expertise is in nuclear envelope dynamics, nuclear pore complexes, and chromatin organization.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary

      The VAP proteins are well established as tail anchored proteins of the ER membrane. VAPs mediates co-operation between the ER and other organelles by creating a transient molecular tether with binding partners on opposing organelles to form a membrane contact site over which lipids and metabolites are exchanged. Proteins which bind VAPs generally contain a short FFAT motif, of varying sequence which binds the MSP domain of VAP. More recently the FFAT motif has been more extensively analysed in multiple different proteins and differential phosphorylation of the FFAT motif has been shown to either enhance or block VAP binding depending on the position of the phosphosite.

      Recent work conducted by the authors demonstrated that a small population of VAPB is not exclusively localised to the ER and can also reach the inner nuclear membrane. They also identified ELYS as a potential interaction partner of VAPB in a screening approach. ELYS is a nucleoporin that can be found at the nuclear side of the nuclear envelope where it forms part of nuclear pore complexes. During mitosis, ELYS serves as an assembly platform that bridges an interaction between decondensing chromosomes and recruited nucleoporin subcomplexes to generate new nuclear pore complexes for post-mitotic daughter cells. In this manuscript, James et al seek to explore this enigmatic potential interaction between ELYS and VAPB to address why VAPB may be found at the inner nuclear membrane.

      Peptide binding assays and some co-immunoprecipitation experiments are used to demonstrate that interactions occur via the MSP-domain of VAPB and FFAT-like motifs within ELYS. In addition, it is demonstrated that, for the ELYS FFAT peptides, the interaction is dependent on the phosphorylation status of serine residues of a particular FFAT-motif that can either promote or reduce its affinity to VAPB. Of most relevance is a serine in the acidic tract (1314) which, when phosphorylated increases VAPB binding. This is completely in line with what is already known about the FFAT motif and so is not surprising, in particular when using a peptide in an in vitro assay.

      The authors then utilise cell synchronisation techniques to provide evidence that both phosphorylation of ELYS and its binding to VAPB are heightened during mitosis. Immunofluorescence and proximity ligation assays are used to demonstrate that the proteins co-localise specifically during anaphase and at the non-core regions of segregating chromosomes.

      The manuscript is concluded by investigating the effect of VAPB depletion on mitosis with some evidence to suggest that transition from meta-anaphase is delayed and defects such as lagging chromosomes are observed.

      Major comments

      Overall, this manuscript is well written and the data presented in Figures 1-3 convincingly show the nature of the interaction between ELYS and VAPB. Clearly the proteins interact via FFAT motifs and this interaction appears to be enhanced during mitosis. However, the work as is, relies heavily on peptide binding assays and would benefit from additional experiments to further support the results. The authors need to more clearly show that this specific phosphorylation happens during mitosis, they may have this data but it is not clearly explained. In addition, the data that VAPB-ELYS interaction contributes to temporal progression of mitosis (as per the title) is not sufficiently clear. VAPB silencing appears to have some impact on mitosis but this is not the same thing. So this section needs to be strengthened before this statement can be made.

      The authors claim that the study "suggests an active role of VAPB in recruiting membrane fragments to chromatin and in the biogenesis of a novel nuclear envelope during mitosis". Given the data presented in Figures 4 and 5, this appears to be rather speculative with little evidence to support it, so data should be provided or this statement toned down. Currently, without additional supporting data the authors may wish to revise the overarching conclusions of the study and change the title.

      Specific points.

      Peptide pull down assays clearly show which FFAT-like motifs are important in facilitating binding. The co-immunoprecipitation systems used in Figure 2 also provide useful information on the interaction in a cell context. The authors should combine these findings by introducing full length ELYS mutants with altered FFAT-like motifs into their stably expressing GFP-VAPB HeLa cell line and then performing Co-IPs to help identify which FFAT motif/s drive the mitotic interaction. Other mutants of ELYS harbouring either phosphomimetic or phospho-resistant residues may also be introduced to further investigate mechanisms of the molecular switch in a cellular environment to support the work currently done with peptides alone. This is an obvious gap in the work which, based on the other data the authors have shown, should presumably be straightforward and would also lead directly into the next major point.

      • Whilst silencing VAPB does appear to delay mitosis, no reference is made to ELYS throughout Figure 5 nor as part of its associated discussion. Given that VAPB has more than 250 proposed binding partners, the observed aberration of mitotic progression could result from a huge number of indirect processes. Further work is needed to link the experiment specifically to the VAPB-ELYS interaction and not just loss of VAPB. We would suggest generating a complementation system where ELYS is either knocked out or silenced and then wild-type ELYS and an ELYS FFAT mutant (which cannot interact with VAPB),and/or a phospho mutant (whose interaction cannot be regulated during mitosis) are introduced. Then the observed effects can be better attributed to the VAPB-ELYS interaction and not just loss of VAPB.
      • The immunofluorescence and PLA results in Figure 4 could be strengthened by including other ER markers. This would show that co-localisation of ELYS at the non-core region is specific to VAPB protein, not any ER protein or rather than an artefact of the ER being pushed out of the organelle exclusion zone during mitosis and therefore 'bunching' at the periphery of the nuclear envelope. It would be worthwhile repeating these experiments with candidates such as VAPA, other ER membrane proteins or at least GFP-KDEL, to make this phenomenon more convincing. As part of this the authors should ideally generate a complemented ELYS KO (see point above) to avoid the residual activity attributed to endogenous background in the PLA Figure 4E.
      • Authors should clarify if the phosphorylation events (in particular S1314) only occur or are increased during mitosis. This may be data they have from the MS experiment in Figure 3 or it could also be shown using a phospho-antibody (although this can be challenging if a suitable antibody cannot be made).
      • The authors should clarify why they need to do these semi in-vitro assays with purified GST-VAPB-MSP on beads and then lysates added and not just a standard co-IP. If this is simply signal intensity due to a very small proportion of VAPB binding to ELYS then this is fine but this should be stated and it should be made clear that ELYS is not a major binding partner - most of VAPB is on the ER. Otherwise, this is misleading.

      I estimate that the suggested alterations above would incur approximately 3-6 months of additional experimental work, depending on if KO cell lines were required.

      Minor comments

      • To show that the observed interactions and potential role of VAPB-ELYS interaction is universal it would be useful to have at least a subset of experiments also shown in another cell line or system - this is now also a requirement for some journals.
      • Consider re-wording the title of the manuscript to better reflect the data presented within the study. Alternatively, provide further evidence that VAPB-ELYS interactions directly affect temporal progression of mitosis to validate this claim, as discussed above.
      • Quantification of blots in Figure 2A could allow measurement of relative binding affinities between VAPB-ELYS throughout the cell cycle. The same could be applied to the effect of phosphorylation on binding affinity in Figure 2D.
      • The cells used are never clearly mentioned in the text - I assume this is always in HeLa but this should be added in all cases for clarity
      • Page 8: "As shown in Fig. 2A,a large proportion of GFP-VAPB was precipitated under our experimental conditions." - I don't understand how this is shown in this figure as the non-bound fraction is not shown?
      • Please provide some controls to demonstrate the extent to which the samples used are asyn, G1/M or M.
      • Page 9 - why are Phos-tag gels not shown as this would make this result more convincing?
      • Figure 3A - I find the SDS-PAGE gel confusing. Why not show the whole gel and why is the band size apparently reduced in the mitotic fraction when previously it was increased (by phosphorylation)? It would also be useful to see if there were any other band shifts.
      • "FFAT-2 of ELYS is regulated by phosphorylation" The way you have setup the experiment leads the reader to think you are going to show which sites are differentially phosphorylated in mitosis, but then this is not the case - so there seems no purpose to doing the experiment this way. If you used TMT MS approach you would be able to potentially quantify the change in phosphorylation at the FFAT motif sites in mitosis. Otherwise what is the purpose of using these 2 samples, mitotic and AS?
      • For all of the antibodies used, in particular for the PLA, please provide evidence of validation of the antibodies.
      • Just a minor point to consider - In the methods for your lysis buffer you use 400mM NaCl - might this slightly reduce the VAPB-FFAT interaction? Worth considering reducing this?
      • "The rather small difference observed between the wild-type and the mutant protein observed in this experiment probably results from the presence of endogenous VAPB in the stable cell lines, which could form dimers with the exogeneous HA-tagged versions." If this is the case then please demonstrate that this is happening, or use the KO approach in the major points above.
      • "we now show that the proteins can indeed interact with each other, without the need for additional bridging factors (Figs. 1 and 3)." You show that the peptides can bind - but this is not the same thing as the peptide in the full context of the protein - so this should be toned down or removed.
      • "Remarkably, this region is highly conserved between species, suggesting that it is important for protein functions (data not shown)". Please show the alignments so the reader can judge for themselves. It is conserved in ALL species and the phosphosites are also conserved??
      • "In our experiments, knockdown of VAPA alone did not lead to a delay in mitosis (data not shown). " Why not show this data - as this is a very interesting and potentially important observation? Also add the validation of knockdown of VAPA.
      • I find the end to the discussion to the paper rather abrupt. It would be interesting to discuss further how VAPB, but not apparently VAPA reaches the INM and if so why this function is required of an ER adaptor and not another more obvious adaptor protein. In short - why would VAPB be performing this role?

      Referees cross-commenting

      I agree with the comments of the other reviewers, and they are very much in line with my own review. We all seem convinced that VAPB binds ELYS via a pFFAT, and that this interaction is enhanced during mitosois. However the role of this interaction in mitotic progression remains unclear and based on this data should not be claimed in the title or discussion of the paper.

      Significance

      Overall, if the manuscript could be improved with the suggested changes, then this could be a considerable conceptual advance in how we understand the VAP proteins, showing functions beyond those as an ER adaptor. This would be significant for the field.

      In the context of the existing literature the work does not advance our knowledge of FFAT-VAP interactions, this has already been shown, but it would give a nice example of how this can be regulated during mitosis and how VAP can contribute beyond just as an ER adaptor at membrane contact sites.

      There would be a wide audience in the cell biology field and more widely as mutations in VAPB cause a form of ALS, and many people are working in this area.

      My field of expertise is in organelle cell biology and membrane contact sites.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      The authors do not wish to provide a response at this time.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      This work uses state-of-the art cell imaging and careful image quantification to study early secretory pathway dynamics during budding yeast gametogenesis. The work builds on previous findings of roles for Sec16 in ER exit site formation, Sec-body formation during specific developmental stages, and for Vps13 in lipid homeostasis. The work appears to be carefully conducted and is nicely presented.

      Much of the early work - here relating to meiosis in budding yeast, reflects that from studies of mitosis in other systems. This work therefore adds nicely to our understanding of membrane dynamics during cell division. Figures 1 and 2 are useful additions to the literature in this regard. I would have preferred images to be presented in magenta/green rather than red/green for wider accessibility.

      The advance is therefore not conceptual but functional. It would, in my view, be unfair to dismiss this as incremental.

      Major

      My major comment here relates to the FRAP data - the difference in half-life of recovery is clear but there are also substantial differences in the immobile fraction. It is vital that this is expanded on and discussed - it has direct relevance to the conclusions relating to the ongoing functional activity of ERES and the comparison to Sec bodies. Is it not possible that the immobile element here is a functional "reserve" like with Sec bodies? This might be consistent with multiple pools of COPII proteins acting at different stages to maintain then promote secretory activity. Some consideration needs to be given to expanding this and possibly including these data in the main figure. Further analysis and controls should also be included here, other COPII proteins and other markers that one might predict would not alter dynamics in these conditions.

      The key mechanistic advance in the manuscript relates to the role of Gip1 with clearly defined outcomes showing its role in ERES remodelling in nascent spores, regeneration of the Golgi and PSM elongation. The context of this part of the work is most important. Specifically, the comparison to VPS13 mutant needs to be expanded on and better explained. The analysis in Fig.4 needs a clearer explanation within the figure of how localization to the PSM is defined. The detail in the methods is also insufficient and the "custom R script" should be published with the work (or on a publicly accessible repository such as Git/Zenodo etc).

      The development of this work with the delta-sep mutant gives useful insight and the analysis of Sec16 does indeed support a model where this is an early marker for the process. Despite the link to septins no direct analysis of YSW1 is included (which suppresses the sporulation defect in gip1 ts alleles.

      Minor:

      Figure 3 introduces new data on reticulons and their impact on ER membrane shape. Again, this reflects findings in other systems but does not add much to the specific narrative of this story but is useful for those in the field. Similar to this, the data on Sec4 are of interest to the specialist but add little to the overall story.

      The discussion is quite lengthy and speculative dealing with themes and ideas that are not addressed directly by this work. My comments on the FRAP data relate directly to the models in Figure 8 and this discussion. Given the emphasis on nutrient starvation in the final discussion more detail is needed on the relative experimental conditions used here and in flies/mammals.

      Some relevant prior work should also be cited e.g. on the role of Sec16 on exit from mitosis PMID: 21045114, other work relating to gip1 mutants (PMID: 19465564).

      Consider presenting images as magenta/green.

      Referees cross-commenting

      I agree broadly with the other reviewers comments.

      While there are elements that could be developed much further. I am not familiar with the role of GIP1 in transcriptional regulation - is this from work in yeast or solely Arabidopsis (is GIP1 here - GBF1 interacting protein, a true equivalent?).

      I agree with the comments on the need for further - and well explained - statistical analyses.

      Significance

      Overall, the work is solid and adds nicely to our understanding. It is likely to be of most interest to a quite specialist audience. The work on PSM formation and spore formation is a clear advance with significant sections of the work being of interest to a wider audience working on early secretory pathway (notably COPII dynamics). Deeper mechanistic insight is missing but non-trivial. More depth would be added by studying further deletion mutants but I am not entirely convinced that this will rapidly advance the field further than this current presentation.

      My expertise is in early secretory pathway function.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Suda et al. conducted an in-depth investigation of gametogenesis in budding yeast, focusing on the formation of the prospore membrane (PSM) through membrane trafficking rearrangement. They made an interesting observation that the number of endoplasmic reticulum exit sites (ERES) fluctuates during PSM formation, transitioning from decreasing to increasing. The study proposed that ERES regeneration, facilitated by protein phosphatase-1 and its specific subunit, Gip1, plays a crucial role in this process. However, the mechanism by which Gip1 regulates ERES numbers remains unclear, and the authors primarily used Gip1 mutants that may affect transcriptional regulation through Glc7, raising concerns about potential indirect effects. It is essential for the authors to experimentally validate the key mechanisms underlying their findings to strengthen their conclusions.

      Major Points:

      1. The conclusion that the loss of ERES causes a transient stall in membrane trafficking and leads to Golgi loss is based on the phenotype of GIP1 KO and SED4 KO cells. However, how Gip1 regulates ERES numbers remains unclear. The authors need to define whether Gip1 mediates this regulation through Glc7 dephosphorylation or via transcriptional regulation.
      2. The claim of ERES fluctuation during gametogenesis lacks statistical validation (Figure 1D). Since the difference is very small, the authors should perform a statistical analysis to determine if there is a significant difference in ERES numbers during different stages of gametogenesis.
      3. The conclusion regarding the loss and regeneration of the Golgi apparatus is based on qualitative observations of Mnn9, Sys1, and Sec7 signals. A quantitative analysis is necessary to strengthen these findings, as some cells may retain these signals despite their disappearance in representative images.
      4. Based on phenotypic similarity between GIP1 KO and SED4 KO cells, they concluded that Gip1 regulates the ERES number required for PSM expansion. They demonstrated that the number of ERES and Golgi dramatically decreased in GIP1 KO cells. The authors also need to do this experiment in SED4 KO cells? Since Sed4 affects ER function in general, the authors should demonstrate that SED4 KO cells are appropriate to make a conclusion about ERES regulation and PSM expansion.

      Referees cross-commenting

      Consistent with the other two reviewers, we feel our comments should be addressed prior to publication of this manuscript.

      Significance

      Overall, the study presents a high-quality imaging analysis of gametogenesis in budding yeast. However, the authors should experimentally validate the mechanisms underlying ERES regulation by Gip1 and conduct rigorous statistical analyses to support their observations. Additionally, since gametogenesis and Gip1 are yeast specific, the significance of this study might be limited.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary: Yeast gametogenesis requires major membrane reorganization to ensure proper spore formation for survival during starvation, but many questions remain for how this process occurs. This current manuscript by Suda et al. uses fluorescence live imaging to visualize the dynamics of secretory pathway components which are critical for contributing lipids to the prospore membrane (PSM) in the developing spore. The authors find that ER exit sites (ERES) initially decrease and then gradually increase, coinciding with their appearance inside the PSM, suggesting that new lipids for PSM growth are trafficked through the secretory pathway from within the PSM. By screening through known genetic mutants that cause meiosis defects, the authors identify Gip1p, an adaptor for protein phosphatase 1, as a master upstream factor for prospore-associated ERES formation. Interestingly, the authors additionally identify a non-essential component of ERES in vegetative cells, sed4, to be important for sporulation and ERES PSM localization.

      Overall, the biological question is interesting and the imaging quality is appropriate. The general conclusion that ERES foci localize inside developing PSMs in a gip1- and sed4- dependent manner is supported by the data. However, the manuscript is purely descriptive; not much molecular insight is gleaned into how secretory pathway components localize to the inside of the PSM, nor is it clear how important this localization is in contributing new lipids to the PSM. Additionally, there are multiple points within the writing and presentation of the results, some specified below, that require clarification; more details in the quantifications also need to be included to ascertain whether the data robustly support the authors' current conclusions.

      Major comments:

      • The loss of gip1 affects multiple aspects of sporulation and leads to an early termination of spore formation, giving little insight into how ERES are established inside the PSM. The most intriguing result is that loss of sed4, a nonessential paralog of the membrane-bound Sar1 GEF, leads not only to sporulation defects, but also affects the localization of Sec13/ERES to the spores. Given that some spores still form in the sed4 cells, more experiments detailing ERES and golgi localization within the forming spores could be done. Does the golgi no longer localize within the PSM in sed4 cells? Is there are a PSM size difference between those that do and do not have ERES foci in this genetic background? Where does Sed4 localize in gip1 cells?
      • While Vps13 is introduced as an additional pathway for supplying lipids, this manuscript does not address the relative contribution of vesicular trafficking versus vps13 lipid transport in PSM formation. Where does Vsp13 localize in the sed4 cells? Are they enriched around/within those spores that do form?
      • The clarity of writing in the results and discussion section could be improved, some of which I point out below. The discussion could also be shortened.

      Specific comments:

      • For all quantifications, more information is necessary, including sample sizes, mean/median values, and number of biological replicates. It may be helpful to include these values in a separate supplemental table.
      • Relatedly, for 1D, 3C, and 4C graphs: It is difficult to judge whether the changes of ERES # are significantly different across the various genetic backgrounds as displayed, and given the large spread and small changes, statistical analyses are required to make such conclusions. Could the authors comment on why there is a minor yet noticeable percent of cells with very high ERES numbers?
      • To make specific conclusions that ERES 'regenerate' inside PSMs, more detailed quantifications of ERES foci # inside the developing prospores should be included, with appropriate statistical analysis.
      • Figure 6A, B: The localization of Glc7 does not look different to me, as claimed. The septin-like cable localization presumably occurs during elongation, as seen in 6A, and gip1D cells do not enter this phase, then it should be expected that there would be no septin-like localization. In 6B, the lower panels seem to show mature, closed PSMs; can the authors label the phases and explain why this is?
      • Figure 8, Top panels, indicate the purple coverage is PSM. It is unclear why the authors suddenly say that ERES are 'transiently inactivated' here and in the discussion to describe the lower # of ERES foci, whereas the appearance of PSM-associated ERES foci is considered 'regenerated' (which implies de novo assembly). In general, from the present data, one cannot conclude inactivation vs. formation/regeneration, so some caution in terminology is warranted.

      Minor comments:

      • A schematic showing the different stages of meiosis and of PSM formation would be useful.
      • Scale bar dimensions are missing for most of the figures.
      • It may be helpful to use an alternate color combination for merged images (i.e. cyan/yellow, red/cyan, or magenta/green), to accommodate colorblind readers.
      • For Figure 1C, authors should show orthogonal views along the z plane at timepoint 8 to show that ERES foci are indeed inside the PSM.
      • Figure 1E legend, define closed arrowhead; additionally, include an explanation in the main text of what the Spo20(51-91) marker is.
      • For kymograph displays (Figures 1E, 5A, 7E), please include time points in each frame.
      • Figure 4D, 4E legend, the multiple terms describing PSM circumference length is confusing: 'cell perimeter, 'PSM perimeter' and 'PSM length'. Please choose one term and describe this fully in the text.
      • Fig 5: The images in Fig 5 are dim and the gain should be adjusted accordingly. Figure 5B, is this an intensity trace of one punctum, or punctae from multiple cells as implied in the text?
      • More details, not just references, in methods for sporulation induction and image analysis should be included.

      Referees cross-commenting

      I also agree that our comments should be addressed prior to publication. Of the existing data, the need for further statistical analysis is a high priority.

      Significance

      The data and conclusions presented here are for a specialized, basic audience interested in yeast meiosis, especially focused on how membranes and the secretory pathway are remodeled during this process. The paper's results have some implications for the reproductive aging field, but this area is not directly investigated in this current manuscript. The paper uses mostly established organelle markers and gene mutants previously known to be involved. The finding of Sed4's involvement in sporulation is, to my knowledge, novel and intriguing.

      Reviewer Expertise: organelle morphology, the secretory pathway, protein aggregation, stress responses, aging, fluorescence microscopy, yeast, C. elegans, mammalian cells, biochemistry

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      I - General criticisms

      Reviewer #1: My main criticism is unfortunately inherent to the approach: comparative studies are absolutely critical, but they can only provide a very sparse sampling of diversity. Fortunately, thanks to high-throughput sequencing, bioinformatic analyses can now be performed on a large number of species, but experimental validation is typically restricted to two or three species. The consequence of this for the present manuscript is that while the functional conservation of the Gwl site is convincingly shown, the exact mechanisms responsible for the reduced effect of PKA phosphorylation remain relatively vaguely defined. Indeed, in their Discussion the authors list a number of experimental approaches to address this - but I understand that these would all involve substantial efforts to address. In particular, testing chimeric constructs around the consensus PKA site and from multiple species could be very informative.

      We completely agree with the reviewer that comparative approaches are critical to understanding biological mechanisms, and are excited by the increasing possibilities to perform not only sequence and descriptive comparisons but functional studies across a range of emerging model organisms. We hope that more and more researchers in cell and molecular biology will profit from experimental tools and techniques now available in such species, and to pioneer new ones. Of course, and he/she rightly points out, conclusions are currently limited by the number of species studied, but comparisons between two judiciously chosen species can already be very informative. Thus, in our study, the use of Xenopus and Clytia allowed us to make significant progress towards our main objective of understanding the cAMP-PKA paradox in the control of oocyte maturation; specifically by showing both that PKA phosphorylation of Clytia ARPP19 is lower in efficiency and that the phosphorylated protein has a lower effect on oocyte maturation than the Xenopus protein. As the reviewer points out, unravelling the exact mechanisms underlying these differences will require a large amount of additional work and is beyond the scope of the current study. Actually, we have embarked on several series of experiments to this end using some of the approaches listed in the Discussion. Specifically, we are testing the biochemical and functional properties of chimeric constructs containing the consensus PKA site from various species. This is a substantial undertaking which will require one to two years to complete, but is already giving some very interesting findings.

      Reviewer #1: The figures and text could be slightly condensed down to about 6 figures.

      We have reduced the number of figure panels but we prefer to maintain the number of figures, because the experimental data presented in them is essential to the interpretation of our results and the overall conclusions of the article. If the journal editor would like us to reduce the number of figures, we could do this by displacing Figure 4 and some panels of other figures (to then fuse some of them) to supplementary material, but this would be a pity.

      ____________II - Abstract

      As recommended by Reviewer #2, we have reworked the Abstract to make it more accessible to new readers, attempting to bring out more clearly and simply the main results and conclusions of the study. We correspondingly simplified and shortened the title of the article. Changes: Page 2.

      ____________III- Introduction points

      Reviewer #2: I believe that it would be interesting to include some time-references when introducing the prophase arrest of Clytia and Xenopus oocytes. How long is prophase arrest in Xenopus compared to Clytia or other organisms? How can this affect the prophase arrest mechanisms? It seems that the prophase arrest in Xenopus oocytes is found to be significantly more prolonged compared to Clytia and various other organisms, and also meiotic maturation proceeds much more rapidly in Clytia than in Xenopus. This should be indicated in the introduction with a short introduction of why, and not others, were these species chosen for this study.

      Differences in timing of oocyte prophase arrest and in maturation kinetics across animals are indeed highly relevant in relation to the underlying biochemical mechanisms. Unfortunately, not enough information is currently available concerning the duration of the successive phases of oocyte prophase arrest across species to make any meaningful correlations with PKA regulation of maturation initiation. We have nevertheless expanded the Introduction to cover this issue as follows:

      • We start the introduction by mentioning how the length of the prophase arrest varies across species. Changes: Page 3, lines 5-11.
      • We have added examples of species which likely have similar durations of prophase arrest but show cAMP-stimulated vs cAMP-inhibited release. Changes: Page 4, lines 28-35.
      • We have specified the temporal differences in meiotic maturation in Xenopus (3-7 hrs) and Clytia (10-15 min). Changes: Page 5, lines 32-33.

      Reviewer #2: why, and not others, were these species [Xenopus, Clytia] chosen for this study. A brief justification is included in lines 1-page 5 "..a laboratory model hydrozoan species well suited to oogenesis studies", but it does not explain why this and not other hydrozoan species like Hydra, that has also been used for meiosis studies.

      As requested by Reviewer #2, fuller details are now included about the advantages of Clytia compared to other hydrozoan species, citing several articles and recent reviews here and also in the Discussion. Changes: Page 5, lines 21-32 & 37-39.

      Hydra is a classic cnidarian experimental species and has proved an extremely useful model for regeneration and body patterning, but is not suitable for experimental studies on oocyte maturation because spawning is hard to control and fully-grown oocytes cannot easily be obtained, manipulated or observed. In contrast many hydromedusae (including Clytia, Cytaeis, and Cladonema) have daily dark/light induced spawning and accessible gonads, so provide great material for studying oogenesis and maturation. Of these, Clytia has currently by far the most advanced molecular and experimental tools.

      Reviewer #2: The proteins MAPK is not introduced properly, as it is first mentioned in the results section in line 12. Given the importance of the results provided with it, it should be presented in the introduction prior to the results section.

      As requested by Reviewer #2, the involvement of MAPK activation during Xenopus oocyte meiotic maturation is now introduced, explaining how its phosphorylation serves as a marker of Cdk1 activation. Changes: Page 5, lines 1-5.

      Reviewer #2: These sentences need a more elaborate explanation: Page 4 Lines 16-17 "... no role for cAMP has been detected in meiotic resumption, which is mediated by distinct signaling pathways" Which pathways?

      We now give the example of the well-characterized pathway Gbg-PI3K pathway for oocyte maturation initiation in the starfish. Changes: Page 4, lines 1-15.

      Reviewer #2: Page 4 line 34-39. Introduction indicates that the phosphorylation of ARPP19 on S67 by Gwl is a poorly understood molecular signaling cascade (line 34). However, the positive role of ARPP19 on Cdk1 activation, through the S67 phosphorylation by Gwl, appears to be widespread across all eukaryotic mitotic and meiotic divisions studied (lines 36-37). These two sentences seem a little contradictory. If the general pathway has been identified but the signaling cascade is still not well described, please indicate that in a clearer way.

      We apologise that the wording we used was not clear and implied that the mechanisms of PP2A inhibition by Gwl-phosphorylated ARPP19 were poorly understood. On the contrary, they are very well studied. The part that remains mysterious concerns the upstream mechanisms. We have reworded the paragraph to make this point unambiguous. Changes: Page 5, lines 1-8.

      ____________IV - Results

      Reviewer #2: The text of the results is generally well described; however, all the sections start with a long introductory paragraph. I believe this facilitates the contextualization of the experiments, but please try to summarize when possible. For example, in page 5 lines 12-25, or page 7 lines 30-37, are all introduction information.

      As requested by Reviewer #2, we have shortened or removed the introductory passages of the Results section paragraphs, which were redundant with the information given in the introduction. We did not restrict to the two examples cited by the reviewer, but have shortened all the Results passages that repeat information already provided in the Introduction. Changes: Page 7, lines 3-4 & 14-16 & 36-37 - Page 8, lines 12-15 - Page 8, lines 37-40 & Page 9, lines 1-6.

      Reviewer #2: Page 7, Lines 14-19 present a general conclusion of the findings explained in lines 20-27. I think these results are important and they should be explained better, in my opinion they are slightly poorly described.

      We have followed the reviewer's recommendation. The explanation of the experiments and the results are more detailed and the paragraph ends with a general conclusion which came too early in the previous version. Changes: Page 8, lines 22-24 & 32-34.

      Reviewer #2: Page 8, lines 16-17: "It was not possible to increase injection volumes or protein concentrations without inducing high levels of non-specific toxicity". What are the non-specific toxicity effects? How was this addressed? What fundaments this conclusion?

      Clytia oocytes are relatively fragile. Sensitivity of oocytes to injection varies between batches, while in general increasing injection volumes or protein concentrations increases the levels of lysis observed. We do not know exactly what causes this but lysis can happen either immediately following injection or during the natural exaggerated cortical contraction waves that accompany meiotic maturation, suggesting that it relates to mechanical trauma. We have expanded this paragraph and the legend of Fig. 3C to explain these injection experiments more fully in the text and to clarify these issues. Changes: Page 9, lines 16-29 - Page 32, lines 34-41 & Page 33, lines 1-11 - Supplementary Table 1.

      Same paragraph: Lines 25-27 of page 8. Text reads, "These results suggest that PP2A inhibition is not sufficient to induce oocyte maturation in Clytia, although we cannot rule out that the quantity of OA or Gwl thiophosphorylated ARPP proteins delivered was insufficient to trigger GVBD.". Please provide evidence if higher concentrations of OA or Gwl were tested to state this conclusion.

      As explained above, we could not increase the concentrations of ARPP19 protein beyond 4mg/ml. It is important to note that at the same concentration, both Clytia and Xenopus proteins induce activation of Cdk1 and GVBD in the Xenopus oocyte.

      Concerning OA, it is well documented in many systems including Xenopus, starfish and mouse oocytes as well as mammalian cell cultures, that high concentrations lead to cell lysis/apoptosis as a result of a massive deregulation of protein phosphorylation (Goris et al, 1989; Rime & Ozon, 1990; Alexandre et al, 1991; Boe et al, 1991; Gehringer, 2004; Maton el al, 2005; Kleppe et al, 2015). Specific tests in Xenopus oocytes, have shown that injecting 50 nl of 1 or 2 mM OA specifically inhibits PP2A, while injecting 5 mM also targets PP1 and higher OA concentrations inhibit all phosphatases. For these reasons, we did not increase OA concentrations over 2 mM. When injected in Xenopus oocyte at 1 or 2 mM, OA induces Cdk1 activation, GVBD but then the cell dies because PP2A has multiple substrates essential for cell life. When injected at 2 mM in Clytia oocytes, OA does not induce Cdk1 activation nor GVBD but promotes cell lysis. This supports the conclusion that 2 mM OA is sufficient to inhibit PP2A (and possibly other phosphatases) but that PP2A inhibition is not sufficient to induce oocyte maturation in Clytia.

      We have reworded the relevant text to make these points clearer. The previous statement that “we cannot rule out that the quantity of OA or Gwl thiophosphorylated ARPP proteins delivered was insufficient to trigger GVBD” has been removed because it was unnecessarily cautious in the context of the literature cited above, as now fully explained_._ Changes: Page 9, lines 31-35 - Page 32, lines 34-41 & Page 33, lines 1-11 - Supplementary Table 1.

      References: Alexandre et al, 1991, doi: 10.1242/dev.112.4.971; Boe et al, 1991, doi: 10.1016/0014-4827(91)90523-w; Gehringer, 2004, doi: 10.1016/s0014-5793(03)01447-9; Goris et al, 1989, doi: 10.1016/0014-5793(89)80198-x; Kleppe et al, 2015, doi: 10.3390/md13106505; Maton el al, 2005, doi: 10.1242/jcs.02370; Rime & Ozon, 1990, doi: 10.1016/0012-1606(90)90106-s

      Reviewer #2: Lines 12-13: the sentence "This in vitro assay thus places S81 as the sole residue in ClyARPP19 for phosphorylation by PKA." is overstated. As not all residues had been tested, please indicate that "it is likely that" or "among the residues tested", in contrast to "the sole residue in ClyARPP19".

      We realise that we had not explained clearly enough how the thiophosphorylation assay works. In this assay, γ-S-ATP will be incorporated into any amino acid of ClyARPP19 phosphorylatable by PKA. The observed thiophosphorylation of the wild-type protein, demonstrates that one or more residues are phosphorylated by PKA. This thiophosphorylation was completely prevented by mutation of a single residue, S81. This experiment thus shows that S81 is entirely responsible for phosphorylation by PKA in this assay. We have rewritten this section more clearly. Changes: Page 10, lines 18-28.

      ____________V - Figures and text related to the figures

      Figure 1A

      Reviewer #2: Why is mouse not included in Figure 1A? Although it might be very similar to human, given that mouse is the species that is most commonly use as a mammalian model, I believe it could be included. However, this is optional upon decision by the authors.

      We have replaced the human sequence in Figure 1A with the mouse sequence as suggested. The sequences of each of the mouse and human ENSA/ARPP19 proteins are indeed virtually identical across mammals. Changes: Fig. 1A.

      Figure 1C

      Reviewer #2: There should be a better explanation in the text of the results sections for the image included in in Fig1 C. Note that Clytia is not a commonly used species, therefore images should be properly explained for general readers. Please indicate in the text that ClyARPP19 mRNA is expressed in previtellogenic oocytes and not in vitellogenic, plus any additional information needed to understand the image. In addition, the detection of ARPP19 in the nerve rings is intriguing. This is mentioned in the discussion section, any idea of its function there? Please include some additional information or additional references, if they exist.

      We have expanded the explanations of Fig. 1C in the text and in the figure legend. We have also added cartoons to the figure to help readers understand the organisation of the Clytia jellyfish and gonad. As now explained, ClyARPP19 mRNA is detected in oocytes at all stages, but the signal is much stronger in pre-vitellogenic oocytes because all cytoplasmic components including mRNAs are significantly diluted by high quantity of yolk proteins as the oocytes grow to full size. Changes: page 7, line 40 & page 8, lines 1-9 - Fig. 1C - Legend page 31, lines 19-31.

      Nothing is known about the function of ARPP19 in the Clytia nervous system. The only data linking ARPP19 and the nervous system concerns mammalian ARPP16, an alternatively spliced variant of ARPP19. ARPP16 is highly expressed in medium spiny neurons of the striatum and likely mediates effects of the neurotransmitter dopamine acting on these cells (Andrade et al, 2017; Musante et al, 2017). This point is included in the Discussion in relation to the hypothesis that PKA phosphorylation of ARPP19 proteins in animals first arose in the nervous system and only later was coopted into oocyte maturation initiation. Changes: page 16, lines 12-13 & 17-20 - page 19, lines 6-9.

      Figure 2A

      Reviewer #1: Fig. 2A (and similar plots in subsequent figures): is it really necessary to cut the x axis? Would it be possible to indicate the number of oocytes for each experiment (maybe in the legend in brackets)?

      As requested by reviewer #1, the x-axis is no longer cut. The number of oocytes for each experiment is now provided in the legend of Fig. 2A and in similar plots of Fig. 5A and 5D. Changes: Fig. 2A - Legends page 31, lines 37-38 (Fig. 2A), page 33, line 25 (Fig. 5A) - page 33, line 34 (Fig. 5D).

      Figure 2D-E (as well as Figure 6C-D and Figure 8B-C)

      Reviewer #1: Fig. 2D (and all similar plots below): I am lacking the discrete data points that were measured. Without these it is impossible to evaluate the fits. The half-times shown in 2E are somewhat redundant, and the information could be combined on a single plot.

      We added all the data points to the concerned plots: 2D, 6C and 8B. As recommended by reviewer #1, we combined on a single plot the phosphorylation levels and the half-times. 2D-E => 2D, 6C-D => 6C and 8B-C => 8B. Changes: Figs 2D, 6C and 8B - Legends page 32, lines 9-14 (Fig. 2D), page 34, lines 24-30 (Fig. 6C) - page 35, lines 13-18 (Fig. 8B).

      Figure 3A and 3B

      Reviewer #1: Fig. 3: why is the blot for PKA substrates cut into 3 pieces? It would be clearer to show the entire membrane.

      In western blot experiments using Clytia oocytes, the amount of material was limited so the membranes were cut into three parts. The central part was incubated sequentially in distinct antibodies. We finally incubated all three parts of the membrane with the anti-phospho-PKA substrate antibody to reveal the full spectrum of proteins recognized by this antibody. The 3 pieces in Fig. 3A therefore together make up the same original membrane. We had separated them on the figure to make it clear that the membrane had been cut. In the new presentation, the 3 pieces are shown next to each other, making it clear that all the membrane is present, with dotted lines indicating the cut zone as explained in the legend. Changes: Fig. 3A and 3B - Legend page 32, lines 22-25 (Fig. 3A), lines 30-33 (Fig. 3B) - Page 24, lines 3-6 (Methods).

      Figure 3C

      Reviewer #2: Fig. 3C needs a better explanation in the text. The way these graphs are presented is somehow confusing. The meaning of the dots is not self-explanted in the graph, and it seems that each experiment was done independently but then the complete set of results is presented. Legend says that "each dot represents one experiment" but this is difficult to read as in every analysis the figure also indicates the average and the total number of oocytes. If authors wish so, they can keep the figure as it is, but then please explain this graph better in the text, and please include statistical analysis. These results are very robust, but a comparison between the number of oocytes that go through spontaneous GVBD of lysis in the different conditions will benefit their understanding.

      This figure is intended to provide an overview of all the Clytia oocyte injection experiments that we performed, for which full details are given in Supplementary Table 1. Since these experiments were not equivalent in terms of exact timing and types of observation (or films) made and oocyte sensitivity to injection -as ascertained by buffer injections-, it is not justified to make statistical comparisons between groups. We apologise that the presentation was misleading in this respect and hope that the new version is easier to understand. We removed from the figure the average percentage of maturation for each condition between experiments to avoid any misunderstanding of the nature of the data, and rather represent the values of each experiment independently. We also now explain the data included in the figure fully in the text and figure legend. Changes: Page 9, lines 16-39 - Fig. 3C and Supplementary Table 1 - Legend page 32, lines 34-41 & page 33, lines 1-11.

      Reviewer #2: Also, please provide in the text a plausible explanation for the cause of oocyte lysis for all experimental conditions (Fig 3C). Given that in the control experiments with buffer this effect is also observed in some oocytes, please explain if this is caused by a mechanical disruption of the oocyte during the injection. In contrast, okadaic acid induces the lysis in all the 14/14 oocytes analyzed, is this due also to the mechanical approach? Or is there other reason more related to the PP2A inhibition? Please explain.

      These points are treated above in the response to this reviewer concerning the Results section.

      Figure 5

      Reviewer #2: In Figure 5 D-F, cited in page 9 lines 35-35. Can you provide an explanation of why the time course of meiosis resumption was delayed?

      The binding partners/effectors of XeARPP19-S109D that are involved in maintaining the prophase arrest have not yet been identified. The most probable explanation of the delay in meiotic maturation induced by ClyARPP19-S109D is that Clytia protein recognizes less efficiently these unknown ARPP19 effectors that mediate the prophase arrest. As a result, maturation would be delayed, but not blocked. This explanation was provided in the Discussion (page 17, lines 14-17) and is now mentioned in the Results section. Changes: page 11, lines 16-19.

      ____________VI - Discussion

      Reviewer #2: Although it presents highly interesting suggestions, discussion may border on being overly speculative, especially from line 37 of page 15 till the end.

      We agree and have reduced the speculation in this part of the discussion, in particular regrouping and reformulating ideas about evolutionary scenarios in a single paragraph. Changes: page 17, lines 37-41 - page 18, lines 1-41 - page 19, lines 1-18.

      SUMMARY - Point by point responses to individual reviewers’ comments in their order of appearance.

      Reviewer 1

      • The figures and text could be slightly condensed down to about 6 figures.

      We have reduced the number of figure panels but we prefer to maintain the number of figures, because the experimental data presented in them is essential to the interpretation of our results and the overall conclusions of the article. If the journal editor would like us to reduce the number of figures, we could do this by displacing Figure 4 and some panels of other figures (to then fuse some of them) to supplementary material, but this would be a pity.

      • The exact mechanisms responsible for the reduced effect of PKA phosphorylation remain relatively vaguely defined. Indeed, in their Discussion the authors list a number of experimental approaches to address this - but I understand that these would all involve substantial efforts to address. In particular, testing chimeric constructs around the consensus PKA site and from multiple species could be very informative.

      As the reviewer points out, unravelling these exact mechanisms will require a large amount of additional work and is beyond the scope of the current study.

      • 2A (and similar plots in subsequent figures): is it really necessary to cut the x axis? Would it be possible to indicate the number of oocytes for each experiment (maybe in the legend in brackets)?

      Fig. 2A has been changed in line with the reviewer's request (as well as similar plots in Fig. 5A and 5D). Changes: Fig. 2A - Legends page 31, lines 37-38 (Fig. 2A), page 33, line 25 (Fig. 5A) - page 33, line 34 (Fig. 5D).

      • 2D (and all similar plots below): I am lacking the discrete data points that were measured. Without these it is impossible to evaluate the fits. The half-times shown in 2E are somewhat redundant, and the information could be combined on a single plot.

      Fig. 2D has been changed in line with the reviewer's request (as well as similar plots in Figs 6C-D and 8B-C). Changes: Fig. 2D, 6C and 8B - Legends page 32, lines 9-14 (Fig. 2D), page 34, lines 24-30 (Fig. 6C) - page 35, lines 13-18 (Fig. 8B).

      • 3: why is the blot for PKA substrates cut into 3 pieces? It would be clearer to show the entire membrane.

      In western blot experiments using Clytia oocytes, the amount of material was limited so the membranes were cut into three parts. The central part was incubated sequentially in distinct antibodies. We finally incubated all three parts of the membrane with the anti-phospho-PKA substrate antibody to reveal the full spectrum of proteins recognized by this antibody. The 3 pieces in Fig. 3A therefore together make up the same original membrane. In the new presentation, the 3 pieces are shown next to each other, making it clear that all the membrane is present, with dotted lines indicating the cut zone as explained in the legend. Changes: Fig. 3A and 3B - Legend page 32, lines 22-25 (Fig. 3A), lines 30-33 (Fig. 3B) - Page 24, lines 3-6 (Methods).

      Reviewer 2

      • Abstract needs to be simplified if wants to reach a broader range of readers.

      We have reworked the Abstract to make it more accessible to new readers. Changes: Page 2.

      • It would be interesting to include some time-references when introducing the prophase arrest of Clytia and Xenopus oocytes. This should be indicated in the introduction with a short introduction of why, and not others, were these species chosen for this study.

      We have expanded the Introduction to cover the issue of time-references. Fuller details are now included about the advantages of Clytia compared to other hydrozoan species. Changes: Page 3, lines 5-11, page 4, lines 28-35, page 5, lines 32-33, page 5, lines 21-32 & 37-39.

      • The proteins MAPK is not introduced properly, as it is first mentioned in the results section.

      The involvement of MAPK activation during Xenopus oocyte meiotic maturation is now introduced. Changes: Page 5, lines 1-5.

      • Page 4 Lines 16-17 "... no role for cAMP has been detected in meiotic resumption, which is mediated by distinct signaling pathways" Which pathways?

      We now give the example of the well-characterized pathway Gbg-PI3K pathway for oocyte maturation in starfish, also mentioning that in many species the pathways are still unknown. Changes: Page 4, lines 1-15.

      • Page 4 line 34-39. Introduction indicates that the phosphorylation of ARPP19 on S67 by Gwl is a poorly understood molecular signaling cascade (line 34). However, the positive role of ARPP19 on Cdk1 activation, through the S67 phosphorylation by Gwl, appears to be widespread across all eukaryotic mitotic and meiotic divisions studied (lines 36-37). These two sentences seem a little contradictory.

      The mechanisms of PP2A inhibition by Gwl-phosphorylated ARPP19 are very well studied. The part that remains mysterious concerns the upstream mechanisms. We have reworded the paragraph to make this point unambiguous. Changes: Page 5, lines 1-8.

      • Why is mouse not included in Figure 1A?

      We have replaced the human sequence in Figure 1A with the mouse sequence. Changes: Fig. 1A.

      • 1C: There should be a better explanation in the text of the results sections for the image included in in Fig1 C. Please indicate in the text that ClyARPP19 mRNA is expressed in previtellogenic oocytes and not in vitellogenic.

      We have expanded the explanations of Fig. 1C in the text. We have also added cartoons to the figure to help readers understand the organisation of the Clytia jellyfish and gonad. As now explained, ClyARPP19 mRNA is detected in oocytes at all stages, but the signal is much stronger in pre-vitellogenic oocytes because all cytoplasmic components are significantly diluted by high quantity of yolk proteins. Changes: page 7, line 40 & page 8, lines 1-9 - Fig. 1C - Legend page 31, lines 19-31.

      • In addition, the detection of ARPP19 in the nerve rings is intriguing. Any idea of its function there?

      The only data linking ARPP19 and the nervous system concerns a mammalian variant of ARPP19 that is highly expressed in the striatum. This point is included in the Discussion_. Changes: page 16, lines 12-13 & 17-20 - page 19, lines 6-9._

      • Figure 3C. The way these graphs are presented is somehow confusing. If authors wish so, they can keep the figure as it is, but then Also, please provide in the text a plausible explanation for the cause of oocyte lysis for all experimental conditions. please explain this graph better in the text, and please include statistical analysis.

      This figure is intended to provide an overview of all the Clytia oocyte injection experiments, for which full details are given in Supplementary Table 1. We have modified the figure and now clarified this fully in the text and figure legend. Clytia oocytes are relatively fragile. Sensitivity of oocytes to injection varies between batches, while in general increasing injection volumes or protein concentrations increases the levels of lysis observed. We do not know exactly what causes this but it probably relates to mechanical trauma. We now explain these injection experiments more fully in the text. Changes: Page 9, lines 16-39 - Fig. 3C and Supplementary Table 1 - Legend page 32, lines 34-41 & page 33, lines 1-11.

      • In Figure 5 D-F, cited in page 9 lines 35-35. Can you provide an explanation of why the time course of meiosis resumption was delayed?

      The most probable explanation is that Clytia protein recognizes less efficiently the unknown ARPP19 effectors that mediate the prophase arrest in Xenopus. This explanation is provided in the Results section. Changes: page 11, line 16-19.

      • All the sections start with a long introductory paragraph. I believe this facilitates the contextualization of the experiments, but please try to summarize when possible.

      As requested, we have shortened or removed the introductory passages of the Results section paragraphs, which were redundant with the information given in the introduction. Changes: Page 7, lines 3-4 & 14-16 & 36-37 - Page 8, lines 12-15 - Page 8, lines 37-40 & Page 9, lines 1-6.

      • Page 7, Lines 14-19 present a general conclusion of the findings explained in lines 20-27. I think these results are important and they should be explained better, in my opinion they are slightly poorly described.

      The explanation of the experiments and the results are now more detailed and the paragraph ends with a general conclusion which came too early in the previous version. Changes: Page 8, lines 22-24 & 32-34.

      • Page 8, lines 16-17: "It was not possible to increase injection volumes or protein concentrations without inducing high levels of non-specific toxicity". What are the non-specific toxicity effects? How was this addressed? What fundaments this conclusion?

      As explained above, increasing injection volumes or protein concentrations increases the levels of lysis observed due probably to mechanical trauma. But it is important to note that at the same concentration, both Clytia and Xenopus proteins induce activation of Cdk1 and GVBD in the Xenopus oocyte. Changes: Page 9, lines 16-29 - Page 32, lines 34-41 & Page 33, lines 1-11 - Supplementary Table 1.

      • Lines 25-27 of page 8. "These results suggest that PP2A inhibition is not sufficient to induce oocyte maturation in Clytia, although we cannot rule out that the quantity of OA or Gwl thiophosphorylated ARPP proteins delivered was insufficient to trigger GVBD." Please provide evidence if higher concentrations of OA or Gwl were tested to state this conclusion.

      High OA concentrations lead to cell lysis/apoptosis as a result of a massive deregulation of protein phosphorylation. For these reasons, we cannot increase OA concentrations over 2 µM. When injected in Xenopus oocyte at 1 or 2 µM, OA induces Cdk1 activation, but then the cell dies because PP2A has multiple substrates essential for cell life. When injected at 2 µM in Clytia oocytes, OA does not induce Cdk1 activation but promotes cell lysis. This supports the conclusion that 2 µM OA is sufficient to inhibit PP2A but that PP2A inhibition is not sufficient to induce oocyte maturation in Clytia. We have reworded the relevant text. Changes: Page 9, lines 31-35 - Page 32, lines 34-41 & Page 33, lines 1-11 - Supplementary Table 1.

      • Lines 12-13: the sentence "This in vitro assay thus places S81 as the sole residue in ClyARPP19 for phosphorylation by PKA." is overstated. As not all residues had been tested, please indicate that "it is likely that" or "among the residues tested", in contrast to "the sole residue in ClyARPP19".

      The observed thiophosphorylation of the wild-type protein demonstrates that one or more residues are phosphorylated by PKA. This thiophosphorylation was completely prevented by mutation of a single residue, S81. This experiment thus shows that S81 is entirely responsible for phosphorylation by PKA in this assay. We have rewritten this section more clearly. Changes: Page 10, lines 18-28.

      • Some parts of the discussion are a bit speculative.

      We have reduced the speculation in this part of the discussion, in particular regrouping and reformulating ideas about evolutionary scenarios into a single paragraph. Changes: page 17, lines 37-41 - page 18, lines 1-41 - page 19, lines 1-18.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary of the main findings of the study.

      This work presents very interesting data about the maintenance and release of the prophase arrest of oocytes during sexual reproduction. Authors approach some of the remaining questions about oocyte maturation in animals by taking a comparative approach between two species (Clytia and Xenopus) that use opposing cAMP/PKA signaling pathways to trigger oocyte maturation. To do it they focused on phosphorylation characteristics and function of the regulatory protein ARPP19 from the amphibian Xenopus and its orthologue in the hydrozoan Clytia. Results suggest that the low capacity of Clytia ARPP19 to be phosphorylated by PKA. Moreover, Clytia ARPP19 is inherently a poorer PKA substrate than Xenopus ARPP109 both in vivo and in vitro, despite the presence of a functional PKA site. In addition, the absence of functional interactors mediating its negative effects on Cdk1 activation may provide a double security allowing induction of meiosis resumption in Clytia by elevated PKA activity despite the presence of ARPP19, while additional and yet unidentified mechanisms ensure the Clytia oocyte prophase arrest.

      Minor comments: read detailed review below. Figure 1 and Figure 3 need a better explanation of the results. Abstract needs to be simplified if wants to reach a broader range of readers. Some parts of the discussion are a bit speculative.

      Overall, this work used a robust set of molecular experiments that strongly support the conclusions of the study.

      Significance

      Strengths and limitations of this work:

      The primary strength of this work lies in its innovative use of two distinct species and the integration of molecular experiments to extract conclusions from their different signaling pathways. The well-designed and executed experiments, particularly those of figures 5-9, contribute to an elaborated exploration of the topic, elucidating the underlying mechanisms with clarity. The explanation of each experiment in the results section further adds to the clarity and depth of the study.

      The abstract requires improvement, particularly from lines 10 to 21, as it becomes fully understood only after reading the entire manuscript. To make the work more accessible to new readers, it would be good to present the abstract in a more approachable manner. Figures 1C and 3C need a better explanation in the text. Additionally, some sentences would benefit from citations or further clarification in the results or discussion section. Although is presents highly interesting suggestions, discussion may border on being overly speculative, especially from line 37 of page 15 till the end.

      Detailed review

      Introduction:<br /> I believe that it would be interesting to include some time-references when introducing the prophase arrest of Clytia and Xenopus oocytes. How long is prophase arrest in Xenopus compared to Clytia or other organisms? How can this affect the prophase arrest mechanisms? It seems that the prophase arrest in Xenopus oocytes is found to be significantly more prolonged compared to Clytia and various other organisms, and also meiotic maturation proceeds much more rapidly in Clytia than in Xenopus. This should be indicated in the introduction with a short introduction of why, and not others, were these species chosen for this study. A brief justification is included in lines 1-page 5 "..a laboratory model hydrozoan species well suited to oogenesis studies", but it does not explain why this and not other hydrozoan species like Hydra, that has also been used for meiosis studies.<br /> The proteins MAPK is not introduced properly, as it is first mentioned in the results section in line 12. Given the importance of the results provided with it, it should be presented in the introduction prior to the results section.

      These sentences need a more elaborate explanation:<br /> Page 4 Lines 16-17 "... no role for cAMP has been detected in meiotic resumption, which is mediated by distinct signaling pathways" Which pathways?

      Page 4 line 34-39. Introduction indicates that the phosphorylation of ARPP19 on S67 by Gwl is a poorly understood molecular signaling cascade (line 34). However, the positive role of ARPP19 on Cdk1 activation, through the S67 phosphorylation by Gwl, appears to be widespread across all eukaryotic mitotic and meiotic divisions studied (lines 36-37). These two sentences seem a little contradictory. If the general pathway has been identified but the signaling cascade is still not well described, please indicate that in a clearer way.

      Results section: this review will first comment the figures, and then the text.<br /> Figure 1<br /> Why is mouse not included in Figure 1A? Although it might be very similar to human, given that mouse is the species that is most commonly use as a mammalian model, I believe it could be included. However, this is optional upon decision by the authors.<br /> There should be a better explanation in the text of the results sections for the image included in in Fig1 C. Note that Clytia is not a commonly used species, therefore images should be properly explained for general readers. Please indicate in the text that ClyARPP19 mRNA is expressed in previtellogenic oocytes and not in vitellogenic, plus any additional information needed to understand the image. In addition, the detection of ARPP19 in the nerve rings is intriguing. This is mentioned in the discussion section, any idea of its function there? Please include some additional information or additional references, if they exist.

      Figure 3<br /> The way these graphs are presented is somehow confusing. The meaning of the dots is not self-explanted in the graph, and it seems that each experiment was done independently but then the complete set of results is presented. Legend says that "each dot represents one experiment" but this is difficult to read as in every analysis the figure also indicates the average and the total number of oocytes. If authors wish so, they can keep the figure as it is, but then please explain this graph better in the text, and please include statistical analysis. These results are very robust, but a comparison between the number of oocytes that go through spontaneous GVBD of lysis in the different conditions will benefit their understanding.

      Also, please provide in the text a plausible explanation for the cause of oocyte lysis for all experimental conditions (Fig 3C). Given that in the control experiments with buffer this effect is also observed in some oocytes, please explain if this is caused by a mechanical disruption of the oocyte during the injection. In contrast, okadaic acid induces the lysis in all the 14/14 oocytes analyzed, is this due also to the mechanical approach? Or is there other reason more related to the PP2A inhibition? Please explain.

      Figure 5<br /> In Figure 5 D-F, cited in page 9 lines 35-35. Can you provide an explanation of why the time course of meiosis resumption was delayed?

      • The text of the results is generally well described; however, all the sections start with a long introductory paragraph. I believe this facilitates the contextualization of the experiments, but please try to summarize when possible. For example, in page 5 lines 12-25, or page 7 lines 30-37, are all introduction information.<br /> Page 7, Lines 14-19 present a general conclusion of the findings explained in lines 20-27. I think these results are important and they should be explained better, in my opinion they are slightly poorly described.

      Page 8, lines 16-17: "It was not possible to increase injection volumes or protein concentrations without inducing high levels of non-specific toxicity". What are the non-specific toxicity effects? How was this addressed? What fundaments this conclusion?

      Lines 25-27 of page 8. Text reads, "These results suggest that PP2A inhibition is not sufficient to induce oocyte maturation in Clytia, although we cannot rule out that the quantity of OA or Gwl thiophosphorylated ARPP proteins delivered was insufficient to trigger GVBD.". Please provide evidence if higher concentrations of OA or Gwl were tested to state this conclusion.

      Lines 12-13: the sentence "This in vitro assay thus places S81 as the sole residue in ClyARPP19 for phosphorylation by PKA." is overstated. As not all residues had been tested, please indicate that "it is likely that" or "among the residues tested", in contrast to "the sole residue in ClyARPP19".

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      In their present manuscript Meneau and coworkers investigate the evolutionary conserved functions of ARPP19 in regulation of meiotic maturation of oocytes. During meiotic maturation, the maturation hormone induces a signaling cascade ultimately leading to the activation of the master regulator, cdk1-cyclin B. within this signaling network, the phosphatase PP2A prevents cdk1 activation in immature oocytes. Upon the action of the maturation hormone, ARPP19 is activated through phosphorylation by the kinase Gwl, and then functions as a potent inhibitor of PP2A, thereby contributing to cdk1 activation. Additionally, ARPP19 is subject to a second layer of regulation: a second site is phosphorylated by the kinase PKA. Interestingly, in vertebrates this cAMP/PKA pathway prevents maturation, while in many other species the same pathway has an opposite effect and cAMP/PKA is indeed sufficient to drive maturation -- referred to as the cAMP paradox.

      The authors' major aim was to reveal the molecular basis of these diverse functions of ARPP19 in triggering meiotic maturation. Firstly, they show that the Gwl site is extremely well-conserved all across eukaryotes. They then functionally validate this by comparing the functions of Xenopus ARPP19 to its orthologue in the jellyfish Clytia hemisphaerica. They confirm that the jellyfish ARPP19 is phosphorylated on the conserved Gwl site in vitro and in frog and jellyfish oocytes, acting as a PP2A inhibitor and contributing to cdk1 activation. However, while this is sufficient to drive maturation in Xenopus, PP2A inhibition alone is not sufficient to trigger entry to meiosis in Clytia oocytes, indicating the existence of additional mechanisms. Secondly, they show that the PKA site exists and is phosphorylated both in Xenopus and Clytia. However, the Clytia protein appears to be a much worst substrate for PKA and other interactors, which explains why PKA-phosphorylated ARPP19 does not inhibit maturation either in jellyfish oocytes or when exogenously injected into Xenopus oocytes.

      I find the manuscript well-written and easy to follow. The experiments are carefully performed, well-controlled and well-documented. The data shown on the figures fully supports the conclusions drawn -- although the figures and text could be slightly condensed down to about 6 figures. Overall, I would highly recommend the manuscript for publication.

      My main criticism is unfortunately inherent to the approach: comparative studies are absolutely critical, but they can only provide a very sparse sampling of diversity. Fortunately, thanks to high-throughput sequencing, bioinformatic analyses can now be performed on a large number of species, but experimental validation is typically restricted to two or three species. The consequence of this for the present manuscript is that while the functional conservation of the Gwl site is convincingly shown, the exact mechanisms responsible for the reduced effect of PKA phosphorylation remain relatively vaguely defined. Indeed, in their Discussion the authors list a number of experimental approaches to address this - but I understand that these would all involve substantial efforts to address. In particular, testing chimeric constructs around the consensus PKA site and from multiple species could be very informative.

      In addition, I would have a few small suggestions for improving the figures:

      • Fig. 2A (and similar plots in subsequent figures): is it really necessary to cut the x axis? Would it be possible to indicate the number of oocytes for each experiment (maybe in the legend in brackets)?
      • Fig. 2D (and all similar plots below): I am lacking the discrete data points that were measured. Without these it is impossible to evaluate the fits. The half-times shown in 2E are somewhat redundant, and the information could be combined on a single plot.
      • Fig. 3: why is the blot for PKA substrates cut into 3 pieces? It would be clearer to show the entire membrane.

      Significance

      Overall, I find this study extremely important, because it is only possible to entangle the diversity of cellular mechanisms though such comparative studies. Oocyte maturation perfectly exemplifies this issue: without doubt, oocyte maturation is a fundamental process and its detailed understanding is critical. However, researchers are often discouraged by diversity across species, which indeed complicates and hinders progress, well-reflected by the name "cAMP paradox". Combined with careful bioinformatic analyses, comparative studies can elegantly resolve such "paradoxes" through resolving the evolutionary history of molecular mechanisms.

    1. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary: Centrosomes separate early in mitosis to allow faithful spindle assembly and chromatin segregation. In the current study the author show that in RPE-1 cells the separated centrosomes typically position each other along the shorter axis of the nucleus while in cancer derived U2OS and MDA-MB-486 this is rather random. Mitotic cell rounding is not causal for this effect. Rather, the LINC (linker of nucleoskeleton and cytoskeleton) complex, a protein complex spanning both membranes of the nuclear envelope, is required for this. The data indicate that this is dynactin1 recruitment to the nuclear envelope. The work suggest that proper arrangement of the centrosomes along the short nuclear axis via the LINC complex contributes to chromatin segregation fidelity in RPE-1 cells.

      Major comments: The data, derived mostly by life cell imaging of cell culture lines, are of very high quality, carefully controlled and analyzed. They fully support the claims of the study and are well presented, both in text and figures. Statistical analysis seems adequate, but since the authors show different kinds of data sets including time series and use several kinds of statistical tests, it would make sense to indicate the test used for each p-value in all the figure legends. I have no major criticism or experiments to suggest.

      Minor comments:

      1. Figure legends are quite repetitive and could be shortened. E.g. in Fig. 1 the description for E, F, H and I repeats what has been explained for B and C. Same applies between figure legends. The authors might refer to previous legends if the analysis was done in a similar way.
      2. How is nuclear solidity defined and analyzed in Fig S3D?
      3. The references to Fig S3 in figure legend 3 ("see Fig S3") do not enlighten the message and could be removed. The same applies to Fig5 - here it is not clear why the author refer to Fig S4.
      4. Fig. 3: I suggest to quantify the lamin B1 and LBR overexpression levels.
      5. Fig. 5: Consider reordering the panel: Start with the current panel C (as in the text) as it is the necessary control prior to the experimental data.
      6. Fig 5 I: what means "before"? Can the authors give a time window they use for analysis.
      7. Page 20: "... shortest nuclear axis (Fig. 1C, 5D-G; n.s. - not significant). However, DN-KASH-expressing cells showed compromised separation and positioning of centrosome (Fig. 5D-G, * p=0.0155 and * p=0.0237, respectively). - rather point to the specific panels, i.e. Fig. 1C, 5D and F as well as and Fig. 5E and G.
      8. Fig 6B. The DN- KASH bars are on my pdf not visible - use a darker grey
      9. Fig S6, albeit mentioned in the text, is not included in the supplementary info.
      10. Material and Methods: in general very clear and carefully written
      11. a. GlutaMAX instead of GlutaMAXE (page 29)
      12. b. What means "as described previously"? No reference is given. Do you refer to the upper part of the method section? (page 30)
      13. c. 20 nM HEPES should most probably read 20 mM (page 32)
      14. d. "1:50 protease inhibitor; 1:100 Phenylmethylsulfonul fluoride" - which protease inhibitor (mixture)? Rather phenylmethylsulfonyl fluoride.
      15. e. exact composition of the cytoskeleton buffer used to prepare 4% paraformaldehyde could be given

      Referees cross-commenting

      I also mentioned in teh significance section the two weak points (only one non-cancer cell line (RPE-1); the precise role of the LINC complex). I thus think all three reviewers come to a similar conclusion: technically well done albeit some improvements are possible (reviewer 2). Manuscript is interesting but whether the findings can be generalized remains open and the overall impact is limited. Personally, I think a good strategy for the authors might be to stay with the three cell lines and avoid too general statements.

      Significance

      General assessment: This is a very elaborate analysis of centrosome positioning at the entry of mitosis. The experiments are carefully controlled and the findings supported by multiplied experiments, e.g. the aspect of mitotic cell rounding by analysis of unperturbed cells but also by manipulation accelerating and inhibiting cell rounding. Contribution of the LINC complex is evaluated by shRNA against SUN1/2, i.e. main LINC components, but also by the KASH-DN fragment, which acts as dominant negative. On the downside the study is limited to one untransformed cell line. Given that the treatments interfering with LINC complex function most likely affect all aspects of LINC-centromere interplay, it remains open what precise function of the LINC-complex contributes to chromatin segregation fidelity

      Advance: The work clearly shows that at least in RPE-1 cells the separated centrosomes arrange each other along the shorter axis of the nucleus and that the LINC complex is required for this.

      Audience: The work is certainly interesting for researches interested in mitosis, most precisely in spindle assembly. It enlightens a very specific aspect of spindle assembly but this very convincing. The work is basic research.

      Our experience is basic research of mitosis, nuclear structure and function both using biochemical assays and life cell imaging.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      A nuclear signal in prophase determines centrosome positioning and ensures efficient mitotic spindle assembly.<br /> Lima and Ferreira investigate in this manuscript the regulation of centrosome positioning in early mitosis. The authors first analyze the position of the two centrosomes either relative to the cell length axis or the shortest or longest axis of the nucleus and describe differences between RPE1, U2OS, and MDA-MB cells. Next, they analyze whether mitotic cell rounding determines the position of centrosomes; however, delayed cortical retraction (Rho-kinase inhibition), adhesion disassembly inhibition (Rap1Q63E), and inducing premature rounding (CalA) did not impact centrosome positioning in RPE1, U2OS, and MDM-MB cells. In addition, the nuclear lamina and LBR were also not required for centrosome positioning on the shortest nuclear axis. In contrast, depletion of SUN1 or SUN2 and overexpression of a dominant-negative DN-KASH affected the nuclear positioning of centrosomes in RPE1 cells. Finally, the authors analyze whether the LINC complex impacts mitotic fidelity. This is indeed the case when SUN1 is depleted, but it is not the case for SUN2 depletion or DN-KASH overexpression. This difference between LINC complex components is not discussed in the manuscript. Since SUN1, SUN2, and DN-KASH affect centrosome positioning in a similar way (Figs. 4 and 5), the chromosome segregation defect in SUN1-depleted cells is most likely not caused by a centrosome position defect but probably by another defect caused by SUN1 depletion.

      Major comments

      1. Figure 1 is insufficiently explained. The authors have to describe in an understandable way how they measured centrosome-centrosome angle and centrosome-nucleus angle. They should show a cartoon in which these angles are clearly shown. The small cartoons in Fig. 1C are not helpful at all; they are also not explained. The authors should explain the meaning of the black dots (are these centrosomes?) and the even smaller dots. The short nuclear axis should be indicated, e.g., by a red line.
      2. On the first page of the manuscript: "Consequently, at the NEP, centrosomes are positioned on the shortest nuclear axis (Fig. 1C) as can be seen in Fig. 1A. This means that the centrosome-nucleus angle relative to the shortest nuclear axis should be 0. However, in Fig. 1C, this angle is between 45 and 90 degrees. This is also the case for Fig. 1D. Please clarify.
      3. I find it confusing that in Fig. 1, depending on the subfigure, the short or longest nuclear axis is used as a reference point: Fig. 1C: shortest; D: shortest; F: shortest; G: longest; I: shortest; J: longest. Thus, even within the same cell line, the reference point is changing. What is the rational for this variation?
      4. Fig. 4K, L, M: in figure, y-axis: "shortest nuclear axis". In legend: "relative to the longest nuclear axis". I guess the longest nuclear axis is correct. Same in Fig. 5D and E. Fig. 5C lacks the WT control.
      5. The cells in Fig. 5J are not comparable: one has a monopolar spindle, the other a bipolar. The authors need some other NE protein as a control to show that the reduction of dynein by DN-KASH is a specific defect and not a broad impact on the NE. The dynein data in Figs. 5J-L need to be extended to SUN1/2.
      6. The title of the paper is misleading: the authors do not provide any indication for a nuclear signal in prophase that determines centrosome positioning.

      Minor comment

      1. It would make sense to use the same time scale in Figs. 1A and B (either min.sec. or sec.) to allow direct comparison.
      2. 2nd section: Mitotic cell rounding "The authors state: Given that cancer cells failed... I would be careful with this generalization; only one cancer cell was used in this study.
      3. The authors say: "However, they did not place the centrosomes at the shortest nuclear axis (Figure 4K-M)." Centrosomes are still on the shortest nuclear axis but not as frequent as in control.
      4. The white color in Fig. 6B cannot be seen and needs to be changed to something else.
      5. The paper has neither line nor page numbers.

      Referees cross-commenting

      My comments are more or less reflected by the comments and concerns of reviewer 1 (only one cancer cell line; the role of the LINC complex). This reduced the impact of this manuscript that is certainly intresting and has novel aspects.

      Significance

      The manuscript analysis an early step in spindle assembly: the positioning of the two centrosomes on the NE. As such, the paper is interesting and important. They exclude cell rounding and lamin disassembly as mechanisms for centrosome positioning. The SUN1/2 and KASH data on centrosome positioning are convincing, and they provide a novel finding on the function of the LINC complex in centrosome positioning, probably via dynein recruitment to the NE. It remains unclear whether LINC recruits dynein directly or functions via one of the two known dynein/NE recruitment pathways. LINC-dynein at the NE binds centrosome microtubules and dynein pulls them towards the NE. However, how LINC-dynein spatially positions centrosomes relative to the short axis of the nucleus remains unclear (dynein uniformly decorates the NE (Fig. 5J)). The data on chromosome missegregation are not so clear because the defect only occurs in SUN1-depleted cells. Thus, this phenotype indicates most likly a function of SUN1 but not the LINC complex and is probably not related to centrosome positioning since all LINC components affect centrosome positioning. The paper falls short in explaining how parameters were measured and contains mistakes in the figures, as outlined above. The paper lacks a coherent story (a little bit on cancer, some negative data, LINC-dynein, but it stops on the surface).

      It will be relatively easy to improve some aspects of the manuscript (explaining the angles, correcting the figures: one week). Measuring dynein at the NE in SUN1/2-depleted cells is also easy to do (1-2 months). To get more mechanistic insides into how LINC-dynein positions centrosomes probably will not be possible during revision time.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary: in this study, Lima and colleagues, investigate the mechanisms controlling the position of the two centrosomes at nuclear envelope breakdown. The authors show that in the non-cancerous human epithelial RPE1-cell line, the centrosomes are generally positioned in the short axis of the nucleus; in contrast in two cancer cell lines, they did not find an equivalent pattern. When the authors set out to identify potential molecular players required for this positioning, they find that the LINC complex is required , possibly by recruiting dynein to the nuclear membrane. Finally, the authors show that disruption of the LINC complex is associated with chromosome segregation errors.

      Major Comments:

      In general, the presented experiments are of excellent technical quality. The main conclusions of the manuscript, are however, not always well supported by the experimental data. They should be either interpreted more cautiously or supported by additional experimental evidence. I highlight these here, using the main conclusions of the abstract.

      1. "We show that in untransformed cells, centrosome positioning is regulated by a nuclear signal, independently of external cues. »

      The authors conclude based on three cell lines that the centrosome positioning mechanisms is present in non-transformed cells and not in cancerous cells. The authors have, however, only analysed 1 non-cancerous cell line, and they compare cells originating from vastly different tissues (retina, bones and breast) and origins (epithelial vs. mesenchymal cartilage cells). Such a general statement is not possible, without a systematic comparison of several healthy cells vs cancerous cells from the same tissue.<br /> 2. "This nuclear mechanism relies on the linker of nucleoskeleton and cytoskeleton (LINC) complex that controls the loading of dynein on the nuclear envelope (NE), providing spatial cues for robust centrosome positioning on the shortest nuclear axis, prior to nuclear envelope permeabilization (NEP). »

      While the data showing that centrosome positioning depends on the LINC complex is solid and robust, some of the "negative" examples identified by the authors are less convincing. One the process the authors study is cell rounding. Based on the fact that Rap1 transfection or treatment with Calyculin A does not lead to differences that are statistically different, the authors conclude that cell rounding is not involved. However, absence of statistical difference does not mean that there is no difference. Indeed, when comparing the raw data in Figure 2L and 2Q to the positive hit shSun2 in Figure 4J, one could conclude that cell rounding does make a difference, and that this statistical difference would emerge if the authors would count a high number of cells. Therefore the authors should interpret these results in a more differentiated manner, and also instead of just stating non-significant, state also the real p-values for the different experiment.<br /> The second major concerns emerges when looking at the data in Figure 5, when the authors test for the abundance of the dynein complex on the nuclear envelope in cells treated with DPPPL-KASH or DN-KASH. Yes, there is a statistical difference, but the absolute difference is tiny (I estimated a normalized intensity of 1.44 vs 1.35). This is a difference of less than 10%. How do the authors think that such a small change in dynein could have such a strong effect on centrosome positioning? Would a partial dynactin depletion by 10% give an equivalent result? Does the depletion of other proteins involved in the late recruitment of dynein at the NE also affect centrosome positioning?<br /> 3. « Moreover, we demonstrate this mechanism is altered in cancer cells, leading to increased chromosome segregation errors. »

      Here the authors infer that the identified mechanism is absent in cancer cells and that its absence contributes to chromosome segregation errors. Both conclusions are not supported by the presented data. First, the authors did not test whether any members of the LINC complex or dynactin is present at lower levels on the nuclear membranes of the cancer cells. Such a direct validation would be essential to make such a strong statement. Second, the authors conclude that this mechanism prevents chromosome segregation errors, based on the fact that depletion or impairment of the LINC complex (shSUN1, shSUN2, DN-KASH) results in chromosome segregation errors. These perturbations lead ,however, as noted by the authors themselves to pleiotropic effects, including insufficient retraction of nuclear membrane, which will can all contribute to chromosome segregation errors. It is therefore impossible to estimate the contribution of the centrosome positioning mechanism to these segregation errors using this type of pertubrations. One could even argue that this mechanism might not be that important, since depletion of SUN2, which also impairs centrosome positioning has no significant effect on chromosome segregation.

      Minor comments:

      The author state in the Material and methods that all the figure legends contain the number of replicates. This is, however, not the case, the authors only indicate the total number of analyzed cells.

      Referees cross-commenting

      I agree that all three reviewers come to similar conclusions - strong technical quality, novel results and concepts, but some limitations due to lack of precise tools or the limited number of model cell lines investigated.<br /> I recommend that the authors prioritize which are the suggested experiments that could be done within a few months, and otherwise rephrase their conclusions in less general terms.

      Significance

      This study establishes for the first time that some cell lines set up the mitotic spindle at predefined positions of the nucleus and they identify a first molecular complex controlling this complex.

      The strength of this study is the high technical quality of the data - a limitation is the over-interpretation of the current data (see major comments), and the fact that the authors do not have a tool that specifically only disrupts centrosome positioning, which would allow them to probe the importance of this mechanism.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      The authors do not wish to provide a response at this time.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      In this interesting paper Bideau et al. report an antero-posterior gradient in the plasticity of tail tissues during tail regeneration in the annelid worm Platynereis dumerilii. The experiments are well designed and thoroughly quantified, the figures are of high quality.

      Major comments:

      The microscopic images lack scale bars. These should be added to all figures.

      The authors should provide the source data for all quantifications as txt files. They should also provide as supplement representative confocal stacks for the various stainings.

      The authors use LatrunculinB treatment to investigate the role of cell migration in regeneration. However, since LatB inhibits f-actin, it could also interfere with cell proliferation and other processes. The authors should check if this is the case and provide control data.

      Minor comments:

      Sometimes the language is a bit quite cryptic. For example, the title of Figure 4 is "Cell proliferation and migration, as well as tissue maturity modulate the plasticity of posteriorized gut progenitors through regeneration"<br /> in short: 'cell migration modulates the plasticity of progenitors' - this is just to say that inhibiting cell migration reduces regeneration

      The authors should attempt to simplify the language.

      Language:

      "is an tremendous and essential process in animals" - not clear what 'tremendous' means here - please revise

      "Those regeneration processes have been studied from a long" - for a long time

      "more EdU+ cells in S1 than in S6 or S7 regardless the EdU incubation time" - regardless of the

      "It showed that the gut is composed of" - The stainings showed that

      "Indeed, cell labelled with a rather short EdU" - cells labelled

      "tissues plays a major role on the reformation" - in the formation

      The paper will be of interest to animal developmental biologists and scientists working on the plasticity of tissues during regeneration.

      Referees cross-commenting

      I agree with the comments made by the other reviewers. The authors need to be more clear and careful in interpreting their data. I don't think that new data are needed (unless they would like to demonstrate a 'gradient' with more positions) and the comments could be addressed by substantially rewriting the text and revising the claims.

      Significance

      The paper will be of interest to animal developmental biologists and scientists working on the plasticity of tissues during regeneration.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Bideau et al. studied the origin, plasticity and fate of the cells participating in blastema formation during posterior regeneration in the annelid Platynereis dumerilii. To label and track the fate of proliferative cells, the authors applied EdU/BrDU incorporation coupled with mRNA in situ hybridization and fluorescent beads labelling, on a wide array of regeneration assays. They also performed drug treatments to assess the role of proliferation and cell migration during posterior regeneration. Interestingly, the Authors showed that some proliferative gut cells can participate in the formation of ectodermal and mesodermal tissues during regeneration, in case of two successive posterior regeneration events, suggesting that gut cycling cells residing in a regenerating segment are more plastic than those located in a non-regenerating segment. They also suggested the existence of two different cell populations - "slowly" and "quickly" cycling cells - acting during regeneration. Overall, the experiments are well presented, and methods clearly described. The Authors concluded that posterior regeneration in Platynereis relies on a gradient of cell plasticity and cell proliferation, along the antero-posterior axis of the animal.

      Major comments

      Two of the main conclusions of this study are, in my opinion, not supported by the data:

      As indicated in the title of the manuscript, the Authors put forward a cellular model for posterior regeneration relying on gradients of cell proliferation, cell differentiation and cell plasticity along the the antero-posterior axis of the animal. I am not convinced that the Authors have provided strong enough evidence to prove any of these gradients. They showed that there are differences between the region directly adjacent the most posterior segment and a region located more anteriorly (6 or 7 segments from the posterior end). However, by comparing only two positions, they cannot distinguish between graded or clearly regionalized contexts. To prove the existence of a gradient along the animal's antero-posterior axis, the authors would need to compare cell proliferation, cell dynamics and cell differentiation between multiple regions at increasing anterior positions, and show that their responses are indeed graded. This would represent a quite substantial amount of work. Instead, I would suggest removing the reference to a gradient in the paper entirely.

      Using "short" (5h) and "long" (48h) EdU pulses, the Authors claim they have established the existence of two cell populations, namely "slowly-cycling cells" and "quickly-cycling cells" (first paragraph of the result section - pages 5/6 "We exposed uninjured worms to EdU, either for 5 or 48h to discriminate quickly-cycling cells from cells harboring a slower replication rate"). I am not convinced that the Authors provide strong enough evidence to demonstrate the existence of two such cell populations. Given that about 20% of cells incorporated EdU after 5h of exposure, that almost all of them have done so after 48h, and that only a fraction of proliferative cells are in S-phase at any given time, it is well possible that a majority of cells stained after 5h and 48h are from the same cell cycling population. To show the existence of different populations of cells, cycling at different rates, the Authors would need to compare staining after an equal EdU exposure time, following a period of chase of different duration. Without this set of experiments, I would refrain from distinguishing between several slower and faster cell cycling populations.

      Minor comments

      Page 1. Please correct "is an tremendous" into "is a tremendous".

      Page 7. "The huge majority of the EdU+ cells colocalize with FoxA". Please provide quantification.

      Page 11 "Quickly-cycling gut progenitors.... cannot give rise to neural progenitors and probably not to stem cells from the ectodermal growth zone"; Page 12 "cannot regenerate neural tissues"; Page 20 "posterior gut progenitors cannot produce nervous system or putative posterior stem cells". What the authors show in their experiments, is that labeled gut cycling cells likely do not generate neural cells or stem cells, in the assessed context. However, the Authors do not show that those cells 'cannot' do so. Please rephrase.

      Page 11. "migration, through actin polymerization (LatrunculinB or LatB) widely used inhibitors". Please add a reference to justify the use of LatB as a cell migration inhibitor.

      Significance

      This is a thorough, well executed and interesting study on a tractable annelid regeneration model. The experiments are neatly performed and the manuscript reads well. As stated in my major comments, two of the main conclusions of the study (gradient of cell proliferation/plasticity/differentiation; identification of two types of progenitors differing in their cell cycle rates) have not been demonstrated properly and would need to be either strengthened or deleted from the manuscript. Several other findings, notably the increased plasticity of cells that recently participated in posterior regeneration (notably gut cells) are well demonstrated and of interest. Overall, this manuscript significantly advances our understanding of the cellular mechanisms that occur during posterior regeneration in Platynereis. It will be of interest to anyone working on comparative regeneration, but may be of lesser interest to researchers working outside this field.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Bideau et al. describe posterior regeneration in the annelid Platynereis. The authors aimed to identify patterns of proliferative cells. Pulse-chase and double labeling by EdU/BrdU was used to track cells. Finally, they attempted to reveal the identity of cycling cells and their contribution to regeneration. Platynereis is a relatively new regeneration model. Understanding the cellular source of regeneration in this annelid would be of considerable interest.

      The authors performed EdU labeling of intact worms to find a posterior-anterior decreasing gradient of S-phase cells. Histological sections showed that proliferative cells are located in all tissues in the posterior-most segment, but mostly restricted to the gut epithelium in more anterior segments. Using fluorescent beads that are taken up specifically by gut epithelial cells, they show that gut epithelial cells of the intact animal contribute only to gut regeneration, i.e., they are lineage restricted. The authors also performed immunostaining and mRNA in situ hybridization experiments to better understand the tissue identity of proliferative cells.

      The following are my specific comments:

      1. I am not sure I understand how the authors identify slow- or fast-cycling cells. EdU gets incorporated in S-phase; the longer the incubation time the more cells will be labeled until saturation is reached. The length of the cell cycle and the number of different populations cannot be directly derived from this experiment. I think it would be fair to conclude that there are more cycling cells in posterior segments and in the gut of anterior segments but the conclusion of two distinct populations is unsupported in my opinion.
      2. The Results and Discussion sections will have to be revised to address the above issue. The two supported conclusions are (1) the gradient of proliferative cells (but w/o reference to the number of distinct populations); (2) the fate-restricted nature of gut epithelial cells. The plasticity gradient is unsupported because worms can regenerate if amputated at segment #5. This suggests the presence of either resident stem cells (with broad potential or lineage-specific), cells that can dedifferentiate, or a combination of both. The authors' experiments cannot discriminate between the alternatives.
      3. The text would benefit from copy editing to improve the language and making it more accessible. In its current form, it is rather difficult to read, with descriptions of experiments that are not easy to follow.
      4. The figures and their legends can be improved. Re the legends, one has to read the full text to understand what each panel shows. The figures are very complex. It would already be easier if usage of A', A', A' was avoided. The figures could also be improved by direct annotation. Finally, consider simplifying the main figures and moving some material to the supplement.
      5. How do the authors know that proliferative cells in the gut are "gut progenitors"? They might simply be proliferative gut epithelial cells.
      6. The conclusions drawn from the drug experiments are overstated.

      Referees cross-commenting

      The comments made by the two other reviewers are complementary to mine. Either the authors extensively revise the text to remove unsupported conclusions or they perform additional experiments.

      Significance

      Little is known about the cellular basis of Platynereis posterior regeneration.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      We thank the Reviewers for their helpful and constructive comments. In response to these suggestions we have performed new experiments and amended the manuscript, as we describe in our detailed response below.

      Reviewer #1:

      1. The Reviewer notes that while our analysis of centrosome size was comprehensive, we provided no analysis of centrosomal MTs, pointing out that while centrosome size declines as the embryos enter mitosis, the ability of centrosomes to organise MTs might not. This is a good point, and we now provide an analysis of centrosomal-MT behaviour (Figure 2). We find that there is a dramatic decline in centrosomal MT fluorescence at NEB, although the pattern of centrosomal MT recruitment prior to NEB is surprisingly complex.

      2. The Reviewer questions how PCM client proteins can be recruited in different ways by the same Cdk/Cyclin oscillator. We apologise for not explaining this properly. It is widely accepted that Cdk/Cyclins drive cell cycle progression, in part, by phosphorylating different substrates at different activity thresholds (e.g. Coudreuse and Nurse, Nature, 2010; Swaffer et al., Cell, 2016). Moreover, it is also clear that Cdk/Cyclins can phosphorylate the same protein at different sites at different activity thresholds (e.g. Koivomagi et al., Nature, 2011; Asafa et al., Curr. Biol., 2022; Ord et al., Nat. Struct. Mol. Biol., 2019). Thus, we hypothesise that rising Cdk/Cyclin cell cycle oscillator (CCO) activity phosphorylates multiple proteins at different times and/or at different sites to generate the complicated kinetics of centrosome growth. We now explain this point more clearly throughout the manuscript.

      3. The Reviewer is puzzled as to how we conclude that Cdk/Cyclins phosphorylate Spd-2 and Cnn at all the potential Cdk/Cyclin phosphorylation sites we mutate in our study. The Reviewer is right that we cannot make this conclusion, and we did not intend to make this claim. As we now clarify (p11, para.1), although it is unclear if Cdk/Cyclins phosphorylate Spd-2 or Cnn on all, some, or none of these sites, if either protein can be phosphorylated by Cdk/Cyclins, then these mutants should not be able to be phosphorylated in this way—allowing us to address the potential significance of any such phosphorylation. We now also note that several of these sites have been shown to be phosphorylated in embryos in Mass Spectroscopy screens (Figure S6).

      4. The Reviewer highlights differences in how Spd-2 and Cnn help recruit γ-tubulin to centrosomes (Figure 6). They ask for a more detailed description, and are puzzled as to how this is compatible with direct regulation by a single oscillator. We now explain our thinking on this important point in much more detail. It appears that Spd-2 helps recruit γ-tubulin throughout S-phase, while Cnn has a more prominent role in late S-phase (Figure 6). This is consistent with our overall hypothesis of CCO regulation, as we postulate that low-level CCO activity promotes the Spd-2/γ-tubulin interaction in early S-phase, while higher CCO activity promotes the Cnn/γ-tubulin interaction in late-S-phase, potentially explaining the increase in the rate of γ-tubulin (but not γ-TuRC) recruitment we observe at this point (see minor comment #1, below, for an explanation of the various γ-tubulin complexes in flies). This is consistent with recent literature showing that CCO activity promotes γ-tubulin (but not γ-TuRC) recruitment by Cnn/SPD-5 in worms and flies (Ohta et al., 2021; Tovey et al., 2021).

      5. The Reviewer was not convinced by our model (Figure 8, now Figure 9), raising two major concerns. First, they were unsure how a single oscillator could generate different patterns of protein recruitment. We addressed this in point #2 and #4, above, where we explain how different thresholds of CCO activity trigger different events, so there is no expectation that we should observe steady changes in recruitment over time as CCO activity rises. Second, they questioned how modest levels of Cdk/Cyclin activity can promote recruitment, while high levels of activity can inhibit recruitment. In point #1, above, we cite several examples where such positive and negative regulation by different Cdk/Cyclin activity levels have been described. We also now explain throughout the manuscript why this hypothesis provides a plausible explanation for our results: with moderate CCO activity promoting Spd-2-dependent PCM-client recruitment in early S-phase; higher CCO activity promoting a decrease in Spd-2 recruitment in mid-late-S-phase (so centrosomal Spd-2 levels decline); and even higher levels of CCO activity leading to a decrease in the interactions between the client proteins and the Spd-2/Cnn scaffold as the embryos enter mitosis (so the client proteins are rapidly released from the centrosome).

      The Reviewer also raised the important point here that our model does not explain why the mutant forms of Spd-2 and Cnn accumulate to higher levels at the start of S-phase, and not just at the end of S-phase/entry into mitosis. We apologise for not explaining this properly. The accumulation of the mutant proteins (particularly Spd-2, Figure 5C) in early-S-phase occurs because the excess mutant protein that accumulates at centrosomes in _late-_S-phase/mitosis is not removed properly from centrosomes during mitosis (presumably because there is insufficient time). Thus, centrosomes still have too much mutant Spd-2 at the start of the next S-phase. We show this in Reviewer Figure 1 (attached to this letter), which tracks Spd-2 behaviour further into mitosis, and now explain this in more detail in the text (p12, para.1).

      1. The Reviewer questions how the CCO can both induce centrosome growth and also switch it off, as it is unclear how an oscillator that only phosphorylates sites to decrease centrosome binding could also promote growth. They ask if we can identify and mutate any Cdk/Cyclin sites in centrosome proteins that promote centrosome recruitment. As we now clarify, we did not intend to claim that the CCO only phosphorylates sites that decrease the centrosome binding of proteins, although we do hypothesise that such phosphorylation is important for switching off centrosome growth in mitosis. In addition, we hypothesise that moderate levels of CCO initially promote centrosome growth, and our data suggests that the CCO does this, at least in part, by promoting Polo recruitment (Figure 8). We speculate that the CCO phosphorylates specific Polo-box-binding sites in Ana1 and Spd-2, the main proteins that recruit Polo to centrioles. We agree that identifying these sites is an important next step, but it is complicated as our studies indicate that multiple sites contribute in a complex manner. Importantly, it is well established that the CCO triggers centrosome growth as cells prepare to enter mitosis, so our hypothesis that moderate levels of CCO activity initiate centrosome growth is not new or controversial.

      Minor Comments

      1. The reviewer asks how we explain the different incorporation profiles we observe for the different subunits of the γ-tubulin ring complex. We apologise for not discussing this point. In flies there is a “core” γ-tubulin-small complex (γ-TuSC) and a larger γ-tubulin-ring complex (γ-TuRC) that contains the Grip71, Grip75 and Grip128 subunits we analyse here (Oegema et al., JCB, 1999). The γ-TuSC functions independently of the γ-TuRC so γ-tubulin and γ-TuRC components can behave differently.

      2. The Reviewer questions why we claim an “inverse-linear” relationship between S-phase length and the centrosome growth rate when the relationship is not linear (Figure 3, now Figure S3). I was originally confused by this as well but, mathematically, a linear relationship means y is proportional to x, whereas an inverse-linear relationship means y is proportional to 1/x. Thus, an inverse-linear relationship between x and y does not plot as a straight line, but rather as the curves we show on the graphs. We now explain this in text (p9, para.2).

      Reviewer #2:

      This Reviewer found the manuscript hard to follow, so we are very grateful that they took the time to try to understand it. We agree that the subject matter is complicated, and that our presentation was not always helpful. The Reviewer’s comments have been very useful in helping us to identify (and hopefully improve) areas of particular difficulty.

      Major points:

      1. The Reviewer highlights that the two experimental approaches underpinning our main conclusions are problematic: (1) Experiments with mutants of Spd-2 and Cnn that theoretically cannot be phosphorylated by Cdk/Cyclins are hard to interpret as these mutations may have other effects; (2) It is unclear whether reducing Cyclin B levels reduces peak CDK activity or simply slows the time it takes to reach peak levels. They suggest a more direct test of our model would be to analyse PCM recruitment in embryos arrested in S-phase or mitosis. (1) We agree that the mutations designed to prevent Cdk/Cyclin phosphorylation could perturb function in other ways, but this is true for any such mutation, and there are many papers that infer a function for Cdk/Cyclin phosphorylation from such experiments. Importantly, the centrosomal accumulation of the phospho-null mutants actually slightly increases compared to WT (Figure 5C and I), and we now show that the centrosomal accumulation of a phosphomimicking Spd-2-Cdk20E mutant slightly decreases (Figure S8). We now acknowledge the potential caveat of a non-specific perturbation of protein function, but feel that the reciprocal behaviour of the phospho-null and phospho-mimicking mutants somewhat mitigates this concern (p12, para.2). (2) Fortunately, and as we now clarify, it has recently been shown that reducing Cyclin levels does not reduce peak Cdk activity, but rather slows the time it takes to reach peak activity (Figure 2A, Hayden et al., Curr. Biol., 2022). Thus, the cyclin half-dose experiments provide an excellent alternative test of our hypothesis as they show that the WT proteins can exhibit similar behaviour to the mutants if the rate of Cdk/Cyclin activation is slowed. We feel the evidence supporting our hypothesis is strong enough that it warrants serious consideration.

      The suggestion to look at PCM recruitment in embryos arrested in either S-phase or M-phase is a good one, but these experiments produce complicated data. In M-phase arrested embryos, for example, Cnn levels continue to rise (see Figure 1G, Conduit et al., Dev. Cell, 2014), but the other PCM proteins do not (unpublished); in S-phase arrested embryos (arrested by mitotic cyclin depletion) centrosomes continue to duplicate, but now do so asynchronously, greatly complicating the analysis (McCleland and O’Farrell, Curr. Biol.., 2008; Aydogan et al., Cell, 2020). The centrosomes that don’t duplicate, however, reach a constant steady-state size (where the rate of centrosome protein addition is balanced by the rate of loss). These observations are consistent with our recent mathematical modelling of mitotic PCM assembly (Wong et al., 2022) if we additionally account for cell cycle regulation (which was not considered in our original model). We believe such analyses are beyond the scope of the current paper and we plan to publish a second paper incorporating our new hypothesis into our mathematical modelling.

      1. The Reviewer questions whether our methods accurately measure centrosomal protein accumulation, pointing out that γ-tubulin and Grip128 occupy different centrosomal areas—which should not be possible if they are part of the same complex. They suspect that our use of different transgenes with different promotors could explain these differences. As we should have described (see point #1 in our response to the minor comments of Reviewer #1), γ-tubulin exists in two complexes in flies, only one of which contains Grip128, so γ-tubulin and Grip128 exhibit different localisations. Moreover, as we now show (Figure S2), using different promotors does not seem to make a difference to overall recruitment kinetics. Thus, we are confident that our methods measure centrosome protein recruitment dynamics accurately.

      2. The Reviewer is concerned that our measurements of centrosome size based on fluorescence intensity (Figure 1) and centrosomal area (Figure S1) do not always match. They suggest a potential reason for this is that proteins are not uniformly distributed within centrosomes, and this may impact our ability to measure protein accumulation based on 2D projections (noting, for example, that Polo and Spd-2 are concentrated at centrioles and in the PCM, potentially explaining the different shape of their growth curves compared to the client proteins). When the centrosome-fluorescence-intensity and centrosome-area recruitment profiles of a protein do not match, the average “centrosome-density” of that protein must be changing over time. In some cases, we understand why density changes. Cnn, for example, stops flaring outwards on the centrosomal MTs during mitosis so its centrosomal area decreases even as its fluorescence intensity increases (leading to an increase in its centrosomal-density). We agree (and now discuss—p19, para.3) that the prominent accumulation of Spd-2 and Polo at centrioles could help to explain why Spd-2 and Polo accumulation dynamics differ from the client proteins.

      Other points:

      1. The Reviewer suggests it would be good to know how much Polo at the centrosome is active. We agree, but although commercial antibodies against PLK1 phosphorylated in its activation loop work in cultured fly cells, we cannot get them to work in embryos. Moreover, the recruitment of Polo/PLK1 to its site of action by its Polo-Box Domain is sufficient to partially activate the kinase independently of phosphorylation (Xu et al., NSMB, 2013). Thus, it seems likely that all the Polo/PLK1 recruited to centrosomes will be at least partially activated, even if it is not necessarily phosphorylated on its activation loop.

      2. The Reviewer asks if it is clear that less Spd-2 and Cnn are recruited to centrosomes in the half gene-dosage embryos. We apologise for not mentioning that this is indeed the case. We showed this previously for Cnn (Conduit et al., Curr. Biol., 2010) and we now state that this is also the case for Spd-2. We do not show the Spd-2 data as we plan to publish a comprehensive dose-response curve of Spd-2 (and Cnn) recruitment in our next modelling paper.

      3. Would it not be relevant to examine Polo ½ dosage embryos? We do have this data (Reviewer Figure 2), attached to this letter, but it is quite complicated to interpret (as we explain in the legend). We feel it would be more appropriate to include this in our next modelling paper where we can properly explain the behaviours we observe. Publishing this data here would distract from our main message without changing any of our conclusions.

      4. The Reviewer asks why the non-phosphorylatable Spd-2 protein is also present at higher levels on centrosomes at the start of S-phase (not just the end of S-phase). This was also raised by Reviewer #1 (point #5), so please see the second paragraph of our response there.

      Minor/Discussion Points:

      1. We thank the Reviewer for highlighting that absolute and relative centrosome size control are different things and we have amended the manuscript accordingly.

      2. The Reviewer questions whether it is accurate to describe Spd-2 and Polo as scaffold proteins, noting that only Cnn has been shown to have scaffolding properties. There is strong evidence that Spd-2 has Cnn-independent scaffolding properties in flies (e.g. Conduit et al., eLife, 2014), but this is a fair point for Polo. We think it is justified to separate Polo from other client proteins as Polo is essential for scaffold assembly, whereas other client proteins are not. We now define our scaffold/client terminology to avoid confusion (p4, para.3).

      3. The Reviewer highlights several points related to differences in recruitment kinetics (also touched on in points #2 and #3, above), noting we don’t discuss properly the idea of two different modes of PCM recruitment. These are all good points, largely addressed in our response to points #2 and #3, above. We now discuss much more prominently the two different modes of client protein recruitment throughout the manuscript.

      4. As we now clarify, in all our experiments we use centrosome separation and nuclear envelope breakdown (NEB) to define the start and end of S-phase, respectively.

      5. The Reviewer quotes the landmark Woodruff paper (Cell, 2017) as showing that the ability to concentrate client proteins (including ZYG-9, the worm homologue of Msps) is an intrinsic property of the PCM scaffold, so how do we explain that Msps departs prior to NEB while Cnn continues to accumulate? It is indeed a striking observation of our study that all PCM client proteins (not just Msps) start to leave the centrosome prior to NEB, even as Cnn levels continue to accumulate. Our hypothesis is that this ‘leaving’ event is triggered by a threshold level of Cdk/Cyclin activity—explaining why these client proteins all start to leave the PCM at the same time (just prior to NEB) irrespective of nuclear cycle length. This is not incompatible with the Woodruff paper, which did not attempt to reconstitute any potential regulation by Cdk/Cyclins in their in vitro studies.

      6. The Reviewer questions why Spd-2 that cannot be phosphorylated by Cdk/Cyclins (Spd-2-Cdk20A) accumulates abnormally at centrosomes in late S-phase, yet γ-tubulin (which is recruited by Spd-2) seems to leave centrosomes more slowly in the presence of the mutant protein. As we now explain more clearly, there is no contradiction here. Spd-2-Cdk20A accumulates to abnormally high levels in late-S-phase/early mitosis (Figure 5C), and this reduces the γ-tubulin dissociation rate, as we would predict (Figure 7B, right most graph). It does not “prevent” dissociation, however, (as the Reviewer seems to suggest it should?), but this is probably because these experiments have to be performed in the presence of large amounts of the WT Spd-2 (Figure 5A).

      7. The referencing error has been corrected.

      8. The Reviewer asks why in Figure 1 not all of the centrosome proteins could be followed for the full time period (as we mention in the legend, but do not explain). There are different reasons for different proteins: (1) Polo cannot be followed in mitosis as it binds to the kinetochores, making it impossible to accurately track centrosomes (so the data for mitosis is missing for Polo); (2) Cnn exhibits extensive flaring at the end of mitosis/early S-phase (Megraw et al., JCS, 1999), so we cannot track individual separating centrosomes labelled with NG-Cnn in early S-phase until they have moved sufficiently far-apart (so the early S-phase time-points are missing for Cnn); (3) In addition, several of the client proteins bind to the mitotic spindle, so although we can still track and measure the centrosomes in late mitosis in the graphs, we don’t show pictures of these late mitosis centrosomes in the montage in Figure 1A as the images look a bit odd. We now explain these reasons in the Materials and Methods.

      9. We now indicate that nuclear cycle 12 (NC12) is being analysed in Figures 4-8.

      10. The reviewer questions why we don’t show the decrease rate for γ-tubulin in Figure 6 (the Spd-2 and Cnn half-dose experiments), when we do show it in Figure 7 (the Spd-2 and Cnn Cdk-mutant experiments), suspecting that it is slowed in both cases. The reviewer is correct and we now show this data for both sets of experiments.

      11. We have corrected the labelling error in Figure S1.

      12. The Reviewer suggest moving some of the data from the main Figures, and the entirety of Figures 2 and 3 to the Supplemental Information. We understand this point, and agree that the amount of data presented in Figures 1-3 is somewhat overwhelming. We have played around with the Figures a lot—in particular trying to show a few examples of the data and moving the rest to Supplementary—but it is hard to pick a “typical” example, and the power of comparing the behaviour of so many different centrosome proteins is somewhat lost. We have tidied up several Figures and, as a compromise, we keep Figure 2 (now Figure 3) in the main text, but have moved Figure 3 to Supplementary (now Figure S5).

      13. The Reviewer suggests that we should repeat the analysis of Spd-2, Polo and Cnn dynamics that we show here, as we already presented this data in a previous publication (Wong et al., EMBO. J, 2022). We understand this point, but feel this would be a less accurate comparison, as essentially all of the data shown in Figure 1 was obtained several years ago during a contiguous ~6month period. Since then, the lasers and software on our microscope system have been updated, so it would probably be less fair of a comparison to obtain new data for a subset of these proteins (and it seems overkill to perform the entire analysis again). We clearly state that this data has been presented previously, so we hope the Reviewer will agree that it is acceptable to present it again here so readers can more easily compare the data.

      Reviewer #3:

      This Reviewer is broadly supportive of the manuscript, but to publish in a prestigious journal they think additional experimental evidence will be required to support our hypothesis.

      The Reviewer notes that our only evidence that Cdk/Cyclins directly phosphorylate Spd-2 comes from our analysis of the Spd-2-Cdk20A mutant, as the effect of reducing Cyclin B dosage on WT Spd-2 behaviour is very modest. They request that we analyse the behaviour of a Spd-2-Cdk20E phospho-mimicking mutant. The effect of halving the dose of Cyclin B on Spd-2 behaviour is modest, but this is what we would predict as all we are doing in this experiment is slowing S-phase by ~15%, so Spd-2 should accumulate at centrosomes for a slightly longer time and to a slightly higher level (as we observe, Figure 5E). A great advantage of the early fly embryo system is that we can compare the behaviour of many hundreds of centrosomes, so even subtle differences like this are usually meaningful. To illustrate this point, we have now repeated the Spd-2 analysis in WT and CycB1/2 embryos (but now using a CRISPR/Cas9 Spd-2-NG knock-in line) and we see the same subtle differences (Figure S9). In addition, as requested, we have now analysed the behaviour of a Spd-2Cdk20E mutant protein using an mRNA injection assay (as it would have taken too long to generate and test new transgenic lines). In this assay we injected embryos with mRNA encoding either WT Spd-2-GFP, Spd-2-Cdk20A-GFP or Spd-2-Cdk20E-GFP. The mRNA is quickly translated, and we computationally measured the fluorescence intensity of the centrosomes in mid-S-phase (i.e. at the Spd-2 peak) (Figure S8). This analysis confirms that Cdk20A accumulates to slightly higher levels, and reveals that Cdk20E accumulates to slightly lower levels, than the WT protein. Together, these new experiments strongly support our original conclusions.

      The Reviewer notes that we propose that the CCO initially promotes centrosome growth by stimulating Polo recruitment to centrosomes, but states that we only provide indirect evidence for this by showing that centrosomal Polo levels are strongly reduced in Cyclin B half-dose embryos. They suggest we determine Spd-2 levels in Polo half-dose embryos, and/or the centrosome levels of mutant forms of Spd-2 that cannot be phosphorylated by Polo. We believe the Cyclin B half-dose experiment provide direct support for our hypothesis that Cdk/Cyclin activity influences Polo recruitment (Figure 8), although, clearly, we have not identified the mechanism. We do, however, suggest a plausible mechanism: Ana1 and Spd-2 are largely responsible for recruiting Polo to centrosomes, and we have previously shown that several of the potential phosphorylation sites in these proteins that help recruit Polo to centrosomes are Cdk/Cyclin or Polo phosphorylation sites (Alvarez-Rodrigo et al., eLife, 2020 and JCS, 2021; Wong et al., EMBO J., 2022). We are currently testing this hypothesis, but progress is slow as it is clear that multiple sites in both proteins can influence this process.

      As the Reviewer requests, we have now also examined how Spd-2 and Cnn behave in Polo half-dose embryos (Reviewer Figure 2, attached to this letter). As we describe in the Figure legend, this data is informative, but is complicated. With relatively minor, but mechanistically important, tweaks to our previous mathematical modelling we can explain these behaviours, but introducing such a significant mathematical modelling element would be beyond the scope of this paper. As described above, these findings will form the basis of a follow-up paper that is more mathematically oriented.

      It is a great idea to look at mutant forms of Spd-2 that cannot be phosphorylated by Polo, but the consensus Polo phosphorylation site (N/D/E-X-S, with the N/D/E at -2 and the S at 0 being preferences, rather than a strict rule) is less well-defined than the consensus Cdk/Cyclin phosphorylation site (where the Pro at -1 is essentially invariant). Thus, we cannot accurately predict which sites would need to be mutated to generate such a mutant.

      The Reviewer requests that we analyse the behaviour of TACC in embryos expressing the Spd-2-Cdk20A and Cnn-Cdk6A (as we do in Figure 7 for γ-tubulin). This is a reasonable request, but we prefer not to show this data as we have recently identified an interesting interaction between TACC, Spd-2 and Aurora A that will be the subject of another paper we hope to submit shortly. This data is hard to interpret without explaining these interactions properly, which is beyond the scope of the current manuscript.

      We hope the Reviewers will agree that these changes have improved the manuscript substantially, and that it is now suitable for publication. We would like to thank them again for taking the time to read this rather complicated paper so thoroughly.

      We look forward to hearing from you.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      In this manuscript, the authors investigated growth control of PCM at the mitotic centrosomes in late stages of the Drosophila syncytial embryos. They observed that mitotic centrosomes reach to the correct sizes through 13 rounds of nuclear division by reciprocally slowing their growth rate and increasing their growth period. They assumed that the Cdk/Cyclin cell cycle oscillator (CCO) is a main controller, based on their previous works (Aydogan et al., 2018, 2020; 2022). They determined the recruitment dynamics of the key mitotic PCM scaffolding proteins (Spd-2, Polo and Cnn) and PCM-client proteins (γ-tubulin, Msps, TACC, GFP, Grip75, Grip128 and Aurora A) in living embryos, and proposed that moderate levels of the CCO activity promote centrosome growth by stimulating Polo recruitment to centrosomes, while higher levels of activity subsequently inhibit centrosome growth by phosphorylating centrosome proteins, such as Spd-2, to decrease their centrosome recruitment and/or maintenance as the embryos enter mitosis.

      Experiments were cleverly designed and carefully executed. The results are nicely presented, the manuscript is clearly written, and their proposal draws a strong attention. However, in order to publish the manuscript in a prestigious journal, the authors may provide additional experimental evidence to support their proposal.

      • It is very significant that the centrosome levels of Spd-2-Cdk20A-NG is stronger than Spd-2-NG throughout the cell cycle (Figure 5B,C). However, this is only an experimental evidence to support that Cdk/Cyclins directly phosphorylate Spd-2 in the run-up to mitosis to help reduce Spd-2's centrosome recruitment and/or maintenance. As the authors confessed, recruitment of Spd-2-NG to the centrosomes in CycB1/2 embryos (Figure 5D,E) may be moderate or not significant at least in this reviewer's eyes. It is worth to perform the same experiments with a phospho-mimetic Spd2-Cdk20E-NG mutant.
      • The authors proposed that moderate levels of CCO activity promote centrosome growth by stimulating Polo recruitment to centrosomes. They provided an indirect evidence that centrosome levels of polo were strongly reduced in CycB1/2 embryos (Figure 4E,F). It is worth to determine the centrosome levels of Spd-2 in the Polo1/2 embryos and/or the centrosome levels of Polo phospho-resistant Spd-2 (Spd-2-Polo#A-NG).
      • TACC may be an ideal PCM-client protein, apart from its importance in spindle formation in comparison to γ-tubulin (Figure 4C,D). Therefore, it is worth to perform the Figure 7 experiments with TACC.

      Significance

      Experiments were cleverly designed and carefully executed. The results are nicely presented, the manuscript is clearly written, and their proposal draws a strong attention. However, in order to publish the manuscript in a prestigious journal, the authors may provide additional experimental evidence to support their proposal.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Control of organelle size has been an active field of research for many years for a large variety of cellular structures and in a range of experimental models. Here, Jordan Raff and colleagues examine the mechanisms underlying centrosome (PCM) size control in Drosophila syncytial embryos, building on their previous work (Wong, EMBOJ 2022) to propose a role for CDK in both promoting (at intermediate levels) and inhibiting (at high levels) PCM expansion.

      I found this a difficult manuscript to review, not only because the subject matter is complicated, but so is the writing. Having read and re-read the manuscript some clarity eventually emerges, but it shouldn't be that inaccessible. As for the authors' model I find it intriguing, but not fully supported by the data currently presented.

      Major points

      1. Central to the authors' model is the proposed dual function of Cdk (or CCO in the authors' terminology) in both promoting and inhibiting centrosomal protein accumulation. This the authors test by reducing the gene dosage of cyclin B and using putatively non-phosphorylatable versions of Spd-2 and Cnn. Both approaches to me appear quite problematic. The latter perturbation is hard to interpret given that whether these are indeed Cdk phosphosites that they have mutated is unknown and there are plenty of other possibilities how this might perturb protein function, as the authors' lack of success doing the same for gamma-tubulin illustrates. The former perturbation also lacks context. Does reducing cyclin B gene dosage reduce peak CDK activity or does it merely take longer to reach the same maximum, as appears to occur naturally as the cell cycle slows between embryonic cycles 11 and 13 (Edgar, Genes Dev 1994)? A more direct way to test their model would be to arrest the embryo in S phase (which in their model should lead to indefinite growth) or mitosis using suitable drugs/genetic perturbations. Is this not feasible in the fly system?
      2. Similarly critical is that centrosomal protein accumulation is accurately measured. I am not entirely convinced that this is so. If one takes their estimations of centrosome size at face value, then the space occupied by gamma-tubulin (slightly over 1 µm2 peak area according to Fig. S3) is significantly smaller than that occupied by Grip128 (4µm2). How is this possible if these form part of the same gamma-tubulin complex? This likely reflects the fact that the dynamics of many proteins is being assessed using transgenic reporters under the control of heterologous regulatory sequences (not all of which are fully functional, eg Polo), which could result in wildly inappropriate centrosomal protein levels. It may then not be a coincidence that the centrosomal domain for Grip128 (endogenously tagged) is larger than that for gamma-tubulin (transgene).
      3. Another concern is that centrosome size and integated signal intensity do not always match, as demonstrated by Grip71 (increasing as expected during centrosome maturation in cycle 13 based on fluorescence intensity but not area
      4. compare Figs. 1B and S1). A potential reason for this is that proteins are not uniformly distributed within centrosomes. For example, Polo and Spd2 are highly concentrated at centrioles. This impacts the ability to accurately measure protein accumulation based on 2D projections. Such inaccuracies likely will not affect estimation of when peak protein accumulation occurs, but may explain apparent differences in the kinetics of recruitment/dissociation of different components. Thus, the differences in the shape of the PCM client growth curves compared to those of Polo and Spd-2 (p6) may simply reflect the centriole concentration of the latter.

      Other points<br /> 4. In C. elegans much of Polo at centrosomes is apparently inactive, particularly in the vicinity of centrioles (Cabral, Dev Cell 2019). Knowing whether this is also the case in flies would seem like important information to have, particularly when comparing signal intensities across the cell cycle.<br /> 5. Is it clear that there is less Spd2/Cnn at centrosomes in Spd-2/Cnn 1/2 gene dosage embryos, as the authors assume?<br /> 6. Would it not be relevant to also examine Polo 1/2 dosage embryos?<br /> 7. Based on the authors model, Cdk phosphorylation first drives PCM accumulation, then at higher levels inhibits. Yet, their non-phosphorylatable Spd2 mutant exhibits not only a delayed decline in centrosomal levels, but also higher initial levels (Fig. 5B). If Cdk initially promotes Spd2 activity what is their explanation for this?

      Minor/discussion points

      1. p4 "In typical somatic cells the two mitotic centrosomes need to grow to approximately the same size, as mitotic centrosome size asymmetry can lead to asymmetric spindle assembly and so to defective chromosome segregation. How centrosome growth is regulated in somatic cells is unclear, but in early C. elegans embryos, mitotic centrosome size appears to be set by a limiting pool of the PCM-scaffolding protein SPD-2."<br /> The authors here conflate absolute and relative size. Relative size matters to avoid spindle asymmetries, and centriole involvement in PCM recruitment helps to prevent this (Zwicker et al., PNAS 2014). Absolute size, which is what the authors are concerned with in this manuscript, may be important for spindle scaling, but this is not the same thing.
      2. p5 "The centrosomal levels of Polo, Spd-2 and Cnn all started to increase at the start of S-phase, but whereas Cnn levels continued to rise and/or plateau as the embryos entered mitosis, the centrosomal levels of Polo and Spd-2 started to decrease before the entry into mitosis (Wong et al, 2021) (Figure 1A,B). Thus, the components of the mitotic PCM scaffold exhibit different growth kinetics, making it hard to use these proteins to define centrosome "size" at any particular point in the cell cycle."<br /> It is misleading and confusing for the reader to describe Polo and Spd2 as scaffold proteins as opposed to regulators of scaffold assembly. Presently Cnn is the only PCM protein demonstrated to have self-assembly/scaffolding properties based on the authors' own work (conduit, Dev Cell 2014; Feng, Cell 2017). There is little evidence that Polo and Spd2 form anything other than a nucleus for PCM growth.
      3. p7 "The centrosomal levels of Grip71, Grip75, Grip128, and Aurora A tended to increase steadily through most of NC13, whereas TACC, Msps and γ-tubulin exhibited a noticeable increase in their recruitment rate towards the end of S-phase, shortly before their recruitment levels peaked (compare NC13 graphs in Fig. 1B). This difference was also obvious if we used centrosome area as a measure of centrosome size (Fig. S1). We conclude that PCM client proteins can be recruited to centrosomes in at least two different ways."<br /> As discussed above apparent differences in kinetics may reflect limitations in the way protein accumulation is measured. It is hard to conceive of a reason why the Grips would display a different mode of protein accumulation from gamma-tubulin, nor is the idea of two different modes of protein accumulation picked up again later in the manuscript.
      4. Since the authors mention that the duration of S phase increases between cycles 11 and 13 (p9), are there any measures for the timing of the beginning/end of S phase in each cycle?
      5. One of the main findings in the landmark Woodruff paper from 2017 Cell paper was that PCM scaffold polymer could dynamically concentrate client proteins in the absence of any other factors, to an extent similar to that observed in vivo. This list did not include gamma-tubulin, which was later shown to require PLK1 phosphorylation of SPD-5 (Ohta, JCB 2021). However, it did include ZYG-9, the C. elegans ortholog of Msps. If client protein accumulation is an intrinsic property of the PCM scaffold, how do the authors explain that Msps departs prior to NEBD while Cnn continues to accumulate?
      6. p13 "The expression of the mutant proteins did not appear to dramatically perturb the centrosomal recruitment of γ-tubulin-GFP, except that the rate at which γ-tubulin-GFP left the centrosome as the embryos entered mitosis was reduced in both mutants compared to WT (Figure 7). This phenotype was subtle, but it was statistically significant, and it seems likely that the presence of large amounts of WT Spd-2 and Cnn in the mutant embryos (Figure 5A,F) would help to mask the potential severity of this phenotype."<br /> This does not quite make sense. Fig. 5 shows that Spd2 dissociation is significantly slowed in the mutant condition. If Spd2 drives gamma-tubulin accumulation (as Fig 6 shows), then the continued presence of Spd2 should prevent dissociation. Yet it apparently does not. Why?

      Other

      1. p3 and following. The reference for the authors' prior work on PCM recruitment (Wong et al, 2021) should probably be for the final, published article in EMBO J, not the 2021 preprint.
      2. Fig. 1. legend "Note that for technical reasons not all of the centrosome proteins could be followed for the full time period." Why not?
      3. Figs 4-6. Which cycle is being assessed here?
      4. Fig 6. Not plotted here is the rate of dissociation of gamma-tubulin, unlike eg in Fig 7. It is notable that both accumulation and dissociation appear to be slowed in the Spd2 1/2 gene dosage condition.
      5. Fig S1B. Some of the graphs in this figure are not labeled (based on Fig.1 presumably gamma-tubulin and Msps).
      6. Some of the data in the main figures, including the entirety of Figs. 2 and 3, could be moved to Supplemental to present a more crisp and accessible manuscript.
      7. While I sympathize with the authors needing to repeat entire sets of experiments I am not entirely sure it is appropriate to recycle entire sets of data from a previous publication of theirs (Cnn, Spd-2 and Polo recruitment kinetics, reproduced from Wong et al., EMBOJ 2022), since this manuscript is largely concerned with apparent differences between the kinetics of those components and the PCM client proteins now being analysed.

      Significance

      Control of organelle size has been an active field of research for many years for a large variety of cellular structures and in a range of experimental models. Here, Jordan Raff and colleagues examine the mechanisms underlying centrosome (PCM) size control in Drosophila syncytial embryos, building on their previous work (Wong, EMBOJ 2022) to propose a role for CDK in both promoting (at intermediate levels) and inhibiting (at high levels) PCM expansion.

      I found this a difficult manuscript to review, not only because the subject matter is complicated, but so is the writing. Having read and re-read the manuscript some clarity eventually emerges, but it shouldn't be that inaccessible. As for the authors' model I find it intriguing, but not fully supported by the data currently presented.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      The manuscript by Wong et al. investigates how cells regulate the increase in the size of the centrosomes (more specifically the size of the pericentriolar material or PCM) that occurs during preparation for mitosis. They use the Drosophila syncytial embryo as a model, focusing on nuclear cycles 11-13, during which cell cycle progression gradually slows. The authors find that centrosomes grow to a consistent size at each cycle by adjusting to the slowed cell cycle, reducing the growth rate and increasing the growth period. This adjustment is proposed to be regulated by the Cdk/Cyclin cell cycle oscillator. Curiously, Cdk/Cyclin activity seems to both promote and inhibit the increase in centrosome size, depending on whether its activity is moderate or very high, respectively. Both effects are proposed to depend on the phosphorylation of centrosome proteins by Cdk/Cyclin.

      1. While being comprehensive in the number and type of markers that are being analyzed, there is no analysis of the centrosome's MTOC activity. In my opinion this is missing since centrosome size alone is not necessarily indicative of its MTOC activity, but MTOC activity is what ultimately matters for its role during mitosis. For example, it was observed that centrosome size declines already before mitotic entry, but it is possible that centrosome MTOC activity does not (similar to differences in the timing of the decline of PCM scaffold vs PCM client proteins). While not strictly related to size control, centrosome activity is biologically more relevant than solely size. I would consider it optional, if the authors decide to talk only about centrosome size, but then it should be made clear that size here may not be the most relevant factor.
      2. The authors say that during NC13 PCM client proteins can be recruited in "at least two different ways" (p. 7), including a way (rapid increase before peak) that does not resemble PCM scaffold recruitment (steady increase during NC13). How can these two different ways and kinetics be determined by the same Cdk/Cyclin oscillator?
      3. I am puzzled by the conclusion that Cdk/Cyclin directly phosphorylates Spd-2 or Cnn at the sites used for mutagenesis. This cannot be concluded based on the presented data.
      4. Fig. 6: Doesn't the data show that Cnn does not affect the initial rate of g-tub recruitment, but only the later rapid recruitment shortly before mitosis? In contrast Spd-2 seems to affect the initial phase. This should be described more precisely. Again, I am wondering how this is compatible with direct regulation by a single oscillator, as suggested by the authors (see also point 2 above.
      5. I don't find the proposed model very convincing and not fully supported by the presented data.<br /> First, the recruitment kinetics of different centrosome proteins are not all the same, arguing against a simple relationship based on phosphorylation by Cdk/Cyclin. For example, kinases (or phosphatases) may be recruited (or displaced) by Cdk/Cylclin at the centrosome and then locally regulate binding or maintenance of certain centrosome proteins. This could explain profiles that do not display a steady change over time, as would be expected by direct regulation by Cdk/Cyclin.<br /> Second, it is not clear from the description in the text or from Fig. 8 how moderate Cdk/Cyclin activity can promote recruitment and high activity induce loss of proteins at centrosomes. In fact, the experiments with Spd-2 and Cnn phospho-mutants suggest that phosphorylations at the mutated sites also reduce centrosome binding during S phase (at moderate activity) and not only shortly before mitosis (at high activity), since alanine mutants of both Spd-2 and Cnn are increased at centrosomes also during S phase. The model seems to ignore this observation. If these sites are already phosphorylated to decrease centrosome binding in S phase, then what triggers the rapid decrease shortly before mitosis?
      6. Can the authors identify and mutate CdK/Cyclin dependent phospho-sites in centrosome proteins that promote centrosome recruitment at moderate Cdk/Cyclin activity? As an alternative to the "protein availability" model for regulation of centrosome size, the proposed model needs to explain how a steadily increasing activity (Cdk/Cyclin) can first induce growth and then turn growth off, when the desired size is reached. This is obvious in the "protein availability" model, where the available protein steadily decreases as centrosomes grow, but this is not at all obvious for an oscillator that behaves in the opposite way during the same period and that can only phosphorylate sites that decrease centrosome binding.

      Minor:

      1. The authors observe differences in the intensity profiles for different subunits of the gamma-tubulin complex. How do they explain this? Are they not in the same complex? The authors should mention and comment on this.
      2. The authors refer at various points in the manuscript to an "inverse-linear" relationship between S phase length and centrosome growth rate, but according to the graphs the rate does not change linearly.

      Significance

      This is an interesting manuscript that reaches somewhat different conclusions regarding centrosome size control when compared to previous studies in other organisms. In particular, work in C. elegans has proposed that centrosome growth regulation is controlled by the limited cytoplasmic availability of PCM building blocks, whereas the current study proposes a different model based on the activity of a cell cycle oscillator. The model system and approaches are well presented and the data is of good quality. The authors monitor a large number of centrosome markers, each with detailed quantifications of intensity and distribution over time during the different cycles. They also employ two different ways of quantifying centrosome size with similar results, making their quantifications more robust. While the authors include phospho-mutants in their analyses that presumably cannot be phosphorylated by Cdk/Cyclin, the study is largely descriptive. Still, the authors present interesting observations and propose the "oscillator model" an alternative to the "limited availability model" for the regulation of centrosome size, and perhaps that of other organelles. Assuming the authors can clarify inconsistencies and/or provide additional data to support the proposed model, this could be an important finding that expands cell biologists' understanding of organellar size control.

      I have expertise in centrosome biology and the role of centrosomes as MTOCs, as well as more general expertise regarding the function of the microtubule cytoskeleton in cell division and differentiation.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity):

      In their manuscript „Live-cell super-resolution nanoscopy reveals modulation of cristae<br /> dynamics in bioenergetically compromised mitochondria", Golombek et al. tested the effects of different mitochondrial toxins on cristae dynamics. The main focus of their work lies on live STED imaging, which they use to visualize cristae merging and splitting. They found swelling of mitochondria and reduced cristae density in response to most toxins, but cristae dynamics remained largely unaffected. Depletion of the membrane potential by administration of CCCP increased cyristae dynamics, while inhibition of ANT had a negative effect on cristae dynamics at least in a subset of mitochondria.

      1. The authors state that the used concentrations of mitochondrial toxins commonly result in a change in oxygen consumption. While this is believable, it is not guaranteed that the specific chemicals used for the experiments were working properly (freeze/thawing or simply incorrect storage or aliquotation may have an effect on the compounds). This is even more important in the case of results where no significant change after the administration of the toxins is seen. In Figure 5, the authors report no change in membrane potential after oligomycin administration, this is unexpected. I therefore suggest to include a supplementary figure, in which the functionality of the compounds is verified. This could be done by respiratory measurements (e.g. Seahorse). A Mito Stress Test was performed for Figure 6, but this was done using the Seahorse kit chemicals, which were probably different from the chemicals used in the microscopy experiments.

      Response: We appreciate the valid concerns of the reviewer in this point.

      A) In order to show the functionality of compounds which were used for performing our experiments including STED imaging, we now performed respiratory measurements employing the concentrations of mitochondrial toxins (Oligomycin A, CCCP, rotenone/antimycin A) which were used during imaging conditions as well as commercially available mitochondrial toxins (Oligomycin A, FCCP, rotenone/antimycin A) with respective concentrations used as a standard for the Mito stress Kit. The new figures are included in Fig S1A & B. HeLa cells treated with seahorse compounds or those used during imaging conditions showed similar results including basal, maximal and spare respiratory capacity. Further, in order to overcome the inefficiency of mitochondrial toxins employed, due to freeze/thaw cycles, we used fresh aliquots (stored at -20°C) as a general strategy. This is clearly observed by a drastic reduction of ΔΨm upon treating HeLa cells with CCCP, antimycin A as well as rotenone (Fig S6A & B). A reduction of mitochondrial ATP levels was also observed upon employing rotenone, antimycin A and oligomycin A confirming that active mitochondrial toxins were used. These experiments demonstrate that the mitochondrial toxins employed throughout our manuscript are functional as expected.

      New Figure S1A & B

      B) The Fig 6 (now Fig 5 due to Reviewer # 2, Point 7) respirometry experiments which initially employed seahorse compounds and BKA has now been replaced with new experiments where we used mitochondrial toxins similar to STED imaging. Needless, to say, the results are similar to what were observed with seahorse compounds. The new figures are replaced in Fig 5A & 5B.

      New Figure 5A & B

      C) Oligomycin A inhibits ATP synthase which results in decreased ATP synthesis as observed (Fig 4A & B). Further, oligomycin A is expected to hyperpolarise mitochondria (2). In Fig S6, despite some cells having more ΔΨm, there was no overall significant change when compared to untreated cells. Previous publications also show that there is no significant difference in ΔΨm upon treatment with oligomycin (1) demonstrating that the ΔΨm depends on the concentration of oligomycin, treatment time and cell type.

      1. Figure 1 would benefit from a more detailed description of merging/splitting events. Maybe a cartoon plus a zoomed in image of an exemplary event?

      Response: Thank you for the suggestion. In order to clearly explain/simplify the understanding of cristae merging and splitting events, we added a cartoon in Fig 1B. The green and magenta arrows show sites of imminent merging and splitting with the green and magenta asterisks representing them respectively in the subsequent frames. The zoomed in images in Fig1A (leftmost panel) are shown to the right as time-lapse images.

      New Figure 1B

      1. Could the reduced cristae density be an effect of mitochondrial swelling? It is curious that all toxins appear to have the same effect on mitochondrial architecture. What is the fait of an enlarged mitochondrion over time? Mitophagy? And does the percentage of enlarged mitochondria change with increasing treatment time?

      Response: Thank you for the comment.

      A) We agree that the reduced cristae density is due to mitochondrial swelling. We added the relevant text in the results section ‘Cristae structure is altered in a subset of mammalian cells treated with mitochondrial toxins’. Treatment of HeLa cells, with all the mitochondrial toxins mentioned, uniformly result around 50 % of mitochondria undergoing enlargement (Fig 2B). In enlarged mitochondria where the mitochondrial width is ≥ 650 nm, there is no change in cristae area occupied per mitochondria (Fig S3C & D) and as a result reduced cristae density (Fig 2H). Therefore, it indicates that reduced cristae density occurs due to mitochondrial enlargement.

      Figure 2B-F

      Figure S3C and D

      B) In order to address the fate of mitochondria with increasing time upon treatment with various mitochondrial toxins, we treated the HeLa cells for 4 hrs with mitochondrial toxins. Untreated cells maintained normal mitochondrial morphology while cells treated with various mitochondrial toxins displayed fragmented and swollen mitochondrial morphology. The new Fig S5 is included in the supplementary. Cristae morphology was abnormal displaying interconnected cristae in swollen mitochondria. Since mitochondrial fragmentation is already observed at 4 hours and accompanied by interconnected cristae, the number of cristae merging and splitting were severely reduced.

      Our imaging performed within 30 mins of addition of respective toxins overcomes the additional aberrancy of mitochondrial fragmentation which would not allow a reliable analysis of cristae dynamics as too few cristae would be visible within one mitochondrion.

      New Figure S5

      1. Figure 4C: How was the mitochondrial width determined in the LSM images? Especially in the perinuclear area it will be difficult to determine this parameter without the super-resolution provided by STED. Was this parameter determined manually for selected mitochondria? In the methods part it says that only a maximum of two mitochondria per cell were analyzed. How were these chosen? Was the process blinded?

      Response: Thank you for the comment. We could imagine the reason for the ambiguity in understanding.

      A) For LSM confocal images involving FRET-based microscopy to determine the ATP levels, we calculated the cell population as belonging to either normal or enlarged category. The confocal images of HeLa cells displayed clear separation of mitochondria even in the perinuclear area (representative images are shown in Fig 4A) and thus it was possible to measure the width of individual mitochondria. The methods section ‘FRET-based microscopy to measure ATP levels’ describes that ‘the cut off for swollen mitochondria was set to 650 nm in congruence with STED SR nanoscopy. If 85% of the mitochondrial population featured enlarged mitochondria, the cells were designated as swollen. Similarly, if 85% of the mitochondrial population featured mitochondria whose width was less than 650 nm, the cell was considered as having normal mitochondria’.

      Figure 4A

      B) The cristae morphology of various mitochondria is fairly uniform in individual cells. Thus, the mitochondria are representative of the individual cells. Therefore, in order to increase the coverage of various cells, we considered a maximum of two mitochondria from each cell which were randomly chosen. This part is modified in the methods section ‘Quantification of various parameters related to cristae morphology’ to make it clear. Thus, while the quantification of various parameters including dynamics involved individual mitochondria, various cells were classified as belonging to normal or enlarged category while measuring ATP levels.

      1. What is the average size of all mitochondria per cell? Is this addressed in Figure 2B or are only analyzed mitochondria included? Please carify. Were the mitochondria chosen for analysis representative for the respective cell?

      Response: The data obtained by super-resolution imaging of mitochondria is used for quantifying cristae dynamics which is a very challenging and time-consuming method done in a blind-manner. As mentioned in response 4B, the cristae morphology is fairly uniform in individual cells, therefore, we only included the mitochondria which were analysed for various cristae parameters in our analysis which are really huge data-sets already. Thus, the average size of individual mitochondria per cell are not represented while analysing images obtained with STED SR imaging. Please also check response 4B.

      1. explain the mt-Go-AT team2, what is GFP (green fluorescent protein) and OTP (?)

      Response: GFP is Green Fluorescent Protein and OFP is Orange Fluorescent protein and included in the revised text.

      1. the graphs show in principle, e.g. Fig.1B, 3B-E show events/mitochondrion as far as I understand, not per cristae.

      Response: Thank you for pointing this out. It is actually the average number of events per cristae per mitochondria. We have changed the Y-axis to events/cristae/mito in Fig 1C (previous 1B), Fig 3B-E and wherever applicable for other figures throughout the manuscript.

      Figure 1C

      Figure 3B-E

      1. I would recommend changing the legend of the x-axis of Fig.2B-F to mito-width (y-axis could be probability density function, PDF).

      Response: We have now changed the X-Axis to mito width (originally width) in Fig 2B-F. The Y-axis are still retained as percentage mitochondria where cells treated with few mitochondrial toxins do not show a gaussian distribution of mitochondrial width.

      Figure 2B-F

      Referees cross-commenting

      both expert opinions address similar concerns and therefore a revision should be requested

      Reviewer #1 (Significance):

      The study is thorough and the experiments and results are well described. Overall, however, it remains a descriptive study and does not provide mechanisms. There is also no discussion of how MMP-dependent proteins, such as Opa1, which was previously studied by the Reichert group, might be affected. For swelling mechanisms, the opening of the mitochondrial permeability transition pore was discussed. This could be tested using inhibitors, but perhaps not within the scope of this publication. Nevertheless, the information provided by the study is of interest to the bioenergetics community and should be made available.

      Response: Thank you for the overall inputs.

      We tested the processing of OPA1 forms and found that after 30 mins, only CCCP treatment led to the processing of long isoforms to short forms (Fig S6C). We now included in the discussion that it is possible that short OPA1-forms are correlative to increased cristae merging as well as splitting events upon treatment with CCCP.

      New Figure S6C

      Reviewer #2 (Evidence, reproducibility and clarity):

      Summary:<br /> The authors investigated cristae merging and splitting events using ultra-resolution STED. The goal was to test if cristae membrane remodeling is dependent on OXPHOS complexes, mitochondrial membrane potential (ΔΨm), and the ADP/ATP nucleotide translocator. To do this the authors utilized several mitochondrial toxins with known mechanisms of action. Interestingly, many changed overall cristae density but did not change the cristae remodeling events. Inhibition of ANT did change cristae morphology and cristae dynamics.

      Major Concerns

      1. Many conclusions and concepts need more clarification. For example, a major take home from the abstract is that various ETC inhibitors and protonophores reduce cristae density but not did not change cristae remodeling events. If cristae density is reduced, how can this occur without cristae remodeling events? Remodeling events need to be clearly defined in the introduction and abstract.

      Response: Thank you for pointing out this lack of sharpness in our terminology which indeed can cause a misunderstanding. To avoid this, we have now included ‘changes in cristae morphology’ as well as ‘dynamic merging and splitting events of cristae’ under the broader term cristae remodelling. Thus, we had changed the wording ‘cristae remodeling’ to cristae dynamics in the abstract and wherever appropriate in the manuscript text.

      The cristae morphology analysis showed no change in cristae area (Fig S3C) which was accompanied by mitochondrial enlargement. Therefore, cristae density was reduced. For the purpose of clarity, we added a sentence in the introduction section while giving a peek into our results that ‘cristae dynamic events are ongoing despite reduced cristae density’. In addition, we have now included in the results section the following statement: ‘Cristae membrane remodeling has been used to describe cristae dynamic events (i.e. cristae merging and splitting) as well as overall changes in cristae morphology within a single mitochondrion in this manuscript’.

      Figure S3C and D

      1. Other interpretations are also unclear such as how ETC inhibitors which reduce ATP levels did not impact cristate remodeling events, yet inhibiting ATP/ADP exchange did greatly impact this phenomenon. It seems likely that the inhibition of ANT has nothing to do with ATP/ADP exchange since most of the ETC inhibitors no doubt greatly impact overall ATP/ADP exchange. This interpretation needs clarification.

      Response: We agree that further clarification is needed, in particular to explain why ATP/ADP exchange is actually ongoing even when OXPHOS inhibitors are applied and to explain why reduced ATP levels do not mean that there is no ATP/ADP exchange occuring. Treatment of HeLa cells with various mitochondrial toxins inhibiting the function of OXPHOS complexes leads to decreased ATP levels due to ongoing ATP consumption within the cell (Fig 4). One should also consider that two things can and do happen when most of these toxins are applied regarding ATP exchange. First, the ATPase can act in reverse mode which is a (partial) compensatory mechanism to restore ΔΨm and which will further decrease ATP levels (Note: not in the presence of oligomycin). Second, under these conditions ADP/ATP exchange is still ongoing in order to transport ATP derived from glycolysis in the cytosol to the mitochondrial matrix which also causes an (partial) compensatory increase in membrane potential. After ATP import ATP is hydrolysed to ADP for reverse proton pumping via the F1FO-ATPase or alternatively by the F1-part alone without proton pumping. In all these cases it is essential and possible to exchange ADP with ATP constantly. Therefore, the overall exchange of ADP and ATP is not necessarily grossly expected to be different when compared to untreated cells (due to compensatory glycolysis and subsequent ATP import and hydrolysis in the matrix). On the other hand, BKA treatment which clearly impairs the exchange of ADP and ATP will lead to a completely different situation compared to only treating with OXPHOS inhibitors. With BKA the mitochondrial matrix cannot anymore be resupplemented with ATP derived from glycolysis and metabolite flux is grossly hampered. Consistent with this a strong reduction in ΔΨm and oxygen consumption is accompanied with BKA treatment (Fig. 5AB & SFig 7F). Thus, w.r.t cristae dynamic events, in the time-frame we used for imaging, a reduction of ATP levels does not impede occurrence of cristae merging and splitting events while BKA treatment does (Fig S7). We discuss this indeed interesting and unexpected finding in the discussion section. We propose that rather ongoing metabolite flux (ATP/ADP exchange) is critical for maintaining cristae dynamics and blocking it is detrimental for it. We adapted the discussion in this direction to make it more clear.

      Figure S7A, B and D

      1. Why did the authors wait 30 min to image after the addition of mitochondrial toxins? I would have guessed there is a more rapid change in response to these inhibitors. Is there is a chance he authors missed the most dramatic events?

      Response: Since we were inclined to observe early responses, cells were imaged within the first 30 mins after addition of the respective mitochondrial toxins (Please see methods ‘cell culture transfection and mitochondrial toxin treatment’). Thus, to answer this question we want to emphasize that we did not wait 30 minutes but we restricted our time frame of analysis to 30 min. Therefore, we think that we did not miss out on any rapid changes occurring early on. Regarding this point, Reviewer #1 (Query 3) asked for responses at a later time-point. Please read the Reviewer #1, response 3B.

      1. How do these mitochondrial toxins that are known to cause mitochondrial swelling not induce changes in cristate density?

      Response: Thank you for the question. Probably, there is a misunderstanding. In Fig S3E, we clearly show that as the mitochondrial width increases in cells after treatment with mitochondrial toxins, there is a clear decrease in cristae density. In fact, the reduced cristae density is observed exclusively in enlarged mitochondria. Figure S3E-I

      5. It's interesting that inhibition of the ANT translocator by BKA treatment led to increased percentage of mitochondria with abnormal cristae morphology. It's accepted that inhibition of ANT profoundly reduces mitochondrial swelling. Do the authors have any data suggesting that abnormal cristae morphology actually is a mechanism for reducing cell death events such as permeability transition? Did the authors utilize cyclosporin A concomitantly with any of the mitochondrial toxins?

      Response: This is a very interesting question! As the reviewer might be aware, there is evidence connecting cristae remodelling to induction of apoptosis (3). Cristae transitioned to a highly interconnected state after tBID treatment within minutes. However, it is unclear what is the contribution of cristae dynamics in this regard. Within 30 mins, there were no visual signs of cell death in our experiments as observed under a microscope. Hence, we did not use cyclosporin A in our experiments. In our opinion, this question will form part of a very interesting future study and is currently beyond the scope of this manuscript.

      1. Are the authors confident in the data given many of the experiments utilized quantification of 10-20 mitochondria? How are you sure this sampling is sufficient for phenomenon being studied?

      Response: Please see Reviewer 1, Response 4B. As pointed in the response to reviewer #1, the cristae morphology is fairly uniform in individual cells. Therefore, in order to maximise the cell population covered, we randomly used a maximum of two mitochondria from each cell. In addition, we included cristae analysis from at least three biological replicates in order to observe the reproducibility of the data. Taking these factors into consideration, we are confident that our results reflect a sufficient sample size. Further, we would like to point out while our group performs STED super-resolution imaging routinely, the quantification of cristae merging and splitting events done in a blind yet manual manner is a really laborious and time-consuming process. In the future, we are also looking to optimise this at least in a semi-automated manner.

      1. Figure 4 and 5 merely confirm current dogma and don't really contribute to the overall conclusions and can be moved to supplemental data.

      Response: We agree that Fig 5 is confirming to the current dogma. Therefore, we moved it to Fig S6. Regarding Fig 4, we would like to highlight that there is a decrease of ATP levels before mitochondria enlarge. Thus, we would like to retain it as part of the main figure.

      1. It's interesting that BKA dose dependently decreased ATP-linked respiration and all doses limited maximal respiratory capacity. It would be interesting to know if the BKA normal vs. abnormal mitochondria have differential membrane potential?

      Response: Thank you for the interesting question. Overall, BKA treatment leads to a significant decrease of ΔΨm in the whole cell population (Fig S7). Further, the abnormal cristae morphology is only seen in one-third of the population of mitochondria (Fig shown in Response 2). Thus, a drop in ΔΨm seems to be a very early response upon exposure to BKA and independent of cristae morphology. An ideal experiment to address this question would be to image cristae dynamics and ΔΨm using super-resolution imaging which is challenging according to the state-of-art and available chemicals.

      Figure S7E and F

      1. Overall, this is an interesting study and seems appropriately performed but the results and conclusions are unclear. More discussion should include physiological relevance and impact and how this data influences previous work. Some physiological perturbations beyond the mitochondrial toxins and or utilization of genetic models would strengthen the interpretation and overall impact.

      Response: Thank you. We added an OPA1 blot showing the different L-OPA1 and S-OPA1. (Reviewer #1, response in significance section) where we observed that S-OPA1cleavage is selectively enhanced in CCCP-treated cells which could be correlated with enhanced cristae dynamics. We also included these results in the main text.

      New Figure S6C

      Referees cross-commenting

      Yes, I conclude that given the significant overlap in reviwer comments and general need for clarification of concepts and data that a revision is in order.

      Reviewer #2 (Significance):

      Overall, a highly specialized study with audience limited to mitochondriacs. Although, I'll note tis is a hot area of study and there is high interest in the field. Some of the data interpretation is difficult to understand and overall more context is needed to explain the results, impact and relevance. Defining exactly what a cristae remodeling event is and how this differs from cristae density and how the two aren't directly connected is unclear.

      Review by a mitochondrial biologist specializing in mitochondrial signaling and connection to physiology.

      References:

      1. Baker MJ, Lampe PA, Stojanovski D, Korwitz A, Anand R, et al. 2014. Stress-induced OMA1 activation and autocatalytic turnover regulate OPA1-dependent mitochondrial dynamics. EMBO J 33: 578-93
      2. Farkas DL, Wei MD, Febbroriello P, Carson JH, Loew LM. 1989. Simultaneous imaging of cell and mitochondrial membrane potentials. Biophys J 56: 1053-69
      3. Scorrano L, Ashiya M, Buttle K, Weiler S, Oakes SA, et al. 2002. A distinct pathway remodels mitochondrial cristae and mobilizes cytochrome c during apoptosis. Dev Cell 2: 55-67
    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary:

      The authors investigated cristae merging and splitting events using ultra-resolution STED. The goal was to test if cristae membrane remodeling is dependent on OXPHOS complexes, mitochondrial membrane potential (ΔΨm), and the ADP/ATP nucleotide translocator. To do this the authors utilized several mitochondrial toxins with known mechanisms of action. Interestingly, many changed overall cristae density but did not change the cristae remodeling events. Inhibition of ANT did change cristae morphology and cristae dynamics.

      Major Concerns

      1. Many conclusions and concepts need more clarification. For example, a major take home from the abstract is that various ETC inhibitors and protonophores reduce cristae density but not did not change cristae remodeling events. If cristae density is reduced, how can this occur without cristae remodeling events? Remodeling events need to be clearly defined in the introduction and abstract.
      2. Other interpretations are also unclear such as how ETC inhibitors which reduce ATP levels did not impact cristate remodeling events, yet inhibiting ATP/ADP exchange did greatly impact this phenomenon. It seems likely that the inhibition of ANT has nothing to do with ATP/ADP exchange since most of the ETC inhibitors no doubt greatly impact overall ATP/ADP exchange. This interpretation needs clarification.
      3. Why did the authors wait 30 min to image after the addition of mitochondrial toxins? I would have guessed there is a more rapid change in response to these inhibitors. Is there is a chance he authors missed the most dramatic events?
      4. How do these mitochondrial toxins that are known to cause mitochondrial swelling not induce changes in cristate density?
      5. It's interesting that inhibition of the ANT translocator by BKA treatment led to increased percentage of mitochondria with abnormal cristae morphology. It's accepted that inhibition of ANT profoundly reduces mitochondrial swelling. Do the authors have any data suggesting that abnormal cristae morphology actually is a mechanism for reducing cell death events such as permeability transition? Did the authors utilize cyclosporin A concomitantly with any of the mitochondrial toxins?
      6. Are the authors confident in the data given many of the experiments utilized quantification of 10-20 mitochondria? How are you sure this sampling is sufficient for phenomenon being studied?
      7. Figure 4 and 5 merely confirm current dogma and don't really contribute to the overall conclusions and can be moved to supplemental data.
      8. It's interesting that BKA dose dependently decreased ATP-linked respiration and all doses limited maximal respiratory capacity. It would be interesting to know if the BKA normal vs. abnormal mitochondria have differential membrane potential?
      9. Overall, this is an interesting study and seems appropriately performed but the results and conclusions are unclear. More discussion should include physiological relevance and impact and how this data influences previous work. Some physiological perturbations beyond the mitochondrial toxins and or utilization of genetic models would strengthen the interpretation and overall impact.

      Referees cross-commenting

      Yes, I conclude that given the significant overlap in reviewer comments and general need for clarification of concepts and data that a revision is in order.

      Significance

      Overall, a highly specialized study with audience limited to mitochondriacs. Although, I'll note tis is a hot area of study and there is high interest in the field. Some of the data interpretation is difficult to understand and overall more context is needed to explain the results, impact and relevance. Defining exactly what a cristae remodeling event is and how this differs from cristae density and how the two aren't directly connected is unclear.

      Review by a mitochondrial biologist specializing in mitochondrial signaling and connection to physiology.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      In their manuscript „Live-cell super-resolution nanoscopy reveals modulation of cristae<br /> dynamics in bioenergetically compromised mitochondria", Golombek et al. tested the effects of different mitochondrial toxins on cristae dynamics. The main focus of their work lies on live STED imaging, which they use to visualize cristae merging and splitting. They found swelling of mitochondria and reduced cristae density in response to most toxins, but cristae dynamics remained largely unaffected. Depletion of the membrane potential by administration of CCCP increased cyristae dynamics, while inhibition of ANT had a negative effect on cristae dynamics at least in a subset of mitochondria.

      Major comments

      • The authors state that the used concentrations of mitochondrial toxins commonly result in a change in oxygen consumption. While this is believable, it is not guaranteed that the specific chemicals used for the experiments were working properly (freeze/thawing or simply incorrect storage or aliquotation may have an effect on the compounds). This is even more important in the case of results where no significant change after the administration of the toxins is seen. In Figure 5, the authors report no change in membrane potential after oligomycin administration, this is unexpected. I therefore suggest to include a supplementary figure, in which the functionality of the compounds is verified. This could be done by respiratory measurements (e.g. Seahorse). A Mito Stress Test was performed for Figure 6, but this was done using the Seahorse kit chemicals, which were probably different from the chemicals used in the microscopy experiments.
      • Figure 1 would benefit from a more detailed description of merging/splitting events. Maybe a cartoon plus a zoomed in image of an exemplary event?
      • Could the reduced cristae density be an effect of mitochondrial swelling? It is curious that all toxins appear to have the same effect on mitochondrial architecture. What is the fait of an enlarged mitochondrion over time? Mitophagy? And does the percentage of enlarged mitochondria change with increasing treatment time?
      • Figure 4C: How was the mitochondrial width determined in the LSM images? Especially in the perinuclear area it will be difficult to determine this parameter without the super-resolution provided by STED. Was this parameter determined manually for selected mitochondria? In the methods part it says that only a maximum of two mitochondria per cell were analyzed. How were these chosen? Was the process blinded?
      • What is the average size of all mitochondria per cell? Is this addressed in Figure 2B or are only analyzed mitochondria included? Please carify. Were the mitochondria chosen for analysis representative for the respective cell?

      Minor comments

      1. explain the mt-Go-AT team2, what is GFP (green fluorescent protein) and OTP (?)
      2. the graphs show in principle, e.g. Fig.1B, 3B-E show events/mitochondrion as far as I understand, not per cristae.
      3. I would recommend changing the legend of the x-axis of Fig.2B-F to mito-width (y-axis could be probability density function, PDF).

      Referees cross-commenting

      both expert opinions address similar concerns and therefore a revision should be requested

      Significance

      The study is thorough and the experiments and results are well described. Overall, however, it remains a descriptive study and does not provide mechanisms. There is also no discussion of how MMP-dependent proteins, such as Opa1, which was previously studied by the Reichert group, might be affected. For swelling mechanisms, the opening of the mitochondrial permeability transition pore was discussed. This could be tested using inhibitors, but perhaps not within the scope of this publication. Nevertheless, the information provided by the study is of interest to the bioenergetics community and should be made available.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity):

      Major comments<br /> In the paper "Microtubules under mechanical pressure can breach dense actin networks", the authors showed clear evidence that pressure plays an important role in microtubule breaching into dense actin networks using elegant in vitro reconstitution assays. They have argued that the pressure results from polymerization force of microtubules, which builds up when microtubules are immobilized in the opposite end of breaching, by the means of actin microtubule crosslinking factor Tau.

      Authors answer:

      We thank the reviewer for his/her positive comments on our manuscript.

      It would definitely be interesting to see lack of breaching in the presence of crosslinking deficient Tau construct in order to rule out the off -target effect of Tau on microtubule and actin architecture which may possibly facilitate breaching.

      Authors answer:

      This is an interesting suggestion. Unfortunately, we do not have in hand such crosslinking deficient Tau construct. However, please note that we showed two independent ways to demonstrate the role of pressure. One is indeed by crosslinking microtubule to actin bundle with Tau, but the other is by blocking the two opposite ends of microtubules with two dense actin networks. So, we think our conclusion about the role of pressure is solid.

      The authors have also observed microtubule breaching into dense actin networks in living cells. However, in Figure 1C, better cell/ image processing might have been chosen to increase the visibility of actin structures that microtubules encounter on their way to breaching. In Figure S1D, for example, the similar actin structures in lamellipodia are very nicely visible.

      Authors answer:

      We apologize but we don’t understand reviewer’s comment. In figure 1C images of actin networks are shown in black and white and are more visible than in figure S1D where they are shown in magenta and overlaid with microtubules. In any case, we increased the contrast of images to make fine actin structures at the cell edge clearer.

      It is also interesting that on Figure 6A, actin bundles look different than the rest of the figures on the paper. It almost looks like actin bundles become branched, whereas in the other Figures actin bundles are either singular or two-three bundles joined together at the point very close to the edge of micropatterned lipid bilayer.

      Authors answer:

      This is correct. In this experiment several bundles co-aligned. As mentioned by the reviewer this could also be visible in other conditions without Tau (such as in Figure 4E), and, as shown below, this structure of bundle was not visible in all fields we looked at. So we don’t think this structure is responsible for the changes we measured in the ability of microtubules to penetrate the actin network in the presence of Tau.

      Minor comments<br /> In the legend of Figure 4E, it should be written "white arrow" instead of "yellow arrow".<br /> In the Results section "crosslinking between microtubules and actin bundles increase piercing frequency", in the sentence number 7, it should be written "backwards" instead of "reaward".

      Authors answer: We modified the text and legend according to the reviewer suggestions.

      Reviewer #1 (Significance):

      The experimental setup of the paper is quite significant in the field, given the difficulty of observing dynamics of dense cytoskeletal structures in living cells. Moreover, the paper gives insight into how microtubule behavior can vary depending on different morphological states of actin network.

      Authors answer: We thank the reviewer for his/her overall very positive feedback on our manuscript.

      Reviewer #2 (Evidence, reproducibility and clarity):

      The authors developed a novel in vitro system to investigate the interaction of dynamic microtubules with the F-actin network. While this system does produce some interesting results, it is unclear how exactly this replicates or explains what might happen near a cell's leading edge. There is a limited characterization of the produced F-actin networks. For example, it is unclear to what extent the F-actin networks are similar or different to cell lamellipodial networks. What is the density / expected mesh size of these networks and could that be varied / manipulated? The bottomline observation that microtubules can grow into F-actin networks if they have nowhere else to go does not seem particularly ground-breaking, and the discussion is very shallow. Overall writing could be improved; there are lots of typos and grammatical inconsistencies. The second paragraph of the introduction is a bit convoluted.

      Authors answer:

      We thank the reviewer for his/her comments. Figure 1 was used to illustrate the behavior of microtubules encountering actin networks in cells and the fact that they struggle to penetrate actin network. This is only a way to argue that the penetration of actin network is a relevant question, that cannot be easily addressed in cells. However, it is correct that our in vitro systems, as it is the case for all in vitro reconstituted systems, cannot tend exactly to reproduce a lamellipodial cellular network. But it offers a better way to modulate actin network architecture. We have used in vitro systems to characterize the different behavior of microtubules when they encounter dense actin networks in different conditions, guided or not by actin bundles, constraint or not at the two ends.

      The observation that microtubule can penetrate actin network when pressurized might not be “ground breaking”, still it contradicts previous works showing that microtubule under pressure tend to depolymerize (Janson et al, J Cell Biol, 2003), which would obviously prevent them from penetrating actin networks. So, our conclusion was somehow unexpected.

      We found important to discuss the fact that although the microtubule polymerizing forces is sufficient to breach dense actin network, it must be counteracted by another mechanism immobilizing microtubules. This means that in cells, expression level of actin-microtubule crosslinker modulate the penetration of microtubule into the lamellipodium.

      However, we agree that the second paragraph of the introduction is not absolutely necessary and removed it.

      Specific comments:

      Fig. 1 seems a bit anecdotal. The authors revisit an observation that has been made before. I can see how it is used as rationale for the in vitro system, but not sure that this adds much to the overall story. Clearly different cell types are different, but without some sort of quantification this remains meaningless. It should also be noted in the discussion maybe that there are large differences between cells in 2D and 3D. Microtubules much more frequently grow to the cell edge compared with 2D (see Akhmanova SLAIN2 paper from some years ago).

      Authors answer:

      We agree with these comments. Indeed, Figure 1 is used only as an illustration of the behavior of microtubules encountering actin network in cells. As the reviewer said, microtubule penetration and actin architectures will both vary a lot from one cell type to another. So we believe that quantification for these particular cases will not extend the illustrative purpose of this figure where it is already clear that some microtubules can penetrate and other can’t.

      Fig. 2: While Arp2/3 certainly promotes branched F-actin networks, from the data provided it is not clear to me to what extent the produced F-actin networks replicate F-actin organization at the cell edge. If this a the point the authors are trying to make, the ultrastructure of their in vitro networks needs some additional characterization. As far as it is possible to discern from the data provided, the F-actin meshwork on the stripes in E looks pretty much identical in both panels (and not really like a dendritic network that in a cell also would have a certain polarity with barbed ends facing out), and the bundles on the left don't look like anything that normally occurs in a cell.

      Authors answer:

      We also agree with these comments. The networks we assembled are not lamellipodial-like networks, there are branched network of various densities, with or without bundles. It is true that bundles of filaments do not grow out of lamellipodial network in cells. However, bundles of aligned and linear filaments exist in cells, in the form of radial fibers or transverse arcs, along which microtubule tend to align. And these structures might guide microtubules toward cell protrusions, as it is the case in growth cone for example.

      Fig. 4 It is unclear what is going on here. Given that without F-actin bundles, polymerizing microtubules are freely moving around, it does not come as a surprise that they would never penetrate the F-actin network because as the authors correctly state the growing end will push back from the barrier. So, then why do they sometimes penetrate when bundles are present? In 4A it appears that microtubule growth into f-actin only happens once the microtubule minus ends gets stuck between F-actin bundles on the other side. 4D is the same as 4A; so that makes me think this really does not occur that often. Does the microtubule plus end only penetrate the F-actin meshwork when the minus end gets stuck on the other side? This seems important and also means microtubule penetration may not have anything to do with the F-actin network architecture at the plus end. This needs to be quantified.

      Authors answer:

      This is perfectly correct. In figure 4 the two actin networks are distant, and the microtubules only rarely penetrate them because they are rarely in contact with them at both ends. This occurs only when bundles orient microtubules perpendicular to the edges of the actin network, since in this configuration the distance between the two actin networks is shorter. Hence our motivation to bring actin networks closer to each other in figure 5.

      Fig. 5 I guess that sort of solves my confusion with Fig. 4. The quantification graphs in 5B and 5C are flipped with respect to the figure legend (?).

      Authors answer:

      Indeed, in this figure we distinguished the role of pressure (when both microtubule ends are in contact with actin networks) and the role of alignment with actin bundles. And found that the presence of bundles is useless and that only pressure matters.

      I understand the rationale for not considering microtubules that grow at a shallow angle, but there does not seem to be that much of a difference between 5B and 5D. Wouldn't a better quantification simply compare microtubules that contact F-actin at both ends compared with microtubules were the minus end is free. In this case, I would expect a very large difference in penetration.

      Authors answer:

      This is also correct. The difference is so important that when one end is free the microtubule never penetrate. We mention it in the text but did not plot these data. This is why we measured only microtubule with both ends contacting an actin network and did not consider the one at shallow angles.

      We added the illustration of the condition with short distance and actin bundles (shown below) to make this more clear in the figure.

      The small difference between 5B and 5D shows that by eliminating those microtubules there is no more difference between the conditions with or without bundles. And thus that their contribution in favoring microtubule penetration was to favor optimal orientation to get pressurized at the two ends rather than offering a sort of favorable network organization at their base. However, we agree with the reviewer that the absence of difference between the two populations, with or without actin bundle, when considering only microtubule interacting with actin at angles higher than 30° is not quite striking. We tested all angles (see below) and found that actually the absence of difference is more obvious when considering microtubules interacting with more than 60°. And the analysis of angle distribution, now reported in Figure 5D, showed that in both conditions most microtubules interact with more than 60°, so we only exclude few outliers by considering those that interact with more than 60°. So we changed the presentation of our data in Figure 5C by changing the threshold from 30 to 60°.

      Do microtubules under pressure ever bend/buckle in this in vitro situation. As the authors state, in cells, that happens frequently. This difference is interesting. Why?

      In vitro microtubules buckle homogeneously between their two ends. These long buckling wavelengths are not very spectacular. In cells, microtubules are crosslinked to actin filaments or other structures over shorter distances (see quantification below). This leads to buckling with shorter wavelength, which is more striking.

      It is customary to refer to polymerized actin as F-actin.

      The supplementary videos are not referenced in the manuscript.

      Authors answer:

      We apologize and have now referenced the supplementary video in the manuscript.

      Reviewer #2 (Significance):

      The manuscript describes results from a novel assay to study interactions between F-actin networks and dynamic microtubules in vitro. While of interest to a specialized audience, the overall finding that microtubules can grow into an F-actin meshwork is somewhat incremental especially because of the limited characterization of the F-actin networks used. It remains unclear to what extent this is relevant to a physiological context in cells.

      My field of expertise is related to cytoskeleton dynamics and quantitative microscopy in live cells.

      Authors answer:

      Although intuitive, the demonstration that the density of actin network can prevent microtubule penetration is novel. More importantly, the demonstration that anchoring of microtubule is sufficient to increase the pressure to such a point that microtubule can then penetrate those networks is also novel and significant to appreciate when and how they do so in cells.

      Reviewer #3 (Evidence, reproducibility and clarity):

      In this paper, the authors present an in vitro assay designed to explore the dynamic interaction between growing microtubules and pre-existing actin networks. Notably, the presence of linear actin bundles facilitated the movement of polymerizing microtubules along actin filaments. When microtubules were immobilized to two spatially separated actin networks, they exhibited the ability to breach and penetrate dense actin meshworks. This penetration was attributed to the mechanical pressure generated by microtubule polymerization. The authors tested tau as a microtubule-actin crosslinking protein in this process and found that tau promoted microtubule penetration into dense actin meshwork. Although the findings in this paper are potentially significant, the work is still in its preliminary stage and the scope is limited.

      Authors answer:

      We thank the reviewer to summarize properly the main findings of our manuscript.

      1. The authors observed that the inclusion of tau, a microtubule-associated protein known for its role in promoting microtubule polymerization, significantly facilitated microtubule penetration into dense actin meshworks. This enhancement is likely attributed to tau's ability to promote microtubule polymerization, generating stronger forces within the microtubules that enable them to breach the actin meshworks. To validate the involvement of the crosslinking function in the process, the authors should explore the effects of other microtubule-actin crosslinking proteins in their assay.

      Authors answer:

      We thank the reviewer for this interesting suggestion regarding the role of Tau in our experiments. To address this comment, we have analyzed the rate of growth in our experiments in presence and absence of Tau (see quantification below). We found that the construction of Tau we used reduced microtubule growth rate. Therefore, we believe that microtubule growth was not responsible for the improved penetration of microtubule in dense actin networks in our assay, and that it was rather the crosslinking ability of Tau that played a significant role.

      1. The paper highlights the importance of anchoring both ends of microtubules to two adjacent actin networks for successful penetration into the actin meshworks. However, the precise mechanisms by which these microtubule ends are anchored to actin filaments are not elaborated upon. Providing detailed insights into this anchoring process would enhance the readers' comprehension of the experimental setup and its relevance to the observed results.

      Authors answer:

      We apologize for this lack of clarity. We don’t think that microtubule ends are “anchored” to the actin network. They are simply embedded into it. This embedding prevents them from moving rearward and thus lead to pressure increase as they polymerize.

      1. Additional information on the experimental methods is warranted to improve the reproducibility and clarity of the study. Specifically, the authors should elucidate the process through which nucleation-promoting factors were grafted onto lipid bilayers. This detail is crucial for researchers seeking to replicate or build upon the study's findings.

      Authors answer:

      We apologize for this lack of clarity. There was indeed an error in our description of SUV preparation with lipid-biotin. We have now revised our material and method section. In particular we have described more accurately the various steps we used to micropattern WA-streptavidin onto lipid-biotin.

      1. In Fig. 5D, the authors observed no significant difference in the breaching probability between microtubules that contacted the actin meshwork at an angle higher than 30°, with or without actin bundles. To ensure a better comparison, it is advisable to focus on quantifying the microtubules that are contacting two actin meshworks at both ends (the immobilized microtubules), as they would have similar probabilities of being pressurized by their growth. Moreover, further justification is required to explain the choice of 30° as the threshold angle and its significance in the context of microtubule behavior.

      Authors answer:

      We thank the reviewer for this comment. We apologize for the confusion. The quantification we made is precisely the one described by the reviewer. We made this more clear by adding further illustration of the two conditions and the measurement made.

      1. Fig. 5C appears to depict the "Distribution of the angle of the interaction of microtubules in the presence (10nM of Arp2/3 complex) or absence (100 nM of Arp2/3 complex) of actin bundles" instead of the "proportion of microtubules piercing the branched actin network." The alphabet labels in the figure should be updated accordingly. Additionally, the authors should clarify whether a comparison was conducted between the means of the angles in the two conditions and whether any observed differences were statistically significant.

      Authors answer:

      We apologize for this confusion. We updated the figure legend in which 5C and 5D were inverted.

      1. Investigating the potential significant difference in the mean interaction angles between the absence and presence of actin bundles would be intriguing. The presence of actin bundles might indeed influence the interaction angle or contact position, potentially increasing penetration frequency. This insight would further enrich the findings and provide valuable context for understanding the interplay between microtubules and actin networks.

      Authors answer:

      We apologize for this confusion. We now report the statistical difference. And indeed, it accounts for the difference it the penetration frequency, as shown by the absence of difference when we consider only microtubules that are more or less perpendicular to the network. This is indeed one of the most significant conclusion of our work. We added some schematics to make this clearer.

      1. More comprehensive information about the statistical analyses should be provided. This'd be important for the validity and reliability of the study's conclusions.

      Authors answer:

      We apologize for this lack of clarity. The statistical analysis we performed were not described in the Materials and Methods section but in each figure legend.

      Reviewer #3 (Significance):

      The work represents an advance in understanding the mechanism by which microtubules navigate dense actin meshworks.

      Authors answer:

      We thank the reviewer for this positive evaluation of our work.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      In this paper, the authors present an in vitro assay designed to explore the dynamic interaction between growing microtubules and pre-existing actin networks. Notably, the presence of linear actin bundles facilitated the movement of polymerizing microtubules along actin filaments. When microtubules were immobilized to two spatially separated actin networks, they exhibited the ability to breach and penetrate dense actin meshworks. This penetration was attributed to the mechanical pressure generated by microtubule polymerization. The authors tested tau as a microtubule-actin crosslinking protein in this process and found that tau promoted microtubule penetration into dense actin meshwork. Although the findings in this paper are potentially significant, the work is still in its preliminary stage and the scope is limited.

      1. The authors observed that the inclusion of tau, a microtubule-associated protein known for its role in promoting microtubule polymerization, significantly facilitated microtubule penetration into dense actin meshworks. This enhancement is likely attributed to tau's ability to promote microtubule polymerization, generating stronger forces within the microtubules that enable them to breach the actin meshworks. To validate the involvement of the crosslinking function in the process, the authors should explore the effects of other microtubule-actin crosslinking proteins in their assay.
      2. The paper highlights the importance of anchoring both ends of microtubules to two adjacent actin networks for successful penetration into the actin meshworks. However, the precise mechanisms by which these microtubule ends are anchored to actin filaments are not elaborated upon. Providing detailed insights into this anchoring process would enhance the readers' comprehension of the experimental setup and its relevance to the observed results.
      3. Additional information on the experimental methods is warranted to improve the reproducibility and clarity of the study. Specifically, the authors should elucidate the process through which nucleation-promoting factors were grafted onto lipid bilayers. This detail is crucial for researchers seeking to replicate or build upon the study's findings.
      4. In Fig. 5D, the authors observed no significant difference in the breaching probability between microtubules that contacted the actin meshwork at an angle higher than 30{degree sign}, with or without actin bundles. To ensure a better comparison, it is advisable to focus on quantifying the microtubules that are contacting two actin meshworks at both ends (the immobilized microtubules), as they would have similar probabilities of being pressurized by their growth. Moreover, further justification is required to explain the choice of 30{degree sign} as the threshold angle and its significance in the context of microtubule behavior.
      5. Fig. 5C appears to depict the "Distribution of the angle of the interaction of microtubules in the presence (10nM of Arp2/3 complex) or absence (100 nM of Arp2/3 complex) of actin bundles" instead of the "proportion of microtubules piercing the branched actin network." The alphabet labels in the figure should be updated accordingly. Additionally, the authors should clarify whether a comparison was conducted between the means of the angles in the two conditions and whether any observed differences were statistically significant.
      6. Investigating the potential significant difference in the mean interaction angles between the absence and presence of actin bundles would be intriguing. The presence of actin bundles might indeed influence the interaction angle or contact position, potentially increasing penetration frequency. This insight would further enrich the findings and provide valuable context for understanding the interplay between microtubules and actin networks.
      7. More comprehensive information about the statistical analyses should be provided. This'd be important for the validity and reliability of the study's conclusions.

      Significance

      The work represents an advance in understanding the mechanism by which microtubules navigate dense actin meshworks.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      The authors developed a novel in vitro system to investigate the interaction of dynamic microtubules with the F-actin network. While this system does produce some interesting results, it is unclear how exactly this replicates or explains what might happen near a cell's leading edge. There is a limited characterization of the produced F-actin networks. For example, it is unclear to what extent the F-actin networks are similar or different to cell lamellipodial networks. What is the density / expected mesh size of these networks and could that be varied / manipulated? The bottomline observation that microtubules can grow into F-actin networks if they have nowhere else to go does not seem particularly ground-breaking, and the discussion is very shallow. Overall writing could be improved; there are lots of typos and grammatical inconsistencies. The second paragraph of the introduction is a bit convoluted.

      Specific comments:

      Fig. 1 seems a bit anecdotal. The authors revisit an observation that has been made before. I can see how it is used as rationale for the in vitro system, but not sure that this adds much to the overall story. Clearly different cell types are different, but without some sort of quantification this remains meaningless. It should also be noted in the discussion maybe that there are large differences between cells in 2D and 3D. Microtubules much more frequently grow to the cell edge compared with 2D (see Akhmanova SLAIN2 paper from some years ago).

      Fig. 2: While Arp2/3 certainly promotes branched F-actin networks, from the data provided it is not clear to me to what extent the produced F-actin networks replicate F-actin organization at the cell edge. If this a the point the authors are trying to make, the ultrastructure of their in vitro networks needs some additional characterization. As far as it is possible to discern from the data provided, the F-actin meshwork on the stripes in E looks pretty much identical in both panels (and not really like a dendritic network that in a cell also would have a certain polarity with barbed ends facing out), and the bundles on the left don't look like anything that normally occurs in a cell.

      Fig. 4 It is unclear what is going on here. Given that without F-actin bundles, polymerizing microtubules are freely moving around, it does not come as a surprise that they would never penetrate the F-actin network because as the authors correctly state the growing end will push back from the barrier. So, then why do they sometimes penetrate when bundles are present? In 4A it appears that microtubule growth into f-actin only happens once the microtubule minus ends gets stuck between F-actin bundles on the other side. 4D is the same as 4A; so that makes me think this really does not occur that often. Does the microtubule plus end only penetrate the F-actin meshwork when the minus end gets stuck on the other side? This seems important and also means microtubule penetration may not have anything to do with the F-actin network architecture at the plus end. This needs to be quantified.

      Fig. 5 I guess that sort of solves my confusion with Fig. 4. The quantification graphs in 5B and 5C are flipped with respect to the figure legend (?). I understand the rationale for not considering microtubules that grow at a shallow angle, but there does not seem to be that much of a difference between 5B and 5D. Wouldn't a better quantification simply compare microtubules that contact F-actin at both ends compared with microtubules were the minus end is free. In this case, I would expect a very large difference in penetration. Do microtubules under pressure ever bend/buckle in this in vitro situation. As the authors state, in cells, that happens frequently. This difference is interesting. Why?<br /> It is customary to refer to polymerized actin as F-actin.<br /> The supplementary videos are not referenced in the manuscript.

      Significance

      The manuscript describes results from a novel assay to study interactions between F-actin networks and dynamic microtubules in vitro. While of interest to a specialized audience, the overall finding that microtubules can grow into an F-actin meshwork is somewhat incremental especially because of the limited characterization of the F-actin networks used. It remains unclear to what extent this is relevant to a physiological context in cells.

      My field of expertise is related to cytoskeleton dynamics and quantitative microscopy in live cells.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Major comments

      In the paper "Microtubules under mechanical pressure can<br /> breach dense actin networks", the authors showed clear evidence that pressure plays an important role in microtubule breaching into dense actin networks using elegant in vitro reconstitution assays. They have argued that the pressure results from polymerization force of microtubules, which builds up when microtubules are immobilized in the opposite end of breaching, by the means of actin microtubule crosslinking factor Tau.<br /> It would definitely be interesting to see lack of breaching in the presence of crosslinking deficient Tau construct in order to rule out the off -target effect of Tau on microtubule and actin architecture which may possibly facilitate breaching.

      The authors have also observed microtubule breaching into dense actin networks in living cells. However, in Figure 1C, better cell/ image processing might have been chosen to increase the visibility of actin structures that microtubules encounter on their way to breaching. In Figure S1D, for example, the similar actin structures in lamellipodia are very nicely visible.

      It is also interesting that on Figure 6A, actin bundles look different than the rest of the figures on the paper. It almost looks like actin bundles become branched, whereas in the other Figures actin bundles are either singular or two-three bundles joined together at the point very close to the edge of micropatterned lipid bilayer.

      Minor comments

      In the legend of Figure 4E, it should be written "white arrow" instead of "yellow arrow".<br /> In the Results section "crosslinking between microtubules and actin bundles increase piercing frequency", in the sentence number 7, it should be written "backwards" instead of "reaward".

      Significance

      The experimental setup of the paper is quite significant in the field, given the difficulty of observing dynamics of dense cytoskeletal structures in living cells. Moreover, the paper gives insight into how microtubule behavior can vary depending on different morphological states of actin network.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      General Statements

      We are happy to resubmit our manuscript “The protective function of an immunity protein against the cis-toxic effects of a Xanothomonas Type IV Secretion System Effector” by Gabriel Oka et al. This paper shows that the cohort of immunity proteins associated with the cocktail of toxic effectors secreted by the Xanthomonas citri T4SS are not required to protect against toxins injected by neighboring cells but rather provide protection against endogenous toxins of the cell in which they were produced. To our knowledge, this the first description of an antibacterial secretion system in which the immunity proteins are dedicated to protecting cells against cis-intoxication, a point we emphasize in the revised introduction.

      We thank the reviewers for their thorough revision of the manuscript. Two of the three reviewers clearly expressed the opinion that the manuscript would be of general interest and should be published. We have carried out a number of new experiments and data analyses to respond to most of the suggestions of all three reviewers and believe that the manuscript is significantly improved as a result.

      Reviewer #1 (Evidence, reproducibility and clarity):

      The manuscript by Oka et al. shows that one effector of X. citri is likely translocated into the periplasm where it cleaves PG unless inhibited by its cognate immunity protein. Interestingly, this effector is required for killing of target cells like E. coli in T4SS-dependent manner but it does not seem to be delivered into X. citri cells by T4SS. Authors show using various assays that cells lacking the immunity protein have various phenotypes including lysis and defect in biofilm formation, however, despite "cis-intoxication" the ability to kill other bacteria or infect plants remains unaffected. The manuscript is well written and in general all the experiments have proper controls and thus the conclusions seem solid. The results described here are novel and interesting as they are unexpected.

      Major issues that should be addressed:

      - Test various deletion variants of the toxin to identify which part of the protein is responsible for its translocation into the periplasm. This may help to identify the possible mechanism of translocation of the toxin into the periplasm. Alternatively, the authors may attempt to select for non-toxic point mutants of the toxin. This could be done by a random PCR mutagenesis of the toxin and a selection of the surviving mutants in the absence of the immunity protein.

      Thank you for your insightful experimental suggestions using PCR mutagenesis to investigate the molecular mechanisms of the alternative translocation of X-Tfes to the periplasm. However, I regret to inform you that the first five authors of this manuscript are no longer a member of my lab. Therefore, please consider accepting the results shown in Figure 5A where we observe that the N-terminal domain of X-TfeXAC2609 that lacks the XVIPCD domain still abolishes biofilm formation in the absence of X-TfiXAC2610. Also, note that the E48A point mutation in the active site of the GH19 domain that abolishes the in vitro activity of X-TfeXAC2609 (Souza et al. 2015), also abolishes X-TFEXAC2609 toxicity in vivo in the absence of X-TfiXAC2610(Figure 5). Furthermore, in the predicted structure of the X-TfiXAC2610(54-267)-X-TfeXAC2609(1-194) complex (Figure 6), A tyrosine side chain in the conserved loop in X-TfiXAC2610 interacts directly with Glu48 in the X-TfeXAC2609 active site. One possibility for further investigation is the remaining region of X-TfeXAC2609(195-306) as a putative translocation domain. Sequence analysis of this region indicates that it encodes a canonical peptidoglycan binding domain. Another possibility is the existing intrinsic leakage of cytoplasmic proteins to the periplasm. As we understand it, the leakage of cytoplasmic proteins to the periplasm is not a well-documented phenomenon, although there is some evidence that suggests it may occur (PMID: 28808000, PMC3016450 references cited in the revised manuscript). This poorly characterized T4SS-independent pathway of translocation as indicated in Figure 1B (pathway 2).

      - Test if localization of the immunity protein to the cytoplasm blocks its activity. An immunity protein mutant that lacks its secretion signal should not protect against cis-intoxication.

      To address the question, we conducted new assays on colony opacity, as shown in Figure S2C. The X. citri ∆X-TfiXAC2610 strain was transformed with a plasmid expressing a cytoplasmic version of X-TfiXAC2610 that lacks the signal peptide and lipobox (X-TfiXAC2610(His-22-267)). Figure S2C shows that this cytosolic version of X-TfiXAC2610 protects X. citri from the toxic effects of X-TfeXAC2609 in the ∆X-TfiXAC2610 background. This suggests that X-TfiXAC2610(His-22-267) may directly interact with X-TfeXAC2609 in the cytoplasm, leading to the inhibition of X-TfeXAC2609 hydrolase activity and/or inhibition of its translocation into the periplasm. This is now mentioned in the results section of the revised manuscript

      While many experiments support the conclusion that the toxin is responsible for "cis-intoxication, the test of "trans-intoxication" should be done again but with the same setup as was used for testing of killing of E. coli. The CPRG based assay is far more sensitive than counting survival by plating to count CFUs. This test should be done at a relatively high initial OD so that there is an immediate contact between the "killer" and the "prey" bacteria (lacking immunity/effector). If needed, LacZ should be over-expressed in X. citri to make use of the CPRG based assay. In addition, such assay could be used also for "cis-intoxication" to supplement the potentially hard to quantify biofilm experiments shown in Fig. 4 (e.g. test all the T4SS mutants for "cis-intoxication").

      We are confident that the X-Tfis do not play a role in protecting against T4SS-mediated trans-intoxication since we continue to observe X-TfeXAC2609-dependent intoxication even in the absence of a functional XT4SS (see experiments using strains lacking X-T4SS subunits in Figures 2, S2, 3, 4 and 5. This is not to say that trans-intoxication does not occur. In fact, it does, and there is an independent mechanism that protects against it. We will provide details of the mechanism that protects against trans-intoxication in a forthcoming manuscript. In the present manuscript, we are addressing the phenomenon of cis-intoxication. To support our conclusion that the immunity proteins are not involved in the prevention of trans-intoxication does not occur in X. citri, we have included one additional supplementary video: Movie S7 shows that wild-type Xanthomonas citri does not kill and X. citri Δ8Δ2609-GFP. The absence of killing events in these experiments indicates that the X-T4SS-associated X-Tfi immunity proteins are not required for protection against X-T4SS-mediated sibling attack.

      In addition, such assay could be used also for "cis-intoxication" to supplement the potentially hard to quantify biofilm experiments shown in Fig. 4 (e.g. test all the T4SS mutants for "cis-intoxication").

      - Fig. 2A needs a positive control. For example, test killing of E. coli under the same conditions.

      Figure 2A of the revised manuscript now shows a CPRG assay competition assay that clearly demonstrates X-T4SS-dependent killing of E.coli MG by X. citri. We have now included the results of CFU experiments of X. citri vs E. coli competitions in a new Supplementary Figure (Figure S1) that are consistent with the CPRG assays. We note that our group has published similar results in the past (Souza et al. 2015; Oliveira et al. 2016; Oka et al. 2022). CFU measurements of X. citri vs E. coli competition assays are performed under slightly different conditions from the X. citri vs X. citri assays shown in Fig 2B. This is because E. coli grows at a significantly faster rate than X. citri so the initial cell ratios in these experiments have to be modified.

      - Authors should look at the paper by Ho et al. PNAS 2017, which describes trafficking of VgrG of V. cholerae into the periplasm of E. coli without an obvious secretion signal. The effector of X. citri may behave similarly.

      We thank the reviewer for this observation and now mention the paper by Ho et al. in the Discussion of the revised manuscript. Using a number of different algorithms (TatP, SignalP 6.0) we do not find any evidence of putative signal sequences. In the Discussion, we also mention the manuscript by Dong et al., 2013 that showed that the immunity protein TsiV3 that neutralizes VgrG3 is critical to prevent trans-intoxication.

      - Provide some form of quantification of the phenotypes (cell rounding and cell death) observed using live-cell imaging.

      As suggested by the reviewer, we performed a quantitative analysis of the propidium iodide (PI) permeability by calculating the percentage of PI permeable cells observed in movies S1-S5. This data is now presented in Figure 3 and Table S4 of the revised manuscript.

      - Provide quantification of biofilm related phenotypes as well as of the citrus canker development assay

      As suggested by the reviewer, we have carried out experiments to quantify the amount of biofilm using a crystal violet assay (absorbance at 570 nm). The results are presented in Figure S5 of the revised manuscript.

      Reviewer #1 (Significance):

      The study provides an interesting insight into immunity proteins against anti-bacterial toxins. It points to a need to protect against "cis-intoxication". This is novel and interesting to a wide audience of microbiologists interested in bacterial competition as this could be true also for other toxins.

      We thank the reviewer for his/her positive recommendation.

      It would be however important to identify how is the toxin translocating to the periplasm of the producing bacterium. Some insight into the mechanism would vastly improve the study. My expertise is in understanding bacterial interactions and competition but I lack a direct experience with assays specific for X. citri.

      We agree that an understanding of the mechanism of translocation into the periplasm would be interesting but is beyond the scope of the present manuscript. However, we do point out that this has been observed previously by other groups in the fourth paragraph of the Discussion of the revised manuscript: “... In the case of X-TfeXAC2609, the toxin somehow makes its way into the cell periplasm where, in the absence of X-TfiXAC2610, it degrades the peptidoglycan layer. Analysis of the X-TfeXAC2609 sequence by the SignalP 6.0 (Teufel et al., 2022) and other algorithms failed to detect any putative N-terminal signal peptide. Although the mechanism responsible for X-TfeXAC2609 transfer into the periplasm is at the moment unknown, we have shown that it is independent of a functional X-T4SS and of the XVIPCD secretion signal. Other bacterial proteins have been shown to transfer into the periplasm without any obvious secretion signal, for example VgrG3 from Vibrio cholerae (Ho et al. 2017) and recombinant forms of HdeA and chymotrypsin inhibitor 2 (Banes and Pielak, 2011).”

      Reviewer #2 (Evidence, reproducibility and clarity):

      This manuscript explores the role of an immunity protein of the Xanthomonas type IV secretion system (X-T4SS). In contrast to most T4SSs that conjugate plasmids or transfer effectors into host cells, this system is able to kill other bacteria similar to the role of T6SSs. Here, the authors tested whether the immunity protein XAC2610 functions to prevent cis-intoxication (by self) and/or trans-intoxication (by sister cells). They provide data that the XAC2610 immunity protein functions to protect cis intoxication, but not trans-intoxication, by the T4SS effector XAC2609 (which functions as a peptidoglycan hydrolase). Based on AlphaFold modeling, they went on to identify a residue in XAC2610 that is critical for inhibiting the activity of the XAC2609 toxin. Overall the data is fairly solid and generally support the conclusions the authors made.

      Major comments:

      One of the major conclusions of the manuscript is that XAC2610 does not prevent trans-intoxication and the data in the manuscript support this conclusion. However, I wonder if this is an oversimplification. Notably, the authors observed that wild type Xantho was unable to kill a target cell lacking 8 different toxin/immunity systems (Fig. 1A). One could conclude that none of these immunity proteins function in preventing trans-intoxication ... or ... perhaps it appears that none perform this role because wild-type Xantho never attacks its siblings? For example, it is conceivable that Xantho uses a general mechanism, perhaps somewhat similar to phage exclusion or plasmid incompatibility, to prevent sibling attack? To me this seems more likely than none of the eight immunity proteins play a role in preventing trans-intoxication. Moreover, the phenotype observed for the ∆2610 mutant in preventing cis-intoxication is somewhat subtle, likely because the toxin and the immunity protein are topologically restricted to the cytoplasm and the periplasm, respectively. This would make sense if this were not the primary role for 2610.

      Ideally the authors will be able to test this theory by demonstrating that a wild-type Xantho strain can attack (but likely not kill) its siblings. Alternatively, could the authors test if related, but not identical, Xantho strains that express 2609/2610 are able to kill their ∆2610 mutant, i.e. do "cousins" attack each other? Not sure about the semantics but this could be described as preventing trans-intoxication. If they are unable to do either experiment, that is ok but they should at least describe this concept in their discussion (assuming they agree).

      We thank the reviewer for his insightful comments. Indeed, this manuscript is focussed solely on the role of X-Tfi immunity proteins which we show to be principally involved in avoiding cis-intoxication (self-intoxication). The question of trans-intoxication will be left to an upcoming manuscript by our group. In fact we have identified a key factor (not an X-Tfi) that is responsible for inhibiting trans-intoxication. As suggested by the reviewer, we have now added the following text to the end of the third paragraph of the Discussion: “Nevertheless, the fact that wild-type X. citri is unable to kill strains lacking immunity proteins is intriguing. That cells in some way avoid trans-intoxication is revealed by the fact that X. citri wild-type cells carrying an X-T4SS and full cohort of X-Tfes do not kill the X. citri Δ8Δ2609-GFP, the X. citri ∆X-TfeXAC2609∆X-TfiXAC2610, or any other X-T4SS-deficient strain tested points to a still-to-be-characterized mechanism of protection against trans-intoxication (fratricide) that will be addressed in future studies by our group.”

      Minor comments:

      1. Figure 1 is a bit confusing in terms of the layout. It would be beneficial if the authors separated parts A and B by a few spaces.

      As suggested, we have modified the layout of Figure 1 to more clearly distinguish between the two mechanisms tested.

      1. Figure 2A should start off by showing that the Xantho T4SS can kill other bacteria (e.g. Fig S4A). This would set up the paper better.

      As suggested, old Figure S4A has now been transferred to Figure 2A in the revised manuscript.

      1. Fig. 2A should include p values.

      As suggested, p values have been provided in the legend of Figure 2B (old Figure 2A).

      1. Fig. 2B is really hard to see and should be removed from the manuscript (although I do appreciate the novelty of the technique using Marilyn Monroe).

      We have transferred the old Fig. 2B to the Supplementary Material (Fig S2) of the revised manuscript. We agree that the effect is subtle, but we want to maintain the figure since the transparency of the ΔXAC2610 X. citri colonies over time were the first observations that led us to investigate this phenomenon. Additionally, to reduce potential human bias and to enhance the objectivity of the assay, we employed a Convolutional Neural Network (CNN) to analyze all the colonies presented in Fig S2. This method provides a confidence tendency index for opacity and transparency variations. A detailed description of this new methodology is in the "Materials and Methods" section (Convolutional Neural Network (CNN) analysis).

      1. Instead all of the data in Fig. 2B should be shown in a new version of Fig. 2C. Fig. 2C should include additional controls including:
      2. A wild type strain containing 2609 and 2610 mutants
      3. A complete virB operon deletion in combination with 2609 and 2610 mutants
      4. ∆8 strain
      5. 2609 lacking its T4SS signal sequence
      6. 2609 targeted to the periplasm with a sec signal sequence
      7. etc.

      We sincerely value the comprehensive suggestions for improving what was previously presented as Fig. 2D. (Current version of Figure 2B is the Fig S2 as mentioned in the previous observation). We encounter a practical challenge here: the primary authors responsible for these experiments, especially the first five, have since departed from our lab. This situation limits our immediate capacity to execute the extensive set of experiments you've proposed.

      Recognizing the significance of the controls you've outlined for a quantitative analysis of the colony phenotypes (Fig. 2C (current version)) we have instead supplemented our study with a rigorous quantitative analysis of the microscopy assays referenced in Movies S1-S5, Figure 3, Table S4. These analyses further emphasize our observations concerning colony transparency (Fig S2).

      1. Figure 2C. The VirB7 western band looks like in the 2610 complemented strain.

      Thank you for pointing out the discrepancy in our previous manuscript at line 366, which pertains to the description of the mutants in old Fig. 2. The double mutant, ΔX-TfiXAC2610ΔvirB7 strain, was actually complemented with X-TfiXAC2610 (as stated in the current version (Fig S2B), and not with VirB7. Additionally, we have corrected the legend of the figure (line 684 previous version) from (∆X-TfiXAC2610∆VirB7c) to (∆X-TfiXAC2610c∆VirB7). We apologize for the mix-up in our earlier description and are grateful for your meticulous review and feedback in this matter. Furthermore, we agree that, in this particular experiment, the VirB7 band seems weaker but it is clearly visible in the 2610 complemented strain.

      1. Figure 3C should include a comparison of exponential vs. stationary phase cells. In addition, the results for the ∆2610 mutant and the ∆2610 ∆B7 double mutant appear to be different(?). P values should be provided. If it is statistically significant, then this should be explained in the manuscript. It was not clear how the % damaged cells were calculated? # of cells? Stats?

      The statistical analysis that the reviewer suggested has been provided in the new version of the Figure 4C and its legend. In addition, we have also included a supplementary Table S5 that presents the total number of cells analyzed in these experiments.

      1. The majority of Figure 4 should be replaced by assaying the effect of a virB operon deletion rather than showing the individual mutants.

      We believe that retaining old Figure 4 (Figure 5 of the revised manuscript) is important. By showcasing results from this specific set of single mutants, we are able to rule out the possibility that X-TfeXAC2609 translocation into the periplasm is mediated by a distinct X-T4SS subunit or subcomplex. We've expanded on this rationale at the start of the paragraph to provide a more comprehensive justification for our approach.

      1. Discussion:
      2. The last one to two paragraphs of the results belong in the Discussion.
      3. A more detailed description of cis-intoxication would be useful.

      As suggested, the last two paragraphs the Results section of the original manuscript have now been moved to the end of the Discussion.

      As suggested by the reviewer, third paragraph of the Discussion describes cis-intoxication in more detail.

      Reviewer #2 (Significance):

      This work provides a conceptual advance in understanding the protective function of a T4SS immunity protein, X-Tfe XAC2610, against the cis-toxic effects of the T4SS effector, X-Tfi XAC2610. It will likely be of interest to scientists interested in T4SSs & T6SSs and interbacterial competition. Overall this is a thought-provoking manuscript and should be published in a respectable journal.

      We sincerely thank Reviewer #2 for the thoughtful appraisal and positive feedback regarding our work. We are gratified to hear that the reviewer recognizes the conceptual advance our research brings to the understanding of T4SS immunity proteins and are encouraged by the acknowledgment that this manuscript will be of interest to our peers. We truly appreciate the endorsement for publication in a reputable journal.

      Reviewer #3 (Evidence, reproducibility and clarity):

      In this study, the authors suggest that TfeXAC2609-TfiXAC2610 represent a novel deviation from the established paradigm in contact-dependent interbacterial secretion systems. X. citri strains lacking the predicted immunity protein, TfiXAC2610, do not suffer a competitive disadvantage when grown in T4SS-inducing conditions against a wild-type strain. Furthermore, cells lacking the immunity develop aberrant morphology and auto-lyse. The mechanism for self-intoxication by TfeXAC2609 is independent of a functional T4SS, and intoxication is exacerbated when the toxin's T4SS-signal sequence is removed.

      Major Points

      1. The authors of the study do not provide sufficient evidence that TfeXAC2609 contributes to T4SS mediated killing. Does the toxin behave in a synergistic way, rather than mediate killing independently? Does removing the toxin and immunity change the competitive advantage of X. citri?

      We have shown in a previous publication that X-TfeXAC2609 does contribute to X-T4SS mediated killing (Oka et al, 2022). In that published paper we show that even in the absence of seven other toxin/antitoxin pairs, X-T4SS mediated transfer of only one effector (X-TfeXAC2609 or X-TfeXAC3634) can kill E. coli cells.

      Removing only X-TfeXAC2609 and X-TfiXAC2610 does not significantly reduce the ability of X. citri cells to kill E. coli (Fig. 2A of the revised manuscript). This is expected since this double mutant still retains seven other toxin/immunity pairs.

      Suggested Experiments: Competing, against E. coli, both WT X. citri and X. citri ΔXAC2609 ΔXAC2610, and determining whether there is a change in relative competitive advantage, or expressing TfeXAC2609 in a heterologous system and marking any observed toxic phenotype.

      The results of the experiment suggested by the reviewer have now been included in part A of the revised version of Figure 2. The effect of deleting only one toxin such as X-TfeXAC2609 results in no detectable difference in killing efficiency, most likely due to the presence of the eight other X-Tfes, three of which have been shown (XAC3634) or are predicted (XAC0466 and XAC1918) to have pedptidoglycan hydrolase activity (Oka et al, 2022, Souza, 2015, Sgro et al, 2019).

      1. Authors should directly answer where the toxin is active and localized in the cell.

      Suggested Experiments: Western blot subcellular fractionation (cytoplasm, periplasm, etc) to determine the localization of each protein.

      In response to the query about the toxin's activity and localization within the cell, we acknowledge the importance of such experiments to shed light on these aspects. However, I would like to highlight that the five first authors of this work are no longer affiliated with our lab. Consequently, we are facing constraints in terms of manpower and expertise to undertake comprehensive experiments such as the suggested subcellular fractionation.

      Also, our earlier work demonstrated the importance of the XVIPCD for secretion via X-T4SS (Souza, 2015) and in vivo activity of X-TfeXAC2609 (Oka et al., 2022). Moreover, using heterologous proteins expressed in E. coli (Souza, 2015) and our current observation that the absence of X-TfiXAC2610 induces spheroplast formation (Fig 4A-B, Movie S6) strongly suggest that the peptidoglycan glycohydrolase activity of the N-terminal domain of X-TfeXAC2609 acts in the periplasm.

      1. There is no evidence that TfeXAC2609 plays any role in inter-bacterial killing besides that is predicted from its genetic arrangement and in vitro assays from a previous publication.

      Suggested Experiments: Again, with the available antibodies, detecting whether TfeXAC2609 is being secreted, either in competition settings against X. citri or E. coli; given that there is no killing observed in Fig. 2B, it may also be a suitable control for this experiment.

      We have published in vivo evidence in the past:

      Souza et al, 2015 showed that X-TfeXAC2609 is secreted when in contact with E. coli cells.

      Oka et al, 2022 showed that an X. citri strain expressing X-TfeXAC2609/X-TfiXAC2610 but lacking seven other toxin/antitoxin pairs can still kill E. coli.

      1. The structural and co-evolutionary analysis seems to miss an essential point - that the lack of fratricide protection is not due to a novel protein-protein interaction.

      We do not understand this comment. As we point out in the manuscript, X-TfiXAC2610 does not protect against fratricide (trans-intoxication) but instead does protect against suicide (cis-intoxication). This protection requires a X-TfeXAC2609-X-TfiXAC2610 protein-protein interaction supported by the structural and co-evolutionary analysis as well as the experimental data using the X-TfiXAC2610 Y170A mutant (Fig. 6D of the revised manuscript). Moreover, we believe that the structural and sequence analysis significant expand the knowledge of the broader family of immunity proteins to which X-TfiXAC2610 belongs (Fig. S10 and Fig. S11 of the revised manuscript).

      1. The role of the immunity in biofilm formation is confusing. Cells lacking the immunity die within 96 hours (the auto-lysis phenotype). Given that the immunity is required for viability in this time frame, wouldn't it also be required for viability after five days?

      Suggested Amendments: Remove or de-emphasize.

      In the manuscript we use several different techniques to show that cells lacking the X-TfiXAC2610 immunity protein are less viable than the wild-type strain under certain conditions (growth on LB agar plates, biofilm formation) but perhaps not under others (ie in direct short-term competition experiments against E. coli and in long-term (2 week) in planta citrus canker assays). This is consistent with the fact that ultrastructural analysis by transmission electron microscopy shows that when grown in liquid media, only around 2% of X. citri cells lacking X-TfiXAC2610 present significant damage to their cell envelope (only 0.1% of wild-type cells show damage).

      1. Why does cell permeability increase with the loss of the T4SS signal sequence? Without there being greater evidence to support that an alternative secretion system is secreting or transporting the toxin into the periplasm, which may compete with the T4SS, additional hypotheses should be experimentally probed.

      The reviewer is comparing the propidium iodide permeability results observed for the ΔX-TfiXAC2610 mutant (carrying an empty pBRA plasmid) that expresses full-length X-TfeXAC2609 from its chromossomal gene with the ΔX-TfeXAC2609/ΔX-TfiXAC2610 double mutant carrying the pBRA-X-TfeXAC2609NT plasmid that expresses the X-TfeXAC2609 protein lacking the T4SS signal sequence from a very strong inducible promoter. Therefore, it can be expected that the levels of the truncated effector could be significantly greater than that of the full-length effector, leading to more damage.

      Note that, in the absence of X-TfiXAC2610, cell permeability increases only if X-TfeXAC2609 is present, with or without its XVIPCD T4SS signal sequence. This is consistent with a cis-intoxication mechanism which is independent of the X-T4SS-mediated transfer of the toxin from one cell to another. As we mention in the revised manuscript, and as pointed out by reviewer 1, Ho et al have also observed that when a lysozyme-containing domain of the T6SS effector VgrG3 is expressed in E. coli or in Vibrio cholerae, it can be detected in the periplasm in spite of the lack of a detectable signal sequence and in the absence of a functional T6SS. Ho et al attributed this observation to a “cryptic” secretion mechanism.

      1. Unclear if the the loss of cell envelope integrity is a direct effect of TfeXAC2609 activity and not an artifact of cell death. The microscopy also does not show a consistent change in morphology amongst intoxicated cells as there are healthy cells adjacent to lysed cells. This needs to be investigated in much more mechanistic detail.

      We observed that the X-TfeXAC2609 toxicity is dependent on its lysozyme domain since a point mutation in the active site residue (E48A) abolishes the toxicity-related phenotype in the biofilm assay (Figure 5).

      1. The role for immunity proteins in cis-intoxication is not novel as proposed by the authors. For example, see PMID:22511866 and PMID:26456113 where the authors used an inducible degradation system to show that in a T6SS null strain, cis-intoxication occurs when immunity is depleted.

      We thank the reviewer for pointing out these observations which are now mentioned and cited in the Introduction and in the Discussion of the revised manuscript.

      Minor Revisions

      1. Inconsistent use of the term "self-killing"; either refers to the killing of kin cells, or self (interchangeably used to refer to trans and cis killing).

      The term “self-killing” no longer appears in the manuscript.

      1. Terms trans-intoxication and cis-intoxication are convoluted and not constructive to the points being communicated. Self-killing vs kin-killing seem more intuitive and clearer. We prefer to maintain the use of the terms cis-intoxication and trans-intoxication which we defined in the Introduction, at the beginning of the Results section and in the Discussion as well as in Figure 1.
      2. Readability would be improved by the removal of double negatives.

      We have tried to avoid these whenever possible.

      1. Bacterial competition assay in methods only refers to the E. coli competition, not the one between the different genotypes of X. citri.

      Both methods were described in the same paragraph in the original manuscript. For clarity, this has now been divided into two sections in the revised manuscript: “X. citri vs E. coli competition assays” and “X. citri vs X. citri competition assays”.

      1. Strain naming scheme presented on pg. 16 doesn't conform to traditional, and clearer, nomenclature typically used.

      We have checked the manuscript to make sure that strain naming was consistent throughout the manuscript.

      1. On Pg 25, there is a typo "X-TfiXAC2609" as opposed to X-TfeXAC2609

      Thank you for the observation. This has now been corrected.

      1. Line 619 - "or several other immunity proteins in competition assays"... where was this data shown? No immediate connection to any figures from this paper nor are there any references.

      This is shown in Figure 2B and in Movie S7 which is now cited directly in the revised manuscript.

      Reviewer #3 (Significance):

      Overall it is difficult to take paradigm-conflicting conclusions at face-value when they are not presented alongside concrete experimental evidence. Without directly showing that the toxin localizes to the periplasm, the explanation that "the toxin somehow makes its way into the cell periplasm [independent of the T4SS] where it degrades the peptidoglycan layer" hinders the other conclusions presented by the authors. Consequently, my enthusiasm for this work is minimal.

      We deeply appreciate the insightful feedback from Reviewer #3, particularly regarding the concerns about paradigm-conflicting conclusions. We are steadfast in our commitment to ensuring that our findings are both rigorous and scientifically relevant.

      Evidence for Toxin Localization: We understand the criticality of concrete experimental evidence for toxin localization to the periplasm. While our data suggest an yet to be discovered translocation pathway of X-TfeXAC2609 from the cytoplasm to the periplasm, we recognize the importance of providing direct evidence. We are actively working on methodologies to understand this phenomenon. However, we do not believe that answering this question is absolutely necessary to understand the main conclusions of the present manuscript.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      In this study, the authors suggest that TfeXAC2609-TfiXAC2610 represent a novel deviation from the established paradigm in contact-dependent interbacterial secretion systems. X. citri strains lacking the predicted immunity protein, TfiXAC2610, do not suffer a competitive disadvantage when grown in T4SS-inducing conditions against a wild-type strain. Furthermore, cells lacking the immunity develop aberrant morphology and auto-lyse. The mechanism for self-intoxication by TfeXAC2609 is independent of a functional T4SS, and intoxication is exacerbated when the toxin's T4SS-signal sequence is removed.

      Major Points

      1. The authors of the study do not provide sufficient evidence that TfeXAC2609 contributes to T4SS mediated killing. Does the toxin behave in a synergistic way, rather than mediate killing independently? Does removing the toxin and immunity change the competitive advantage of X. citri?<br /> Suggested Experiments: Competing, against E. coli, both WT X. citri and X. citri ΔXAC2609 ΔXAC2610, and determining whether there is a change in relative competitive advantage, or expressing TfeXAC2609 in a heterologous system and marking any observed toxic phenotype.
      2. Authors should directly answer where the toxin is active and localized in the cell.<br /> Suggested Experiments: Western blot subcellular fractionation (cytoplasm, periplasm, etc) to determine the localization of each protein.
      3. There is no evidence that TfeXAC2609 plays any role in inter-bacterial killing besides that is predicted from its genetic arrangement and in vitro assays from a previous publication.<br /> Suggested Experiments: Again, with the available antibodies, detecting whether TfeXAC2609 is being secreted, either in competition settings against X. citri or E. coli; given that there is no killing observed in Fig. 2B, it may also be a suitable control for this experiment.
      4. The structural and co-evolutionary analysis seems to miss an essential point - that the lack of fratricide protection is not due to a novel protein-protein interaction.
      5. The role of the immunity in biofilm formation is confusing. Cells lacking the immunity die within 96 hours (the auto-lysis phenotype). Given that the immunity is required for viability in this time frame, wouldn't it also be required for viability after five days?<br /> Suggested Amendments: Remove or de-emphasize.
      6. Why does cell permeability increase with the loss of the T4SS signal sequence? Without there being greater evidence to support that an alternative secretion system is secreting or transporting the toxin into the periplasm, which may compete with the T4SS, additional hypotheses should be experimentally probed.
      7. Unclear if the the loss of cell envelope integrity is a direct effect of TfeXAC2609 activity and not an artifact of cell death. The microscopy also does not show a consistent change in morphology amongst intoxicated cells as there are healthy cells adjacent to lysed cells. This needs to be investigated in much more mechanistic detail.
      8. The role for immunity proteins in cis-intoxication is not novel as proposed by the authors. For example, see PMID:22511866 and PMID:26456113 where the authors used an inducible degradation system to show that in a T6SS null strain, cis-intoxication occurs when immunity is depleted.

      Minor Revisions

      1. Inconsistent use of the term "self-killing"; either refers to the killing of kin cells, or self (interchangeably used to refer to trans and cis killing).
      2. Terms trans-intoxication and cis-intoxication are convoluted and not constructive to the points being communicated. Self-killing vs kin-killing seem more intuitive and clearer
      3. Readability would be improved by the removal of double negatives.
      4. Bacterial competition assay in methods only refers to the E. coli competition, not the one between the different genotypes of X. citri.
      5. Strain naming scheme presented on pg. 16 doesn't conform to traditional, and clearer, nomenclature typically used.
      6. On Pg 25, there is a typo "X-TfiXAC2609" as opposed to X-TfeXAC2609
      7. Line 619 - "or several other immunity proteins in competition assays"... where was this data shown? No immediate connection to any figures from this paper nor are there any references.

      Significance

      Overall it is difficult to take paradigm-conflicting conclusions at face-value when they are not presented alongside concrete experimental evidence. Without directly showing that the toxin localizes to the periplasm, the explanation that "the toxin somehow makes its way into the cell periplasm [independent of the T4SS] where it degrades the peptidoglycan layer" hinders the other conclusions presented by the authors. Consequently, my enthusiasm for this work is minimal.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      This manuscript explores the role of an immunity protein of the Xanthomonas type IV secretion system (X-T4SS). In contrast to most T4SSs that conjugate plasmids or transfer effectors into host cells, this system is able to kill other bacteria similar to the role of T6SSs. Here, the authors tested whether the immunity protein XAC2610 functions to prevent cis-intoxication (by self) and/or trans-intoxication (by sister cells). They provide data that the XAC2610 immunity protein functions to protect cis intoxication, but not trans-intoxication, by the T4SS effector XAC2609 (which functions as a peptidoglycan hydrolase). Based on AlphaFold modeling, they went on to identify a residue in XAC2610 that is critical for inhibiting the activity of the XAC2609 toxin. Overall the data is fairly solid and generally support the conclusions the authors made.

      Major comments:

      One of the major conclusions of the manuscript is that XAC2610 does not prevent trans-intoxication and the data in the manuscript support this conclusion. However, I wonder if this is an oversimplification. Notably, the authors observed that wild type Xantho was unable to kill a target cell lacking 8 different toxin/immunity systems (Fig. 1A). One could conclude that none of these immunity proteins function in preventing trans-intoxication ... or ... perhaps it appears that none perform this role because wild-type Xantho never attacks its siblings? For example, it is conceivable that Xantho uses a general mechanism, perhaps somewhat similar to phage exclusion or plasmid incompatibility, to prevent sibling attack? To me this seems more likely than none of the eight immunity proteins play a role in preventing trans-intoxication. Moreover, the phenotype observed for the ∆2610 mutant in preventing cis-intoxication is somewhat subtle, likely because the toxin and the immunity protein are topologically restricted to the cytoplasm and the periplasm, respectively. This would make sense if this were not the primary role for 2610.

      Ideally the authors will be able to test this theory by demonstrating that a wild-type Xantho strain can attack (but likely not kill) its siblings. Alternatively, could the authors test if related, but not identical, Xantho strains that express 2609/2610 are able to kill their ∆2610 mutant, i.e. do "cousins" attack each other? Not sure about the semantics but this could be described as preventing trans-intoxication. If they are unable to do either experiment, that is ok but they should at least describe this concept in their discussion (assuming they agree).

      Since the cis-intoxication phenotype of the ∆2610 mutant is subtle, it would strengthen the authors' conclusions on cis-intoxication if they artificially targeted XAC2609 to the periplasm with a sec signal sequence. If the authors are correct, this should be a lethal event in the absence of the 2610 immunity protein. This might be useful in terms of figuring out how the 2609 toxin normally gets into the periplasm, a major unanswered question in this manuscript.

      Minor comments:

      1. Figure 1 is a bit confusing in terms of the layout. It would be beneficial if the authors separated parts A and B by a few spaces.
      2. Figure 2A should start off by showing that the Xantho T4SS can kill other bacteria (e.g. Fig S4A). This would set up the paper better.
      3. Fig. 2A should include p values.
      4. Fig. 2B is really hard to see and should be removed from the manuscript (although I do appreciate the novelty of the technique using Marilyn Monroe).
      5. Instead all of the data in Fig. 2B should be shown in a new version of Fig. 2C. Fig. 2C should include additional controls including:
      6. a. A wild type strain containing 2609 and 2610 mutants
      7. b. A complete virB operon deletion in combination with 2609 and 2610 mutants
      8. c. ∆8 strain
      9. d. 2609 lacking its T4SS signal sequence
      10. e. 2609 targeted to the periplasm with a sec signal sequence
      11. f. etc.
      12. Figure 2C. The VirB7 western band looks like in the 2610 complemented strain.
      13. Figure 3C should include a comparison of exponential vs. stationary phase cells. In addition, the results for the ∆2610 mutant and the ∆2610 ∆B7 double mutant appear to be different(?). P values should be provided. If it is statistically significant, then this should be explained in the manuscript. It was not clear how the % damaged cells were calculated? # of cells? Stats?
      14. The majority of Figure 4 should be replaced by assaying the effect of a virB operon deletion rather than showing the individual mutants.
      15. Discussion:
      16. a. The last one to two paragraphs of the results belong in the Discussion.
      17. b. A more detailed description of cis-intoxication would be useful.

      Significance

      This work provides a conceptual advance in understanding the protective function of a T4SS immunity protein, X-Tfe XAC2610, against the cis-toxic effects of the T4SS effector, X-Tfi XAC2610. It will likely be of interest to scientists interested in T4SSs & T6SSs and interbacterial competition. Overall this is a thought-provoking manuscript and should be published in a respectable journal.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      The manuscript by Oka et al. shows that one effector of X. citri is likely translocated into the periplasm where it cleaves PG unless inhibited by its cognate immunity protein. Interestingly, this effector is required for killing of target cells like E. coli in T4SS-dependent manner but it does not seem to be delivered into X. citri cells by T4SS. Authors show using various assays that cells lacking the immunity protein have various phenotypes including lysis and defect in biofilm formation, however, despite "cis-intoxication" the ability to kill other bacteria or infect plants remains unaffected. The manuscript is well written and in general all the experiments have proper controls and thus the conclusions seem solid. The results described here are novel and interesting as they as unexpected.

      Major issues that should be addressed:

      • Test various deletion variants of the toxin to identify which part of the protein is responsible for its translocation into the periplasm. This may help to identify the possible mechanism of translocation of the toxin into the periplasm. Alternatively, the authors may attempt to select for non-toxic point mutants of the toxin. This could be done by a random PCR mutagenesis of the toxin and a selection of the surviving mutants in the absence of the immunity protein.
      • Test if localization of the immunity protein to the cytoplasm blocks its activity. An immunity protein mutant that lacks its secretion signal should not protect against cis-intoxication.
      • While many experiments support the conclusion that the toxin is responsible for "cis-intoxication, the test of "trans-intoxication" should be done again but with the same setup as was used for testing of killing of E. coli. The CPRG based assay is far more sensitive than counting survival by plating to count CFUs. This test should be done at a relatively high initial OD so that there is an immediate contact between the "killer" and the "prey" bacteria (lacking immunity/effector). If needed, LacZ should be over-expressed in X. citri to make use of the CPRG based assay. In addition, such assay could be used also for "cis-intoxication" to supplement the potentially hard to quantify biofilm experiments shown in Fig. 4 (e.g. test all the T4SS mutants for "cis-intoxication").
      • Fig. 2A needs a positive control. For example, test killing of E. coli under the same conditions.
      • Authors should look at the paper by Ho et al. PNAS 2017, which describes trafficking of VgrG of V. cholerae into the periplasm of E. coli without an obvious secretion signal. The effector of X. citri may behave similarly.
      • Provide some form of quantification of the phenotypes (cell rounding and cell death) observed using live-cell imaging.
      • Provide quantification of biofilm related phenotypes as well as of the citrus canker development assay.

      Significance

      The study provides an interesting insight into immunity proteins against anti-bacterial toxins. It points to a need to protect against "cis-intoxication". This is novel and interesting to a wide audience of microbiologists interested in bacterial competition as this could be true also for other toxins. It would be however important to identify how is the toxin translocating to the periplasm of the producing bacterium. Some insight into the mechanism would vastly improve the study. My expertise is in understanding bacterial interactions and competition but I lack a direct experience with assays specific for X. citri.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      We have addressed all the queries and suggestions put forth by the reviewers. Major changes include:

      1. Expansion of PILOT Functionality and Analysis: We have substantially extended the functionality and analysis capabilities of PILOT, particularly in relation to sample clustering. This enhancement now encompasses the incorporation of statistical tests aimed at identifying cell types and genes associated with distinct patient groups. We applied this expanded feature in an exploratory analysis of sub-clusters within pancreas ductal adenocarcinoma data (PDAC).
      2. Clarification of Benchmarking Methods: We have provided clear elucidations of the methods employed for benchmarking PILOT alongside competing methodologies. Our benchmarking approach is notably comprehensive, encompassing twelve different datasets and evaluating four to five competing methods through statistical assessment across three problem domains: clustering, distance measurement, and trajectory estimation. The outcomes of these evaluations consistently demonstrate the superior performance of PILOT's Wasserstein metric across all three problem domains. It is noteworthy that previous studies have often limited their analyses to exploratory evaluations on individual datasets, lacking the level of comprehensive benchmarking undertaken in this study.
      3. Examination of Experimental Factors: We have conducted a thorough investigation into the impacts of batch correction, cluster/cell type resolution, and parameter choices used within the PILOT framework.
      4. Enhancement of Text Description: We have enhanced the textual descriptions to provide a high-level overview of the PILOT methodology, along with justifications for the methodological decisions made.
      5. Improvement of Code and GitHub Repository: To enhance accessibility and promote reproducibility, we have made improvements to the codebase and the associated GitHub repository.

      In summary, PILOT stands as a distinctive and all-encompassing framework. It holds the unique distinction of being the sole method offering comprehensive tools for both clustering and trajectory analysis of samples within multiscale single-cell and pathomics data. Moreover, it incorporates statistical methodologies for the interpretation of results. The effectiveness of these tools has been thoroughly validated through the most extensive benchmarking study performed to date on sample-level analysis. The versatility of PILOT is demonstrated through its successful application in exploratory analyses of three distinct datasets: elucidating trajectories in myocardial infarction single-cell RNA-seq data, uncovering trajectories within pathomics data from kidney IgNA patients, and facilitating the clustering of pancreas adenocarcinoma samples. We firmly believe that these contributions hold significant value for the fields of bioinformatics, single-cell genomics, and pathology.

      Reviewer #1 (Evidence, reproducibility and clarity):

      The paper describes a computational method, PILOT, that uses optimal transport to compute the Wasserstein distance between two individual single-cell samples. It uses PILOT to detect sample (patient) level trajectories and clusters associated with diseases. The method was applied separately to single-cell genomics data and to digital pathology data. The method was applied to several datasets and compared against other tools.

      Major comments:

      The paper is not easy to follow and should be improved considerably to make it readable and reproducible. Consequently, I was not convinced that the PILOT method is much better than other methods.

      We extend our appreciation to the reviewer for their valuable suggestion. We have further refined the manuscript by incorporating a comprehensive and high-level description of our method. This expansion encompasses methodological justifications and clarifications to enhance the overall clarity. Additionally, we wish to emphasize that, to the best of our knowledge, our benchmarking analysis stands as the most comprehensive within the current literature. The results of this analysis unequivocally demonstrate that PILOT surpasses all competing methods in at least one of the various computational analysis tasks, namely clustering, trajectory estimation, and distance evaluation.

      Furthermore, we have undertaken significant enhancements in the codebase of PILOT, coupled with a reorganization of the associated GitHub repository. This effort includes the development of in-depth and improved tutorials that faithfully replicate the analyses conducted on datasets related to myocardial infarction, pancreas adenocarcinoma, and pancreas pathomics (https://pilot.readthedocs.io/en/latest/). This changes guarantee the reproducibility of the PILOT framework.

      See below for specific changes and additional clarifications.

      At first read of the title and abstract, I got the impression that the method analyzes single cell and pathomics data concurrently rather than separately. This should be fixed.

      We have changed the text of the abstract and introduction to make clear that PILOT is either applied to single cell or pathomics data independently.

      The usage of Wasserstein distance to compute distance between single-cell samples is elegant and is the main strength of this study. Given that PILOT is the main achievement, it should be described more carefully and in a detailed manner.

      For example, in the first Results paragraph, "The test indicates for features explaining the predicted pseudotime by fitting either linear or quadratic models" - I could not understand this sentence. Also, which test do the authors refer to? A few sentences down, there is a reference to a Wald test, is that it?

      PILOT has three major parts: (1) a method for measuring distance of samples with optimal transport; (2) an patient level unsupervised analysis part (clustering or trajectory analysis) and (3) a part for explaining predicted trajectories/clustering. The sentence mentioned before, refers to the interpretation approach after trajectory analysis. Here, we fit linear, quadratic or linear quadratic models to find association of predicted sample pseudo-time with data features (gene expression values in scRNA or morphological features in pathomics data). This fit can be done for all cells in the data or only for cells from a specific type. In the case of a cell specific fit, we use a Wald test to check if the cell type fit differs from all other cell types in the data, i.e. the gene is associated with the trajectory and the expression changes are specific to the cluster at hand.

      While these details were found in the method section, we agree with the referee that they can be better introduced in the main manuscript. We have therefore improved the first subsection of the results and Figure 1 to reflect this.

      One of the key aspects of the Wasserstein distance is the cost metric. The determination of the cost metric should be detailed as part of the Results. Have the authors considered and estimated other ways to define the distance?

      This is an interesting question. Currently, PILOT uses the Cosine metric. In our revision, we evaluate other metrics (Euclidean, Manhattan, and Chebyshev). This benchmarking indicates that the Cosine and Manhattan performed best regarding the clustering problem (ARI), while Cosine was better than all metrics for the Silhouette statistic; and Cosine and Euclidean performed best regarding AUPR. Therefore, we adopt the Cosine metric in the paper. We include these results in the revised manuscript and in Sup. Fig. 5F-H.

      Figure 1 provides a schematic view of PILOT. However, there is no explanation of the notation, which makes it confusing rather than helpful. Also, what is the relationship between J and j, if any?

      We understand that the figure 1 was problematic, as it did not introduce the formulation. We have now improved the first sub-section of the results page and figure 1 to improve this.

      The motivation and usage of adjusted Rand index (ARI) and Friedman-Nemenyi tests should be provided. Currently, they are not clear, including why those tests are suitable in the cases shown.

      The adjusted Rand index is a well known metric to evaluate clustering results when labels are known. Among others this metric has many interesting features as it does not require an association of clusters with class labels. Moreover, it has a correction for random clustering solutions, therefore values lower than zero indicate poor solutions and values of 1 a perfect solution.

      The Friedman-Nemenyi test allows us to compare the performance of several algorithms whenever evaluated in the same data sets. Here, the null hypothesis is that all algorithms have the same performance (same ARI statistic). The test is nonparametric and is based on the rank of the algorithm at each data set. This is important, as ARI values (or any other evaluation statistic) are data set specific, e.g. some clustering problems are more difficult than others. By evaluating the rank, the test indicates which methods perform relatively better than others. Moreover, it follows a rigorous statistical framework including correction for multiple testing. This test has an increasing adoption in the machine learning community (Demsar et al., JMLR, https://jmlr.org/papers/v7/demsar06a.html).

      We have added phrases with these justifications in the main text (subsection Evaluation of patient-level clustering and trajectory analysis) and included a new section in the materials and methods with more information in the experimental design of the benchmarking analysis.

      Fig. 2 the use of method colors should be constant across panels.

      We have changed the colors of panels in figure 2A-C (and equivalent panels everywhere else) to avoid confusions.

      The proportions method works at least as well as PILOT in 2B and 2C (silhouette and AUPR). Explain why PILOT is better.

      The benchmarking analysis shows that PILOT has the highest ARI value (clustering performance) at absolute and ranking levels (Fig. 2A). Moreover the Friedman-Neymeni test indicates this PILOT has significantly higher ranking than all evaluated methods. Regarding Silhouette (distance evaluation) and AUPR (trajectory evaluation) both proportion and PILOT have similar absolute values (Fig. 2B and 2C; panel left), while PILOT has a superior ranking in both cases (Fig. 2B and 2C panel right). Friedman-Neymeni test indicates higher ranking of PILOT than PhEMD for Silhouette and higher ranking of PILOT than PhEMD and pseudo-Bulk regarding trajectory evaluation. The difference in the results on absolute and ranking values can be understood by looking at the statistics in table Table S1. PILOT has highest AUPR in 8 out 12 data sets; proportion has highest values in 5 (including 4 ties with PILOT); proportion-PHATE had 3 best results (including 3 ties with both PILOT and proportions), while PhEMD is best in one data set and Pseudo-bulk in 3 (including 1 tie with PILOT). Altogether, PILOT obtained a higher or equal AUCPR in 9 out of the 12 data sets. We have also changed Fig.2A, 2B and 2C to include all data points and to show the mean, as this provides a better visualization of the previously reported results.

      Altogether, these results indicate that PILOT outperforms all competing methods in at least one of the evaluated problems (clustering, trajectory and distance estimation) and ranks favorably in all evaluated scenarios. We have changed the manuscript text to reflect these results.

      Likewise, Figure 2C,D and Figures S1 and S2 don't show a clear and consistent advantage for PILOT over other methods. Explain what advantage of PILOT do the fraction panels highlight in Fig. 2E and Fig. S3. Fig. 2C is not mentioned in the text.

      Figure 2D, 2E, and now figures S2 and S3 represent visualizations of the results, which were statistically evaluated in panels of Fig.2A-2C. As discussed in the previous point, PILOT does perform better than all methods for the clustering problem and performs better or as good as the proportion test on 9 of the 12 evaluated data sets in the trajectory problem. We also have improved the text to include references to all figures in the main text.

      I assume Kidney IgAN (text) and Kidney IgA (fig. 2) are the same.

      The correct name is IgAN and this has been corrected in Figure 2.

      Fig. 3B fix the p-value notation (what is p=1.05E?) and R2 (R square?). Nrte tha both this problem also occurs in other figs. Fig. 3B shows the major cellular changes.

      We now adopt the term “R-squared” in the figures. Also, the previous version did not display p-values properly. We apologize for this. This has been fixed now.

      Are these changes consistent with known ones? Explain and provide references. Are there cell types that were expected to show a change and did not? Same questions for Fig. 3C wrt genes. Is this an exploratory analysis highlighting interesting candidate genes? If so, it should be described as such.

      Cardiac remodeling after myocardial infarction is characterized by loss of cardiomyocytes, infiltration by immune cells (myeloid and lymphocytes) and increase in myofibroblast populations (doi.org/10.1038/s41392-022-00925-z;doi.org/10.3389/fcvm.2019.00026). PILOT indicates these populations, with the exception of lymphocytes, are most relevant at both clustering levels (see Sup. Fig, 6). Particularly important are results from the low granularity analysis, as this indicates particular macrophage/fibroblast sub-populations (SPP1+ Mac. and Myofibroblast) with increase in disease. PILOT could not detect changes in lymphocyte cells, but this is explained by the poor coverage of these cells in the data set (>3%). We have updated the main manuscript to reflect this.

      We also explicitly mention that the analysis of genes and cells are exploratory analysis.

      The point of Fig. S6 and its major findings should be mentioned in the text (or it can be removed).

      We now make the reference to the gene ontology analysis presented in the new Figure S7 more explicit in the text.

      Fig. 4B legend - eGFR not GFR. What do the high-low values of Fig. 2B refer to?

      We have fixed these points.Hhigh and low values of panel 4B refer to the eGFR.

      Fig. S12 is out of order in supp file.

      This has been fixed.

      AUCPR - explain.

      The AUCPR stands for area under the curve of the precision recall (AUCPR) curve. We have now improved the explanation of the evaluation metric in the main text and methods section.

      The github looks like work in progress with many internal comments (eg, add ,edit, etc). I could not find the tutorials.

      We have removed all the comments, improved the repository organization and code. The tutorials are explicitly mentioned in the main github page (https://github.com/CostaLab/PILOT/) and in readthedocs webpage (https://pilot.readthedocs.io/). It include tutorials replicating analysis with trajectory inference and clustering problems, which are discussed in the manuscript.

      In the process of code review, we have noticed that while we could replicate all the analysis, the procedure for selection of healthy cardiomyocyte genes was distinct (gene were ranked by regression model fit p-value) than the analysis of the myofibroblast genes (genes were ranked by the Wald test p-value). As explained before, the Wald test, which compares the expression of the regression model fits across samples, is a more appropriate criteria, as it finds cluster and trajectory specific genes. We have changed the analysis of the cardiomyocyte to make the gene selection to be based on the Wald-test p-value. New results recover other sarcomere related genes (MYBPC3 and MYOM1) as being dysregulated during disease progression. These findings are in accordance with observations made in the original study presenting the data (Kuppe et al. 2022). We have updated Fig.3 and respective genes accordingly.

      Minor comments:

      Introduction: "Alternatively, trajectory analysis can be performed to uncover disease progression allowing the characterization of early disease events." Citations should be added (some appear later in the text).

      We included a reference to PhEMD.

      "Currently, there are no analytical methods to compare two single cell experiments from the same tissue from two distinct individuals." There have been several comparisons among data from patients, (e.g. Cain et al, 2023), so the authors should be more careful/accurate in their statements.

      We assume that the referee mentions https://www.nature.com/articles/s41593-023-01356-x. Indeed, we were not aware of this recently published study. The manuscript focuses on comparing cell proportion changes (estimated by deconvolution) between distinct phenotypes and does not provide any approach for sample level analysis of single cell data. This is in our view a different problem than the one addressed by PILOT or PhEMD. We refer to it in our manuscript, as its cell community based analysis is an interesting approach for the interpretation of PILOT results.

      "Except for PhEMD, all related methods9, 11, 12 require labels of patients for their analysis and cannot be used in the unsupervised analysis " - this sentence comes immediately after describing ref 13, which can be used in unsupervised analysis and accordingly is not cited in this sentence. The authors did well in describing ref 13 (a bioRxiv paper), and its description should come after this sentence.

      We changed the text to reflect this.

      "These can be clusters", clustered?

      Done.

      " acquire an injury cell states" remove an.

      Done.

      "As for scRNA-seq, there is no analytical method which is able to compare two or more histological slides based on morphometric properties of their structures." The sentence seems to refer to pathomics, not to sc data as suggested in "As for scRNA-seq"

      This has been rephrased.

      "Thus PILOT represents the first approach to detect unknown patient trajectories and clusters" patient clusters were also observed by others (eg ref 13, Cain et al).

      This has been rephrased.

      Equation 7 - Cosine(Mi,Mi) should be Cosine(Mi,Mj)

      Done.

      In the beginning of the Results, PILOT is not referred to as a package but as a researcher ("PILOT explores").

      This has been rephrased.

      Reviewer #1 (Significance):

      In general, the paper is a Methods paper. Hence, likely audience includes computational biologists interested in methodologies, not to biologists interested in the actual findings.

      Although I am among the likely audience, I was not convinced by the merits of the method, potentially due to the way the paper was written.

      I do not have sufficient expertise to check the math.

      In this revision, we have significantly enhanced the text to incorporate high-level descriptions of methods tailored for non-computational experts. Additionally, we have refined the description of the benchmarking process, which, as far as our knowledge extends, stands as the most comprehensive in the literature. This comprehensive analysis strongly underscores the statistical superiority of PILOT when compared to other methods. Lastly, PILOT presents an unique framework, encompassing methods for trajectory analysis, clustering, and interpretation of sample-level analyses within the realm of multiscale single-cell genomics and pathomics data.

      Reviewer #2 (Evidence, reproducibility and clarity):

      Joodaki et al. propose PILOT, a computational method for analysing single-cell genomics and pathomics data. PILOT is a method that enable clustering, trajectory analysis, and feature detection at a patient level using scRNA-seq data. This is an important task and represent the growing application of scRNA-seq to understand diseases and other perturbations to other biological systems. In particular, PILOT enables unsupervised analysis which alleviate the need of patient labels required by many alternative methods. We have the following comments for the authors' consideration.

      1. A key consideration in dealing with scRNA-seq data at a patient level is the batch effect in the data. Typically, each patient sample may be treated as a "batch" especially when they are processed separately to obtain a scRNA-seq dataset that are subsequently combined with scRNA-seq datasets from other patients to form a single dataset. Analysing these data without batch correction may lead to the identification of "cell types" and "states" that are driven by batch effect. In Figure 1. PILOT takes a clustered and integrated scRNA-seq data as input for analysis. I wonder how PILOT would behave if there is a strong batch effect in the data and how would the authors propose to handle them?

      This is an interesting question. Currently, PILOT is using the batch procedure used in the paper proposing the original data. We evaluate now the impact of batch correction methods implemented in scanpy (Harmony, bknn and Scanorama). We focus here on single cell data, which we have access to the original count matrix (Lupus, COVID, and Diabetes). We observe no impact of the batch correction algorithm in these data sets (see Sup. Fig. 5C-E). These results are now included in the manuscript.

      We have noticed however that strong batch effects in the lung cell atlas or the kidney cell atlas.For the lung cell atlas, we observed that single cell data measured from distinct techniques (Seq-well, Drop-seq, 10x 5’ and 10x 3’) or distinct tissue sampling approaches confounded results for all evaluated approaches. Therefore, we restricted the analysis to the technology with more samples (10x genomics 3’) and to lung tissues. This sample selection was previously described in the material and methods. Of note, the use of samples from distinct 10x genomic version kits (v1, v2 or v3) did not impact results. For the kidney cell atlas, we also observed a strong batch between single nuclei and single cell protocols. Here, we opted to focus on the largest cohort of single cell RNA experiments (see Review Fig. 1). Altogether, PILOT and other evaluated methods do require samples to be analyzed with an uniform technique and sampling approaches. We now include a brief discussion about this open point in the “Discussion” section. This is an important topic of future research.

      Review Fig. 1. - Data of the Kidney Precision Medicine Project was measured using either single cell or single nucleus protocols. All evaluated methods were affected by the differences in these technologies and could not separate disease status in this data.

      1. It appears that the Wasserstein distance (W) matrix of the samples was used for patient clustering and also trajectory analysis. However, most of the figures presented in the manuscript are for trajectory analysis. Since the patient clustering were performed prior to trajectory analysis, could the authors visualize the patient data based on the W before performing disease trajectory estimation?

      Indeed, despite the clustering-based analysis (ARI statistics; Fig. 2A) the current manuscript focuses on results of the trajectory analysis. We now include additional features for clustering analysis. This includes heatmap visualizations of the OT distance matrices together with Leiden clustering (Sup. Fig. 1). See points 4 and 5 below for further changes regarding clustering analysis.

      1. In trajectory analysis in figure 2D and E, why not use Multi-scale PHATE which appears to be specifically designed for trajectory analysis? The authors also mentioned SCANCell. While these methods require labels of patients for their analysis, it would be interesting to know how well they perform in comparison to PILOT if such information is available.

      This is an interesting point. Multiscale-PHATE is based on doing a multi-resolution clustering of the cells. It then applies PHATE (instead of diffusion map) to find a non-linear embedding on the cell proportions across samples and resolutions. While this analysis is presented at Multiscale-PHATE manuscript (Fig. 5), we could not find any code or functionality in their github to replicate this (https://github.com/KrishnaswamyLab/Multiscale_PHATE). Moreover, we were not able to find a function to find the cluster/resolution associations of cells to reimplement the above mentioned analysis following the descriptions of the manuscript. We also contacted authors, but obtained no reply. It is also important to state that Multi-PHATE used a supervised filter to select cell types for further analysis.

      Alternatively, we now include an evaluation of the use of cell proportions followed by a PHATE embedding in the trajectory based evaluation, which is close to the method proposed in Multiscale-PHATE. Our benchmarking indicates that Multiscale-PHATE is the third best ranked method being overpassed by proportion and PILOT. Regarding SCANCell, it focuses on the interpretation of cell communities and it uses embedding/distances by exploring PhEMD. Therefore, its performance in the trajectory or clustering performance problem is the same as PhEMD. We refer to these points in the text now.

      1. The current design of PILOT appears to assume that there is always a "smooth" trajectory in the data. Is this going to be the case in reality? What if we have a well separated and distinct groupings of the patients and controls data? In the latter case, imposing a trajectory seems artificial. I am also not sure how meaningful the trajectory analysis would be if, biologically, such a smooth transition is not present in the data.

      The EMD based distance can be used both for clustering or trajectory analysis. Also, PILOT performed quite well in the clustering problem benchmarking (Fig. 2A). The choice of application lies on the problem at hand. In our view, both the kidney pathomics and the myocardial infarction data (explored in Fig. 3 and Fig. 4) represent medical data with potential disease trajectories. We now expand the PILOT framework to include new visualizations and statistical methods to improve the interpretation of the clusterings (see point 2; Fig. S1; and point 5 and new Fig. 5).

      1. The feature analysis is also built on trajectory analysis using regression models. Again, how would this work out if there isn't a smooth trajectory/transition in the data (e.g. the data are obtained from a discrete case-control study)?

      We expanded the PILOT framework to also include statistical tests for accessing changes in cell populations and markers for the clustering problem. First, we use a Welch’s t-test to evaluate cell proportion changes associated with detected clusters. Next, we use a differential analysis test from limma to find genes within a cluster, whose gene expression is changing between the two groups of samples for a given cluster of cells. While these are standard approaches in the literature, this further improves the functionality of PILOT as a general framework for patient level analysis. This is now described in PILOT manuscript (Results subsection “Patient level distance with Optimal Transport” and methods).

      We also include in the manuscript an explorative analysis of a sub-cluster found in the PDAC data. This analysis could find a population of PDAC patients displaying higher levels of malignant cells and marked by both increase in hypoxia and fibrosis pathways. This example highlights how PILOT can be used to find potentially interesting groups of samples. These are implemented in the main manuscript (Fig. 5). We also include a new tutorial of PILOT with this analysis (see https://pilot.readthedocs.io/).

      1. It is not clear from the formulation of PILOT (and also Figure 1) if the cell type labels is required/used or the cluster id of a clustering algorithm was used instead. The author also mentioned that the clustering output does not have much impact on the downstream analysis. I wonder why and if so can we group the data in any way we want for downstream analysis? This can be useful when one would like to focus on certain grouping of cells.

      Clustering at the cells (or structure level) is required. For the benchmarking analysis, we have used the cell annotation reported in the original paper, which were derived via clustering analysis. The use of annotated clusters is crucial for interpretation. We also included in the original manuscript an analysis on the impact of the clustering resolution of the Leiden algorithm. This indicates that the change in resolution did not have a high impact in the clustering (ARI) of the samples (Sup. Fig. 5A-B). However, this analysis could not consider any interpretation of results, as cluster labels were not present.

      We believe, however, that the granularity of the clustering will impact the interpretation of the sample analysis. To investigate this, we evaluate how using higher level annotation/clustering of the heart myocardial infarction (also reported in Kuppe et al. 2022) impacts our ability to find cell specific changes. We observe similar changes whenever using low resolution clustering (decrease of cardiomyocytes, increase in fibroblasts and myeloid cells). However, this analysis loses a lot of important nuances found in the high resolution clustering (see Sup. Fig. 6). For example, it does not recover the fact that damaged cardiomyocyte populations have a slower decay than healthy myocytes. Or the fact that myofibroblasts has an increase in the latter disease trajectory stage, while progenitor fibroblast cells (Fibro_Scara5) have an increase previous to myofibroblasts. These results show how low resolution clustering can lead to loss of interesting information contained in cellular sub-states or cell sub-populations. This is now discussed in the results subsection ‘PILOT trajectories detect events associated with cardiac remodeling in myocardial infarction’.

      Reviewer #2 (Significance):

      PILOT is designed for analyzing scRNA-seq data at a patient level. There is a growing application of scRNA-seq to diseases and the development of computational tools for analyzing such data at phenotype level is critical. The key aspect of PILOT compared to other currently available tools is that it enables unsupervised analysis which alleviate the need of patient labels required by many alternative methods.

      Thanks for this very positive feedback and constructive comments.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Joodaki et al. propose PILOT, a computational method for analysing single-cell genomics and pathomics data. PILOT is a method that enable clustering, trajectory analysis, and feature detection at a patient level using scRNA-seq data. This is an important task and represent the growing application of scRNA-seq to understand diseases and other perturbations to other biological systems. In particular, PILOT enables unsupervised analysis which alleviate the need of patient labels required by many alternative methods. We have the following comments for the authors' consideration.

      1. A key consideration in dealing with scRNA-seq data at a patient level is the batch effect in the data. Typically, each patient sample may be treated as a "batch" especially when they are processed separately to obtain a scRNA-seq dataset that are subsequently combined with scRNA-seq datasets from other patients to form a single dataset. Analysing these data without batch correction may lead to the identification of "cell types" and "states" that are driven by batch effect. In Figure 1. PILOT takes a clustered and integrated scRNA-seq data as input for analysis. I wonder how PILOT would behave if there is a strong batch effect in the data and how would the authors propose to handle them?
      2. It appears that the Wasserstein distance (W) matrix of the samples was used for patient clustering and also trajectory analysis. However, most of the figures presented in the manuscript are for trajectory analysis. Since the patient clustering were performed prior to trajectory analysis, could the authors visualize the patient data based on the W before performing disease trajectory estimation?
      3. In trajectory analysis in figure 2D and E, why not use Multi-scale PHATE which appears to be specifically designed for trajectory analysis? The authors also mentioned SCANCell. While these methods require labels of patients for their analysis, it would be interesting to know how well they perform in comparison to PILOT if such information is available.
      4. The current design of PILOT appears to assume that there is always a "smooth" trajectory in the data. Is this going to be the case in reality? What if we have a well separated and distinct groupings of the patients and controls data? In the latter case, imposing a trajectory seems artificial. I am also not sure how meaningful the trajectory analysis would be if, biologically, such a smooth transition is not present in the data.
      5. The feature analysis is also built on trajectory analysis using regression models. Again, how would this work out if there isn't a smooth trajectory/transition in the data (e.g. the data are obtained from a discrete case-control study)?
      6. It is not clear from the formulation of PILOT (and also Figure 1) if the cell type labels is required/used or the cluster id of a clustering algorithm was used instead. The author also mentioned that the clustering output does not have much impact on the downstream analysis. I wonder why and if so can we group the data in any way we want for downstream analysis? This can be useful when one would like to focus on certain grouping of cells.

      Significance

      PILOT is designed for analysing scRNA-seq data at a patient level. There is a growing application of scRNA-seq to diseases and the development of computational tools for analysing such data at phenotype level is critical. The key aspect of PILOT compared to other currently available tools is that it enables unsupervised analysis which alleviate the need of patient labels required by many alternative methods.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      The paper describes a computational method, PILOT, that uses optimal transport to compute the Wasserstein distance between two individual single-cell samples. It uses PILOT to detect sample (patient) level trajectories and clusters associated with diseases. The method was applied separately to single-cell genomics data and to digital pathology data. The method was applied to several datasets and compared against other tools.

      Major comments:

      The paper is not easy to follow and should be improved considerably to make it readable and reproducible. Consequently, I was not convinced that the PILOT method is much better than other methods.

      At first read of the title and abstract, I got the impression that the method analyzes single cell and pathomics data concurrently rather than separately. This should be fixed.

      The usage of Wasserstein distance to compute distance between single-cell samples is elegant and is the main strength of this study. Given that PILOT is the main achievement, it should be described more carefully and in a detailed manner.

      For example, in the first Results paragraph, "The test indicates for features explaining the predicted pseudotime by fitting either linear or quadratic models" - I could not understand this sentence. Also, which test do the authors refer to? A few sentences down, there is a reference to a Wald test, is that it?<br /> One of the key aspects of the Wasserstein distance is the cost metric. The determination of the cost metric should be detailed as part of the Results. Have the authors considered and estimated other ways to define the distance?

      Figure 1 provides a schematic view of PILOT. However, there is no explanation of the notation, which makes it confusing rather than helpful. Also, what is the relationship between J and j, if any?

      The motivation and usage of adjusted Rand index (ARI) and Friedman-Nemenyi tests should be provided. Currently, they are not clear, including why those tests are suitable in the cases shown.

      Fig. 2 the use of method colors should be constant across panels. The proportions method works at least as well as PILOT in 2B and 2C (silhouette and AUPR). Explain why PILOT is better. Likewise, Figure 2C,D and Figures S1 and S2 don't show a clear and consistent advantage for PILOT over other methods. Explain what advantage of PILOT do the fraction panels highlight in Fig. 2E and Fig. S3. Fig. 2C is not mentioned in the text.

      I assume Kidney IgAN (text) and Kidney IgA (fig. 2) are the same.

      Fig. 3B fix the p-value notation (what is p=1.05E?) and R2 (R square?). Norte tha both this problem also occurs in other figs. Fig. 3B shows the major cellular changes. Are these changes consistent with known ones? Explain and provide references. Are there cell types that were expected to show a change and did not?<br /> Same questions for Fig. 3C wrt genes. Is this an exploratory analysis highlighting interesting candidate genes? If so, it should be described as such.<br /> The point of Fig. S6 and its major findings should be mentioned in the text (or it can be removed).

      Fig. 4B legend - eGFR not GFR. What do the high-low values of Fig. 2B refer to?<br /> Fig. S12 is out of order in supp file.

      AUCPR - explain.

      The github looks like work in progress with many internal comments (eg, add ,edit, etc). I could not find the tutorials.

      Minor comments:

      Introduction: "Alternatively, trajectory analysis can be performed to uncover disease progression allowing the characterization of early disease events." Citations should be added (some appear later in the text).<br /> "Currently, there are no analytical methods to compare two single cell experiments from the same tissue from two distinct individuals." There have been several comparisons among data from patients, (e.g. Cain et al, 2023), so the authors should be more careful/accurate in their statements.<br /> "Except for PhEMD, all related methods9, 11, 12 require labels of patients for their analysis and cannot be used in the unsupervised analysis " - this sentence comes immediately after describing ref 13, which can be used in unsupervised analysis and accordingly is not cited in this sentence. The authors did well in describing ref 13 (a bioRxiv paper), and its description should come after this sentence.

      "These can be clusters", clustered?

      " acquire an injury cell states" remove an.

      "As for scRNA-seq, there is no analytical method which is able to compare two or more histological slides based on morphometric properties of their structures." The sentence seems to refer to pathomics, not to sc data as suggested in "As for scRNA-seq"

      "Thus PILOT represents the first approach to detect unknown patient trajectories and clusters" patient clusters were also observed by others (eg ref 13, Cain et al).

      Equation 7 - Cosine(Mi,Mi) should be Cosine(Mi,Mj)

      In the beginning of the Results, PILOT is not referred to as a package but as a researcher ("PILOT explores").

      Significance

      In general, the paper is a Methods paper. Hence, likely audience includes computational biologists interested in methodologies, not to biologists interested in the actual findings.

      Although I am among the likely audience, I was not convinced by the merits of the method, potentially due to the way the paper was written.

      I do not have sufficient expertise to check the math.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity):

      Summary:<br /> In this study, the authors delineate the association of paralog dispensability with the frequency of homozygous deletions (HDs) and thereby show that paralog dispensability can play a significant role in shaping tumor genomes. The authors analyzed the strength of negative selection on the paralogs relative to the singletons using frequencies of the homozygous deletions (HD). The study focused on HDs because they ensure a complete loss of function, unlike other mutational aberrations that can be masked because of haplo-sufficiency. While accounting for potential confounding factors, authors find that paralogs tend to have a relatively high frequency of HDs, suggesting a relaxed negative selection. Furthermore, the authors specifically attribute this association to the dispensable paralogs by analyzing gene inactivation data generated from multiple experimental systems. Overall, the findings of this study can potentially have significant implications in cancer biology field and specifically to the researchers studying cancer genome evolution.

      We thank the reviewer for the careful reading and positive assessment of our manuscript

      Major comments:

      1. To dissect further which dispensable paralogs are more likely to be associated with a high HD frequency, synthetic lethal paralogs could be compared with non-synthetic lethal ones.

      In the section titled 'Homozygous deletion frequency of paralog passengers is influenced by paralog properties' (begins from line #289), authors have shown that paralogs with a high frequency of HDs are more likely to have the properties of dispensability (in Figure 4). It seems that all of those properties are also associated with synthetic lethality as the authors identified in their previous study (DeKegel et al. 2021). Furthermore, as shown in the subsequent section ('Essential paralogs are less frequently homozygously deleted than non-essential paralogs', begins from line #344), the high HD is associated with the dispensable paralogs. Some of those dispensable paralogs are expected to be synthetic lethal. Therefore, the association of paralogs with a high frequency of HDs with experimentally validated or predicted sets of synthetic lethal paralogs could be tested. This may help authors to contextualize their findings in terms of genetic interactions between paralogs.

      We thank the reviewer for highlighting the potential relationship with our previous work. We agree that many of these properties are associated with synthetic lethality, but we note that they are also associated with single gene essentiality. This makes the relationship between synthetic lethality, essentiality, and deletion frequency somewhat difficult to dissect.

      Nonetheless we have tested, in a number of ways, whether there is a relationship between a paralog having a reported/predicted synthetic lethality and being homozygously deleted. We find no obvious connection between the two.

      We first tested using a set of synthetic lethal interactions identified by integrating molecular profiling data with genome wide CRISPR screens in a large panel of cancer cell lines (the data used to train the classifier in De Kegel et al, 2021). As there is an ascertainment bias in this dataset (paralogs must have frequent loss of function alterations / silencing to be tested) we restricted our analysis to only those paralog pairs tested for synthetic lethality. We identified no clear pattern (p>0.05, Fisher's Exact Test).

      We next tested using an integrated set of four combinatorial CRISPR screens (aggregated in De Kegel et al) where we considered a pair to be synthetic lethal if it was a hit in any screen and not synthetic lethal if it was screened at least once and never identified as a hit. Again we restricted our analysis to paralogs that were present in this dataset to prevent issues with ascertainment bias. We found no clear association.

      We further tested using a consensus dataset derived from the same combinatorial screens, where a pair were marked as synthetic lethal if they were identified as a hit in at least two screens and not synthetic lethal if they were screened at least twice and never identified as a hit. Again we restricted our analysis to paralogs that were present in this dataset and found no clear association.

      We finally tested using our predicted synthetic lethal interactions – annotating the top 3% of predictions as synthetic lethal and the remainder as non-synthetic lethal. The 3% threshold is similar to the observed frequency of synthetic lethality in the training set. In this case, as this dataset covers all paralogs considered, no restriction was necessary.

      None of the above analyses revealed a clear relationship between deletion frequency and synthetic lethality. A caveat of these analyses is that none of the experimental datasets are complete (covering only a minority of all paralog pairs) and they are all somewhat noisy. Furthermore, as we show in our modelling analysis (Fig S3) the observed homozygous deletions are far from saturating.

      However we think there may be a simpler explanation, beyond limitations of the data, for why we do not observe a relationship between HDs and synthetic lethality.

      As the reviewer notes, there is evidence in cell lines that one reason paralogs are more dispensable than singletons is because of buffering / redundant relationships as revealed by synthetic lethal interactions. These relationships therefore provide an explanation for why some paralogs are dispensable. As our primary claim is that paralogs are more frequently deleted because they are more dispensable we might anticipate a relationship between deletion frequency and synthetic lethality. However, by definition, synthetic lethal interactions can only be observed for non-essential (dispensable) genes. Therefore when analysing the overlap with synthetic lethal interactions we are primarily restricting our analyses to genes that are already individually dispensable. Consequently we might not anticipate observing any enrichment. The buffering relationship revealed by synthetic lethality provides an explanation for why a paralog is dispensable but once we are restricting our analysis to dispensable paralogs we do not necessarily expect to see further enrichment.

      We think that an ideal way to explore this question further would be to look at selection on deletions of pairs of paralogs – we anticipate that if a gene is dispensable because of paralog buffering then both paralogs should not be deleted simultaneously. However, the current copy number datasets are too small to evaluate such pairwise relationships. This is discussed in manuscript as follows:

      Analyzing the frequency with which two members of a paralog family are lost would provide more direct insight into the contribution of paralog redundancy, but due to the overall rarity of passenger gene HDs, we cannot make a comprehensive assessment of co-deletions here – e.g. among paralog pairs where both genes are non-drivers, and not on the same chromosome, only two pairs are co-deleted in at least one TCGA sample. Larger cohorts would also allow us to search for patterns of mutual exclusivity of HDs to identify genetic interactions – this has been applied for identifying interactions between driver genes [57,58]__, but is more challenging for interactions between non-driver genes, which are much less frequently altered.

      Minor comments:<br /> 1. The number of TCGA and ICGC tumor samples analyzed:<br /> As mentioned in the Results section (line #106), 9966 tumor samples were analyzed. However, the sample size mentioned in Figure 2A is 9951. Is the lower number shown in the figure due to the filtering procedure mentioned in the Methods section (line #455)? The change in sample sizes could be explained. A similar difference in sample sizes exists for the ICGC data also.

      The difference was indeed due to filtering process, but numbers were only provided in the methods. We have now addressed this in the main text :

      After removing a small number of ‘hyper-deleted’ samples (see Methods) we retained 9,951 samples for further analysis.

      1. The rationale behind setting the threshold at 100 HD genes to classify 'hyper-deleted' samples for TCGA (line #462) and ICGC data (line #473) could be explained.

      We excluded hyper-deleted samples to avoid any individual sample having undue influence on the genes observed to be ever deleted or indeed to influence the overall patterns observed. It is also common in analyses of selection in tumours that make use of mutational profiles (rather than copy number profiles) to exclude hypermutated samples (e.g. Martincorena et al, Cell 2017; Lopez et al, Nature 2020). However the exact threshold of 100 samples was somewhat arbitrary and this query prompted us to assess whether it had any significant impact on the results.

      We therefore repeated all analyses using a more stringent threshold (50 samples) and also without thresholding. Although the exact percentages and odds-ratios vary somewhat with the different thresholds, all major conclusions are still supported.

      We appreciate that this was minor comment and that reviewer did not request this new analysis, but in the absence of a strong justification for a single threshold we felt it appropriate to assess multiple thresholds (and none).

      1. Citation for DepMap is missing (caption of Figure 5). We have added the text below to the legend for Figure 5 :

      Essential genes for the DepMap dataset (Meyers et al, 2017) are obtained from a version of the data reprocessed in (De Kegel et al, 2021) to reduce off-target sgRNA effects (see Methods).

      CROSS-CONSULTATION COMMENTS<br /> Along the lines of Reviewer #3's second major comment, I have a suggestion, the possible benefits of which would depend on the target audience to which the authors intend to communicate their study.

      I would suggest including a brief comparison of the findings of this study which deal with human paralogs, with the findings in model organisms such as yeast, perhaps in the discussion section. To facilitate such a comparison, authors could try measuring the enrichments of, for example, molecular functions, gene families, types of genetic interactions, etc., among the paralogs associated with a high frequency of HDs and then discussing their comparison with what is known in the literature for paralogs in other model organisms that tend to be frequently deleted.

      Such a comparison could be of interest to the community of researchers working on other model organisms and put this study in a much broader context. However, as I said before, this would depend on the authors' intended target audience.

      We thank the reviewer for the suggestion. We have added an additional section to the discussion highlighting differences and similarities to the observations from yeast as follows:

      Much of our understanding of the factors that influence gene dispensability comes from studies in model organisms, in particular the budding yeast Saccharomyces cerevisiae [3,9,10,43,44]__. Analyses of the yeast gene deletion collection, a set of gene deletion mutants systematically generated in a single S. cerevisiae strain, revealed that paralogs were less likely to be essential than singleton genes [3,45]__. Furthermore, more detailed analyses of yeast paralogs revealed that paralogs from large families were less likely to be essential as were genes with highly sequence similar paralogs [43,44]__. Previous analyses, including our own, demonstrated that many of these trends are also evident when analyzing gene essentiality from CRISPR screens in cancer cell lines [12,13,15,35]__. Our results here are also consistent with these findings – many of the features that are associated with paralog dispensability in yeast are also associated with gene deletion frequency in tumor genomes.

      The connection between the budding yeast observations and those in cancer is less clear when it comes to the relative dispensability of WGDs and SSDs. Analyses of the yeast gene deletion collection revealed that SSDs are more likely to be essential than WGDs in the single genetic background studied [43,44]__. In our previous analyses of gene essentiality in hundreds of cancer cell lines we found that SSDs were more likely to be broadly essential (essential in most cell lines) than WGDs but that WGDs were less likely to be never essential (i.e. more likely to be essential in at least one cell line)__[13]__. As the analyses of gene essentiality in budding yeast were generated in a single genetic background the concordance with our cancer cell line results was difficult to assess, but as gene deletion collections are now being generated in additional yeast strains it should become possible to perform a more direct comparison__[46–48]__.

      Here we found that WGDs are less likely to be deleted than SSDs in tumors. This is surprising in light of the yeast gene deletion collection results, where SSDs were more likely to be essential than WGDs in the strain studied, but less so in light of the cancer cell line results, where WGDs were less likely to be never essential. It is also worth noting that experimental evolution studies in yeast found that SSDs accumulate protein-altering mutations at a higher rate than WGDs [49,50]__. These results are perhaps especially relevant when analyzing the influence of paralog features on selection in tumors.

      We note that there are many additional differences in the features of WGDs and SSDs in budding yeast that may alter their relative dispensability in tumors. An obvious large scale difference is that in the ancestor of humans there were two rounds of whole genome duplication compared to a single duplication event in yeast__[51,52]__. Less obvious, but potentially of importance for cancer, is that the two classes of paralogs are enriched in pathways in humans that do not have obvious counterparts in yeast. For example, WGDs are highly enriched in signaling pathways involved in development while SSDs are enriched in immune response genes__[53]__. How the membership of these pathways influences the dispensability and selection of genes in tumors and cancer cell lines warrants further study.

      Reviewer #1 (Significance):

      As the authors note in their manuscript, it is expected that paralog dispensability could be associated with the relaxed negative selection in tumor genomes because (1) paralogs are prevalent in the human genome, and (2) many of them are dispensable, as apparent from the large-scale gene inactivation screens in hundreds of cancer cell lines (Blomen et al. 2015, Wang et al. 2015, Dandage and Landry 2019, De Kegel and Ryan 2019). However, direct mapping of this association, while importantly accounting for potential confounding factors, has been lacking.<br /> As a researcher with prior experience in the research topics such as gene duplication and genetic interactions, it appears to me that this study presents formal proof of the important association between paralog dispensability and tumor genome evolution which could be of major implication for the research community of cancer biology field and specifically to the researchers dealing with the topics such as cancer evolution, copy number alterations in cancer genomes, and synthetic lethality-based precision oncology therapeutics.

      Thank you again for the positive assessment.

      Reviewer #2 (Evidence, reproducibility and clarity):

      Summary

      Here, De Kegel & Ryan analyse thousands of tumour samples from the TCGA and ICGC projects to identify homozygously deleted genes, finding that about 40% of protein-coding genes are deleted in at least one sample. They find homozygously deleted genes to be enriched for paralogous genes, and further, more frequently deleted genes are increasingly likely to be paralogs. The authors then test the influence of several factors on the likelihood of being deleted, including gene length, distance to a fragile site or chromosomal region, and distance to a recurrently deleted tumour suppressor gene (TSG). They find that proximity of a TSG, telomere, centromere, and fragile site all increase likelihood of being deleted in a sample, as does gene length. Having a paralog also remains an important predictor of deletion after accounting for these other factors. Additionally, the more similar in sequence the closest paralog is to the gene and having a larger gene family size are also predictive of deletion. Conversely, if a gene is a whole genome duplicate as opposed to a small-scale duplicate, it is less likely to be deleted. Finally, the authors test the hypothesis that paralogs that are deleted in cancer are less likely to be essential and find that this is indeed the case.

      Comments

      The authors have done a good job of identifying trends of paralog deletion in cancer samples and the factors influencing them. The results are well described and presented and support the conclusions. I like the inclusion of the saturation analysis as an estimate of what to expect given current and potential future sample sizes, and I appreciate the inclusion of a WGD/SSD paralog distinction. The data and methods are sufficiently detailed. I have a few minor comments below.

      We thank the reviewer for the careful reading and positive assessment of our manuscript

      1. Around line 160 in the text and supplemental figure 4A, the authors test if the trends they see are observed across individual cancer types. With 9 of 33 cancer types reaching a sample size threshold, 8 of 9 comparisons are significant. The authors do not state correcting for multiple testing.

      We have now also assessed the significance of the results after performing a Holm-Bonferroni correction for multiple hypothesis testing and find that all 8/9 cancer types remain significant.

      1. I initially misunderstood the hemizygously deletion analysis, thinking the analysis in supplement figure 4B/C was asking if a sample has any singleton or any paralog deleted and comparing the number of samples with any deletion of either - given the number of genes deleted per sample this wouldn't make sense as a good test. I think the authors are actually comparing the number of loss-of-hemizygosity events per gene and grouping by paralog/singleton. I think this is a good analysis, but I think it would be helpful to clarify this in the text and figure legend e.g. "Samples w/ gene LOH" could be "LOH events per gene" or something similar.

      As suggested we have now updated the y-axis label in these charts to ‘LOH events per gene’. We note that there are now two additional panels in this figure to address copy neutral LOH, per Reviewer 3’s request.

      1. Occasionally, I wanted some more detail in the text for context, which was sometimes later provided - e.g. I noted when reading about line 125 that I was curious at this point how often TSGs occurred on segments, and this was later provided on line 241. Similarly, around line 114 I was curious how many genes are typically deleted per HD segment, for which the median value was provided on line 206 (and distribution in supplemental figure 1), and again for hemizygous deletions. I think sometimes it would be helpful to provide this context earlier in the text to aid interpretation of the results.

      We thank the reviewer for these suggestions which we have now incorporated into the text.

      On line 115 (previously 114) the relevant sentence now reads:

      Typically an HD that results in the loss of a protein coding gene also results in the loss of several chromosomally adjacent genes – in the TCGA dataset a median of three genes are lost per gene-deleting HD segment

      On line 124 the relevant sentence now reads:

      We found that almost half (49%) of the HDs that result in the loss of at least one protein coding gene overlap a known tumor suppressor.

      1. In the discussion, on line 420, the authors include the point that a paralog might not be required at all in a tumour cell and therefore easily deleted. I think this possibility could be expanded on here and in the introduction/results section, as it is an important point. I think it would be helpful to include more about the possibility that a paralog might be deleted in a tumour cell because it is simply not required or that is more likely to have less of a phenotypic impact compared to a singleton, and that this could be a reason for the observed enrichment of paralogs in deleted genes. A citation to support this point could be Áine N O'Toole, Laurence D Hurst, Aoife McLysaght, Faster Evolving Primate Genes Are More Likely to Duplicate, Molecular Biology and Evolution, Volume 35, Issue 1, January 2018, Pages 107-118, https://doi.org/10.1093/molbev/msx270. Duplicate genes can be duplicates because copy number variation of them has minimal impact.

      We thank the reviewer for raising this important point.

      We have briefly addressed this in the introduction as follows:

      In multiple model organisms, paralogs have been demonstrated to be more dispensable than singletons (genes without a paralog) [3–5]__. There are a number of reasons for why a paralog might be more dispensable than a singleton gene, including preferential retention of duplications of non-essential genes [6,7]__, but perhaps the most obvious explanation is buffering between paralogs.

      Where references 6 and 7 are as follows:

      1. O’Toole ÁN, Hurst LD, McLysaght A. Faster Evolving Primate Genes Are More Likely to Duplicate. Mol Biol Evol. 2018;35: 107–118.
      2. He X, Zhang J. Higher duplicability of less important genes in yeast genomes. Mol Biol Evol. 2006;23: 144–151.

      We discuss this more comprehensively in the discussion as follows:

      In both yeast and cancer there are a number of reasons for why paralogs might be more dispensable than singleton genes. Perhaps the most obvious is the existence of buffering relationships between paralog pairs, such that when one paralog is lost the other paralog can compensate for this loss. Such buffering relationships between paralogs can be revealed through synthetic lethality screens and a number of recurrently deleted paralogs in cancer have already been reported to display synthetic lethal interactions with their paralog (recently reviewed in [54]__). Supporting this model, in previous work analysing essentiality in cancer cell lines we found that buffering relationships between paralogs could explain 13-17% of cases where a paralog was essential in some cell lines but not others__[13]__. This suggests that at least some of the increased dispensability of paralogs in cancer cells can be attributed to buffering relationships between paralog pairs. However this is not the only explanation for paralogs displaying increased dispensability in tumour cells. An additional explanation is that paralogs may perform essential functions in specific contexts (e.g. within specific tissues or at specific developmental stages) but are not required within the specific context of a tumour. Consistent with this model, human paralogs are more likely to display tissue-specific expression patterns [55]__. Finally we note that there is evidence to suggest that genes whose perturbation has a lower phenotypic impact may more ‘duplicable’ – i.e. rather than paralogs being under weaker selection because they are duplicated, their duplication was tolerated because they were already under weaker selection__[6,7]__. Teasing apart the relative contributions of these factors to the increased dispensability of paralogs in cancer will require further research and potentially new data resources such as gene essentiality profiles in diverse non-cancer cell types [56]__.

      CROSS-CONSULTATION COMMENTS<br /> I agree, that's a helpful suggestion from reviewer 1.

      Reviewer 3's suggestion regarding age of the two whole genome duplication events is quite difficult to unpick as the duplication events seem to have happened relatively close in time to each other while rediploidisation of the first was occurring. Additionally, paralogs from SSDs tend to be more similar in sequence simply because the two WGD events are relatively old while SSDs can occur at any time up to present. They're therefore biased by young duplicates that have not had the opportunity to diverged much and decrease in sequence similarity.

      We appreciate these comments.

      Reviewer #2 (Significance):

      This is a novel study as it examines the frequency of paralog deletion in cancer samples and the factors influencing it, building upon work already conducted in cancer cell lines. This study extends the knowledge of the field confirming previous trends observed in cell lines, this time in actual cancer samples. It confirms that paralogs are more dispensable than singletons, likely because they have a similar counterpart that can provide some level of functional redundancy. The more similar the closest paralog, the more likely it is to be deleted provides support for this.<br /> It is certainly limited by the number of samples currently available in the two cancer sample projects included but the authors attempt to quantify how limiting this sample size is by conducting a saturation analysis using down-sampling to estimate how many gene deletions one can expect from different numbers of samples. This is important as the lack of observance of many gene deletions is likely due to the limited sample size and not due to negative selection. This low observance of gene deletions disappointingly limits further testing beyond single paralogs to consider the deletion effects of multiple gene family members and more directly test evidence of functional redundancy between paralogs. The authors provide a good discussion of the limitations of their study.

      The results are of interest to evolutionary biologists and cancer biologists. Those with an interest in duplicate genes, and/or factors affecting gene loss in tumours will be interested in this work.

      My field of expertise is molecular evolution, gene duplication and copy number variation.

      We thank the reviewer for the positive assessment of the significance of our work.

      Reviewer #3 (Evidence, reproducibility and clarity):

      Thank you review "Paralog dispensability shapes homozygous deletion patterns in tumor genomes" by DeKegel et al. This manuscript uses TCGA and ICGC tumor data to show evidence for paralog dispensability. They analyze the rate of homozygous deletions and show that it is higher for paralogs compared to singletons. Their findings are robust to a number of confounding variables that they take into account e.g. distance to tumor suppressor, telomere, centromere or fragile site. They show that paralogs that belong to large families and have higher sequence identity tend to show more dispensability and these dispensable paralogs are less likely to be WGD.

      We thank the reviewer for the time taken to review our manuscript.

      Major comments.<br /> 1. Does the finding pertaining to lack of enrichment of paralogs in regions LOH take into account whether LOH is copy neutral or not i.e. how does dosage affects this finding? Is it possible that there is a difference in paralog rate in LOH that results in total copy 1 and that the presence of copy neutral LOH masks the effect? Also, Integration of gene expression dataset would be helpful to resolve the difference between dosage paralog that compensate of the lack of their sister by upregulating their gene expression.

      In the submitted manuscript we focussed solely on LOH events where the copy number of one allele was 0 and the other allele was ≥1. These include copy loss events (total copy number = 1), copy neutral events (total copy = 2), as well as amplifications (total copy number > 2). The rationale for this approach was that we were interested in understanding whether the mechanism that was generating deletions was preferentially generating deletions in paralog-rich regions.

      However, we agree that understanding the influence of dosage is of interest here. We have therefore expanded the analysis in the paper to separately assess the enrichment of paralogs in copy neutral LOH regions (total copy number = 2) and copy loss LOH regions (total copy number = 1).

      As shown in the new updated Figure S4B we do not find an enrichment of paralogs in genes subject to either copy neutral LOH or copy loss LOH.

      The relevant section of the text on page 6 now reads :

      We do not find that paralogs are more frequently subject to LOH than singletons in either the TCGA or ICGC cohort (Fig. S4B-C); when considering all LOH segments we even see that singletons are slightly more frequently subject to LOH in the ICGC cohort (Fig. S4C, left), but when considering only focal LOH segments – i.e. segments whose length is less than half of the chromosome arm’s length, which is the case for all HD segments – there is no significant difference between paralog and singleton LOH frequency in either cohort. To assess whether gene dosage influenced the observed LOH frequency we further restricted our analysis to copy neutral LOH events (total copy number = 2) and copy loss LOH events (total copy number = 1) and again found no significant increase in deletion frequency of paralogs compared to singletons (Fig. S4B-C).

      Regarding the integration of gene expression to identify dosage compensation between paralogs – we agree that this is extremely interesting. However, it is quite challenging to address properly. Most paralogs are only observed to be homozygously deleted a single time and so statistically identifying how loss of one gene impacts the mRNA abundance of another is challenging. In the minority of cases where a paralog is recurrently deleted, often these deletions occur in samples from different cancer types and so integrating transcriptomic data still presents some technical challenges. Given this complexity, and as the question of dosage compensation is not central to our key observations, we have not integrated transcriptomic data here.

      1. Is the finding that paralogs are depleted among WGD is influenced by the age of WGD since there are 2 WGD events? Do SSD tend to be more or less similar by seq than WGD? This should be explored further since this observation is the opposite of what is observed in model organisms such as yeast whereby SSD are less functionally similar than WGD and often show properties similar to singletons than WGD.

      As noted by reviewer 2 in the cross commentary, it is extremely challenging to age the duplicates that arose from the WGD due to the close temporal proximity of the two whole genome duplication events. In the dataset of paralogs analysed used here, SSDs have lower average sequence identity than WGDs. However we note that both sequence identity and duplication type are included in our regression analysis (Figure 4D) and both are significantly associated with homozygous deletion frequently.

      This should be explored further since this observation is the opposite of what is observed in model organisms such as yeast whereby SSD are less functionally similar than WGD and often show properties similar to singletons than WGD.

      We do not actually think that our results are in opposition to the findings from model organisms. The bulk of studies on the functional consequences of deletions of SSDs/WGDs in model organisms are derived from analyses of the budding yeast gene deletion collection, which is generated in a single strain and grown in lab conditions. Consequently these studies report on which genes can be lost in a single genetic background when grown in rich media. We think it is not fully clear how these findings will apply in the context of a panel of genetically heterogenous tumours derived from multiple different cell types. We note that there are additional complexities when analysing human genes (tissue types, two rounds of WGD, metazoan specific pathways enriched in either WGDs/SSDs) that make a straightforward comparison with yeast challenging. We also note that although the results of analyses of the yeast gene deletion collection suggest that SSDs are more likely to be essential than WGDs, experimental evolution studies have demonstrated that SSDs are more likely to accumulate protein altering mutations than SSDs (Keane et al, Genome Research 2014; Fares et al, PLoS Genetics 2013). This is not what would expect based on the analyses of the yeast gene deletion collection, but is closer to what we observe for tumour genomes where SSDs are more likely to be homozygously deleted.

      We agree that we did not adequately discuss these issues in the previous version of our manuscript and so have added a new section to the discussion where we compare our results here with those from budding yeast:

      Much of our understanding of the factors that influence gene dispensability comes from studies in model organisms, in particular the budding yeast Saccharomyces cerevisiae [3,9,10,43,44]__. Analyses of the yeast gene deletion collection, a set of gene deletion mutants systematically generated in a single S. cerevisiae strain, revealed that paralogs were less likely to be essential than singleton genes [3,45]__. Furthermore, more detailed analyses of yeast paralogs revealed that paralogs from large families were less likely to be essential as were genes with highly sequence similar paralogs [43,44]__. Previous analyses, including our own, demonstrated that many of these trends are also evident when analyzing gene essentiality from CRISPR screens in cancer cell lines [12,13,15,35]__. Our results here are also consistent with these findings – many of the features that are associated with paralog dispensability in yeast are also associated with gene deletion frequency in tumor genomes.

      The connection between the budding yeast observations and those in cancer is less clear when it comes to the relative dispensability of WGDs and SSDs. Analyses of the yeast gene deletion collection revealed that SSDs are more likely to be essential than WGDs in the single genetic background studied [43,44]__. In our previous analyses of gene essentiality in hundreds of cancer cell lines we found that SSDs were more likely to be broadly essential (essential in most cell lines) than WGDs but that WGDs were less likely to be never essential (i.e. more likely to be essential in at least one cell line)__[13]__. As the analyses of gene essentiality in budding yeast were generated in a single genetic background the concordance with our cancer cell line results was difficult to assess, but as gene deletion collections are now being generated in additional yeast strains it should become possible to perform a more direct comparison__[46–48]__.

      Here we found that WGDs are less likely to be deleted than SSDs in tumors. This is surprising in light of the yeast gene deletion collection results, where SSDs were more likely to be essential than WGDs in the strain studied, but less so in light of the cancer cell line results, where WGDs were less likely to be never essential. It is also worth noting that experimental evolution studies in yeast found that SSDs accumulate protein-altering mutations at a higher rate than WGDs [49,50]__. These results are perhaps especially relevant when analyzing the influence of paralog features on selection in tumors.

      We note that there are many additional differences in the features of WGDs and SSDs in budding yeast that may alter their relative dispensability in tumors. An obvious large scale difference is that in the ancestor of humans there were two rounds of whole genome duplication compared to a single duplication event in yeast__[51,52]__. Less obvious, but potentially of importance for cancer, is that the two classes of paralogs are enriched in pathways in humans that do not have obvious counterparts in yeast. For example, WGDs are highly enriched in signaling pathways involved in development while SSDs are enriched in immune response genes__[53]__. How the membership of these pathways influences the dispensability and selection of genes in tumors and cancer cell lines warrants further study.

      Minor comments<br /> 1. There is a missing reference on line 55.

      We thank the reviewer for catching this oversight. We have now added a reference to Zerbino et al, NAR 2018 on this line.

      CROSS-CONSULTATION COMMENTS<br /> That's a good suggestion by reviewer 1. Homozygous deletion collection is available in yeast so these data can be used directly in addition tot he haploid gene deletion collection data.

      Since authors of this manuscript included in their analysis the comparison of WGD and SSD then they should do it more thoroughly. It is not sufficient what they presented here especially given that it contradicts the findings from model organisms.

      As noted above we have now added a significant discussion of the yeast findings and also of the SSD/WGD observations

      Reviewer #3 (Significance):

      This work provides the first systematic assessment of paralog dispensability specifically looking at homozygous deletions of paralogs across primary tumor samples and builds on the existing findings in cancer cell lines. It will be broadly interesting to those studying duplicated gene evolution and genome robustness. My expertise is in complex genetic networks in yeast and human cancer as well as genome evolution.

      We thank the reviewer for the positive assessment of our manuscript.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Thank you review "Paralog dispensability shapes homozygous deletion patterns in tumor genomes" by DeKegel et al. This manuscript uses TCGA and ICGC tumor data to show evidence for paralog dispensability. They analyze the rate of homozygous deletions and show that it is higher for paralogs compared to singletons. Their findings are robust to a number of confounding variables that they take into account e.g. distance to tumor suppressor, telomere, centromere or fragile site. They show that paralogs that belong to large families and have higher sequence identity tend to show more dispensability and these dispensable paralogs are less likely to be WGD.

      Major comments.

      1. Does the finding pertaining to lack of enrichment of paralogs in regions LOH take into account whether LOH is copy neutral or not i.e. how does dosage affects this finding? Is it possible that there is a difference in paralog rate in LOH that results in total copy 1 and that the presence of copy neutral LOH masks the effect? Also, Integration of gene expression dataset would be helpful to resolve the difference between dosage paralog that compensate of the lack of their sister by upregulating their gene expression.
      2. Is the finding that paralogs are depleted among WGD is influenced by the age of WGD since there are 2 WGD events? Do SSD tend to be more or less similar by seq than WGD? This should be explored further since this observation is the opposite of what is observed in model organisms such as yeast whereby SSD are less functionally similarthan WGD and often show properties similar to singletons than WGD.

      Minor comments

      1. There is a missing reference on line 55.

      Referees cross-commenting

      That's a good suggestion by reviewer 1. Homozygous deletion collection is available in yeast so these data can be used directly in addition tot he haploid gene deletion collection data.

      Since authors of this manuscript included in their analysis the comparison of WGD and SSD then they should do it more thoroughly. It is not sufficient what they presented here especially given that it contradicts the findings from model organisms.

      Significance

      This work provides the first systematic assessment of paralog dispensability specifically looking at homozygous deletions of paralogs across primary tumor samples and builds on the existing findings in cancer cell lines. It will be broadly interesting to those studying duplicated gene evolution and genome robustness. My expertise is in complex genetic networks in yeast and human cancer as well as genome evolution.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary

      Here, De Kegel & Ryan analyse thousands of tumour samples from the TCGA and ICGC projects to identify homozygously deleted genes, finding that about 40% of protein-coding genes are deleted in at least one sample. They find homozygously deleted genes to be enriched for paralogous genes, and further, more frequently deleted genes are increasingly likely to be paralogs. The authors then test the influence of several factors on the likelihood of being deleted, including gene length, distance to a fragile site or chromosomal region, and distance to a recurrently deleted tumour suppressor gene (TSG). They find that proximity of a TSG, telomere, centromere, and fragile site all increase likelihood of being deleted in a sample, as does gene length. Having a paralog also remains an important predictor of deletion after accounting for these other factors. Additionally, the more similar in sequence the closest paralog is to the gene and having a larger gene family size are also predictive of deletion. Conversely, if a gene is a whole genome duplicate as opposed to a small-scale duplicate, it is less likely to be deleted. Finally, the authors test the hypothesis that paralogs that are deleted in cancer are less likely to be essential and find that this is indeed the case.

      Comments

      The authors have done a good job of identifying trends of paralog deletion in cancer samples and the factors influencing them. The results are well described and presented and support the conclusions. I like the inclusion of the saturation analysis as an estimate of what to expect given current and potential future sample sizes, and I appreciate the inclusion of a WGD/SSD paralog distinction. The data and methods are sufficiently detailed. I have a few minor comments below.

      1. Around line 160 in the text and supplemental figure 4A, the authors test if the trends they see are observed across individual cancer types. With 9 of 33 cancer types reaching a sample size threshold, 8 of 9 comparisons are significant. The authors do not state correcting for multiple testing.
      2. I initially misunderstood the hemizygously deletion analysis, thinking the analysis in supplement figure 4B/C was asking if a sample has any singleton or any paralog deleted and comparing the number of samples with any deletion of either - given the number of genes deleted per sample this wouldn't make sense as a good test. I think the authors are actually comparing the number of loss-of-hemizygosity events per gene and grouping by paralog/singleton. I think this is a good analysis, but I think it would be helpful to clarify this in the text and figure legend e.g. "Samples w/ gene LOH" could be "LOH events per gene" or something similar.
      3. Occasionally, I wanted some more detail in the text for context, which was sometimes later provided - e.g. I noted when reading about line 125 that I was curious at this point how often TSGs occurred on segments, and this was later provided on line 241. Similarly, around line 114 I was curious how many genes are typically deleted per HD segment, for which the median value was provided on line 206 (and distribution in supplemental figure 1), and again for hemizygous deletions. I think sometimes it would be helpful to provide this context earlier in the text to aid interpretation of the results.
      4. In the discussion, on line 420, the authors include the point that a paralog might not be required at all in a tumour cell and therefore easily deleted. I think this possibility could be expanded on here and in the introduction/results section, as it is an important point. I think it would be helpful to include more about the possibility that a paralog might be deleted in a tumour cell because it is simply not required or that is more likely to have less of a phenotypic impact compared to a singleton, and that this could be a reason for the observed enrichment of paralogs in deleted genes. A citation to support this point could be Áine N O'Toole, Laurence D Hurst, Aoife McLysaght, Faster Evolving Primate Genes Are More Likely to Duplicate, Molecular Biology and Evolution, Volume 35, Issue 1, January 2018, Pages 107-118, https://doi.org/10.1093/molbev/msx270. Duplicate genes can be duplicates because copy number variation of them has minimal impact.

      Referees cross-commenting

      I agree, that's a helpful suggestion from reviewer 1.

      Reviewer 3's suggestion regarding age of the two whole genome duplication events is quite difficult to unpick as the duplication events seem to have happened relatively close in time to each other while rediploidisation of the first was occurring. Additionally, paralogs from SSDs tend to be more similar in sequence simply because the two WGD events are relatively old while SSDs can occur at any time up to present. They're therefore biased by young duplicates that have not had the opportunity to diverged much and decrease in sequence similarity.

      Significance

      This is a novel study as it examines the frequency of paralog deletion in cancer samples and the factors influencing it, building upon work already conducted in cancer cell lines. This study extends the knowledge of the field confirming previous trends observed in cell lines, this time in actual cancer samples. It confirms that paralogs are more dispensable than singletons, likely because they have a similar counterpart that can provide some level of functional redundancy. The more similar the closest paralog, the more likely it is to be deleted provides support for this.

      It is certainly limited by the number of samples currently available in the two cancer sample projects included but the authors attempt to quantify how limiting this sample size is by conducting a saturation analysis using down-sampling to estimate how many gene deletions one can expect from different numbers of samples. This is important as the lack of observance of many gene deletions is likely due to the limited sample size and not due to negative selection. This low observance of gene deletions disappointingly limits further testing beyond single paralogs to consider the deletion effects of multiple gene family members and more directly test evidence of functional redundancy between paralogs. The authors provide a good discussion of the limitations of their study.

      The results are of interest to evolutionary biologists and cancer biologists. Those with an interest in duplicate genes, and/or factors affecting gene loss in tumours will be interested in this work.

      My field of expertise is molecular evolution, gene duplication and copy number variation.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary:

      In this study, the authors delineate the association of paralog dispensability with the frequency of homozygous deletions (HDs) and thereby show that paralog dispensability can play a significant role in shaping tumor genomes. The authors analyzed the strength of negative selection on the paralogs relative to the singletons using frequencies of the homozygous deletions (HD). The study focused on HDs because they ensure a complete loss of function, unlike other mutational aberrations that can be masked because of haplo-sufficiency. While accounting for potential confounding factors, authors find that paralogs tend to have a relatively high frequency of HDs, suggesting a relaxed negative selection. Furthermore, the authors specifically attribute this association to the dispensable paralogs by analyzing gene inactivation data generated from multiple experimental systems. Overall, the findings of this study can potentially have significant implications in cancer biology field and specifically to the researchers studying cancer genome evolution.

      Major comments:

      1. To dissect further which dispensable paralogs are more likely to be associated with a high HD frequency, synthetic lethal paralogs could be compared with non-synthetic lethal ones.<br /> In the section titled 'Homozygous deletion frequency of paralog passengers is influenced by paralog properties' (begins from line #289), authors have shown that paralogs with a high frequency of HDs are more likely to have the properties of dispensability (in Figure 4). It seems that all of those properties are also associated with synthetic lethality as the authors identified in their previous study (DeKegel et al. 2021). Furthermore, as shown in the subsequent section ('Essential paralogs are less frequently homozygously deleted than non-essential paralogs', begins from line #344), the high HD is associated with the dispensable paralogs. Some of those dispensable paralogs are expected to be synthetic lethal. Therefore, the association of paralogs with a high frequency of HDs with experimentally validated or predicted sets of synthetic lethal paralogs could be tested. This may help authors to contextualize their findings in terms of genetic interactions between paralogs.

      Minor comments:

      1. The number of TCGA and ICGC tumor samples analyzed:<br /> As mentioned in the Results section (line #106), 9966 tumor samples were analyzed. However, the sample size mentioned in Figure 2A is 9951. Is the lower number shown in the figure due to the filtering procedure mentioned in the Methods section (line #455)? The change in sample sizes could be explained. A similar difference in sample sizes exists for the ICGC data also.
      2. The rationale behind setting the threshold at 100 HD genes to classify 'hyper-deleted' samples for TCGA (line #462) and ICGC data (line #473) could be explained.
      3. Citation for DepMap is missing (caption of Figure 5).

      Referees cross-commenting

      Along the lines of Reviewer #3's second major comment, I have a suggestion, the possible benefits of which would depend on the target audience to which the authors intend to communicate their study.

      I would suggest including a brief comparison of the findings of this study which deal with human paralogs, with the findings in model organisms such as yeast, perhaps in the discussion section. To facilitate such a comparison, authors could try measuring the enrichments of, for example, molecular functions, gene families, types of genetic interactions, etc., among the paralogs associated with a high frequency of HDs and then discussing their comparison with what is known in the literature for paralogs in other model organisms that tend to be frequently deleted.

      Such a comparison could be of interest to the community of researchers working on other model organisms and put this study in a much broader context. However, as I said before, this would depend on the authors' intended target audience.

      Significance

      As the authors note in their manuscript, it is expected that paralog dispensability could be associated with the relaxed negative selection in tumor genomes because (1) paralogs are prevalent in the human genome, and (2) many of them are dispensable, as apparent from the large-scale gene inactivation screens in hundreds of cancer cell lines (Blomen et al. 2015, Wang et al. 2015, Dandage and Landry 2019, De Kegel and Ryan 2019). However, direct mapping of this association, while importantly accounting for potential confounding factors, has been lacking.<br /> As a researcher with prior experience in the research topics such as gene duplication and genetic interactions, it appears to me that this study presents formal proof of the important association between paralog dispensability and tumor genome evolution which could be of major implication for the research community of cancer biology field and specifically to the researchers dealing with the topics such as cancer evolution, copy number alterations in cancer genomes, and synthetic lethality-based precision oncology therapeutics.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      We would like to thank all reviewers for taking the time to evaluate our manuscript. Many helpful suggestions and discussion points were raised. These comments were instrumental to provide more data that strengthen our conclusion about the relevance of centrin condensation in vivo, expand our findings to other organisms, and improve the manuscript in general. Details are given in the following individual replies.

      Reviewer #1 (Evidence, reproducibility and clarity):

      Voss and colleagues show calcium-dependent assembly of Plasmodium falciparum centrins in vitro and in parasites. This assembly is dependent on the EF-hands of centrin and an N-terminal disordered region.

      Major concerns:

      1. The very definitive title is not wholly supported by the data. This should be qualified by specifying the conditions under which the centrins can accumulate in this way.

      We understand this comment by the reviewer. There are multiple dimensions to the potential of centrins to condensate, such as the specific centrin family member, in vivo vs in vitro situation, and media conditions. Naturally it is difficult to represent these various conditions in a concise and compelling title but in line with the suggestion by Reviewer 2 we are changing the title to “Malaria parasite centrins can assemble by Ca2+-inducible condensation” to reflect the conditionality of this process.

      1. A major concern is whether this behaviour of centrins represents a biologically relevant mechanism in centriolar plaque formation. Is this limited to high overexpression conditions or in vitro high concentrations? Or is it a result of the tagging of the P. falciparum centrins?...

      Centrin accumulation at the centriolar plaque and assembly of the centriolar plaque itself must be differentiated. Although compelling we are already very careful in the text about extrapolating our findings about centrin accumulation in cells to centriolar plaque or centrosomal assembly in general. We, however, thank the reviewer for this important comment and now have carried out hexanediol treatment of wild type parasites to test the effect on centrin in a native context. After IFA staining we failed to detect any centrin foci at the centriolar plaques, suggesting that they can be resolved by inhibiting weak hydrophobic interactions that are typical for phase separation (now Fig. 6, lines 283ff).

      Concerning the effect of tagging we have generated new data of cells overexpressing an untagged version of PfCen1 in parasites, which still shows formation of ECCAs as revealed by IFA (now Fig. 4H-K, lines 243ff). This significantly alleviates the concern that the observed phenomenon is only a consequence of GFP-tagging. Our in vitro data already showed that native and tagged PfCentrin1 & 3 can undergo condensation.

      Concerning the critical concentration of our in vitro assay we find it to be around 10-15 µM without the addition of crowding agents such as PEG (now Fig. S3C, lines 120ff). To our understanding it is challenging to select an in vitro concentration that is adequate to define a threshold for “biological relevance” due to so many additional factors playing a role in vivo. Those factors can also favor a phase separation locally when total saturation concentration is not reached as we now discuss in more detail (lines 440ff). For reference the critical concentration of FUS, which is one of the most studied phase separating proteins in model system, is around 2 µM, but concentrations below 15 µM are well within the range of what is observed for in vitro LLPS. Additionally, it is important to consider that we find Cen1/3 and HsCen2 LLPS is inducible and reversible and that very homologous proteins i.e. Cen2 and 4 serve as an adequate internal control.

      … A convincing approach to addressing this issue would be to knock-in a fluorescent tag to the centrin loci. Roques et al. (ref. 12 in this submission) report the GFP tagging of centrin-4 in P. berghei, although they note that centrins-1 to -3 were refractory to tagging in this organism. It is unclear whether Voss et al. attempted this tagging in P. falciparum. This should be clarified and relevant data presented.

      We indeed attempted several unsuccessful iterations of tagging Cen1/3 with HA and GFP tag and now explain this in the text more clearly (lines 81ff). We did not attempt tagging Cen2 and 4 as they do not display phase separation in vitro or carry IDRs.

      If the tagged molecules used in the biochemical parts of this study are functional, it is challenging to understand why the centrins cannot be tagged in P. falciparum. If the tags render the P. falciparum centrins dysfunctional, the study becomes significantly less useful.

      Our data shows that in vitro Cen1-GFP can undergo Ca2+-inducible and reversible LLPS and that GFP-tagged centrins can still localize to the centriolar plaque. Centrin function, however, certainly goes beyond its ability to condensate and localize. It is easily conceivable that interaction with critical binding partners at the centriolar plaque is inhibited by tagging a protein as small as centrin, which prohibits tagging the endogenous version, while its ability to phase separate remains unaltered. To dynamically study a protein in cells tagging is, however, unavoidable. Even though tagging affects any proteins function to highly variable degree we are still convinced that studying those proteins still provides useful information. Our mutant versions of PfCen1 in vivo shows that non-condensating version display different localization. Importantly, as mentioned above, we now provide images of cells overexpressing an untagged Cen1 version, which still causes ECCA formation (Fig. 5H-K). Ultimately, even though tagged versions might not be fully functional, our observations are compatible with the ability of centrins to condensate in vivo.

      1. If a knock-in cannot be achieved, it must be shown that the transgenic expression of tagged Plasmodium centrins does not confound the analysis of centrin behaviour. It is known that these proteins can behave anomalously when overexpressed (Yang et al. 2010, PMID: 20980622; Prosser et al. 2009, PMID: 19139275), at least in other species.

      Thank you for this comment. Transgenic expression of proteins can in principle influence their behavior. In the context of this study the overexpression is, however, used intentionally since protein concentration correlates with the phase separation. Here, transgenic overexpression is used as a tool, rather than being a confounding factor, and ECCA formation can be used as quantifiable phenotype. The observation that ECCAs appear significantly earlier the higher they are expressed is in our opinion one of the stronger points of evidence that this result from phase separation in vivo. Yet centrins maintain their centriolar plaque localization and no significant impact on growth is observed. To definitely answer whether phase separation of endogenous centrin is occurring during centriolar plaque accumulation is challenging. These challenges and limitations are now addressed in the significantly extended discussion. As explained above untagged Cen1 also forms ECCAs.

      A previous description of centriolar plaque from the authors' lab (Simon et al. 2021, PMID: 34535568) shows an organized structure of an established size. It should be demonstrated whether the structures formed with the GFP tagged centrins show the same dimensions and dynamics as those in wild-type parasites. The extent of the overexpression of the GFP-tagged centrins should also be demonstrated.

      We thank the reviewer for this suggestion. We have now added spatial measurements of the centrin signal dimensions at the centriolar plaque of mitotic spindle containing nuclei in PfCen1-GFP overexpressing vs non-induced cell lines. We found that the width of the centrin-signal at the centriolar plaque was unaltered while the height only increased by 11% (Fig. S9). Further, we found no significant growth phenotype in overexpressing parasites, which indicates that the centriolar plaque is functional.

      Due to several confounding factors, we were, unfortunately, unable to clearly quantify the extent of overexpression. Most notably the induction of overexpression only works in about 50% of the cells (Fig. S6). The mean intensity after induction further displays quite some variability. Furthermore, the expression kinetics along the IDC of endogenous centrin and our overexpression system that we use as a tool differ. Lastly, our centrin antibodies display crossreactivity (see also Fig. S12) making it impossible to identify how much of the endogenous pool we are labeling in comparison to the GFP- tagged Cen1 protein.

      1. It would also be useful to remove the His tag from the recombinantly expressed and purified centrins for the in vitro analyses, particularly if concern remains about the impact of tags on Plasmodium centrin behaviour.

      Based on the published in vitro studies on other centrins, we did not anticipate the His-tag to change LLPS properties. Also, Cen1 and 3 and Cen2 and 4 would need to be differentially affected by the tag. We further have experimented with N-terminally tagged 6His-Cen3 protein and found no significant differences in our turbidity assays. Nevertheless, we expressed new versions of the recombinant PfCen1-4 proteins with a TEV cleavage site inserted after the His-tag to purify untagged proteins and found no fundamental differences in our LLPS assay aside some slight variation in the kinetics (Fig. S3E).

      1. The discussion is very short and does not consider the findings presented here in the context of the literature, with respect to centrins, Plasmodium MTOC assembly mechanisms, or to general considerations around biological condensates. Andrea Musacchio's recent commentary (ref. 44 in the current submission) advocates caution in ascribing phase separation as an assembly mechanism for organelles in vivo, particularly on the basis of in vitro experiments with high concentrations of homogeneous protein. It is not clear that the concentration dependence of extracentrosomal centrin accumulations (ECCAs) at the onset of schizogony provides sufficient justification of a phase separation model in vivo. The authors' recent description of the involvement of an SFI1-like protein, SIp (Wenz et al. 2023 PMID: 37130129), in the centriolar plaque makes a case for non-homotypic interactions also driving assembly and alternative models for ECCA are not convincingly excluded. The absence of a robust discussion of such considerations is unhelpful to the reader.

      We very much thank the reviewer for this suggestion, which helped to significantly improve the manuscript. We have purposefully included the commentary by Andrea Musacchio to highlight a different (possibly the most antipodal) point of view on the role of biomolecular condensation in membraneless organelle formation for the unfamiliar readers that might be just getting to know the field of phase separation. In the absence of word limitations, the reviewer is right to point out the lack of more extensive discussion. We now have significantly extended this section and address the suggested points including the potential role of the novel centriolar plaque protein Slp, which was not published upon submission of our previous version (lines 450ff.)

      1. It is also unclear whether the analysis of human centrin is suggested to indicate a phase separation mechanism for centrins in human cells. As this is readily testable, this notion could be considered further. Although its experimental examination may lie outside the theme of this study, one would expect some discussion of the significance of the data presented in the study.

      Since it is the first description of phase separation of centrin, it would indeed be interesting to explore the functional relevance in other organisms such as humans. We are considering approaching this in the future. We have, as requested above, significantly extended the discussion and now also include this aspect. Earlier reports have e.g. shown centriole overduplication in human cells upon centrin overexpression.

      Minor points

      1. There are only three centrins in humans. Centrin 4 is a pseudogene (Gene ID: 729338 on NCBI).

      Thank you for detecting this error, which we now corrected (line 60). Centrin 4 seems only to be an expressed gene in mice.

      1. Line 175 should say 'temporally', rather than 'temporarily. The Abstract should say 'evolutionarily conserved', rather than 'evolutionary conserved'. 'To condensate' is not ideal as a phrase- 'to form a condensate' would be clearer.

      Thank you for those suggestions. The text has been modified accordingly.

      Referees cross-commenting

      I think the other 2 reviewers have made fair, cogent and constructive points. There is good convergence between the reviewers on the significant issues around the study. These concern in vivo and in vitro effects of tagging and of high concentrations.

      Reviewer #1 (Significance):

      The biology of the Plasmodium centriolar plaque is of great interest as an alternative MTOC structure, with obvious additional interest deriving from the role of this organism in malaria. Much remains to be learned about this structure, so the topic of this paper is likely to attract a broad readership. Furthermore, the centrins are a widely-expressed and evolutionarily conserved family of eukaryotic proteins, with multiple roles; a new model for their behaviour, such as is suggested here, would be of interest to many cell biologists.

      With that in mind, significant additional data should be provided to substantiate the model proposed by the authors.

      We appreciate that the reviewer considers our manuscript of interest for a broad audience. We feel that our modifications of the text including a more thorough contextualization and addition of some new experimental data now sufficiently supports our claims.

      Reviewer #2 (Evidence, reproducibility and clarity):

      The authors analyzed the properties of the four Centrin proteins of the malaria parasite using a combination of in vitro and in vivo approaches. Their findings indicate that two of the four Plasmodium Centrin proteins, PfCen1 and PfCen3, as well as the human Centrin protein HsCen2, exhibit features of biomolecular condensates. Moreover, analysis of cells overexpressing PfCen1 indicates that such biomolecular condensates become more numerous as cells approach mitosis and are dissolved thereafter.

      Major comments

      A) A critical point that requires clarification is how the protein concentrations used in the in vitro and in vivo assays (20-200 microM in vitro, and not estimated in vivo) compare to that of the endogenous components. This is important because it may well be that 6His-tagged PfCen1, PfCen3 and HsCen2 can form biomolecular condensates when present in vast excess, but not when present in physiological concentrations. The authors should report the estimated cellular concentration of PfCen1-4, as well as that achieved upon PfCen1-GFP overexpression (on top of endogenous PfCen1), for instance using semi-quantitative immunoblotting analysis. Given this limitation, the authors may also want to temper their title by introducing the word "can" after "centrins".

      In the context of phase separation, protein concentration is of course a critical metric. However, in vitro and in vivo concentrations cannot be directly compared as the composition of the surrounding solute has a significant impact on the effective saturation concentration. In vitro we find a saturation concentration for Cen1 of 10-15 µM (Fig. S3C), which is within a range that is frequently found other in vitro studies as listed in the in vitro LLPS data base (PMID: 35025997). We now more explicitly discuss this in the text (lines 422ff). At this point, unfortunately, we have no means of investigating the absolute concentrations of centrin in vivo and to our knowledge no such data is available for apicomplexan. Additionally, one has to keep in mind the presence of other centrin family members in the cell which can interact and co-condensate as well as other centriolar plaque proteins, like PfSlp, but are difficult to separate through analysis. Further we now discuss several contexts that modify the saturation concentration in vivo (lines 440ff).

      As explained above in a response to Reviewer 1, we were not able to produce a satisfactory quantification of the overexpression levels. We are repasting the previous response here:

      “Due to several confounding factors we were, unfortunately, unable to clearly quantify the extent of overexpression. Most notably the induction of overexpression only works in about 50% of the cells (Fig. S6). The mean intensity after induction further displays quite some variability. Lastly the expression kinetics along the IDC of endogenous centrin and our overexpression system that we use as a tool differ. Lastly, our centrin antibodies display crossreactivity (see also Fig. S12) making it impossible to identify how much of the endogenous pool we are labeling in comparison to the GFP- tagged Cen1 protein. “

      Concerning the title, as explained above, we followed the suggestion and added the word “can”.

      B) Movies S1 and S2 (and the related Fig. 1D and 1E) are not the most convincing to support the notion that the observed assemblies are biomolecular condensates, as not much activity is going on during the recordings. Likewise, Movies S3, and even more so Movie S4, as out of focus for a large fraction of the time, making it difficult to assess what happens at the beginning of the process. Moreover, it appears that fusion events, while occurring, are rather rare. The movies should be exchanged for ones that are in focus, and ideally a rough quantification of fusion events as a function of biomolecular condensate size provided.

      We thank the reviewer for requesting clarification. Movies S1 and S2 are by no means direct evidence for biomolecular condensation and we do not claim them to be but rather say that they are “…reminiscent of biomolecular condensates…”. We think that this is an appropriate entry into the subsequent analyses. For Movie S1 it is noteworthy that the shape of the accumulation, which can only be resolved by super-resolution microscopy in live cells, is round as would be expected for a liquid condensate in the absence of forces and on these short time scales. Nevertheless, the centriolar plaque must be duplicated which might be the process partly depicted in Movie S2. The observation that centrin can be still change its shape at least suggests that it is not a solid aggregate. In the context of centriolar plaque biology and the technological advance of applying live cell STED in P. falciparum, we think these data are still worth reporting.

      Concerning Movies S3 and S4 we have carefully selected the focal plane to highlight all the hallmarks of LLPS. Since the protein droplets freely move in 3D throughout the entire imaged liquid volume there is no z-plane that is in focus. Our positioning of the focal plane presents the best compromise between showing round droplet shape, droplet fusion events, and surface wetting. All those observations demonstrate the liquid nature of the condensates. Fusion events are indeed relatively rare, and we do not go beyond this qualitative statement that it can be seen.

      C) An important control is missing from Fig. 2, namely assaying PfCen1-4 without the 6His tag, to ensure that the tag does not contribute to the observed behavior (although it can of course not be sufficient as evidenced by the lack of biomolecular condensates for PfCen2 and PfCen4).

      Thank you for this suggestion. Since reviewer 1 made a similar comment, I’m reiterating our previous reply here: Generally speaking, and based on the published in vitro studies on other centrins, we didn’t anticipate the very small His-tag to change LLPS properties. Also, Cen1 and 3 and Cen2 and 4 would need to be differentially affected by the tag. We further have experimented with N-terminally tagged 6xHis-Cen3 protein and found no significant differences in our turbidity assays. However, we expressed new versions of the recombinant PfCen1-4 proteins with a TEV cleavage site inserted after the His-tag to purify untagged proteins and found no significant differences in our LLPS assay (Fig. S3E).

      D) The authors should test whether the assemblies formed by PfCen1 and PfCen3 are sensitive to 1,6-hexanediol treatment, as expected for biomolecular condensates.

      This is an interesting and helpful suggestion. We now tested 1,6-hexanediol addition to recombinant PfCen1 and wildtype parasites (now Fig. 6). Interestingly the dissolving effect of hexanediol on PfCen1 in vitro was moderate, which we attribute to the polar component in centrin assembly, which has been documented earlier (Tourbez et al. 2004). In vivo, however, only 5 min of treatment caused a striking dissolution of most centrin foci in wild type parasites, which is compatible with the interpretation that centrin or centriolar plaque assembly could be driven by biomolecular condensation.

      E) The fact that HsCen2 also forms biomolecular condensates is very intriguing, but further investigation would be needed to assess the generality of these findings. For instance, the authors could test in vitro also S. cerevisiae Cdc31, the founding member of the Centrin family of proteins to further enhance the impact of their study.

      We thank the reviewer for this suggestion. It would of course be exciting to investigate in more detail how widely this biochemical property of some centrins is conserved. To take a first step in that direction, we have recombinantly expressed centrins containing some N-terminal IDRs from C. reinhardtii, T. brucei and S. cerevisiae to represent organism of significant evolutionary distance. Using our in vitro phase separation assays, we found a very similar behavior to PfCen1 for two centrins while yeast Cdc31, although forming droplets, had a much higher saturation concentration, which could be explained by the significantly lower intrinsic disorder in its sequence (now new Fig. 3).

      Minor comments

      1) For the experiments reported in Fig. 3D, the same concentrations as those used in Fig. 3A-C (namely 10 microM, and not 30 microM as in Fig. 3D) should be used. Moreover, it would be informative to test whether PfCen2 and PfCen4 as PfCen3 when added to PfCen1.

      Unfortunately, this experiment is not feasible since Cen3 does not produce droplets at 10 µM. Hence, in Fig. 3D we aimed to test if Cen1 is incorporated into preformed droplets i.e. whether there is still some interaction between them. We have, however, tested the addition of Cen2 to Cen1 and Cen3 and as expected from the inability PfCen2 to condensate we did not find the same synergistic effect as for Cen1 and 3 together (now Fig. S6). The combination of Cen1/2/3 still enabled co-condensation while addition of Cen4 did not further improve droplet formation. Taken together this strongly suggests that only Cen1 and 3 contribute to the phase separation in vitro (lines 184ff).

      2) The authors mention that the effect of Calcium in inducing biomolecular condensates is specific, as Magnesium was not effective (lines 94-95). However, an examination of Fig. S3B indicates that the Magnesium also exhibits some activity, albeit less potent than Calcium. The authors should discuss this point and rectify the wording in the main text.

      Thank you for pointing this out. While PfCen1 is not reactive to Magnesium, PfCen3 and HsCen2 do display a small reaction, which we now more clearly mention in the text (lines 118ff). Of note Mg2+ and other divalent cation are known to generally promote phase separation.

      3) Do the authors think that PfCen2 and PfCent4 localize to the centriolar plaque in vivo using another mechanism that deployed by PfCen1 and PfCent3? It would be good to discuss this point.

      This is indeed a point worth discussing. Centrins can of course still interact in the absence of biomolecular condensation and their localization to the centriolar plaque is not dependent on their ability to phase-separate as seen for PfCen2 and 4. We have recently described a novel centriolar plaque protein PfSlp that interacts with centrins and might assist recruitment (Wenz et al. 2023). Cellular condensates are, however, often separated into scaffold proteins, which actually phase separate and client protein which get recruited into those condensates. It is easily conceivable that Cen1 and 3 participate in formation of the biomolecular condensate into which Cen2 and 4 as well as other centriolar plaque proteins might be recruited. Unfortunately, we were not yet able to establish a recruitment hierarchy by e.g. dual-labeling of centrins to test whether PfCen1 and 3 might appear prior to PfCen2 and 4. We now include those aspects in the extended discussion.

      4) Given that the EFh-dead mutant exhibits no activity in vitro and fails to localize in vivo, one potential concern is that the protein is misfolded. The authors should conduct a CD spectrum to investigate this.

      Thank you for suggesting this relevant control experiment. We have carried out CD spectroscopy of wild type and EFh-dead PfCen1 and find no difference in secondary structure distribution. We now added these data to the supplemental information (now Fig. S14).

      5) It is not entirely clear from the main text in lines 103-104, as well as from the legend, what Fig. S3B shows. When was EDTA added in this case?

      Thank you for requesting clarification. We will assume the reviewer is referring to Fig S4B. We wanted to show that contrary to PfCen3 that PfCen1 droplets can still be resolved after an elongated period of incubation with calcium but forgot to mark the timepoint of EDTA addition at 180 min in the graph. We have now corrected this and further reworded the sentence for more clarity (lines 132ff).

      6) Fig. S7: the correlation between PfCen1-GFP expression levels and ECCA appearance is modest at best. What statistical test was applied? This should be spelled out. Moreover, the authors should combine the two data sets, as this will provide further statistical power to assess whether a correlation is truly present.

      Indeed, the correlation is modest but statistically significant, which is why we decided to place this data in the supplemental information. The used statistical test was an F-test provided by Prism, which compares two competing regression models, which we now mention in the legend. Combining the two data sets is unfortunately not possible since they arise from two independent sets of measurements where different imaging settings had to be used to adjust for the very different fluorescent protein levels in both lines after induction.

      7) The authors may want to discuss how their findings can be reconciled with the notion that Centrin assemble into a helical polymer on the inside of the centriole (doi: 10.1126/sciadv.aaz4137).

      This is an interesting point. Although centrin does localize to the inside of the centriole (https://doi.org/10.15252/embj.2022112107), more precisely one pool at the distal part and one pool at the core, there is no evidence that it is itself part of the helical inner scaffold described by the authors even though it might localize in close proximity to it. Further, there are several examples where polymers such as microtubules act as seeding point for biomolecular condensates or the other way around, and our work suggest this could be a potential working model for centrins. We have discussed our results extensively with the two corresponding authors of the aforementioned study (i.e. Virginie Hamel and Paul Guichard) and agreed that our data are not conflicting. Nevertheless, we include the inner centriole localization and potential association with polymer structures of centrin in our extended discussion.

      9) Likewise, the authors may want to speculate regarding what their findings signify for the role of Centrin proteins in detection of nucleotide excision repair (doi: 10.1083/jcb.201012093).

      We appreciate the comment by the reviewer. Centrins seem to have many different potential roles that remain to be clarified. While we are excited about this, we think it is too early to speculate about the impact of centrin condensation on less well studied aspects of centrins such as nucleotide excision repair. We, however, now cite this study in the discussion to highlight the functional diversity of centrins.

      Small things

      • Fig. 1A: change color for microtubules as red on red is difficult to discern.

      Throughout our publications we use this shade of magenta to label microtubules in schematics and have therefore opted to use a slightly brighter shade of red for the RBCs instead to improve visibility.

      • Fig. 1C: the indicated boxes in the top row do not seem to correspond exactly to the insets shown in the bottom row.

      We have verified the position of the boxes and found them to be accurate. Possibly the different imaging modality used for both panels (confocal vs STED) creates this impression.

      • line 266: typo, promotor > promoter.

      Has been corrected.

      • line 360: a reference should be provided for the GFP-booster, including the concentration at which it was used.

      Has been added.

      • line 363: "an" missing before "HC".

      Has been corrected.

      • line 428: it would be best to deposit the macros on Github or an analogous repository.

      Macros have been deposited on https://github.com/SeverinaKlaus/ImageJ-Macros (line 737)

      • line 461: "to the" is duplicated.

      Has been corrected.

      • Fig. S5A: maybe draw the lines in red (as red in Fig. S5B correspond to the proteins that do not have IDRs).

      Since we cannot easily change the line colors of the IDR graphs, we have inverted the font color for Fig. S5B instead.

      • Movie S7, legend: left frames shows PfCen1-GFP, not microtubules as currently stated.

      Has been corrected.

      Reviewer #2 (Significance):

      This is a provocative study that extends initial observations regarding self-assembly properties of Centrin proteins, and posits that some members of this evolutionarily conserved family can form biomolecular condensates. After the above outstanding issues have been properly addressed, these data could have important implications for understanding Centrin function in centriole biology and DNA repair. Therefore, these findings will be of interest to a cell biology audience.

      Field of expertise: cell biology.

      Reviewer #3 (Evidence, reproducibility and clarity):

      Summary:

      The authors have provided a comprehensive characterisation of centrin proteins in Plasmodium falciparum. Through expression of episomal GFP-tagged centrin for in vitro, they were able to observe co-localisation of centrin with centriolar plaques during the replicative stage of the parasite. They also utilised live cell STED microscopy to track dynamic changes in centrin morphology. They have also demonstrated calcium-dependent phase separation dynamics in bacterially-expressed P. falciparum centrin and human centrin 2. The formation of liquid-liquid phase separation in PfCen1, 3 and HsCen2 tied well with IUPred3 predictions of intrinsically disordered regions in these proteins. Using an inducible DiCre overexpression system with two promoters of varying strengths, the authors have shown accumulation of centrin1 outside of centrosomes and premature appearance of centriolar plaques. Finally, changes on the centrin1 protein, i.e., N-terminal deletion, and mutations in calcium binding sites in the EFh domains, have shown a reduction in the formation of ECCAs during overexpression and inability to form LLPS in vitro, respectively.

      Major comments:

      1. Given that parasites cannot tolerate endogenous C-terminal tagging of some centrins (but not all, as PbCen4 was successfully tagged), has N-terminal tagging been attempted either by the authors or in previous publications? Note that this is not a request for further experimentation; rather, maybe this can be noted in the manuscript; and line 62 can be rephrased for transparency.

      We have not attempted N-terminal tagging ourselves but through personal communication with Rita Tewari we were informed that neither N- nor C-terminal tagging for PbCen1-3 was successful in the context of the study published by Roques et al 2018. We have only unsuccessfully attempted C-terminal tagging in several iterations. Due to importance of N-terminus for interaction and function in other organisms it is plausible that N-terminal tagging is even more unlikely to work. Since we have not exhaustively attempted every tagging strategy on every centrin we, as suggested, rephrased the text accordingly (lines 81ff).

      1. Is there a possibility that by adding a C-terminal tag, centrin may lose a specific function or cause change in the physicochemical properties of the protein (thus making C-terminal tagging lethal)? Was His tag removal attempted so the native protein can be used in the LLPS experiments? IUPred3 analysis showed potential IDR at the C-terminal end of PfCen4. Could the C-terminal tag have caused the protein to not form droplets in the presence of Ca2+?

      As we could show for PfCen1-GFP, the tag did not impair its ability to undergo LLPS which is at least partly mediated by the N-terminus, and that it could still properly localizes to the centriolar plaque. The fact that some endogenous centrins cannot be tagged suggest that there is a functional relevance to the C-terminus that could e.g. be an interaction with other essential centriolar plaque components. As suggested in a reply to Reviewer 1, we consider a substantial and centrin-specific effect of the small His-tag on phase separation unlikely. To be sure, we have repeated our turbidity assays with tag-free versions of PfCen1-4 and found no change in phase separation properties (now Fig. S3E).

      1. It has been shown by the authors that different tagged centrins co-condense which may support the localisation data (Figure 1C). However, is there a way to show that the episomally- and endogenously-expressed centrin co-localise with each other (e.g., confocal microscopy with anti-centrin vs anti-gfp in PfCen-GFP lines, that is if the authors have access to anti-centrin antibodies)? Has endogenous centrin been demonstrated to form ECCAs (in previous publications or by the authors)?

      These are important questions by the reviewer. Due to the high sequence homology centrin antibodies, even if raised against a specific centrin (such as PfCen3 in this study), will likely cross-react with other centrins. So far, we have not been able to produce a staining were the anti-GFP-positive foci are devoid of anti-centrin3 staining, which limits the interpretation of these data. The outer centriolar plaque compartment containing centrin is, however, well defined by now and the localization pattern of endogenous centrin and Centrin1 and 4-GFP seems identical. In a more recent study from our lab Cen1-GFP IP has identified other endogenous centrins as interaction partners (Wenz et al 2023), like the Roques et al. 2018 study did for PbCen4-GFP indicating that the tag does not abolish interaction between centrins. So far, we have never detected any ECCAs, nor have we identified any similar structure in the literature. This suggest that this is indeed a consequence of excessive centrin concentration. Importantly we now have added data from a new parasite line overexpressing untagged PfCen1 using the T2A skip peptide (pFIO+_GFP-T2A-Cen1) which displays ECCAs upon induction, showing that this effect is not a mere consequence of tagging (now Fig. 5H-K).

      Minor comments:

      1. How were the times (post addition of Ca2+) presented in Figure 2A determined?

      We noted down the time of calcium addition and cross-referenced it with the timestamps available in the metadata of the movie files (e.g. file creation timepoint marks the start of the movie). We now mention this in the legend.

      1. Line 126: Figure 1B should be Figure 1C

      2. Line 145: Figure 1C-D should be Figure 1D-E

      3. Line 151: Figure 3A should be Figure 4A

      Thank you for spotting these mistakes, which now have been corrected.

      1. Line 152: Suggest rephrasing "placing the gene of interest in front of the promoter" to "placing the gene of interest immediately downstream of the promoter" or something similar

      Thank you for this good suggestion.

      1. Any growth phenotype changes observed in the overexpressors?

      The parasite lines seem to silence the Cen1-4-GFP expression plasmids readily, which suggest that there might be a growth disadvantage. However, repeated attempts to quantify a growth phenotype were unsuccessful due to high variability in the data, which might be partly connected to the fact that the fraction of GFP positive cells after induction can vary between lines and replicas.

      1. How often are ECCAs observed in pARL strains, or are they not observed at all? This might be good to mention.

      ECCAs in the pArl strains have been observed on very limited instances but are too rare to be quantified. We now mention this in the text (lines 217ff).

      1. Line 192 and Figure S8: n {less than or equal to} 33 (either a typographical error and should have been {greater than or equal to}, otherwise, it may be expressed as a range)

      It was indeed a typographical error that was now corrected.

      1. Line 258: Methods on the generation of FIO/FIO+ was a bit difficult to understand. Maybe a simple plasmid schematic with the restriction sites (at least for the original plasmid) in the supplementary may help clarify this.

      Cloning strategy has been expanded with additional information for clarity.

      1. Line 295: include abbreviation of cRPMI here rather than in Line 303

      Has been corrected.

      1. Line 322: typographical error on WR99210 working concentration?

      Has been corrected.

      1. Line 372: Last sentence on area and raw integrated density measurement is unclear.

      We have reformulated the sentence for more clarity.

      1. Line 461: typographical error in last sentence

      Has been corrected.

      1. Line 532: Figure 4E should be Figure 4F

      Has been corrected.

      Reviewer #3 (Significance):

      DNA replication is vital to the survival of malaria parasites. A deeper understanding on their unusual form of replication may be exploited to find drug targets uniquely directed to the parasite. Biological insights from this work can also provide a jump-off point for unravelling unusual replication in other organisms. Data on the physicochemical analysis of centrin is not just of great interest for those in the field of parasitology, but also for those in the much wider fields of biology, physics and chemistry. Techniques presented in this work (e.g., DiCre overexpression with different promoters) can definitely be utilised for the elucidation of protein function within and outside the field of parasitology.

      My field of expertise is in Plasmodium spp., particularly in parasite replication, molecular and cellular biology, and epigenetics.

      We thank the reviewer for the appreciation of our work in terms of insight and technology development.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary:

      The authors have provided a comprehensive characterisation of centrin proteins in Plasmodium falciparum. Through expression of episomal GFP-tagged centrin for in vitro, they were able to observe co-localisation of centrin with centriolar plaques during the replicative stage of the parasite. They also utilised live cell STED microscopy to track dynamic changes in centrin morphology. They have also demonstrated calcium-dependent phase separation dynamics in bacterially-expressed P. falciparum centrin and human centrin 2. The formation of liquid-liquid phase separation in PfCen1, 3 and HsCen2 tied well with IUPred3 predictions of intrinsically disordered regions in these proteins. Using an inducible DiCre overexpression system with two promoters of varying strengths, the authors have shown accumulation of centrin1 outside of centrosomes and premature appearance of centriolar plaques. Finally, changes on the centrin1 protein, i.e., N-terminal deletion, and mutations in calcium binding sites in the EFh domains, have shown a reduction in the formation of ECCAs during overexpression and inability to form LLPS in vitro, respectively.

      Major comments:

      1. Given that parasites cannot tolerate endogenous C-terminal tagging of some centrins (but not all, as PbCen4 was successfully tagged), has N-terminal tagging been attempted either by the authors or in previous publications? Note that this is not a request for further experimentation; rather, maybe this can be noted in the manuscript; and line 62 can be rephrased for transparency.
      2. Is there a possibility that by adding a C-terminal tag, centrin may lose a specific function or cause change in the physicochemical properties of the protein (thus making C-terminal tagging lethal)? Was His tag removal attempted so the native protein can be used in the LLPS experiments? IUPred3 analysis showed potential IDR at the C-terminal end of PfCen4. Could the C-terminal tag have caused the protein to not form droplets in the presence of Ca2+?
      3. It has been shown by the authors that different tagged centrins co-condense which may support the localisation data (Figure 1C). However, is there a way to show that the episomally- and endogenously-expressed centrin co-localise with each other (e.g., confocal microscopy with anti-centrin vs anti-gfp in PfCen-GFP lines, that is if the authors have access to anti-centrin antibodies)? Has endogenous centrin been demonstrated to form ECCAs (in previous publications or by the authors)?

      Minor comments:

      1. How were the times (post addition of Ca2+) presented in Figure 2A determined?
      2. Line 126: Figure 1B should be Figure 1C
      3. Line 145: Figure 1C-D should be Figure 1D-E
      4. Line 151: Figure 3A should be Figure 4A
      5. Line 152: Suggest rephrasing "placing the gene of interest in front of the promoter" to "placing the gene of interest immediately downstream of the promoter" or something similar
      6. Any growth phenotype changes observed in the overexpressors?
      7. How often are ECCAs observed in pARL strains, or are they not observed at all? This might be good to mention.
      8. Line 192 and Figure S8: n {less than or equal to} 33 (either a typographical error and should have been {greater than or equal to}, otherwise, it may be expressed as a range)
      9. Line 258: Methods on the generation of FIO/FIO+ was a bit difficult to understand. Maybe a simple plasmid schematic with the restriction sites (at least for the original plasmid) in the supplementary may help clarify this.
      10. Line 295: include abbreviation of cRPMI here rather than in Line 303
      11. Line 322: typographical error on WR99210 working concentration?
      12. Line 372: Last sentence on area and raw integrated density measurement is unclear.
      13. Line 461: typographical error in last sentence
      14. Line 532: Figure 4E should be Figure 4F

      Significance

      DNA replication is vital to the survival of malaria parasites. A deeper understanding on their unusual form of replication may be exploited to find drug targets uniquely directed to the parasite. Biological insights from this work can also provide a jump-off point for unravelling unusual replication in other organisms. Data on the physicochemical analysis of centrin is not just of great interest for those in the field of parasitology, but also for those in the much wider fields of biology, physics and chemistry. Techniques presented in this work (e.g., DiCre overexpression with different promoters) can definitely be utilised for the elucidation of protein function within and outside the field of parasitology.

      My field of expertise is in Plasmodium spp., particularly in parasite replication, molecular and cellular biology, and epigenetics.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      The authors analyzed the properties of the four Centrin proteins of the malaria parasite using a combination of in vitro and in vivo approaches. Their findings indicate that two of the four Plasmodium Centrin proteins, PfCen1 and PfCen3, as well as the human Centrin protein HsCen2, exhibit features of biomolecular condensates. Moreover, analysis of cells overexpressing PfCen1 indicates that such biomolecular condensates become more numerous as cells approach mitosis and are dissolved thereafter.

      Major comments

      • A) A critical point that requires clarification is how the protein concentrations used in the in vitro and in vivo assays (20-200 microM in vitro, and not estimated in vivo) compare to that of the endogenous components. This is important because it may well be that 6His-tagged PfCen1, PfCen3 and HsCen2 can form biomolecular condensates when present in vast excess, but not when present in physiological concentrations. The authors should report the estimated cellular concentration of PfCen1-4, as well as that achieved upon PfCen1-GFP overexpression (on top of endogenous PfCen1), for instance using semi-quantitative immunoblotting analysis. Given this limitation, the authors may also want to temper their title by introducing the word "can" after "centrins".
      • B) Movies S1 and S2 (and the related Fig. 1D and 1E) are not the most convincing to support the notion that the observed assemblies are biomolecular condensates, as not much activity is going on during the recordings. Likewise, Movies S3, and even more so Movie S4, as out of focus for a large fraction of the time, making it difficult to assess what happens at the beginning of the process. Moreover, it appears that fusion events, while occurring, are rather rare. The movies should be exchanged for ones that are in focus, and ideally a rough quantification of fusion events as a function of biomolecular condensate size provided.
      • C) An important control is missing from Fig. 2, namely assaying PfCen1-4 without the 6His tag, to ensure that the tag does not contribute to the observed behavior (although it can of course not be sufficient as evidenced by the lack of biomolecular condensates for PfCen2 and PfCen4).
      • D) The authors should test whether the assemblies formed by PfCen1 and PfCen3 are sensitive to 1,6-hexanediol treatment, as expected for biomolecular condensates.
      • E) The fact that HsCen2 also forms biomolecular condensates is very intriguing, but further investigation would be needed to assess the generality of these findings. For instance, the authors could test in vitro also S. cerevisiae Cdc31, the founding member of the Centrin family of proteins to further enhance the impact of their study.

      Minor comments

      1. For the experiments reported in Fig. 3D, the same concentrations as those used in Fig. 3A-C (namely 10 microM, and not 30 microM as in Fig. 3D) should be used. Moreover, it would be informative to test whether PfCen2 and PfCen4 as PfCen3 when added to PfCen1.
      2. The authors mention that the effect of Calcium in inducing biomolecular condensates is specific, as Magnesium was not effective (lines 94-95). However, an examination of Fig. S3B indicates that the Magnesium also exhibits some activity, albeit less potent than Calcium. The authors should discuss this point and rectify the wording in the main text.
      3. Do the authors think that PfCen2 and PfCent4 localize to the centriole plaque in vivo using another mechanism that deployed by PfCen1 and PfCent3? It would be good to discuss this point.
      4. Given that the EFh-dead mutant exhibits no activity in vitro and fails to localize in vivo, one potential concern is that the protein is misfolded. The authors should conduct a CD spectrum to investigate this.
      5. It is not entirely clear from the main text in lines 103-104, as well as from the legend, what Fig. S3B shows. When was EDTA added in this case?
      6. Fig. S7: the correlation between PfCen1-GFP expression levels and ECCA appearance is modest at best. What statistical test was applied? This should be spelled out. Moreover, the authors should combine the two data sets, as this will provide further statistical power to assess whether a correlation is truly present.
      7. The authors may want to discuss how their findings can be reconciled with the notion that Centrin assemble into a helical polymer on the inside of the centriole (doi: 10.1126/sciadv.aaz4137).
      8. Likewise, the authors may want to speculate regarding what their findings signify for the role of Centrin proteins in detection of nucleotide excision repair (doi: 10.1083/jcb.201012093).

      Small things

      • Fig. 1A: change color for microtubules as red on red is difficult to discernn.
      • Fig. 1C: the indicated boxes in the top row do not seem to correspond exactly to the insets shown in the bottom row.
      • line 266: typo, promotor > promoter.
      • line 360: a reference should be provided for the GFP-booster, including the concentration at which it was used.
      • line 363: "an" missing before "HC".
      • line 428: it would be best to deposit the macros on Github or an analogous repository.
      • line 461: "to the" is duplicated.
      • Fig. S5A: maybe draw the lines in red (as red in Fig. S5B correspond to the proteins that do not have IDRs).
      • Movie S7, legend: left frames shows PfCen1-GFP, not microtubules as currently stated.

      Significance

      This is a provocative study that extends initial observations regarding self-assembly properties of Centrin proteins, and posits that some members of this evolutionarily conserved family can form biomolecular condensates. After the above outstanding issues have been properly addressed, these data could have important implications for understanding Centrin function in centriole biology and DNA repair. Therefore, these findings will be of interest to a cell biology audience.

      Field of expertise: cell biology.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Voss, Reinert and colleagues show calcium-dependent assembly of Plasmodium falciparum centrins in vitro and in parasites. This assembly is dependent on the EF-hands of centrin and an N-terminal disordered region.

      Major concerns:

      1. The very definitive title is not wholly supported by the data. This should be qualified by specifying the conditions under which the centrins can accumulate in this way.
      2. A major concern is whether this behaviour of centrins represents a biologically relevant mechanism in centriolar plaque formation. Is this limited to high overexpression conditions or in vitro high concentrations? Or is it a result of the tagging of the P. falciparum centrins? A convincing approach to addressing this issue would be to knock-in a fluorescent tag to the centrin loci. Roques et al. (ref. 12 in this submission) report the GFP tagging of centrin-4 in P. berghei, although they note that centrins-1 to -3 were refractory to tagging in this organism. It is unclear whether Voss et al. attempted this tagging in P. falciparum. This should be clarified and relevant data presented.

      If the tagged molecules used in the biochemical parts of this study are functional, It is challenging to understand why the centrins cannot be tagged in P. falciparum. If the tags render the P. falciparum centrins dysfunctional, the study becomes significantly less useful.<br /> 3. If a knock-in cannot be achieved, it must be shown that the transgenic expression of tagged Plasmodium centrins does not confound the analysis of centrin behaviour. It is known that these proteins can behave anomalously when overexpressed (Yang et al. 2010, PMID: 20980622; Prosser et al. 2009, PMID: 19139275), at least in other species.

      A previous description of centriolar plaque from the authors' lab (Simon et al. 2021, PMID: 34535568) shows an organized structure of an established size. It should be demonstrated whether the structures formed with the GFP tagged centrins show the same dimensions and dynamics as those in wild-type parasites. The extent of the overexpression of the GFP-tagged centrins should also be demonstrated.<br /> 4. It would also be useful to remove the His tag from the recombinantly expressed and purified centrins for the in vitro analyses, particularly if concern remains about the impact of tags on Plasmodium centrin behaviour.<br /> 5. The discussion is very short and does not consider the findings presented here in the context of the literature, with respect to centrins, Plasmodium MTOC assembly mechanisms, or to general considerations around biological condensates. Andrea Musacchio's recent commentary (ref. 44 in the current submission) advocates caution in ascribing phase separation as an assembly mechanism for organelles in vivo, particularly on the basis of in vitro experiments with high concentrations of homogeneous protein. It is not clear that the concentration dependence of extracentrosomal centrin accumulations (ECCAs) at the onset of schizogony provides sufficient justification of a phase separation model in vivo. The authors' recent description of the involvement of an SFI1-like protein, SIp (Wenz et al. 2023 PMID: 37130129), in the centriolar plaque makes a case for non-homotypic interactions also driving assembly and alternative models for ECCA are not convincingly excluded. The absence of a robust discussion of such considerations is unhelpful to the reader.<br /> 6. It is also unclear whether the analysis of human centrin is suggested to indicate a phase separation mechanism for centrins in human cells. As this is readily testable, this notion could be considered further. Although its experimental examination may lie outside the theme of this study, one would expect some discussion of the significance of the data presented in the study.

      Minor points

      1. There are only three centrins in humans. Centrin 4 is a pseudogene (Gene ID: 729338 on NCBI).
      2. Line 175 should say 'temporally', rather than 'temporarily. The Abstract should say 'evolutionarily conserved', rather than 'evolutionary conserved'. 'To condensate' is not ideal as a phrase- 'to form a condensate' would be clearer.

      Referees cross-commenting

      I think the other 2 reviewers have made fair, cogent and constructive points. There is good convergence between the reviewers on the significant issues around the study. These concern in vivo and in vitro effects of tagging and of of high concentrations.

      Significance

      The biology of the Plasmodium centriolar plaque is of great interest as an alternative MTOC structure, with obvious additional interest deriving from the role of this organism in malaria. Much remains to be learned about this structure, so the topic of this paper is likely to attract a broad readership. Furthermore, the centrins are a widely-expressed and evolutionarily conserved family of eukaryotic proteins, with multiple roles; a new model for their behaviour, such as is suggested here, would be of interest to many cell biologists.

      With that in mind, significant additional data should be provided to substantiate the model proposed by the authors.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      The authors do not wish to provide a response at this time.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary:

      The manuscript by Schauer et al. uses embryonic explants to study the coordination of Nodal and BMP signaling for embryo morphogenesis. They show that Nodal signaling triggers explant elongation by inducing mesendodermal progenitors that undergo cell intercalation. Looking at the role of BMP signaling, they show that BMP overactivation ventralizes the explants, reducing cell intercalation and therefore explant elongation. Looking at pSmad5, they then establish that BMP signaling in the explant is attenuated by Nodal signaling, through activation of chordin expression, and through some unidentified chordin-independent mechanisms. Moving to the entire embryo, using combinations of BMP overexpression and Nodal inhibition, authors show that Nodal signaling limits BMP signaling on the dorsal side of the embryo, which is key to proper embryo elongation.

      Major comments:

      • The authors used the sebox::EGFP line to show that the growing region of the explant consists mostly of mesendodermal cells. Although this transgenic line had not been used to do so, the authors and others, had previously demonstrated that the extending part of the explant is mostly made of mesoderm and even shows some patterning (1,2). This should be stated and not presented as a new finding.
      • Explant elongation is driven by cell intercalation. The authors analyzed the shape of the mesendodermal tissue to conclude that cells intercalate. While I do not question this conclusion, as it is well known in the embryo, direct observation of cell intercalation, as was done in the embryo (3), would be a better demonstration.
      • Explant elongation is driven by mesendodermal cell intercalation. I certainly agree from the movies and images that the extending region is mostly made of mesendoderm. However, it seemed clear to me that in Movie 1, starting at about 140 minutes, most of the convergence movement is taking place in a non-green region of the explant, fueling the extension of the mesendodermal region. Also, to demonstrate that cell intercalation is occurring in the mesendoderm, the authors performed clone dispersal analysis, comparing clones of mesendodermal and ectodermal cells. However, the selected ectodermal clone is very far from the extending region. To show that the cell intercalation is specific to mesendoderm, I think the authors should try to compare the behavior of mesendodermal and non-mesendodermal cells that are located at the same distance from the extending region. For example, from the image in Figure 1E (235 mpe), it appears that the right side of the base of the extended region is not green and could be compared to the left side. Currently, the quantification shown in 1G mostly demonstrates that the extending region is extending, and that the non-extending region is not.
      • Based on their observations in explants, the authors propose that Nodal signaling maintains an area of low BMP signaling on the dorsal side of the gastrula for robust axis elongation. While I acknowledge that the experiments performed by the authors have not been previoulsy reported, I did not understand how this differs from the very well established fact that Nodal inhibits BMP signaling, in particular through chordin expression. Von der Hardt for instance already reported that overexpression of bmp and inhibition of chordin leads to severe elongation defects (4). More insight could probably be gained by analyzing the effect in more detail: Is the elongation defect due to cell intercalation defects? How are cell fates affected? Is this Nodal effect mediated by Chordin?...

      Minor comments:

      • Fig6B. Are the curves significantly different? If so, how were they compared?
      • Fig6D-E, I found the quantification a bit confusing. The reader is left with the impression of an all-or-nothing answer (effect only with BMP overexpression and strong Nodal inhibition), whereas the effect on the pSmad5 gradient is gradual. Plotting and comparing the pSmad5 intensity gradients would be better.
      • Fig6G. 'Axis length/embryo height' should appear on the x-axis, not the y-axis.

      Referees cross-commenting

      I feel that the three reviews are very much in agreement, recognising that the experiments carried out are well done and calling for a reasonable amount of additional data. The three reviews also agree that the results obtained here in explants were already known from intact embryos, limiting the relevance to ex vivo research.

      Significance

      Overall, the experiments appear carefully carried out, and very precisely quantified. The paper is well written and easy to read. The results add to our understanding of the morphogenetic events occurring in embryonic explants. I therefore support their publication.

      My main concern is with the significance of the results. I am convinced that embryonic explants are great tools, to reduce the complexity of the embryo and to address questions that cannot be addressed in the embryo, as the authors and others have done, for instance, to address the role of extraembryonic tissues and patterning by maternal contributions. Here, however, I felt that most, if not all, of the experiments essentially demonstrate in embryonic explants, results that have been known for years in the intact embryo. While gathering detailed information on what happens in embryonic explants will certainly prove useful in further understanding the self-organizing abilities of these explants, and is worth publishing, the significance of the results reported here seems limited. Specifically, that elongation is driven by cell intercalation, that BMP-mediated dorsoventral patterning affects cell intercalation, that BMP signaling is attenuated by Nodal through Chordin, that Chordin is required for elongation, has been well established in the embryo over the last 20 years. Again, showing that it works the same way in embryonic explants is of interest, but at this point, does not add to our understanding of embryonic development.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      This paper from the Heisenberg lab takes a reductionist approach to understanding how BMP and Nodal signaling interact to coordinate morphogenesis. They mostly use blastoderm explants that they culture in vitro. These explants elongate over time, with Nodal signaling that induces mesendoderm driving the cell intercalations that explain the elongation. They show that increased BMP signaling inhibits this process, but reducing BMP signaling has no effect. They see that reducing Nodal signaling results in an upregulation of BMP activity as read out by phosphorylated Smad5 staining and increasing Nodal signaling has the opposite effect. They explain this mostly by the observation that Nodal induces the expression of the BMP antagonist, Chordin, and validate this idea by demonstrating that a reduction in Chordin expression reduces explant elongation. Returnign to the embryo, the authors show that manipulation of Nodal signaling levels influences the size of the BMP activity gradient as expected from the in vitro results. Finally, they show that reduction of Nodal signaling with SB505124 sensitises the embryos the effects of bmp2b overexpression, and that BMP overeactivation at 90% epiboly reduced C&E movements.

      Major comments

      In general I think the work is well done and the data justify the conclusions.

      I have several suggestions for additional experiments and discussion that I think would improve the paper.

      1. In Figure S1 they present data on elongation of explants treated with a Nodal inhibitor. It would be good to show some examples of images of the explants.
      2. In Figure 1G and 3A, the same wildtype images are shown. This is mentioned and I assume therefore that the results were all part of the same experiment. How many times were these experiments performed? It would be much better to use different biological replicates in the two figures.
      3. It is important for the authors to make clear how many biological replicates each of the experiments correspond to.
      4. In Figure 4E, it would be good to show the levels of P-Smad2 in the Oep and MZ lefty1, 2 explants.
      5. On page 11 the authors mention chordin-independent inhibition of BMP signaling. The most likely candidate would be noggin as it too is expressed dorsally and is at least in part activated by Nodal. This should be tested in their model.
      6. The authors focus on Chordin as downstream of Nodal signaling, and discuss the role of Nodal signaling in inducing chordin as being due to peak Nodal signaling. However, Chordin has been shown to also be downstream of Fgf signaling and Bozozok (PMIDs 23499658 and 16873584), which likely explains its dorsal expression domain. Furthermore, Rogers et al, (PMID 33174840) who the authors refer to, also show that to disrupt BMP signaling in embryos, inhibition of Nodal and Fgf is required. These issues need to be discussed in more detail. It is the combinatorial signaling that is thought to be responsible for the dorsal location of the chordin (and noggin) expression domains.

      Minor comments

      I think in general the manuscript is well written and the figures are clear. Previous data is generally well cited. My only comment is that there is a wealth of data from Xenopus and zebrafish that BMP antagonists are induced as a result of combinatorial Nodal signaling and other pathways (dorsal wnt and fgf) that inhibit BMP signaling. I think this could be better referenced.

      Significance

      The paper is well done and provides important information about the interactions between Nodal and BMP signaling to induce axis elongation. I think the work would be improved if the authors revise it along the lines suggested above. In terms of novelty, many of the component parts of the paper are known (Nodal signaling is important for elongation via cell intercalation and Nodal and BMP can antagonize on another by the induction of BMP antagonists by Nodal), but it is novel to put them together to investigate axis extension using explants. The paper will be of interest to those interested in how these signaling pathways operate in early vertebrate development and to those interested in morphogenesis.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary

      The authors presented intriguing observations on the molecular mechanisms regulating morphogenic cell movement, with a particular focus on convergent-extension (CE) movement associated with cell type specification in the zebrafish blastoderm explant. In this manuscript, Schauer et al. identified the CE movement of the mesendoderm as triggering the elongation of the zebrafish embryonic explant. In this process, the Nodal signal represses the BMP signal, which negatively regulates the movement of the mesendoderm precursors, through the induction of its inhibitor chordin. This suggests that the Nodal signal is the key factor coordinating cell fate specification and morphogenesis in the zebrafish blastoderm explant. Finally, suppression of Nodal signalling increases sensitivity to BMP signalling in the CE movement of intact embryos. This suggests that promotion of mesendoderm cell intercalation via BMP suppression by Nodal may be involved in conferring robustness to morphogenic cell movement in vivo.

      Major comments

      1. While one of the main conclusions of this manuscript is that "Nodal signaling regulates CE movement of mesendodermal cells by promoting their intercalation through inhibition of BMP signaling". However, this was predicted by changes in individual cell morphology and cell dispersal, and the authors didn't directly examine the behavior of individual cells. It would be better to confirm intercalation during the process of explant elongation by cell tracking analysis.
      2. Although the authors discuss that Nodal signaling inhibits BMP signaling in the later gastrulation stage, this has not been experimentally tested. If possible, the time window in which Nodal signaling acts should be investigated by temporal inhibition of Nodal signaling using chemical inhibitors.
      3. Only the signal gradient of pSmad5 and axis elongation were examined in the intact embryo part of the study (Fig. 6 and Fig. S7). The information on the domain of pSmad2 and the expression of chordin would be helpful for the comparison of the blastoderm explant and the intact embryos.

      Minor concerns

      The first letter of a gene name should be in lowercase. ( ex. Fig.S3C; Smad5 MO)

      Significance

      The zebrafish blastoderm explant assay has the potential to elucidate the molecular mechanisms regulating the complex processes of morphogenesis during vertebrate gastrulation, as the authors demonstrate in this paper. In this manuscript, the authors addressed the molecular mechanism coordinating cell fate specification and morphogenic cell movement in the blastoderm explant. All of the experiments are well-designed, the interpretation of the results is convincing and the paper is well-written. Also, the conclusion is very clear and well supported by the presented data. These findings provide fundamental and important insights for studying morphogenic cell movements in early vertebrate embryos using zebrafish blastoderm explants. On the other hand, most of the molecular mechanisms reported in this manuscript are already predicted by previous studies using intact embryos. Therefore, the impact of this work may be limited to ex vivo research.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity):

      Summary: This manuscript describes molecular mechanisms by which ACBD3 is recruited to the Golgi complex. ACBD3 recruits PI4KIIIb which is required to generate PI4P, a phosphoinositide which is key for the recruitment of essential Golgi proteins and hence is key to Golgi identity. The authors have used a combination of mass spectrometry, high quality fluorescence imaging, transient CRISPR knockdowns, and biochemical approaches such as IPs to identify the key determinant for recruitment of ACBD3 to the Golgi complex. They map the interaction between ACBD3 and the Golgi as a unique region (UR) upstream of its GOLD domain, identifying, in particular, an MWT motif as key for this recruitment. Using mass spectrometry they identify several novel interactors of ACBD3 as well as some established binding partners. Knockdown of these interactors reveal a key role for the SNARE, SCFD1, where reduced levels lead to complete loss of ACBD3 localisation to the Golgi without apparent disruption of Golgi structure. They further validate this interaction and that of another SNARE (Sec22b), which is part of the same SNARE complex as SCFD1, mapping the interaction to the longin domain of Sec22b. Surprisingly however they demonstrate that the UR domain does not mediate the interaction between ACBD3 and these SNAREs suggesting an alternative mechanism of recruitment. Previously identified ACBD3 interactors, Golgi proteins giantin and golgin-45 were also identified in the mass spectrometry screen and the authors demonstrate that these two proteins can recruit ACBD2 to the Golgi and this is dependent on the MWT motif identified in the UR domain. By knocking down SCFD1, they show reduced recruitment of ACBD3 leading them to propose a model of sequential recruitment of ACBD3 by SCFD1 followed by interactions with the golgins.

      Major points: This study is a well-executed and rigorous study of the molecular requirements for the recruitment of ACBD3 to the Golgi. The experimental approaches are state-of-the-art and the data are clean and convincing. The only caveat, raised by the authors themselves, is their interpretation that there are two sequential steps for Golgi recruitment of ACBD3. While they show that loss of SCFD1 reduces the interaction of ACBD3 with giantin and golgin 45, their model depends on doing the reverse experiment, i.e. assessing the effects of knocking down either giantin or golgin-45. This is especially relevant given the demonstration that golgin-45 is sufficient to recruit ACBD3 to mitochondria. It may well be that recruitment involves a tripartite complex, which is not uncommon in vesicular transport mechanisms Giantin is not an essential protein do it should be feasible to perform this experiment. The authors are equipped in the quantitative fluorescence microscopy which would be required and which would help resolve whether sequential or redundant mechanisms are required for ACBD3 recruitment.

      We thank the reviewer for the positive comments and are glad that they consider our study "well-executed and rigorous". We totally agree with the reviewer that our conclusions regarding the sequential aspect of the recruitment of ACBD3 in the original submission could be better supported. We have worked to strengthen this in our resubmission. As the reviewer states, this limitation was already discussed in the original submission. To further support our model, we have performed the experiment suggested by the reviewer, in which we test the effects of knocking down both giantin and golgin45 (double knockdown) on the binding of ACBD3 to SCFD1.

      The results of this experiment further support our sequential model with little to no effect of loss of the Golgins on ACBD3. As we already knew, a large effect of SCFD1 KO on the binding of the Golgins to ACBD3 was also observed here. We should note that this was performed in a different cell line than before (HeLa cells rather than HEK cells), as the efficiency of multiple knockdowns was much lower in HEK cells, as determined by qPCR. Taken together, the new data in Figure 7 supports a sequential model for Golgi recruitment. We also agree that other, less likely models could explain our data and have included this openly in the discussion. In conclusion, we thank the reviewer for their comments and have revised the manuscript with a new experiment with the relevant repeats, which supports our model.

      Reviewer #1 (Significance):

      Significance PI4P is a phosphoinositide that is important for the recruitment of Golgi proteins. As with most PIs it is likely to act by coincidence detection in that Golgi associated proteins will recognise PI4P as well as other factors on Golgi membranes. This results in different local membrane environments which will be specific for particular functions. PI4KIII__b_ is key for PI4P production although the absolute levels of PI4P are likely to be determined by a balance of lipid kinases and phosphatases. However, since ACBD3 is key for the recruitment of PI4KIII__b, it is important to understand the molecular mechanisms by which it is recruited. The manuscript thus makes a significant contribution to understanding one of the underlying mechanisms for PI4KIII__b _recruitment although, as indicated above, stops short of establishing a clear model for the roles SCDF1 and Sec22b versus golgin 45 and giantin. For the future it will be of interest to determine why either a sequential or a redundant mechanism is required for the recruitment of ACBD3 as a scaffold protein.

      We thank the reviewer for this set of positive comments on the manuscript and for agreeing that this is a significant contribution. Our revised version further supports our sequential model of ACBD3 recruitment to the Golgi apparatus, and the comments here have helped us further to strengthen the quality and clarity of the manuscript.

      Reviewer #2 (Evidence, reproducibility and clarity):

      Summary This is a very interesting and potentially important paper for the field of membrane biology and membrane trafficking, in which the authors have studied the molecular mechanisms by which ACBD3 (and consequently PI4KIIIb) is recruited to the cis-Golgi membranes. The authors suggest that this recruitment is based on a two-step process, mediated by interactions to, on the one hand, SCFD-1 (SLY1) and, on the other hand, two redundant golgins (golgin-45 and giantin).

      We once again thank the reviewer for the positive comments and are glad that they consider our manuscript important.

      Comments:- Pg.1 : arfaptins, as far as I know, have not been shown to be involved in intra-golgi trafficking but rather in Golgi export (see e.g. ref. 12)

      We thank the reviewer for pointing this out. We have corrected the text accordingly.

      • Pg. 1: reigon --> region

      We thank the reviewer for noticing this typo. We have corrected the text accordingly.

      • Arf1 also recruits PI4KIIIb right?

      This is correct. The De Matteis lab has shown that PI4KIIIβ associates with the Golgi complex in an Arf1-dependent manner (Godi et al. 1999). We think this is excellent work. However, Arf1 is somewhat of a master regulator of the Golgi, affecting the recruitment and localisation of many different Golgi proteins. It has also previously been reported that Arf1 does not directly interact with PI4KIIIβ (Klima et al. 2016). Overall, the molecular relationship between Arf1 and the kinase remains unclear. We do not exclude, however, that there are factors other than ACBD3 important for recruiting and regulating PI4KIIIβ levels at the Golgi. We have changed the wording in the manuscript to reflect that there are multiple ways that PI4KIIIβ is recruited to the Golgi apparatus.

      Fig. S1: the information about the number of cells per experiment is missing. Also, please add the information about what exactly is represented in the box plots (is it the distribution of the mean value of R per experiment? or the total distribution on a cell-by-cell basis of a representative experiment?)

      For each experiment, a minimum of 100 cells per condition were imaged. The Pearson's correlation was then calculated, and the average was taken for each biological repeat. The plot in Fig. S1B represents 3 independent biological repeats. We have included this information in the revised manuscript.

      • The definition of Avg. Golgi int/avg. cell int. (a.u.) in Fig 1E,F is a bit difficult to understand to me. If I understand correctly, the total fl. int in the Golgi mask was computed and divided by the area of the Golgi mask (this is the av. Golgi intensity). A similar computation is done for the entire cell (including the Golgi), i.e., total fl. intensity in the cell mask is computed and divided by the area of the cell mask. Then the two av. intensities are divided (ratio = av. Golgi int / av. cell int.). This ratio, for a protein that is enriched in the Golgi area, should be larger than 1. For a protein that is equally distributed all over the cell, it should be 1, and for a protein that is excluded from the Golgi area, smaller than 1. Then to this value, the authors subtract the value of the ratio found for an inert construct (GFP of Halo alone), which I imagine should have an original ratio value of the order of 1, and hence, after this subtraction, norm. ratio values larger than 0 mean that they are more enriched at the Golgi area than GFP/HaloTag themselves. Is this correct? In principle, I don't see anything entirely wrong with this way of thought, but I just found it a bit difficult to understand, and in general one has to be careful when computing rations (quotients) and then subtract another ratio. Also, the units are not a.u., the value is dimensionless, what is "arbitrary" is the definition of 0 value and the based on this definition, also the actual value. I think it would probably be much clearer for the readers to compute somthing like the relative enrichment in the Golgi area as compared to the rest of the cell (excluding the Golgi area). That is, a value r'=(Int. Golgi mask / Area Golgi mask) / [(Int. Cell mask - Int. Golgi mask)/(Area cell mask - Area Golgi mask)]. This can be computed directly or defining a mask that is the cell mask - the Golgi mask. Also, some maths (unless I made a mistake) give that this r'= r (1-aG)/(1-r aG); where r is the ratio (before subtraction) defined by the authors, and aG=Area cell mask/Area Golgi mask. In any case, I'd suggest the authors to either adopt this other quantitation (without subtraction of the GFP/HAloTAG), which gives directly the fold-enrichment in the intensity density in the Golgi area with respect to the rest of the cell; or explain in more detail the maths of the value they are plotting now.

      We thank the reviewer for these well-reasoned and thoughtful suggestions for our imaging analysis. These are issues that we have also considered when quantifying this dataset. At the heart of it, the second method of calculation (Golgi/outside of Golgi), results in a non-linear distribution, as the pool of proteins re-distribute from inside the Golgi to the cytosol. This is why we have chosen to use the first method of Golgi/total, as it provides a linear distribution.

      The reviewer is also correct that the GFP (inert protein) ratio is 1 without adjustment. We have chosen to normalise to GFP/HaloTag (inert protein) as we think this is the clearest way of conveying our conclusions from these experiments. We have included the non-normalised graph here for the reviewer to see; however we thought that this conveys the key result less clearly. Overall, we agree this was poorly communicated in the manuscript and we have clarified it in the revised version.

      • Fig. 1C&F: Besides the MWT mutant, the FKE mutant also seems to have a somewhat compromised Golgi localization. Have the authors followed on that, or what is the reason that they have just focused on the MWT mutant?

      In contrast to the MWT mutant, the FKE mutant does not affect ACBD3 localisation significantly. In addition, when having a close look at the pdb structure of the GOLD domain of ACBD3 with 3A protein of Aichivirus A (5LZ3), the MWT patch, in particular residues M and T, make clear contact with protein 3A, which is not the case for FKE residues. Therefore we focused on the MWT residues, which we hypothesised to interact with a Golgi resident protein which competes with protein 3A to interact with ACBD3.

      • Very minor point, and without wanting to sound pedant at all, but I think (I might be wrong of course, so apologies if I am) that the plural of apparatus in latin is not apparati, but apparatus (fourth declination). So, I'd change the word in page 2 (or just rephrase the sentence: e.g. "resulting in Golgi fragmentation"). But of course, I'd leave this to the authors' discretion.

      We thank the reviewer for this precision, do not consider it pedantic, and have made the suggested change to the text.

      • Fig. 3A: have the authors tried or been able to perform IF of the endogenous SCFD1 protein?

      As suggested by the reviewer, we attempted to perform IF of endogenous SCFD1, as shown below. Despite trying several different antibodies, we were not satisfied that we were detecting real SCFD1 signal as there was no change in this staining upon SCFD1 CRISPR KO. Please see an example of this IF below (ProteinTech, 12569-1-AP). We have contacted the antibody manufacturers to inform them of this issue.

      • Similarly to what has been done for other panels, could you quantify Fig. 3C? Are PI4KIIIb protein levels affected upon the different KOs?

      As suggested by the reviewer, we are now showing in Figure S2D the percentage of cells with a partial or total loss of PI4KIIIβ at the Golgi in CRISPR-Cas9 KO cells of either PI4KIIIβ, ACBD3 or SCFD1. 3 independent biological repeats were performed and approximately 150 cells were quantified (~50 cells per condition). The results show that the PI4KIIIβ antibody used (BD Bioscience, 611816) is specific (93.22% of cells lose the antibody signal) and that ACBD3 and SCFD1 KO affects PI4KIIIβ recruitment to the Golgi in 88% and 73% of the cells, respectively._-

      The last paragraph of the "SCFD1 and ACBD3 interact upstream of PI4KIIIβ recruitment to the Golgi apparatus" section reads a bit odd placed there. I think it is more appropriate for the discussion or for the intro part on SCFD1.

      Many thanks to the reviewer for pointing this out. We simplified that paragraph to describe the relationship between SCFD1 and SEC22B.

      • I am confused on Fig. 5A/B. The labels in the blots show that 390-528 (without UR) does not bind sec22 or scfd1, but the 368-529 does? Or I guess, judging by the MW seen in the middle blots, that there's some error in the labelling?

      Many thanks to the reviewer for noticing this, which was clearly a labelling error. We corrected this accordingly in Figures 5A and B. We apologise for this oversight.

      also, the IP efficiency of the MWT mutant in the panel A blot is quite low, still sec22 seems to be very efficiently pulled down. Can the authors comment on that please? Would co-IPing against endogenous sec22 and scfd1 would work (so you don't need to rely on HaloTag+ligand?)

      We know that the MWT residues of ACBD3 are important for recruiting ACBD3 to the Golgi (Figure 1C and F). We also know that ACBD3 interacts with SEC22B and SCFD1 (Figure 3B and 4A) and that SCFD1 is important for ACBD3 Golgi recruitment. Therefore we initially speculated that ACBD3 interacts with SEC22B and SCFD1 through the MWT residues. However, as the reviewer points out, Figure 5 shows the opposite. Mutating MWT residues makes the interaction of ACBD3 with SEC22B and SCFD1 stronger. For this reason, we hypothesised that another player(s) also contributes to ACBD3 recruitment through interactions with the MWT residues. We have shown that the second recruitment factors are the 2 golgins, golgin-45 and giantin (Figure 6C). In short, whilst we agree that the IP efficiency is low, the binding is actually stronger, supporting our conclusions. No interaction of ACBD3 with endogenous SEC22B could be detected due to a lack of a sufficiently sensitive antibody (we tried Abcam ab181076 and ProteinTech 14776-1 AP).

      • I really like the experiment 6B. Have the authors tested whether SEC22 is also recruited to mitochondria in those conditions? But not SCFD1?

      We thank the reviewer for the positive comment. We have performed the suggested experiment and are now including this as an additional figure (Figure S3). Ectopic expression of golgin-45 targeted to the mitochondria is not sufficient to redistribute SCFD1-HaloTag or HaloTag-SEC22B to the mitochondria (Figure S3A and B, respectively). We, therefore, speculate that the fraction of ACBD3 that gets redirected in Figure 6B must be the small fraction of ACBD3 that is spontaneously in an open conformation and compatible for interaction with golgin-45.

      • The results shown in Fig 7 might show a partial depletion in the interactions, but to be fully trusted they would need to be quantified and a statistical test used to compare the values. I think this part is important to show very clearly, because even with low binding to golgins (remember, single knockouts do not prevent Golgi localization of ACBD3), one could expect that ACBD3 still localized to the Golgi but it does not in the absence of SCFD1 as shown in this paper. A prediction of the proposed model is that in cells depleted of the two Golgins, SCFD1 and ACBD3 should still bind to one another, right? Did the authors test this?

      We fully agree with the reviewer. As discussed in the replies to reviewer 1, we have repeated this experiment, including both sets of KO. This was not trivial, as a double transient KO is technically challenging and involves validation with qPCR and switching cell types (HEK cells to HeLa). The new data supports our current model and suggests some additional regulatory mechanisms at play.

      • The model presented here (fig 8) seems to suggest that only the conformational variation of ACBD3 that binds Golgins is able to recruit (bind) PI4KIIIb. Is this known, or is there any experimental evidence for that?

      HDX-MS experiments show that the ACBD and GOLD domains undergo conformational changes in the presence of 3A proteins (McPhail et al. 2017). Demonstrating this would require a complicated reconstitution experiment which is technically very challenging and would involve purifying various complex proteins, including SNAREs, SM proteins and golgins. This could perhaps be the subject of several future studies.

      • Have the authors thought about testing the FKE mutant in the experiemnts shown in Fig. 5?

      As mentioned above, since the FKE residues are not making any contact with the protein 3A and since the loss of ACBD3 recruitment to the Golgi is not statistically significant (Figure 1F), we haven't tested the FKE mutant for the binding to SEC22B and SCFD1. We do, however, agree with the reviewer that there might be something interesting happening here. We would like to experimentally interrogate this in future studies and develop more sensitive assays to test if there is a significant effect with the FKE mutant.

      In general, I think the title might be a bit misleading because of the use of PI4Kiiib. I understand what the authors mean, but because they have not thoroughly tested PI4Kiiib recruitment in their experiments, I think they should focuse rather on the mechanism of recruitment of ACBD3 the authors have found.

      We thank the reviewer for their advice regarding the manuscript title, and this is something that we have discussed internally. We chose that title as it highlights the key mechanistic impact of our findings and note that we did include a figure on the recruitment of PI4KIIIβ. However, we remain open to discussing this with advice from the journal editorial team.

      Reviewer #2 (Significance):

      I think, as said above, that this is potentially an important paper for the field of membrane trafficking and membrane biology. Most of the experiments are in general well performed and well controlled, and the paper is clearly written and follows a logical line.

      We once again thank the reviewer for their comments and overall thoughtful and considered review. We believe that the suggestions here have improved the manuscript.

      Reviewer #3 (Evidence, reproducibility and clarity):

      Stalder and colleagues report experiments designed to identify interactors of the Golgi-localized protein ACBD3 (a.k.a. GCP60), and to delineate mechanisms that allow ACBD3 to localize at Golgi compartments. ACBD3 is a 528aa protein with diverse previously reported interactions and functions, both in normal physiology and as a host factor in viral assembly processes. Stalder et al. first map which domains of ACBD3 are required for Golgi localization in HeLa cells, concluding that residues 368-528 are sufficient for localization. This region includes a GOLD (GOLgi Dynamics) domain previously reported to interact with Golgin tethering proteins. Alanine scanning identifies the motif MWT just upstream of the GOLD motif as necessary for Golgi localization. Acute CRISPR knockout identifies two Golgins, Golgin45 and Giantin, as necessary for ACBD3 Golgi localization, and IP indicates that the MWT motif breaks this interaction. These data are a bit scattered around the paper but taken together are reasonably persuasive, particularly when viewed in context with published work. This reader would have found the manuscript easier to follow had the Golgin and MWT motif data been presented en bloc.

      We thank the reviewer for these comments and have considered presenting and rewriting the data as the reviewer suggested. On reflection, we have decided to present it in the original order. We feel that this allows us to highlight the two independent mechanisms individually, bringing them together at the end. In addition, as the experiments were performed in the order presented, it allows for more appropriate controls for each experiment rather than trying to combine them. We hope the reviewer accepts our preferred order.

      In a second set of experiments, IP-mass spec is used to identify ACBD3 interactors that might assist in the protein's localization. The MS data presented are filtered to exclude proteins not already identified as Golgi-localized. This is, I think, a mistake. Even if the authors choose to focus on known Golgi interactors as candidates for a localization function, the biological functions of ACBD3 are far from fully understood, and the full dataset would be of value to both cell biologists and virologists.

      We agree with the reviewer that there are many interesting mysteries surrounding ACBD3 and have therefore included an additional table (table S1) in the revised manuscript, showing the dataset of newly identified ACBD3 interactors before applying the Golgi localisation filter.

      Hits in the filtered dataset include the R-SNARE Sec22B, and the SNARE chaperone Sly1/SCFD1. Acute CRISPR inactivation of Sec22 decreases ACBD3 localization to the Golgi and SCFD1 inactivation more or less abolishes localization. Co-IP experiments are used to argue that ACBD3 interacts with the N-terminal regulatory Longin domain of SEC22B, as well as with SCFD1. The Sec22 data are more detailed and persuasive. No experiments with purified proteins are presented to establish that the detected interactions are direct rather than mediated through a bridging factor or factors. Importantly, SCFD1 is likely to have multiple different client SNARE complexes that operate at different stages of ER and Golgi traffic. Hence its inactivation is likely to be pleiotropic and consequently phenotypes arising must be interpreted with caution.

      We completely agree that studying membrane trafficking in an interconnected system is challenging. We also agree that direct binding experiments in reconstituted systems would be key to proving our model. Our data uses multiple different experimental approaches, including co-localisation, co-immunoprecipitation, CRISPR-KO, and biochemistry, to support our model. In the future, we agree full reconstitution would be necessary to examine this further, and we hope that either ourselves or others can do this in further studies.

      Lastly, the authors perform IP experiments which show that ACBD3-Golgin co-IP efficiency is lower in cells with acute inactivation of SCFD1. This epistatic relationship is used to argue for a sequential model of recruitment with SCFD1 and perhaps client SNARE proteins operating upstream of ACBD3-Golgin interaction. This argument is not persuasive because we do not know whether SCFD1 and its downstream activities increase the rate of ACBD3-Golgin complex asssembly, or alternatively stabilizes ACBD3-Golgin complexes, decreasing the rate of their dissociation.

      We agree with this weakness in our original submission, and it is a comment shared among all reviewers. Overall, we feel that we have chosen the model that best summarises our data. We, of course, accept that there are still components of this pathway that need clarification and are open for further study. This includes the issue raised here by the reviewer, as well as the intriguing observation that both golgins are transcriptionally upregulated upon SCFD1 KO in HeLa cells. In the revised manuscript, we have more clearly laid out the weaknesses of our model in the discussion and suggested future experiments to help clarify some of these issues. We have also modified the model to reflect some of these potential additional regulatory mechanisms.

      In general the methods are fairly clear but that there is room for improvement. The "high throughput" imaging pipeline is not clearly described.

      We agree with the reviewer, and apologise for not clearly explaining this. We feel that this unbiased approach of quantification is particularly rigorous and we have clarified this in the methods section of the updated manuscript.

      Each figure legend should specify the microscopy methods used, and for each result the number of biological replicates and cells analyzed should be specified.

      We agree with the reviewer and have included these details appropriately in the revised manuscript.

      The statistical methods (Student, Tukey, etc.) used for each experiment should be specified. Saying that statistics were calculated using Python 3.7 is useless without additional details. e.g. at least the libraries and codebase used should be indicated or deposited.

      We agree with the reviewer and have updated the manuscript accordingly. In short, all comparisons were made using either Student's t-test or Multiple Comparison of Means - Tukey HSD, FWER=0.05. These were conducted in Python 3.9 using pandas, matplotlib, seaborn and scipy. We used the MultiComparison function in scipy, and the comp.tukeyhsd for the post-hoc adjustment.

      Many figure labels (e.g. Fig. 2) use absurdly small fonts.

      We apologise for this. We believe that this is because we submitted it with in-line formatting. Our resubmission has full-page figures, and we feel the text is clearer now.

      The mass spec hits obtained should be provided both with and without exclusion of non-Golgi-localized proteins.

      We agree with the reviewer. Please see the new Table S1.

      Reviewer #3 (Significance):

      In general I think this is a useful and well controlled set of experiments producing useful insights. However, the interpretations need to be more carefully considered, and alternative interpretations must laid out as clearly as possible. Specifying the limitations of the study will make it more, not less, useful to the field. If the authors want to make the case more robustly that the interactions described are mediated through direct binding, or that the operation of SCFD1 and Golgins operate sequentially to recruit ACBD3, additional wet bench work will be required which will of course take time to complete.

      We once again thank the reviewer for the thoughtful and critical comments. These have helped to strengthen the manuscript. We have performed the additional bench work requested by the reviewer, which has further supported the paper and our model.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Stalder and colleagues report experiments designed to identify interactors of the Golgi-localized protein ACBD3 (a.k.a. GCP60), and to delineate mechanisms that allow ACBD3 to localize at Golgi compartments. ACBD3 is a 528aa protein with diverse previously reported interactions and functions, both in normal physiology and as a host factor in viral assembly processes. Stalder et al. first map which domains of ACBD3 are required for Golgi localization in HeLa cells, concluding that residues 368-528 are sufficient for localization. This region includes a GOLD (GOLgi Dynamics) domain previously reported to interact with Golgin tethering proteins. Alanine scanning identifies the motif MWT just upstream of the GOLD motif as necessary for Golgi localization. Acute CRISPR knockout identifies two Golgins, Golgin45 and Giantin, as necessary for ACBD3 Golgi localization, and IP indicates that the MWT motif breaks this interaction. These data are a bit scattered around the paper but taken together are reasonably persuasive, particularly when viewed in context with published work. This reader would have found the manuscript easier to follow had the Golgin and MWT motif data been presented en bloc.

      In a second set of experiments, IP-mass spec is used to identify ACBD3 interactors that might assist in the protein's localization. The MS data presented are filtered to exclude proteins not already identified as Golgi-localized. This is, I think, a mistake. Even if the authors choose to focus on known Golgi interactors as candidates for a localization function, the biological functions of ACBD3 are far from fully understood, and the full dataset would be of value to both cell biologists and virologists. Hits in the filtered dataset include the R-SNARE Sec22B, and the SNARE chaperone Sly1/SCFD1. Acute CRISPR inactivation of Sec22 decreases ACBD3 localization to the Golgi and SCFD1 inactivation more or less abolishes localization. Co-IP experiments are used to argue that ACBD3 interacts with the N-terminal regulatory Longin domain of SEC22B, as well as with SCFD1. The Sec22 data are more detailed and persuasive. No experiments with purified proteins are presented to establish that the detected interactions are direct rather than mediated through a bridging factor or factors. Importantly, SCFD1 is likely to have multiple different client SNARE complexes that operate at different stages of ER and Golgi traffic. Hence its inactivation is likely to be pleiotropic and consequently phenotypes arising must be interpreted with caution.

      Lastly, the authors perform IP experiments which show that ACBD3-Golgin co-IP efficiency is lower in cells with acute inactivation of SCFD1. This epistatic relationship is used to argue for a sequential model of recruitment with SCFD1 and perhaps client SNARE proteins operating upstream of ACBD3-Golgin interaction. This argument is not persuasive because we do not know whether SCFD1 and its downstream activities increase the rate of ACBD3-Golgin complex asssembly, or alternatively stabilizes ACBD3-Golgin complexes, decreasing the rate of their dissociation.

      In general the methods are fairly clear but that there is room for improvement. The "high throughput" imaging pipeline is not clearly described. Each figure legend should specify the microscopy methods used, and for each result the number of biological replicates and cells analyzed should be specified. The statistical methods (Student, Tukey, etc.) used for each experiment should be specified. Saying that statistics were calculated using Python 3.7 is useless without additional details. e.g. at least the libraries and codebase used should be indicated or deposited. Many figure labels (e.g. Fig. 2) use absurdly small fonts. The mass spec hits obtained should be provided both with and without exclusion of non-Golgi-localized proteins.

      Significance

      In general I think this is a useful and well controlled set of experiments producing useful insights. However, the interpretations need to be more carefully considered, and alternative interpretations must laid out as clearly as possible. Specifying the limitations of the study will make it more, not less, useful to the field. If the authors want to make the case more robustly that the interactions described are mediated through direct binding, or that the operation of SCFD1 and Golgins operate sequentially to recruit ACBD3, additional wet bench work will be required which will of course take time to complete.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary

      This is a very interesting and potentially important paper for the field of membrane biology and membrane trafficking, in which the authors have studied the molecular mechanisms by which ACBD3 (and consequently PI4KIIIb) is recruited to the cis-Golgi membranes. The authors suggest that this recruitment is based on a two-step process, mediated by interactions to, on the one hand, SCFD-1 (SLY1) and, on the other hand, two redundant golgins (golgin-45 and giantin).

      Comments:

      • Pg.1 : arfaptins, as far as I know, have not been shown to be involved in intra-golgi trafficking but rather in Golgi export (see e.g. ref. 12)
      • Pg. 1: reigon --> region
      • Arf1 also recruits PI4KIIIb right?
      • Fig. S1: the information about the number of cells per experiment is missing. Also, please add the information about what exactly is represented in the box plots (is it the distribution of the mean value of R per experiment? or the total distribution on a cell-by-cell basis of a representative experiment?)
      • The definition of Avg. Golgi int/avg. cell int. (a.u.) in Fig 1E,F is a bit difficult to understand to me. If I understand correctly, the total fl. int in the Golgi mask was computed and divided by the area of the Golgi mask (this is the av. Golgi intensity). A similar computation is done for the entire cell (including the Golgi), i.e., total fl. intensity in the cell mask is computed and divided by the area of the cell mask. Then the two av. intensities are divided (ratio = av. Golgi int / av. cell int.). This ratio, for a protein that is enriched in the Golgi area, should be larger than 1. For a protein that is equally distributed all over the cell, it should be 1, and for a protein that is excluded from the Golgi area, smaller than 1. Then to this value, the authors subtract the value of the ratio found for an inert construct (GFP of Halo alone), which I imagine should have an original ratio value of the order of 1, and hence, after this subtraction, norm. ratio values larger than 0 mean that they are more enriched at the Golgi area than GFP/HaloTag themselves. Is this correct? In principle, I don't see anything entirely wrong with this way of thought, but I just found it a bit difficult to understand, and in general one has to be careful when computing rations (quotients) and then subtract another ratio. Also, the units are not a.u., the value is dimensionless, what is "arbitrary" is the definition of 0 value and the based on this definition, also the actual value. I think it would probably be much clearer for the readers to compute somthing like the relative enrichment in the Golgi area as compared to the rest of the cell (excluding the Golgi area). That is, a value r'=(Int. Golgi mask / Area Golgi mask) / [(Int. Cell mask - Int. Golgi mask)/(Area cell mask - Area Golgi mask)]. This can be computed directly or defining a mask that is the cell mask - the Golgi mask. Also, some maths (unless I made a mistake) give that this r'= r (1-aG)/(1-r aG); where r is the ratio (before subtraction) defined by the authors, and aG=Area cell mask/Area Golgi mask. In any case, I'd suggest the authors to either adopt this other quantitation (without subtraction of the GFP/HAloTAG), which gives directly the fold-enrichment in the intensity density in the Golgi area with respect to the rest of the cell; or explain in more detail the maths of the value they are plotting now.
      • Fig. 1C&F: Besides the MWT mutant, the FKE mutant also seems to have a somewhat compromised Golgi localization. Have the authors followed on that, or what is the reason that they have just focused on the MWT mutant?
      • Very minor point, and without wanting to sound pedant at all, but I think (I might be wrong of course, so apologies if I am) that the plural of apparatus in latin is not apparati, but apparatus (fourth declination). So, I'd change the word in page 2 (or just rephrase the sentence: e.g. "resulting in Golgi fragmentation"). But of course, I'd leave this to the authors' discretion.
      • Fig. 3A: have the authors tried or been able to perform IF of the endogenous SCFD1 protein?
      • Similarly to what has been done for other panels, could you quantify Fig. 3C? Are PI4KIIIb protein levels affected upon the different KOs?
      • The last paragraph of the "SCFD1 and ACBD3 interact upstream of PI4KIIIβ recruitment<br /> to the Golgi apparatus" section reads a bit odd placed there. I think it is more appropriate for the discussion or for the intro part on SCFD1.
      • I am confused on Fig. 5A/B. The labels in the blots show that 390-528 (without UR) does not bind sec22 or scfd1, but the 368-529 does? Or I guess, judging by the MW seen in the middle blots, that there's some error in the labelling? also, the IP efficiency of the MWT mutant in the panel A blot is quite low, still sec22 seems to be very efficiently pulled down. Can the authors comment on that please? Would co-IPing against endogenous sec22 and scfd1 would work (so you don't need to rely on HaloTag+ligand?)
      • I really like the experiment 6B. Have the authors tested whether SEC22 is also recruited to mitochondria in those conditions? But not SCFD1?
      • The results shown in Fig 7 might show a partial depletion in the interactions, but to be fully trusted they would need to be quantified and a statistical test used to compare the values. I think this part is important to show very clearly, because even with low binding to golgins (remember, single knockouts do not prevent Golgi localization of ACBD3), one could expect that ACBD3 still localized to the Golgi but it does not in the absence of SCFD1 as shown in this paper.
      • A prediction of the proposed model is that in cells depleted of the two Golgins, SCFD1 and ACBD3 should still bind to one another, right? Did the authors test this?
      • The model presented here (fig 8) seems to suggest that only the conformational variation of ACBD3 that binds Golgins is able to recruit (bind) PI4KIIIb. Is this known, or is there any experimental evidence for that?
      • Have the authors thought about testing the FKE mutant in the experiemnts shown in Fig. 5?
      • In general, I think the title might be a bit misleading because of the use of PI4Kiiib. I understand what the authors mean, but because they have not thoroughly tested PI4Kiiib recruitment in their experiments, I think they should focuse rather on the mechanism of recruitment of ACBD3 the authors have found.

      Significance

      I think, as said above, that this is potentially an important paper for the field of membrane trafficking and membrane biology. Most of the experiments are in general well performed and well controlled, and the paper is clearly written and follows a logical line.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary

      This manuscript describes molecular mechanisms by which ACBD3 is recruited to the Golgi complex. ACBD3 recruits PI4KIII which is required to generate PI4P, a phosphoinositide which is key for the recruitment of essential Golgi proteins and hence is key to Golgi identity. The authors have used a combination of mass spectrometry, high quality fluorescence imaging, transient CRISPR knockdowns, and biochemical approaches such as IPs to identify the key determinant for recruitment of ACBD3 to the Golgi complex. They map the interaction between ACBD3 and the Golgi as a unique region (UR) upstream of its GOLD domain, identifying, in particular, an MWT motif as key for this recruitment. Using mass spectrometry they identify several novel interactors of ACBD3 as well as some established binding partners. Knockdown of these interactors reveal a key role for the SNARE, SCFD1, where reduced levels lead to complete loss of ACBD3 localisation to the Golgi without apparent disruption of Golgi structure. They further validate this interaction and that of another SNARE (Sec22b), which is part of the same SNARE complex as SCFD1, mapping the interaction to the longin domain of Sec22b. Surprisingly however they demonstrate that the UR domain does not mediate the interaction between ACBD3 and these SNAREs suggesting an alternative mechanism of recruitment. Previously identified ACBD3 interactors, Golgi proteins giantin and golgin-45 were also identified in the mass spectrometry screen and the authors demonstrate that these two proteins can recruit ACBD2 to the Golgi and this is dependent on the MWT motif identified in the UR domain. By knocking down SCFD1, they show reduced recruitment of ACBD3 leading them to propose a model of sequential recruitment of ACBD3 by SCFD1 followed by interactions with the golgins.

      Major points:

      This study is a well-executed and rigorous study of the molecular requirements for the recruitment of ACBD3 to the Golgi. The experimental approaches are state-of-the-art and the data are clean and convincing. The only caveat, raised by the authors themselves, is their interpretation that there are two sequential steps for Golgi recruitment of ACBD3. While they show that loss of SCFD1 reduces the interaction of ACBD3 with giantin and golgin 45, their model depends on doing the reverse experiment, i.e. assessing the effects of knocking down either giantin or golgin-45. This is especially relevant given the demonstration that golgin-45 is sufficient to recruit ACBD3 to mitochondria. It may well be that recruitment involves a tripartite complex, which is not uncommon in vesicular transport mechanisms Giantin is not an essential protein do it should be feasible to perform this experiment. The authors are equipped in the quantitative fluorescence microscopy which would be required and which would help resolve whether sequential or redundant mechanisms are required for ACBD3 recruitment.

      Significance

      PI4P is a phosphoinositide that is important for the recruitment of Golgi proteins. As with most PIs it is likely to act by coincidence detection in that Golgi associated proteins will recognise PI4P as well as other factors on Golgi membranes. This results in different local membrane environments which will be specific for particular functions. PI4KIII is key for PI4P production although the absolute levels of PI4P are likely to be determined by a balance of lipid kinases and phosphatases. However, since ACBD3 is key for the recruitment of PI4KIII it is important to understand the molecular mechanisms by which it is recruited. The manuscript thus makes a significant contribution to understanding one of the underlying mechanisms for PI4KIII recruitment although, as indicated above, stops short of establishing a clear model for the roles SCDF1 and Sec22b versus golgin 45 and giantin. For the future it will be of interest to determine why either a sequential or a redundant mechanism is required for the recruitment of ACBD3 as a scaffold protein.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      We would like to thank all reviewers for taking the time to evaluate our manuscript fairly and critically. Many helpful suggestions and discussion points were raised. One important group of comments raised concerns whether our proposed timer and counter models were the appropriate conceptual framework to discuss nuclear multiplication in schizogony, whether they were mutually exclusive, and whether other alternatives should be considered. These comments were instrumental for us to uncover some inconsistencies in our previous modeling approach. In the new manuscript, we now define the counter and timer models much more rigorously in the context of Plasmodium cell division. Based on these refined models we now provide a new statistical analysis that goes beyond the previous analysis, significantly improving the statistical support for our conclusions. Details are given in the following individual replies.

      Reviewer #1 (Evidence, reproducibility and clarity):

      Summary

      Malaria parasites replicating in human red blood cells show a striking diversity in the number of progeny per replication cycle. Variation in progeny number can be seen between different species of malaria parasites, between parasite isolates, even between different cells from the same isolate. To date, we have little understanding of what factors influence progeny number, or how mechanistically it is controlled. In this study, the authors try to define how the mechanism that determines progeny number works. They propose two mechanisms, a 'counter' where progeny number is determined by the measurement of some kind of parasite parameter, and a 'timer' where parasite lifecycle length would be proportional to progeny number. Using a combination of long-term live-cell microscopy and mathematical modelling, the authors find consistent support for a 'counter' mechanism. Support for this mechanism was found using both Plasmodium falciparum, the most prominent human malaria parasite, and P. knowlesi, a zoonotic malaria parasite. Of the parameters measured in this study, the only thing that seemed to predict progeny number was parasite size around the onset of mitosis. The authors also found that during their replication inside red blood cells, malaria parasites drastically increase their nuclear to cytoplasmic ratio, a cellular parameter remains consistent in the vast majority of cell-types studied to date.

      Major Comments

      It is stated a few times in this study that P. knowlesi has an ~24 hour lifecycle, and while this is the case for in vivo P. knowlesi, it was established in the study when P. knowlesi A1-H1 was adapted to human RBCs (Moon et al., 2013) that this significantly extended the lifecycle to ~27 hours, which should be made clear in the text. As much of this study revolves around lifecycle length and timing, the authors should consider some of their findings with the context that in vitro adaption can significantly alter lifecycle length.

      The reviewer raises an important point that we didn’t discuss for P. knowlesi. We now mention this directly in the introduction chapter (line 67) and in the discussion (lines 470ff). We are aware that P. knowlesi takes about 27 hours in the lab, which was also communicated by the Moon lab. We now cite relevant studies again in this context. We further address the issue of modified cell cycle time in vitro in the discussion in the sense that absolute values must be taken with caution and the focus of this study is about the relative ratio and correlation between the different cell cycle metrics.

      • The dichotomous distinction between 'timer' and 'counter' as mutually exclusive mechanisms seems to be a drastic oversimplification. Considering the drastic variation we see in merozoite number across species, between isolates, and between cells, it seems much more likely that there are factors controlled by both time-sensed and counter-sensed mechanisms that both influence progeny number.

      The study of progeny regulation in malaria parasites is very much in the early stages. We can agree that our models are simplifications, as is the case with all models. Our choice of just the two models timer and counter was driven by the number of cellular parameters we measure, i.e., duration of division phase and progeny number. These data essentially allow us to test the two competing models we presented. As we quantify more and more cellular parameters, based on the quantitative live cell imaging protocols established here, we will be able to test more complex cell cycle models. With our current data, we believe more complex models are not warranted.

      However, this valuable criticism, in conjunction with related remarks by other reviewers, made us reevaluate the constraints of our model more precisely. We noticed that the criteria used in the previous version in the manuscript contained unnecessary additional assumptions. Briefly, the previous counter model also required that final merozoite number was tightly controlled, while the previous timer model required the growth rate to be tightly controlled. These side assumptions were not made explicit in the manuscript and could bias the support towards one or the other model.

      We now improved the modeling approach substantially by removing implicit side assumptions, and clearly defining timer and counter models in terms of their correlations. The refined formulation of the timer posits that between individual parasites the target duration and the nuclear multiplication rate vary in a statistically independent way; while in a counter, target number and nuclear multiplication rate are statistically independent. We now explain this extended analysis in more detail in the introduction (lines 86ff). We also now more clearly state the dichotomous nature of the model (line 488). A new results paragraph (lines 213ff) and an entirely new Fig. 2 (and Fig. S4) contains the model predictions and statistical comparison between the models.

      This more rigorous treatment showed that including the variance of the multiplication rate was critical to allow a clean discrimination between the models. Also, with the sole exception of P.knowlesi H2B, where no model was clearly favored (Fig. 2G-H,K), the timer model was found to be inconsistent with the data, while the counter was clearly favored. Our new goodness-of-fit analysis also showed that although the counter is strongly simplified, it produced adequate fits, demonstrating that potential model refinements would need to be justified by new, more extensive data.

      It is also important to consider that the degree of variation in merozoite number could rather be an expression of varying growth conditions and does not directly predict which of the proposed models are true. For instance, a counter where the target merozoite number varies strongly depending on growth conditions, would be consistent with all available data. It is an interesting question for future work whether a counter would indeed describe growth across different isolates.

      The biological reality of growth regulation is certainly complex, and the counter model will likely need to be refined in the future, which we acknowledge in a corresponding statement in the discussion (lines 491ff). Nevertheless, we find it encouraging that a simple model can explain the vast majority of our data very well.

      Additionally, the only parasite parameter measured in this study, size at time of first nuclear division, explained only a small proportion of the variance observed in merozoite number.

      It is indeed the case that amongst the measured parasite parameters i.e. schizont stage duration, nuclear volume, and cell size we only found the latter to correlate with the final progeny number. We did not aim to imply that all variation in progeny number is explained by cell size. It is likely that a putative counter relies on a set of factors, which are somehow linked to cell size. In addition, intrinsic stochasticity in nuclear growth is likely to contribute to final merozoite number variability, which is included in our models via a variable growth rate. Defining the actual limiting factor or combination of factors will be an exciting challenge for the future studies building on this one.

      • For modelling of a timer-based mechanism, the designation of t0 is subjective. The authors chose the time of first nuclear division as their t0. It is possible that a timer-based mechanism could not be supported based on this model the chosen t0 differs from when the "parasite's timer" starts. For example, t could also have been designated as the time from merozoite invasion (t0) to egress (tend). It would be unreasonable to suggest the authors repeat experiments with a longer time-frame to address this, but this possibility should be discussed as a limitation of the model. It may also be possible to develop a different model where t0 = merozoite invasion and tend = egress, and test this model against the data already collected in this study.

      This is a valid point. We indeed, considered the time point of invasion as the other relevant time point in the IDC for a possible timer. Due to necessary compromises in imaging protocols between acquisition length, temporal, and spatial resolution we have not been able yet to combine full-length IDC measurements with quantification of progeny number. Given the choice, however, between time point of invasion and the onset of nuclear division as starting point for a potential timer we would still favor the latter: An argument can be made that a timer that regulates offspring number would be more accurate when activated at the moment of the relevant cellular events rather than “running” for a very prolonged growth phase before any “decision” concerning parasite replication. We are still convinced that the entry into the schizont stage, which we analyze here, marks an important cell cycle transition point that has been highlighted in many different studies. As suggested, we now discuss the limitations of our selection of t0 in the text (lines 146ff).

      • The calculation of the multiplication rate is confusingly defined. In Figure 1 it is stated that it is "...based on t and n", which would imply that the multiplication rate is the number of merozoites formed per hour of schizogony, which would give an average value of ~2 for P. falciparum and ~1.5 for P. knowlesi. The averages rate values shown, however, are in the range of 0.15-3. The authors should clarify how these values were determined.

      Thank you for pointing out the need for more clarity. Since the nuclear multiplication, similar to e.g. cell population growth, follows an exponential law, the multiplication rate used (lambda) is in fact a logarithmic growth rate. Therefore, it occurs in the exponent (not as a coefficient) in the exponential growth function ( ), which explains the range. We now mention this more explicitly in the results (lines 163ff).

      • In Figure 2, the time from tend until egress is calculated, and this is interpreted as the time required for segmentation. In the Rudlaff et al., 2020 study cited in this paper, it is shown that segmentation starts before the final round of nuclear divisions are complete. Considering this, the time from tend until egress is not an appropriate proxy for segmentation time. The authors should consider rewording to something akin to "time from final nuclear division until egress" to more accurately reflect these data.

      Thank you for indicating our imprecise use of the nomenclature. Indeed, some essential segmentation-associated structures such as rhoptries and subpellicular microtubules are clearly forming before the last division. We were referring to “segmentation” as the time window where actual ingression of the plasma membrane occurs between nuclei with the concurrent formation of more prominent IMC-associated sub-pellicular microtubules between nuclei (as in Fig. 1A last panel). We can, however, agree that consistently using the term “merozoite formation” is more adequate here. We have now corrected the terminology according to the suggestions of the reviewer (lines 271ff).

      • There is a significant discrepancy between the data in Figure 5 and Supplementary Figure 8. In Supplementary Figure 8, the authors establish that culturing parasites in media diluted 0.5x has a marginal effect on parasite growth, with no discernible change in parasitaemia over 96 hours. By contrast, in Figure 5a the parasitaemia of parasites cultured in 0.5x diluted media is approximately 5-fold lower than those in 1x media. The authors should explain the significant discrepancy between these results.

      The reviewer correctly points out a difference in parasitaemia between two parasite culture experiments, shown in Figs 5a (now 6A) and S8 (now S11), respectively. There were several differences in the experimental setup used in the two experiments that could explain this discrepancy. In Fig. 5a the parasites were synchronized to early ring stages while in Fig. S8 we used asynchronous cultures (maybe with a slight majority of late stages). One could speculate that by the time the synchronized ring stage culture reached egress the effect of nutrient depletion, which started at t = 0 h is more pronounced. This effect could have been exacerbated by the more frequent media change of 24 h in Fig. 5a vs 48h in Fig. S8. Lastly, the starting parasitemia was differently set being higher at around 0.5% in the Fig. 5a while only 0.2% in Fig. S8. Possibly a lack of nutrient is “felt less” by the culture at lower parasitemias. Generally, in Fig. S8 we were more focused on highlighting the difference between 1x/0.5x and the more diluted conditions on the long-term culture and to show that continuous culture is actually possible in 0.5x medium. We have now expanded the legends to highlight those differences more clearly.

      • In Supplementary Figure 4, the mask on the cell at t0 shows two distinct objects, but it seems very unlikely that they are two distinct nuclei as they vary approximately 5-fold in diameter. The authors should provide more detail on how their masking was performed for their volumetric analysis. Specifically, whether size thresholds were also applied during object detection.

      Thank you for requesting clarification here. Fig S4 (now S7) shows only one z-slice (not a projection) of the entire image stack, to illustrate how the thresholding approach was performed on every single image slice. The two objects in the shown cell are indeed two nuclei, but because they are not in the same z-plane appear to be of different size. In particular, only a slice of the upper part of the nucleus on the lower right is visible in the shown slice. Throughout the study, volume determination was realized by adding up the individual slices, as is explained in detail in the Materials and Methods sections. We have now added a more explanation in the figure legend to clarify the procedure.

      Minor Comments

      • Line 45-48 mentions that merozoite number influences growth rate and virulence, but the corresponding reference (Mancio-Silva et al., 2013) only discusses the relationship between merozoite number and growth rate, not virulence.

      We thank the reviewer for requesting this distinction. Merozoite number and virulence have not been correlated in vivo so far. Certainly, because one can’t retrieve late-stage P. falciparum parasites from patients, but maybe partly because merozoite number has not gotten significant attention as a metric in the previous decades. Even if merozoite number is intuitively connected to growth rate which might causes higher parasitemia which is in turn linked to more severe disease outcome it is important to emphasize that those are certainly not equivalent. We have therefore removed the statement about virulence (line 48).

      • Line 59 states that a 48 hour lifecycle is a baseline from which in vitro cultured parasites deviate. Clinical isolates also show variation in lifecycle length and so it is more accurate to just say that 48 hours is an average, rather than a baseline.

      The word “baseline” has been changed to “average” (line 61).

      • Line 63 cites a study for the lifecycle length of P. knowlesi (Lee et al., 2022), but there seems to be no mention of lifecycle length in this reference

      This reference was meant to serve as an introductory review article to research in P. knowlesi. Actually, to the knowledge of the authors, there is no study presenting quantitative data showing that the in vitro cycle of P. knowlesi is actually around 27 h. Our lab experience is however coherent with a 27 h cycle, which was confirmed by personal communication by the Moon lab. We now also cite in the next sentence the inaugural P. knowlesi adaptation publication (Moon et al. 2013) showing some time course data indicating the duration of the IDC to be around ~27h (lines 67ff).

      • If I am interpreting Figure 3B correctly, this is essentially a paired analysis where the same erythrocytes are measured twice, once at t0 and once at tend. If this is the case, this data may be better represented with lines that connect the t0 and tend values.

      Yes, these are the same erythrocytes measured twice. We have modified Figure 3 (now Fig. 4) accordingly.

      • Figure 3A seems to imply that to calculate diameter of the erythrocytes, three measurements were made and averaged for each cell. I think this is a nice way to get a more accurate erythrocyte diameter, but if this is the case, it should be specified in the figure legend or methods.

      This is already described in the figure legend (line 305).

      • In Figure 4I it is shown that in P. falciparum merozoite number doesn't correlate with nucleus size, but for P. knowlesi in Supplementary Figure 7c, a significant anticorrelation is observed. The authors should state this in the text and discuss this discrepancy.

      Contrary to all other graphs, visual inspection of the distribution of data points in Fig. S10C shows that it contains two outlier data points at the bottom right. Those two specific points are also responsible for the significant anticorrelation. We did not filter or remove any quantification results but also didn’t have sufficient confidence in this data distribution (which is further based on the segmentation of the Histone2B not on an NLS mCherry signal) to make substantial claims about anticorrelation. Because we considered it informative we still decided to show it in the supplements. We now briefly mention the issues with the data set and its interpretation in the text (lines 350ff).

      • The authors show that merozoite number roughly correlates with cell size at t0 but it would be interesting to see whether cell size at tend also corresponds with cell size at t0. This might help answer whether the cell is larger because it has more merozoites, or whether it has more merozoites because it is larger.

      Plotting parasite cell volume at t0 against cell volume at tend (as well as between t-2 and tend) indeed shows a positive correlation (see below). While it is an interesting thought we concluded after some discussion that no convincing causal relationship between cell size and merozoite number can be inferred based on this analysis. Since we consider the possible statement that cells that are bigger in the beginning are also bigger in the end unavailing, we decided not to include the data.

      • I don't feel that "nearly identical" is an appropriate summary of erythrocyte indices in Supplementary Figure 9, considering there is a statistically significant increase in mean cell volume. I think it is unlikely that this change is consequential, and performing these haematology analyses is a nice quality control step, but this change should be stated in the text.

      In the modified text we now express the significant change in MCV in terms of percentage, which is around 1.2% (line 381).

      • In Supplementary Figure 8, parasitaemia only increases ~2-fold compared to >5-fold the previous two cycles. It seems likely that at the final timepoint on this graph the parasites are starting to crash, and therefore it may be best to end the graph with the 96 hour timepoint.

      The reviewer suggests that cultures at those parasitemias might not be in perfect health. Our Giemsa stains did not show signs of an unhealthy culture and kept growing. It was, however, important for us to show that cultures can be maintained in culture over a prolonged period of time in 0.5x medium, even when resulting in reduced growth, while this was not possible with lower dilutions. Therefore, we would like to keep the data point. We have added a cautionary comment in the legend.

      • The error bars in Figure 5C aren't easily visible, moving them in front of the datapoints may help their visibility.

      Error bars were moved in front of the data points.

      • In Figure 6D & E, the y-axis labels should be changed to whole integers as all the values in the graph are whole numbers.

      We have changed the y-axis labels accordingly.

      • My interpretation of Figure 6 C-E, is that these are the same cells measured at three time points (t-2, t0 and tend). If this is the case, 6C is missing the cell that has a merozoite number of 8, which is presumably why the y-axes are not equalised for the three graphs.

      It is correct that the same cells are displayed in all three plots, with the exceptions of three cells in 6C (for the timepoint t-2), which are missing for the following reasons: 1) it was not possible to determine the volume at this respective timepoint due to technical issues or 2) the cell was already just before t0 at the start of the movie so that t-2 had already passed. We now note this in the figure legend and have also equalized the y-axes (now Fig. 7C-E).

      Reviewer #1 (Significance):

      In the asexual blood-stage of their lifecycle, malaria parasites replicate through a process called schizogony. During schizogony an initially mononucleated parasite undergoes multiple asynchronous rounds of mitosis followed by nuclear division without cytokinesis, producing a variable number of daughter nuclei. Parasites then undergo a specialised cytokinesis, termed segmentation to where nuclei are packaged into merozoites that go on to invade new host cells. While nucleus, and therefore merozoite, number are known to be varied between cells, across isolates, and across species, little is known about the mechanisms regulating merozoite number. In this study, the authors use live-cell microscopy to understand how parasites determine their progeny number. They suggest that parasites regulate their progeny number using a 'counter' mechanism, which would respond to the size or concentration of a cellular parameter, as opposed to a 'timer' mechanism. Long-term live-cell microscopy experiments using malaria parasites are extremely technically challenging, and the authors should be commended for their efforts in this regard. While I agree that the data generated from these experiments are technically sound, I have some reservations expressed above about the interpretation of some of these results. I would strongly encourage the authors to consider rewording some of their interpretations taking into account some of the caveats listed above. I would also consider fitting/testing an additional mathematical model where the time-frame proposed for the 'timer' mechanism begins following merozoite invasion.

      We thank the reviewer for the appreciation of our work and hope we have sufficiently reworked the manuscript based on the comments listed above. Furthermore, we think the improved model statement and analysis improves the clarity of our conclusions. Indeed, we would like to test additional models including the full IDC once, as mentioned above, we are technically able to generate these data.

      This work is of specific interest to anybody who grows malaria parasites, as the dynamics of their growth is obviously important to understand. Further, this work is of interest more generally to cell biologists who study the regulation of progeny number or cell size. I have no experience with the application of mathematical modelling to understand biological systems, and so I cannot comment on the interest of this work to that field.

      Reviewer #2 (Evidence, reproducibility and clarity):

      This is a solid study that further characterises the dynamics of nuclear division in Plasmodium falciparum and P. knowlesi. Of two, among potentially several, models for how the number of daughter nuclei, and thus parasites - (called merozoites in this genus), are one that posits nuclei divide until a fixed timer ends, and one that posits that nuclei divide to reach a fixed number that is defined by a cellular counter. I find some practical difficulties in definitive measurement of either model, one issue with the former is that experimental definition of the start of the timer is problematic - we may define the starter's gun (eg by the first nuclear division) but it isn't necessary that the cell is using that same start time.

      We are pleased that the Reviewer found our study ‘solid’. Concerning the timer model, we agree that the selection of the starting point is a critical aspect of this study, as also Reviewer 1 pointed out. We selected this particular “t0” because the entry into the mitotic phase marks an important cell cycle transition. Several studies have suggested a “schizogony entry checkpoint” might be active just before (Matthews et al, 2018; Voß et al, 2023; van Biljon et al, 2018; McLean & Jacobs-Lorena, 2020). Once cells are committed to the schizont stage they are less responsive to stimuli. Alternatively, the timepoint of erythrocyte invasion could be a legitimate starting point. Due to necessary compromises in our imaging protocol between acquisition length, temporal, and spatial resolution we have not been able yet to combine full-length IDC measurements with quantification of progeny number, and therefore we leave exploration of an earlier timer start for future work. Within the confines of the model comparison in the current study, we think the selected t0 is already highly informative. We now explain the selection and limitations more explicitly in the text (line 144ff).

      Additionally, as the authors confirm here, being sure when that first nuclear division has occurred is particularly tricky with Plasmodium parasites, in part because the first few nuclei seem to clump together, preventing one from unambiguously calibrating the first division.

      The Reviewer is concerned about difficulties with precise reporting of the time point of first nuclear division. We suspect there was a misunderstanding here. In the text (line 137) we had written the following:

      “Although separating individual nuclei after the first two rounds of division was challenging due to their spatial proximity, the improvements in resolution and 3D image analysis allowed us to count the final number of nuclei routinely and reliably at the transition into the segmenter stage.”

      To clarify, when analyzing 3D image stacks produced by the LSM900 Airyscan the first nuclear division can consistently and unambiguously be detected. In anaphase the nuclei are pushed apart quite substantially before getting a bit closer together afterwards (see e.g. Fig. 1B and C). Hence the precision of the detection is only limited by the 30 min interval of the time lapse. Later, at the four nuclei stage, crowding makes distinction more difficult. In the final segmenter stage, the reorganization and condensation of nuclei makes reliable counting possible again. We have now reformulated the quoted sentence for more clarity (lines 137ff).

      Furthermore, getting decent replicate numbers is hard because of the difficulties of time lapse microscopy, and most Plasmodium studies (including this one) suffer from low enough numbers that it isn't always clear whether the numbers support one model over another.

      The reviewer points out the difficulty of obtaining enough replicates in Plasmodium time-lapse studies. We agree that depending on technology, sufficient replicates can be challenging. In the present study we obtained Ns between 25 and 35 for all conditions in P. falciparum and P. knowlesi from three independent replicas. To gain confidence in the conclusions from a limited, but not austere, data, it is essential to 1) reduce model complexity to a minimum and 2) perform stringent statistical analysis including accounting for small-sample variation. Motivated by this concern of the Reviewer and a similar point raised by Reviewer 1, we have revisited our modeling approach in the revised manuscript. This led us to a corrected, more rigorous definition of what precisely we mean by ‘counter’ and ‘timer’ models: The timer posits that between individual parasites the target duration and the nuclear multiplication rate and vary in a statistically independent way, while in a counter target number and nuclear multiplication rate are statistically independent. With no further adjustable parameters, the two models are thus both mutually exclusive and minimal. Although biological reality is likely to be more complex, we feel that these minimal models are adequate for the amount and resolution of our current, state-of-the art data. The general result remained the same: The counter model is strongly preferred in almost all our experiments data (new Fig. 2), with the sole exception of P. knowlesi H2B, where indeed more data may be needed to come to a clear conclusion. Furthermore, we have taken care to scrutinize these conclusions accounting for goodness-of-fit for the respective sample size N. This analysis showed, surprisingly, that the counter model was sufficient to account for the data: the real dataset was as similar to the counter prediction as synthetic, counter-generated data. We hope that this improved statistical analysis can help the reader judge the robustness of our conclusions.

      Nonetheless, several recent studies, particularly a study from the same institute (Klaus et al., 2022) employing timelapse imaging of nuclei, and timing the nuclear division of parasites, finds poor correlation between the duration of "schizogeny" (although perhaps using a different definition to the one used by the parasite) and the final number or merozoites. They therefore argue that there is poor evidence for a timer, and conclude by elimination that a counter must exist instead. A review by some of the authors of that study and some of this current study (Voß et al 2023), also concludes that the data from Klaus and colleagues "strongly support" a counter model. This current study also concludes that a counter model controls final nuclear/merozoite number in P. falciparum and P. knowlesi. This much at least is not particularly novel given the recent work on this topic, although the addition of the P. knowlesi data is interesting and consistent with the prior P. falciparum work.

      Our present work, indeed, does confirm the previous report of a counter over a timer, through a more targeted approach. While Klaus et al. used timing data of first nuclear cycle vs. the full duration, we now provide, thanks to an improvement microscopy setup and protocol, simultaneous measurements of timing and final progeny number, i.e. counting of merozoites/nuclei. While the preference for a counter model is not fundamentally novel, the additional information that the counter model holds in different strains, conditions and species is, in our opinion, not trivial and points to some degree of evolutionary conservation. We also demonstrate here that the counter model is not only preferred over the timer, it also fits the data adequately, so that it can be considered ‘correct’ at this level of complexity. Another, possibly more important, value of this study lies in the quantitative and time-resolved assessment of multiple important parasite metrics such a cell volume and nuclear volume together with merozoite number at the single cell level. Although descriptive, this has not been achieved in Plasmodium until now.

      As above, the authors concede that it is difficult to determine with strong confidence when the first nuclear division has occurred, so it may well be that there is substantial noisiness in the time that they define schizogeny to commence. If that were the case, this would contribute to the poor correlation observed between schizogeny duration and number of merozoites produced, so this could be an important confounding experimental factor. This deserves some more discussion by the authors.

      Concerning the confidence with which we identify the first nuclear division we could hopefully clarify in the section above that our precision is only limited by the time resolution of the acquired time-lapse. Therefore, the uncertainty about the start time is not particularly high, and moreover, can expected to affect timer and counter (via the growth rate) to a similar degree. We see no unfair advantage for the counter for this reason.

      Alternative methods to count absolute DNA content (rather than trying to count individual nuclei) might be useful ways of independently confirming this phenomenon. Alternative possibilities for what constitutes the "start" of a possible timer are also warranted - it could be for example, the first division of one of the other organelles.

      This is an interesting suggestion. Next generation fluorogenic DNA dyes have been used by us and the Ganter group (Simon et al. 2021, Klaus et al. 2022, Wenz et l. 2023) to assess DNA content of single cells over time. Our experience shows that there are some caveats to using these Hoechst based dyes, some of which we discussed in the aforementioned publications. While they allow some reasonable absolute quantification of DNA content for the very first S-Phase (and subsequent nuclear division), in later stages only relative quantification can be achieved. One underlying reason is the apparent increase of dye permeability, and therefore higher intensity, at late schizont stages. This issue is exacerbated by the asynchronous DNA replication of multiple nuclei. Further, nuclear division itself can be delayed or even inhibited when increasing the concentration of the dye, which suggest an impact on cell physiology (well documented for Hoechst based dyes in other organisms). When reaching the segmenter stage, the resulting variance in fluorescent intensity would make it challenging to assign a reliable number of nuclei required for analysis, a problem that does not occur when counting individual nuclei. Taken together, unfortunately, all these confounding factors make DNA content analysis in live single cells for the entire schizont stage unachievable at this point.

      These and previous authors in any case conclude that a counter model must exist through exclusion of a timer model. I am less convinced that the evidence discounting the timer is conclusive, and that a straight counter model is the only alternative. Indeed I am unconvinced by the suitability of this strictly dichotomous two-model system to categorise the division of unicellular eukaryotes, and these theories are not universally held to be sufficient to describe division.

      We thank the Reviewer for this insightful comment. As already detailed above, we have clarified and corrected our model definitions in the revised manuscript. Further, we want to make the important distinction between organisms, including unicellular ones that undergo binary fission and the ones like Plasmodium that use schizogony. Our model, although inspired by model organisms, is tailored to a multinucleated division mechanism, and clearly defined within those boundaries. The timer and counter models we consider are defined by their correlation structures. They are at two extremes of a continuum of models which could be characterized, for instance, by the ratio of correlations (growth rate - nuclear number) vs. (growth rate – duration) as an additional parameter. As the reviewer points out, excluding the timer model is not equivalent to proving the counter model, and indeed a partially correlated model, or a more complex model entirely, could yield a better fit. However, within the realm of models without additional parameters, and which are testable with the available data, only timer and counter remain, as different timer start points are not experimentally accessible. Importantly and somewhat surprisingly, the counter model also gave a fit that is as good as can be reasonably expected for the experimental sample size (new Fig. 2). So, we maintain that within the current experimental constraints, the counter model is the only viable option for almost all our tested conditions. The observation that in H2B-GFP expressing P. knowlesi parasites no clear distinction can be made between the models, indeed, suggest that the reality of multiplication rate regulation is more complex and may be limited by different constraints in different growth regimes. We now state these limitations and the room for further model adjustments with more data in the Discussion section.

      Nonetheless, if a counter exists, what is being counted that determines the final number? The authors consider that this might be a physical object or resource inside the parasite, or an extrinsic/extracellular resource. They investigate this by comparing the final cell number to a number of factors. First, the authors investigate the size of the RBC (by musing the diameter as an indicator)- little information is given about the source of the blood used, but it appears to be from a single donor of unknown age, who has approximately typical variance in RBC diameter (at least, after manipulation and storage). The authors observe little correlation between these variables.

      We share the curiosity of the reviewer about what might be “counted” by the parasite. This shall be the subject of future studies, and our present study provides the necessary basis for asking this question and defines a framework to investigate it. Concerning the size of the host cell, the blood used was from a different donor for each of the replicas, which we now specify in the figure legend (line 302). No significant difference between the RBC diameters between the donors was observed. A correlation between RBC diameter and progeny number was indeed not observed.

      Second the authors measure parasite size at the onset of schizogeny, and find that bigger parasites result in more daughter merozoites early in schizogeny (perhaps not surprising, given the earlier mentioned technical problems with measuring the first few steps of schizogeny), but that this different initial cell size doesn't result in a different final merozoite number, or as they describe it "not quite significant anymore". Previous p values were taken as cause for rejecting the timer hypothesis and the timer model. In this case the authors instead interpret the data as suggesting "that the setting of the counter might correlate with parasite cell size". This is inconsistent statistical and analytical handling, and highlights the earlier potential pitfall of rejecting timer-based models based on not gathering data that statistically show a correlation. This needs reworking to highlight that these data are inherently noisy, difficult to measure accurately, and aren't necessarily going strongly reveal a trend even where one biologically exists, and that this ought not be used as grounds for confident rejection of a model.

      The Reviewer raises concerns about the consistency of the statistical interpretation of our data. We care deeply about the well-foundedness of our conclusions and hope to eliminate these concerns in the following. First, we hope that the issue about the “technical problems” in measuring the first division has been solved in our response to previous comments. Next, to clarify an apparent misunderstanding: As stated in the text (lines 329ff) and shown in now Fig. 5D-E, cell size at onset of nuclear division or 2 hours prior does significantly correlate with final merozoite number. The lack of significant p-value (0.08) only pertains to the correlation of cell size at the end of the schizont stage (tend) with merozoite number (now Fig. 5F). We have removed the unfortunate wording “not quite significant anymore” in that context. Finally, regarding potential mechanisms, a potential counter must be set before the first nuclear division is completed because only that way it can be set independent of the speed of nuclear multiplication. This observation gives the statistically significant correlation of volume at the onset of division and progeny number its relevance. We have reformulated the marked sentence for more clarity (lines 331ff). Furthermore, we point out that our rejection of the timer is now based on a revisited statistical analysis (Fig. 2), which is no longer based on a simple correlation between final number and duration, as detailed above.

      Finally, the authors grow the parasites in dilute media, and find that they produce fewer daughter parasites. This is anecdotally unsurprising, as most Plasmodium laboratories are aware that sub-optimal growth conditions result in less healthy schizonts with fewer viable merozoites (and lower magnitudes of single-cycle expansion), but is nonetheless an important result that highlights explicitly how much this occurs in the specific conditions of dilute media. Given the lack of investigation of exactly which nutrient, carbon source, or combination thereof leads to the reduced merozoite number, it is unclear if or how much this is relevant to the scenario of a natural infection and realistic levels of that nutrient in a human or primate parasite environment.

      As rightfully pointed out by the reviewer suboptimal growth conditions affecting parasite growth and multiplication rate have been shown in many instances. The number of studies that actually quantify a reduction in merozoite number under different growth conditions is certainly much lower (Brancucci et al. 2017 (lipids), Mancio-Silva et al. 2017 (calorie-restriction in mice), Tinto-Font et al. 2022 (temperature) come to mind). What our study adds to this body of literature is to which extent duration of the schizont stage and cell volume are affected in relation to progeny number at the single cell level. Importantly, we wanted to test whether the counter model still holds under these more adverse conditions, which we found to be the case. Along the lines of the work on calorie restriction and the likely implication of isoleucine in the process investigated in the laboratory of Maria Mota, it will be exciting to identify a “limiting factor” in future studies. Indeed, any study done in complete RPMI culture medium can be questioned regarding its physiological relevance and we added a sentence addressing this aspect in the discussion (lines 514ff). Yet, our medium dilution experiments suggest that at least to some degree an extracellular resource is implicated, which makes sense from a biological function point-of-view.

      Minor issues

      The manuscript confuses the terms "less" and "fewer". Fewer should be used for countable nouns (fewer daughter cells, fewer nuclei, fewer merozoites), less for uncountable nouns (e.g. less speed, less volume).

      Thank you for pointing this out. The words have been replaced accordingly.

      I didn't understand lines 93-95; "This excluded a timer and thereby confirmed a counter as the mechanism regulating termination of nuclear multiplication (Klaus et al., 2022). A direct correlation between duration of schizont stage and merozoite number is, however, still missing." If I understand the first sentence concludes that there ought not be a direct correlation between schizont duration and merozoite number, but the second sentence, says that that correlation is "however" missing. Isn't this expected? Perhaps reword for clarity?

      Thank you for requesting clarification here. The exclusion of the timer by Klaus et al. 2022 was based on the correlation between duration of the first nuclear division cycle and the total duration of all nuclear replication phases. At no point did Klaus et al. count merozoites in live single cells, which was mainly due to lower spatial resolution of their images (M. Ganter, personal communication). Therefore, they could not directly assess the relation between progeny number and schizont stage duration, which we now report for the first time. The sentence was supposed to convey that this type of data was missing and was now reformulated for more clarity (line 114).

      Lines 104

      "We further uncover that throughout schizogony P. falciparum infringes on the otherwise ubiquitously constant N/C-ratio (Cantwell and Nurse, 2019)" This seems obvious to me, and not something uncovered by this study. In most of the numerous apicomplexans that divide by endoschizogeny, the cells achieve a near final size considerably before the final rounds of nuclear division so the N/C ratio must not remain constant - this is a direct corollary of many previous descriptions and not a novel finding of this study, and this claim here should be made more modest.

      We understand the point raised by the reviewer but still think that our claim is justified due to several aspects. There are examples of eukaryotic cells that undergo multinucleated stages during division were the N/C-ratio is constant (Dundon et al. 2016, Cantwell and Nurse, 2019), while we are not aware of any counter-example in the literature. Studies have also shown that e.g. certain mutant yeast that fail to undergo cytokinesis will increase their volume by factor of up to 16 alongside the still replicating and growing nucleus maintain the N/C-ratio (Neumann et al. 2007, Jorgensen et al. 2007). This demonstrates the tremendous plasticity that cells can reveal with respect to nucleus and cell size regulation. Until the contrary was shown, it was conceivable that nuclear compaction, which does occur (Fig. 5H), compensates for the increase in nuclear number while the cell volume is only increasing slightly. Importantly, we are not aware of any literature where nuclear volume has been quantified for blood stage Plasmodium. Cell volume quantifications remain limited to modelling and the study by Waldecker et al., which provides a few datapoints throughout the IDC. Whether this finding is expected or not, formally speaking, our claim is justified, but for more clarity we replace “uncover” with “demonstrate”. We also introduce the N/C-ratio as cellular parameter in P. falciparum pointing out another divergent aspect of its biology and might in the future understand the functional implication of this usually constant ratio, which is still unclear.

      Dundon SE, Chang SS, Kumar A, Occhipinti P, Shroff H, Roper M, Gladfelter AS. Clustered nuclei maintain autonomy and nucleocytoplasmic ratio control in a syncytium. Mol Biol Cell. 2016 Jul 1;27(13):2000-7.

      Neumann FR, and Nurse P. Nuclear size control in fission yeast. J. Cell Biol. 2007; 179: 593–600. pmid:17998401

      Jorgensen P, Edgington NP, Schneider BL, Rupeš I, Tyers M & Futcher B Molecular Biology of the Cell 18 (2007) The size of the nucleus increases as yeast cells grow.

      Helena Cantwell, Paul Nurse; A homeostatic mechanism rapidly corrects aberrant nucleocytoplasmic ratios maintaining nuclear size in fission yeast. J Cell Sci; 132 (22)

      I lack specialist statistical knowledge to comment on the statistical analyses performed on the correlation data, and in particular, whether the high p values for t-Tests for correlation are sufficient to support the argument that there is not a correlation, and whether these observations are sufficiently powered to robustly test that hypothesis.

      We are confident that our reworked model analysis, as explained above, now sufficiently supports our hypotheses.

      Reviewer #2 (Significance):

      The manuscript purports to find a counting mechanism that determines parasite merozoite numbers, and that this coutner is set by an externally provided and diffusible resource. Many nutrients are in excess in normal culture media, but not all. If that counted nutrient(s) were normally in excess in the bloodstream, it could hardly be said to be the factor that is counted and that therefore defines merozoite number. Conversely, if the amount of that nutrient were increased in normal media, would parasites make even more merozoites? Further, if the "counted" item is a freely diffusible compound in the media, it should be equally accessible to each parasite in a culture condition, and isn't a reasonable explanation for the variable merozoite numbers in the normal media conditions. To me, it is unsurprising that parasites that are healthy and well fed are able to produce more merozoites, but I don't see this as being the same as support for a counter model where the parasite senses and counts a set number of merozoites to produce in response to a specific external counter. I think the shoehorning of this phenomenon into a paradigm used to describe some other eukaryotes may not be appropriate, and that the rejection of one overly simplistic timer model should not automatically lead to us dichotomously accepting a simple counter method as the alternative. The authors need to do more to either identify a countable input whose gradual increase leads to a predictable and gradual increase in merozoite number, to show that they do use a counter, or provide substantially more caveats to their argument that the parasites are using a counter based on an externally provided resource to determine merozoite number.

      The reviewer comments on the feasibility of a counter mechanism based on an externally provided and diffusible resource. In fact this is a limited view of how a counter may arise and not the one we subscribe to. Rather, while a resource may be diffusible in the medium, it would need to be consumed during schizogony, and insufficiently replenished, in order to enable counting by dilution in the host cell. Furthermore, the reviewer has doubts that the fact that “healthy and well fed […] produce more merozoites” implies “support for a counter model”. We fully agree, and we argue in the manuscript that it is the correlations between schizogony durations and merozoite counts that support a counter model.

      As we have argued above, the two alternative models we consider are inspired by paradigm from other eukaryotes, but their definitions in the present context are simple enough for them to be considered natural minimal models of schizogony. As the simplest imaginable phenomenological models of multiplication control, we find it natural to compare them, and we hope our new introductory section introduces them appropriately now. Naturally, we hope to expand on this simple model in the future and identify more precisely the limiting resources and describe a more direct response.

      Audience - relatively specialised - likely interested audience would combine apicomplexan cell biologists, as well as theorists of cell division mechanism

      Advance - limited - confirms phenomenon also described by other researchers in their institute, and extends to another related organism.

      We would like to add that the present data are the first quantitative joint measurements of schizogony dynamics and outcome in P.falciparum and knowlesi. They allowed for the first time a direct correlation of duration and merozoite number, thereby accessing the question of growth control head on. Further they provide a quantitative reference of several key cellular parameters for anybody studying asexual blood stage parasites.

      Reviewer #3 (Evidence, reproducibility and clarity):

      Summary:

      Stürmer and colleagues used super-resolution time-lapse microscopy to probe the mechanism regulating the number of merozoites produced by a single cell in Plasmodium falciparum and P. knowlesi. The authors conclude the followings-

      1. P. knowlesi has similar duration of schizont stage to P. falciparum, although having a 24 h intraerythrocytic developmental cycle (IDC) to 48 h of P. falciparum.
      2. Nuclear multiplication dynamics suggests a counter mechanism of division- which is further suggested by a significant relation of merozoite numbers with schizont size at the onset of division.
      3. Nutritional deprivation caused increase in nuclear volume and decrease in merozoite number. For the most part, the experiments that are presented in this manuscript support the conclusion of the authors. The data are presented in a concise and clear manner. However, some clarification and a couple of experiment (listed below) would improve this manuscript.

      Major comments:

      1. The authors generated at least 3 transgenic lines for this study, But the did not present any genetic validation of the lines in the manuscript. For completeness, I recommend to provide genetic validation (either pcr genotyping or whole genome sequencing) of the lines that were generated and used in this study in the supplement.

      Our study exclusively used episomal expression of the respective fluorescent reporter (H2B-GFP, NLS-mCherry, and cytoplasmic GFP). As is customary in the field resistance to selection drugs and distinct fluorescent signals are assumed to sufficiently validate the presence of the plasmids. We now added the schematic maps of the plasmids in a new Fig. S1 to make our approach more visually clear.

      1. In the H2B-GFP lines, the authors episomally GFP-tagged histone 2B to label the nuclear chromatin for both P. falciparum and P. knowlesi. This provides a very useful parasite line which enables the live time-lapse microscopy. Using these parasite lines, the authors first show that despite having a 24 h IDC in P. knowlesi vs 48 h in P. falciparum, both these parasites have a similar duration of the schizont stage (8.s vs 9.4 h). My concern here is whether this GFP-tagging is influencing the growth dynamics as in slowing down the P. knowlesi parasites. However, if that was the case authors should have seen that for P. falciparum too. Also, for the P. falciparum parasites that episomally express cytosolic GFP and Nuclear mCherry have a higher number of merozoites compared to the H2B-GFP P. falciparum and the authors speculate this is probably because of not tagging Histone 2B. Given this, it is important to show that none of the H2B-GFP parasites show any significant fitness cost due to GFP tagging of histone. I recommend a simple experiment to compare the multiplication rate of H2B-GFP lines to the parental lines in identical growth conditions. This suggested experiment was described in PMID: 35164549 to determine fitness cost of knockout lines. This experiment is vital for validation of the H2B-GFP lines and subsequent interpretation of the data that were presented in this manuscript.

      We thank the reviewer for this excellent suggestion. To validate our lines further we now have carried out multiplication rate measurements similar to the one described in the designated publication for all the used lines alongside their parental strains (Fig. S2). We found no significant differences in between the wild type and the episomally expressing parasite lines (lines 131ff), which gives us confidence that episomal expression of tagged proteins do not significantly alter growth dynamics in these cases.

      1. The authors used the microtubule live cell dye SPY555-Tubulin in P. falciparum to validate the findings presented in 1D and 1E. They did not do that for P. knowlesi. If there is no unsurmountable technical difficulty, I suggest doing the same with P. knowlesi. This will also address the concern that I have pointed out in #1.

      Thank you for this suggestion. We have now generated the requested data with P. knowlesi, added it to what is now Supplemental Figure 3 and included it in our new analysis (Fig. 2I-J). The numerical values align well with the observations made when measuring schizont stage dynamics with the H2B-GFP expressing P. knowlesi line (line 158). A notable difference is that the Tubulin data strongly support the (refined) counter model, while the H2B data alone allow no distinction.

      1. The data in Figure 3 shows that merozoite number does not depend on host cell diameter. My question here is, were these data collected using different donor blood? Or were this measured from different biological replicate? These are not clear from the writing. I am not sure about whether blood from various donor would have on the data, however, different preparation of the cells across various biological replicate will have some effect on host cell diameter hence on data. State if these were collected from independent biological replicates and about the donor blood.

      The data results where indeed collected from three independent biological replicates using different donor blood batches. This is now stated in the figure legend. The batches displayed no difference in RBC diameter.

      1. It is interesting to see that nutrient-limited conditions increase average nuclear volume but less merozoite numbers. In this experiment, as I understand, complete media was diluted 0.5x, which basically diluted every component of the media by half. From this experiment I can see nutritional deprivation as a whole having an effect and supports the counter mechanism, it would be intriguing to see if there is any effect of a particular nutrient have any effect on progeny division. For example, parasites can be grown in amino acid deprived media (except isoleucine) which makes the parasites fully dependent on host cell amino acids. This sort of specific nutrient deprivation will probably allow the authors to probe for specific nutrients that plays role as counter mechanism factor.

      This is indeed a very exciting direction we would like to investigate in more detail in follow-up studies. Our aim for this study was to confirm that nutrient deprivation actually affects “counting” and to provide a workflow to investigate individual nutrients. In the meantime the Mota group, in a study we now cite in the discussion (lines 507ff), actually reported that isoleucine (and possibly methionine) levels are linked to progeny number. A follow-up on this topic using our strains and methodology is certainly worthwhile but requires more detailed analysis in the future.

      Minor comments:

      1. P. knowlesi is sometimes just written as knowlesi. Please, write P. Knowlesi.

      Has been corrected.

      1. Supplemental figure 1D, missing x-axis label.

      We added the x-axis label.

      1. In line 105, define N/C.

      Done.

      1. In line 205, I assume the authors mean episomally, not episomally.

      Thank you for pointing this out. We have replaced “ectopically” with “episomally” throughout the text.

      1. In line 275, Duration of Schizont stage was slightly....

      Has been corrected.

      1. All 'ml' or 'µl' should be 'mL' or 'µL'.

      Changes have been made.

      1. Define iRPMI.

      We added a definition (line 610).

      1. In line 475, replace 'as' with 'and'.

      Done.

      Reviewer #3 (Significance):

      The factors that regulate the number of progenies in malaria parasites remain unknown. While there are few previous studies attempting to answer the question, those studies were done on fixed stained cells. In this study, the authors used genetically modified fluorescent P. falciparum and P. knowlesi parasites that enable live microscopy. These parasites coupled with super-resolution time-lapse microscopy the authors attempt to investigate the mechanism(s) at play in regulating progeny division. This manuscript provides data to suggest that external resources might have some role in progeny division and supports the counter mechanism. More careful validation of the transgenic lines that were used to collect data presented needs to be more systematic and rigorous.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary:

      Stürmer and colleagues used super-resolution time-lapse microscopy to probe the mechanism regulating the number of merozoites produced by a single cell in Plasmodium falciparum and P. knowlesi. The authors conclude the followings:<br /> - a. P. knowlesi has similar duration of schizont stage to P. falciparum, although having a 24 h intraerythrocytic developmental cycle (IDC) to 48 h of P. falciparum.<br /> - b. Nuclear multiplication dynamics suggests a counter mechanism of division- which is further suggested by a significant relation of merozoite numbers with schizont size at the onset of division.<br /> - c. Nutritional deprivation caused increase in nuclear volume and decrease in merozoite number.<br /> For the most part, the experiments that are presented in this manuscript support the conclusion of the authors. The data are presented in a concise and clear manner. However, some clarification and a couple of experiment (listed below) would improve this manuscript.

      Major comments:

      1. The authors generated at least 3 transgenic lines for this study, But the did not present any genetic validation of the lines in the manuscript. For completeness, I recommend to provide genetic validation (either pcr genotyping or whole genome sequencing) of the lines that were generated and used in this study in the supplement.
      2. In the H2B-GFP lines, the authors ectopically GFP-tagged histone 2B to label the nuclear chromatin for both P. falciparum and P. knowlesi. This provides a very useful parasite line which enables the live time-lapse microscopy. Using these parasite lines, the authors first show that despite having a 24 h IDC in P. knowlesi vs 48 h in P. falciparum, both these parasites have a similar duration of the schizont stage (8.s vs 9.4 h). My concern here is whether this GFP-tagging is influencing the growth dynamics as in slowing down the P. knowlesi parasites. However, if that was the case authors should have seen that for P. falciparum too. Also, for the P. falciparum parasites that episomally express cytosolic GFP and Nuclear mCherry have a higher number of merozoites compared to the H2B-GFP P. falciparum and the authors speculate this is probably because of not tagging Histone 2B. Given this, it is important to show that none of the H2B-GFP parasites show any significant fitness cost due to GFP tagging of histone. I recommend a simple experiment to compare the multiplication rate of H2B-GFP lines to the parental lines in identical growth conditions. This suggested experiment was described in PMID: 35164549 to determine fitness cost of knockout lines. This experiment is vital for validation of the H2B-GFP lines and subsequent interpretation of the data that were presented in this manuscript.
      3. The authors used the microtubule live cell dye SPY555-Tubulin in P. falciparum to validate the findings presented in 1D and 1E. They did not do that for P. knowlesi. If there is no unsurmountable technical difficulty, I suggest doing the same with P. knowlesi. This will also address the concern that I have pointed out in #1.
      4. The data in Figure 3 shows that merozoite number does not depend on host cell diameter. My question here is, were these data collected using different donor blood? Or were this measured from different biological replicate? These are not clear from the writing. I am not sure about whether blood from various donor would have on the data, however, different preparation of the cells across various biological replicate will have some effect on host cell diameter hence on data. State if these were collected from independent biological replicates and about the donor blood.
      5. It is interesting to see that nutrient-limited conditions increase average nuclear volume but less merozoite numbers. In this experiment, as I understand, complete media was diluted 0.5x, which basically diluted every component of the media by half. From this experiment I can see nutritional deprivation as a whole having an effect and supports the counter mechanism, it would be intriguing to see if there is any effect of a particular nutrient have any effect on progeny division. For example, parasites can be grown in amino acid deprived media (except isoleucine) which makes the parasites fully dependent on host cell amino acids. This sort of specific nutrient deprivation will probably allow the authors to probe for specific nutrients that plays role as counter mechanism factor.

      Minor comments:

      1. P. knowlesi is sometimes just written as knowlesi. Please, write P. Knowlesi.
      2. Supplemental figure 1D, missing x-axis label.
      3. In line 105, define N/C.
      4. In line 205, I assume the authors mean episomally, not ectopically.
      5. In line 275, Duration of Schizont stage was slightly....
      6. All 'ml' or 'µl' should be 'mL' or 'µL'.
      7. Define iRPMI.
      8. In line 475, replace 'as' with 'and'.

      Significance

      The factors that regulate the number of progenies in malaria parasites remain unknown. While there are few previous studies attempting to answer the question, those studies were done on fixed stained cells. In this study, the authors used genetically modified fluorescent P. falciparum and P. knowlesi parasites that enable live microscopy. These parasites coupled with super-resolution time-lapse microscopy the authors attempt to investigate the mechanism(s) at play in regulating progeny division. This manuscript provides data to suggest that external resources might have some role in progeny division and supports the counter mechanism. More careful validation of the transgenic lines that were used to collect data presented needs to be more systematic and rigorous.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      This is a solid study that further characterises the dynamics of nuclear division in Plasmodium falciparum and P. knowlesi. Of two, among potentially several, models for how the number of daughter nuclei, and thus parasites - (called merozoites in this genus), are one that posits nuclei divide until a fixed timer ends, and one that posits that nuclei divide to reach a fixed number that is defined by a cellular counter. I find some practical difficulties in definitive measurement of either model, one issue with the former is that experimental definition of the start of the timer is problematic - we may define the starter's gun (eg by the first nuclear division) but it isn't necessary that the cell is using that same start time. Additionally, as the authors confirm here, being sure when that first nuclear division has occurred is particularly tricky with Plasmodium parasites, in part because the first few nuclei seem to clump together, preventing one from unambiguously calibrating the first division. Furthermore, getting decent replicate numbers is hard because of the difficulties of time lapse microscopy, and most Plasmodium studies (including this one) suffer from low enough numbers that it isn't always clear whether the numbers support one model over another.

      Nonetheless, several recent studies, particularly a study from the same institute (Klaus et al., 2022) employing timelapse imaging of nuclei, and timing the nuclear division of parasites, finds poor correlation between the duration of "schizogeny" (although perhaps using a different definition to the one used by the parasite) and the final number or merozoites. They therefore argue that there is poor evidence for a timer, and conclude by elimination that a counter must exist instead. A review by some of the authors of that study and some of this current study (Voß et al 2023), also concludes that the data from Klaus and colleagues "strongly support" a counter model. This current study also concludes that a counter model controls final nuclear/merozoite number in P. falciparum and P. knowlesi. This much at least is not particularly novel given the recent work on this topic, although the addition of the P. knowlesi data is interesting and consistent with the prior P. falciparum work. As above, the authors concede that it is difficult to determine with strong confidence when the first nuclear division has occurred, so it may well be that there is substantial noisiness in the time that they define schizogeny to commence. If that were the case, this would contribute to the poor correlation observed between schizogeny duration and number of merozoites produced, so this could be an important confounding experimental factor. This deserves some more discussion by the authors. Alternative methods to count absolute DNA content (rather than trying to count individual nuclei) might be useful ways of independently confirming this phenomenon. Alternative possibilities for what constitutes the "start" of a possible timer are also warranted - it could be for example, the first division of one of the other organelles.

      These and previous authors in any case conclude that a counter model must exist through exclusion of a timer model. I am less convinced that the evidence discounting the timer is conclusive, and that a straight counter model is the only alternative. Indeed I am unconvinced by the suitability of this strictly dichotomous two-model system to categorise the division of unicellular eukaryotes, and these theories are not universally held to be sufficient to describe division. Nonetheless, if a counter exists, what is being counted that determines the final number? The authors consider that this might be a physical object or resource inside the parasite, or an extrinsic/extracellular resource. They investigate this by comparing the final cell number to a number of factors. First, the authors investigate the size of the RBC (by musing the diameter as an indicator)- little information is given about the source of the blood used, but it appears to be from a single donor of unknown age, who has approximately typical variance in RBC diameter (at least, after manipulation and storage). The authors observe little correlation between these variables. Second the authors measure parasite size at the onset of schizogeny, and find that bigger parasites result in more daughter merozoites early in schizogeny (perhaps not surprising, given the earlier mentioned technical problems with measuring the first few steps of schizogeny), but that this different initial cell size doesn't result in a different final merozoite number, or as they describe it "not quite significant anymore". Previous p values were taken as cause for rejecting the timer hypothesis and the timer model. In this case the authors instead interpret the data as suggesting "that the setting of the counter might correlate with parasite cell size". This is inconsistent statistical and analytical handling, and highlights the earlier potential pitfall of rejecting timer-based models based on not gathering data that statistically show a correlation. This needs reworking to highlight that these data are inherently noisy, difficult to measure accurately, and aren't necessarily going strongly reveal a trend even where one biologically exists, and that this ought not be used as grounds for confident rejection of a model.

      Finally, the authors grow the parasites in dilute media, and find that they produce fewer daughter parasites. This is anecdotally unsurprising, as most Plasmodium laboratories are aware that sub-optimal growth conditions result in less healthy schizonts with fewer viable merozoites (and lower magnitudes of single-cycle expansion), but is nonetheless an important result that highlights explicitly how much this occurs in the specific conditions of dilute media. Given the lack of investigation of exactly which nutrient, carbon source, or combination thereof leads to the reduced merozoite number, it is unclear if or how much this is relevant to the scenario of a natural infection and realistic levels of that nutrient in a human or primate parasite environment.

      Minor issues

      The manuscript confuses the terms "less" and "fewer". Fewer should be used for countable nouns (fewer daughter cells, fewer nuclei, fewer merozoites), less for uncountable nouns (e.g. less speed, less volume).

      I didn't understand lines 93-95;<br /> "This excluded a timer and thereby confirmed a counter as the mechanism regulating termination of nuclear multiplication (Klaus et al., 2022). A direct correlation between duration of schizont stage and merozoite number is, however, still missing."<br /> If I understand the first sentence concludes that there ought not be a direct correlation between schizont duration and merozoite number, but the second sentence, says that that correlation is "however" missing. Isn't this expected? Perhaps reword for clarity?

      Lines 104<br /> "We further uncover that throughout schizogony P. falciparum infringes on the otherwise 105 ubiquitously constant N/C-ratio (Cantwell and Nurse, 2019)" This seems obvious to me, and not something uncovered by this study. In most of the numerous apicomplexans that divide by endoschizogeny, the cells achieve a near final size considerably before the final rounds of nuclear division so the N/C ratio must not remain constant - this is a direct corollary of many previous descriptions and not a novel finding of this study, and this claim here should be made more modest.

      I lack specialist statistical knowledge to comment on the statistical analyses performed on the correlation data, and in particular, whether the high p values for t-Tests for correlation are sufficient to support the argument that there is not a correlation, and whether these observations are sufficiently powered to robustly test that hypothesis.

      Significance

      The manuscript purports to find a counting mechanism that determines parasite merozoite numbers, and that this coutner is set by an externally provided and diffusible resource. Many nutrients are in excess in normal culture media, but not all. If that counted nutrient(s) were normally in excess in the bloodstream, it could hardly be said to be the factor that is counted and that therefore defines merozoite number. Conversely, if the amount of that nutrient were increased in normal media, would parasites make even more merozoites? Further, if the "counted" item is a freely diffusible compound in the media, it should be equally accessible to each parasite in a culture condition, and isn't a reasonable explanation for the variable merozoite numbers in the normal media conditions. To me, it is unsurprising that parasites that are healthy and well fed are able to produce more merozoites, but I don't see this as being the same as support for a counter model where the parasite senses and counts a set number of merozoites to produce in response to a specific external counter. I think the shoehorning of this phenomenon into a paradigm used to describe some other eukaryotes may not be appropriate, and that the rejection of one overly simplistic timer model should not automatically lead to us dichotomously accepting a simple counter method as the alternative. The authors need to do more to either identify a countable input whose gradual increase leads to a predictable and gradual increase in merozoite number, to show that they do use a counter, or provide substantially more caveats to their argument that the parasites are using a counter based on an externally provided resource to determine merozoite number.

      Audience - relatively specialised - likely interested audience would combine apicomplexan cell biologists, as well as theorists of cell division mechanism

      Advance - limited - confirms phenomenon also described by other researchers in their institute, and extends to another related organism.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary

      Malaria parasites replicating in human red blood cells show a striking diversity in the number of progeny per replication cycle. Variation in progeny number can be seen between different species of malaria parasites, between parasite isolates, even between different cells from the same isolate. To date, we have little understanding of what factors influence progeny number, or how mechanistically it is controlled. In this study, the authors try to define how the mechanism that determines progeny number works. They propose two mechanisms, a 'counter' where progeny number is determined by the measurement of some kind of parasite parameter, and a 'timer' where parasite lifecycle length would be proportional to progeny number. Using a combination of long-term live-cell microscopy and mathematical modelling, the authors find consistent support for a 'counter' mechanism. Support for this mechanism was found using both Plasmodium falciparum, the most prominent human malaria parasite, and P. knowlesi, a zoonotic malaria parasite. Of the parameters measured in this study, the only thing that seemed to predict progeny number was parasite size around the onset of mitosis. The authors also found that during their replication inside red blood cells, malaria parasites drastically increase their nuclear to cytoplasmic ratio, a cellular parameter remains consistent in the vast majority of cell-types studied to date.

      Major Comments

      • It is stated a few times in this study that P. knowlesi has an ~24 hour lifecycle, and while this is the case for in vivo P. knowlesi, it was established in the study when P. knowlesi A1-H1 was adapted to human RBCs (Moon et al., 2013) that this significantly extended the lifecycle to ~27 hours, which should be made clear in the text. As much of this study revolves around lifecycle length and timing, the authors should consider some of their findings with the context that in vitro adaption can significantly alter lifecycle length.
      • The dichotomous distinction between 'timer' and 'counter' as mutually exclusive mechanisms seems to be a drastic oversimplification. Considering the drastic variation we see in merozoite number across species, between isolates, and between cells, it seems much more likely that there are factors controlled by both time-sensed and counter-sensed mechanisms that both influence progeny number. Additionally, the only parasite parameter measured in this study, size at time of first nuclear division, explained only a small proportion of the variance observed in merozoite number.
      • For modelling of a timer-based mechanism, the designation of t0 is subjective. The authors chose the time of first nuclear division as their t0. It is possible that a timer-based mechanism could not be supported based on this model the chosen t0 differs from when the "parasite's timer" starts. For example, t could also have been designated as the time from merozoite invasion (t0) to egress (tend). It would be unreasonable to suggest the authors repeat experiments with a longer time-frame to address this, but this possibility should be discussed as a limitation of the model. It may also be possible to develop a different model where t0 = merozoite invasion and tend = egress, and test this model against the data already collected in this study.
      • The calculation of the multiplication rate is confusingly defined. In Figure 1 it is stated that it is "...based on t and n", which would imply that the multiplication rate is the number of merozoites formed per hour of schizogony, which would give an average value of ~2 for P. falciparum and ~1.5 for P. knowlesi. The averages rate values shown, however, are in the range of 0.15-3. The authors should clarify how these values were determined.
      • In Figure 2, the time from tend until egress is calculated, and this is interpreted as the time required for segmentation. In the Rudlaff et al., 2020 study cited in this paper, it is shown that segmentation starts before the final round of nuclear divisions are complete. Considering this, the time from tend until egress is not an appropriate proxy for segmentation time. The authors should consider rewording to something akin to "time from final nuclear division until egress" to more accurately reflect these data.
      • There is a significant discrepancy between the data in Figure 5 and Supplementary Figure 8. In Supplementary Figure 8, the authors establish that culturing parasites in media diluted 0.5x has a marginal effect on parasite growth, with no discernible change in parasitaemia over 96 hours. By contrast, in Figure 5a the parasitaemia of parasites cultured in 0.5x diluted media is approximately 5-fold lower than those in 1x media. The authors should explain the significant discrepancy between these results.
      • In Supplementary Figure 4, the mask on the cell at t0 shows two distinct objects, but it seems very unlikely that they are two distinct nuclei as they vary approximately 5-fold in diameter. The authors should provide more detail on how their masking was performed for their volumetric analysis. Specifically, whether size thresholds were also applied during object detection.

      Minor Comments

      • Line 45-48 mentions that merozoite number influences growth rate and virulence, but the corresponding reference (Mancio-Silva et al., 2013) only discusses the relationship between merozoite number and growth rate, not virulence.
      • Line 59 states that a 48 hour lifecycle is a baseline from which in vitro cultured parasites deviate. Clinical isolates also show variation in lifecycle length and so it is more accurate to just say that 48 hours is an average, rather than a baseline.
      • Line 63 cites a study for the lifecycle length of P. knowlesi (Lee et al., 2022), but there seems to be no mention of lifecycle length in this reference
      • If I am interpreting Figure 3B correctly, this is essentially a paired analysis where the same erythrocytes are measured twice, once at t0 and once at tend. If this is the case, this data may be better represented with lines that connect the t0 and tend values.
      • Figure 3A seems to imply that to calculate diameter of the erythrocytes, three measurements were made and averaged for each cell. I think this is a nice way to get a more accurate erythrocyte diameter, but if this is the case, it should be specified in the figure legend or methods.
      • In Figure 4I it is shown that in P. falciparum merozoite number doesn't correlate with nucleus size, but for P. knowlesi in Supplementary Figure 7c, a significant anticorrelation is observed. The authors should state this in the text and discuss this discrepancy.
      • The authors show that merozoite number roughly correlates with cell size at t0 but it would be interesting to see whether cell size at tend also corresponds with cell size at t0. This might help answer whether the cell is larger because it has more merozoites, or whether it has more merozoites because it is larger.
      • I don't feel that "nearly identical" is an appropriate summary of erythrocyte indices in Supplementary Figure 9, considering there is a statistically significant increase in mean cell volume. I think it is unlikely that this change is consequential, and performing these haematology analyses is a nice quality control step, but this change should be stated in the text.
      • In Supplementary Figure 8, parasitaemia only increases ~2-fold compared to >5-fold the previous two cycles. It seems likely that at the final timepoint on this graph the parasites are starting to crash, and therefore it may be best to end the graph with the 96 hour timepoint.
      • The error bars in Figure 5C aren't easily visible, moving them in front of the datapoints may help their visibility.
      • In Figure 6D & E, the y-axis labels should be changed to whole integers as all the values in the graph are whole numbers.
      • My interpretation of Figure 6 C-E, is that these are the same cells measured at three time points (t-2, t0 and tend). If this is the case, 6C is missing the cell that has a merozoite number of 8, which is presumably why the y-axes are not equalised for the three graphs.

      Significance

      In the asexual blood-stage of their lifecycle, malaria parasites replicate through a process called schizogony. During schizogony an initially mononucleated parasite undergoes multiple asynchronous rounds of mitosis followed by nuclear division without cytokinesis, producing a variable number of daughter nuclei. Parasites then undergo a specialised cytokinesis, termed segmentation to where nuclei are packaged into merozoites that go on to invade new host cells. While nucleus, and therefore merozoite, number are known to be varied between cells, across isolates, and across species, little is known about the mechanisms regulating merozoite number. In this study, the authors use live-cell microscopy to understand how parasites determine their progeny number. They suggest that parasites regulate their progeny number using a 'counter' mechanism, which would respond to the size or concentration of a cellular parameter, as opposed to a 'timer' mechanism. Long-term live-cell microscopy experiments using malaria parasites are extremely technically challenging, and the authors should be commended for their efforts in this regard. While I agree that the data generated from these experiments are technically sound, I have some reservations expressed above about the interpretation of some of these results. I would strongly encourage the authors to consider rewording some of their interpretations taking into account some of the caveats listed above. I would also consider fitting/testing an additional mathematical model where the time-frame proposed for the 'timer' mechanism begins following merozoite invasion.

      This work is of specific interest to anybody who grows malaria parasites, as the dynamics of their growth is obviously important to understand. Further, this work is of interest more generally to cell biologists who study the regulation of progeny number or cell size. I have no experience with the application of mathematical modelling to understand biological systems, and so I cannot comment on the interest of this work to that field.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity):

      Summary:<br /> In this original article from Mutscher et al., the authors developed a compartmentalized dissociated mouse myelinating SC-DRG coculture system to investigate the distinct roles of Schwann cells in axon protection and degeneration after injury. The innovation of this approach relies on (i) the use of mouse SCs and mouse DRGS neurons instead of rat cells; (ii) the use of microfluidic chambers, seeded by axons and SCs in different compartment; (iii) the possibility to perform a traumatic injury in vitro. While this novel approach offers new ways to study peripheral nerve regeneration and SC-axon interaction, and technical the study is robust, the paper is currently limited by the exploration of their model.

      Major points:<br /> Reviewer 1. It is unclear is this approach will ever lead to the identification of key mechanism or key candidates. This is a major miss in the current manuscript form. In short: the authors should demonstrate that their in vitro system can lead to significant leap in our understanding of peripheral nerve regeneration by identifying novel targets/pathways or mechanisms.

      Author response: We agree with the reviewer that cell culture approaches have limitations however we would disagree that it is not a viable approach given that a number of seminal studies in the field have already helped identified key cellular and molecular steps using rat SC-DRG cocultures or using mouse DRGs and rat SCs in combination with in vivo study. We have added the following to the introduction to highlight this point in more detail:

      Introduction.

      Dissociated myelinating SC-DRG cocultures from rats were first developed by the Bunge laboratory in the 1980’s to investigate PNS myelination in a more dynamic way (Bunge et al, 1989; Eldridge et al, 1987)__. These cultures have been used to make seminal discoveries in uncovering the cellular and molecular mechanisms of SC myelination alongside in vivo investigation. These include how the inner SC membrane (mesaxon) advances to myelinate axons, and the role of b__-neuregulin-1 (__b__NRG1) and polarity proteins in SC myelination (Bunge et al, 1989; Shen et al, 2014; Chan et al, 2006; Taveggia et al, 2005)__. Similarly, SC-DRG cocultures have been useful in demonstrating how SCs proliferate after axon injury, transfer metabolites, such as pyruvate, to delay axon degeneration, how placental growth factor (Plgf) regulates axon fragmentation by SCs and how SC JUN promotes axon outgrowth after injury (Arthur-Farraj et al, 2011; Babetto et al, 2020; Vaquié et al, 2019; Salzer & Bunge, 1980)__. The use of a coculture system to study axon-SC interactions during axon degeneration and regeneration offers some advantages over in vivo approaches as both neurons and SCs can be genetically manipulated separately and live imaged with ease.

      Discussion.

      Most importantly SCs and DRG neurons from various transgenic mice can be used to perform in vitro analysis to complement findings from in vivo transgenic mouse studies.

      Author response: Furthermore, as this is a methods paper, demonstrating novel molecular mechanisms is outside the scope of this article. However, we have already used this technique with a collaborator to study the role of cdk7 in myelination (see link to conference abstract below) and this manuscript is under preparation to be submitted soon. Additionally, we have ongoing projects within the lab using this technique to help characterise novel molecular targets in nerve injury. https://scholar.google.com/citations?view_op=view_citation&hl=en&user=uEtwAd8AAAAJ&sortby=pubdate&citation_for_view=uEtwAd8AAAAJ:_Qo2XoVZTnwC

      Author response: Despite this, we have shown that the axo-protective effect of SCs is independent of myelination status, which is an advance on what is known in the literature.

      Author response: We have realised through all of the reviewers’ comments that the title and the aims of the manuscript were confusing. We have made this clearer by removing the word novel in the title changing the title to the following:

      A method for mouse myelinating Schwann cell-DRG neuron compartmentalised cocultures: myelination status does not influence the axo-protective effect of Schwann cells after injury.

      Author response: We have also made it much clearer what the purpose of our study is and where and how it fits in with the previous literature by adding the following paragraph to the introduction.

      Introduction.

      Indeed there has only ever been one laboratory detailing convincing myelin formation in dissociated mouse myelinating SC-DRG neuron cocultures, however this was never published as a step by step detailed protocol (Stevens et al, 1998; Stevens & Fields, 2000)__. In the last twenty years, there have been no published studies demonstrating myelination in fully dissociated mouse SC-mouse DRG cocultures. This has largely prevented the use of cells, particularly SCs, from transgenic mice in cocultures and thus restricted the ability to study SC-axon interactions in a system that can be readily manipulated and live imaged and results directly applied back to in vivo findings in the same species.

      Reviewer 1. The use of embryonic DRG neurons or SC isolated from P2 animals are arguably physiologically not the same cells that are affected by traumatic nerve injury, which happen most often than not in adult. This is a problem in the long-term reliance on this approach to study axotomy peripheral nerve regeneration.

      Author response: We agree with the reviewer that one should always be cautious with the use of embryonic/neonatal cells to directly refer them to adult cellular mechanisms. We have added discussion of this point to the discussion:

      Discussion.

      One limitation of our coculture model and indeed all coculture and cell culture models that are used to investigate cellular and molecular mechanisms in nerve injury is that the cells are obtained from embryonic or neonatal animals. This is an important caveat when applying results from cell culture to adult in vivo nerve injury. However, while we would argue that cell culture approaches should always be used in combination with in vivo study it is important to remember that nerve injury is not restricted to adults and brachial plexus injury secondary to birth trauma is unfortunately a significant clinical problem (Pondaag et al, 2007)__. Furthermore, neonatal SCs replicate many of key cellular and molecular mechanisms seen in adult SCs after injury, including JUN upregulation, myelinophagy, promotion of axon growth and expression of key repair program transcripts (Arthur-Farraj et al, 2012; Gomez-Sanchez et al, 2015; Arthur-Farraj et al, 2017; Parkinson et al, 2008)__. A future development would be to try to adapt this protocol to make a coculture model with adult mouse or even human cells.

      Author response: Additionally, we already know Schwann cells in P5 neonatal mice in vivo after nerve transection demyelinate in a similar way to Schwann cells in adult mice and that neonatal cells in vivo and in vitro require the transcription factor c-Jun to do so (Parkinson et al., 2008 JCB Fig.7).

      Moderate points:<br /> Reviewer 1"there are no established protocols in the field describing the use of mouse SCs with mouse DRG neurons in dissociated myelinating cocultures". The use of mouse cells is laudable, but it is not necessarily a technical innovation, or at least the current manuscript does not explain why their approach particularly suitable to mouse Schwann cells.

      Author response: We feel that a detailed working protocol for compartmentalised dissociated mouse myelinating cocultures showing convincing and extensive myelination has been missing from our field for a long time. We agree that it is an incremental technical advance, but it is an important one. We have modified the title as we explained above. We have explained this point more clearly in the introduction, results, and discussion with the following additions:

      Introduction.

      Indeed there has only ever been one laboratory detailing convincing myelin formation in dissociated mouse myelinating SC-DRG neuron cocultures, however this was never published as a step by step detailed protocol (Stevens et al, 1998; Stevens & Fields, 2000)__. In the last twenty years, there have been no published studies demonstrating myelination in fully dissociated mouse SC-mouse DRG cocultures. This has largely prevented the use of cells, particularly SCs, from transgenic mice in cocultures and thus restricted the ability to study SC-axon interactions in a system that can be readily manipulated and live imaged and results directly applied back to in vivo findings in the same species.

      The consensus within the field is that inducing myelination in dissociated mouse SCs is challenging. Certainly, induction of myelin differentiation with cyclic adenosine monophosphate (cAMP) analogues or elevating agents, such as forskolin, is more difficult in mouse SC monocultures compared to rat SC cultures. This is because mouse SCs require additional exogenous b__-neuregulin-1 (__b__NRG1), plating on poly-L-lysine (PLL) instead of poly-D-lysine (PDL), and low concentration horse serum as opposed to foetal calf serum (Stevens et al, 1998; Arthur-Farraj et al, 2011; Päiväläinen et al, 2008)__.

      Author response: We have now explained more clearly that without plating on Matrigel and the regular addition of Matrigel to the myelination medium that mouse cocultures do not myelinate with ascorbic acid or indeed addition of NRG1 nor forskolin. Please see NEW DATA in Supplemental figure 1. We have added the following paragraph to the results section.

      Results

      Importantly, we found that L-ascorbic acid was insufficient to induce substantial myelination in our cultures, unlike in rat SC-DRG cocultures, and in the one previously published dissociated mouse SC-DRG protocol (Stevens et al, 1998)__. In fact, plating cocultures on laminin, adding ascorbic acid (50 m__g ml−1), b__NRG1 (10 ng ml−1) and forskolin (10 m__M) induced very few myelin sheaths (Supp. Fig. 1). Only when cultures were plated on Matrigel__â and further Matrigel__â was added to the myelination medium for each medium change, were we able to visualise robust reproducible myelination in our cocultures (Supp. Fig. 1).

      Author response: We have also added further discussion on how our protocol differs from the Stevens 1998 and other protocols in the discussion.

      Discussion

      Our protocol differs somewhat from the one used by Stevens et al., 1998 to induce myelination in dissociated mouse SC-DRG cocultures__, as they used ascorbic acid and 10% horse serum and presumably plated their cultures on laminin, though they do not explicitly detail this (Stevens et al, 1998)__. In our preliminary experiments we were unable to visualise much myelination with use of laminin, ascorbic acid or indeed if b__NRG1 and high concentration forskolin was added to the medium for up to four weeks. However, if we plated cocultures on Matrigel® and continuously added it to the myelination medium then we saw comparable levels of myelination in our mouse cocultures to that of rat cocultures (Eldridge et al, 1987)__. This approach of using Matrigel® to enhance myelination has previously been successfully employed in cultures of human iPSC sensory neurons with rat SCs and in non-dissociated mouse DRG explant cultures (Clark et al, 2017; Päiväläinen et al, 2008)__. Importantly, we used growth factor depleted Matrigel® as standard Matrigel® preparations contain substantial amounts of Transforming growth factor b (TGF__b__) which is a known inhibitor of myelination (Einheber et al, 1995)__. Additionally, the majority of rat and mouse coculture protocols plate cells on glass whereas we found cultures were healthier and myelinated better when cultured on plastic Alcar® coverslips.

      Discussion

      Furthermore, as this is a dissociated and compartmentalised purely mouse cell culture system, one can utilise the vast array of transgenic and knockout lines available to study neuron-SC interactions in more detail, without concern of contaminating endogenous SCs and other non-neuronal cells that remains a drawback of current mouse dissociated or non-dissociated DRG explant models.

      Reviewer 1: The figures in the paper are largely descriptive. They are very little quantitative measurement. Thus, the readers will have a hard to determine, if they replicate the proposed approach, whether their efficient is on par with the current authors.

      Author response: We have now added quantification of myelin segments per mm2, percentage of SCs that myelinate and quantification of the interperiodic distance of the myelin formed. This is all included in a new version of TABLE 1. We have discussed this data in the results section as follows.

      Results

      Quantifying the number of PRX positive myelin segments we found that there were 325.33 ± 12.3 sheaths per mm2, comparable to what has been originally described in rat SC/DRG cocultures and two fold more extensive myelination than recently described compartmentalised rat cocultures models (n=3; Table 1; Eldridge et al, 1987; Vaquié et al, 2019)__. Furthermore, 25.47 ± 1% of Schwann cells were myelinating in our cultures (n=3; Table 1). To confirm that cocultured myelinated SCs formed compact myelin we performed electron microscopy (EM), which revealed compact myelin formation with multiple myelin wraps and formation of readily visible major dense (MDL) and intraperiod lines (IPL; Fig. 2D). Additionally, we measured the periodicity, i.e., the distance between two adjacent major dense lines, to make sure myelin was compacted. Interperiodic distance was 12.16 ± 0.28 nm, in line with previous reports (n=3; Table 1; Boutary et al, 2021; Fernando et al, 2016; García-Mateo et al, 2018; Giese et al, 1992; Perrot et al, 2007)__.

      Author response: Interestingly, after we performed this analysis, we realised we have double the level of myelination in our mouse cultures (325.33 ± 12.3 per mm2) than in the compartmentalised myelinating rat cocultures in Vaquie et al., 2019 (147 ± 27 internodes per mm2 (n = 3)).

      Author response: We have also quantified the JUN upregulation after injury in both myelinating and aligned cocultures. See Fig. 3B-E).

      Results

      Additionally, we noted a strong upregulation of JUN protein in SCs 12 hours after axotomy (Fig. 3B and C). We also saw significant JUN upregulation 12 hours after axotomy in cocultures with aligned SCs (Fig. 3D and E).

      Author response: The above quantification is in addition to the quantification of the rate of axon degeneration in the presence and absence of aligned and myelinating Schwann cells in Fig. 4B. We have also quantified the % of Schwann cells that contained axonal debris after injury – this data is now quoted in the text as we removed Fig. 4E.

      We thank the reviewer for asking for additional quantification as this has improved the manuscript.

      Minor points:<br /> Fig.4B and E should show individual data points.

      Author response: We have added the individual data points to Fig.4B. We have removed Fig.4E and instead quoted the data in the results section as follows:

      Results

      When we quantified this phenomenon, we found that 97.84 ± 1.462% (n=2) of SCs in our cocultures contained mCherry-labelled axonal fragments.

      Reviewer #1 (Significance):

      In addition, to the demonstration of feasibility of this in vitro approach, the main finding by the authors is that SCs have a role in neuronal protection and support is key for peripheral nerve regeneration. Thus while in vitro approach does not add key information that do not already exists in the field, it somewhat confirms that the effect is SC autonomous. Overall the approach is interesting and has potential, but the study currently lack a demonstration of its usefulness to the community.

      It would have been interesting to have the authors discuss the advantages of their approach in comparison to other innovative approaches to study SC-axon interactions that have been developed in the last decade (i.e., 3D environment, microfluidic approach, transwell systems). There is also a lack of citations about similar studies in the field.

      Author response: We direct the reviewer to the following paragraphs in the introduction and the discussion, which we have now elaborated on further post peer review. We discuss all relevant cocultures studies in mouse as well as all the relevant microfluidic studies and 3D coculture studies as well as the one human nerve organoid study. We found two additional studies, Numata-Uematasu 2023 using DRG mouse explant cultures and Park et al., 2021 using motorneuron-SC cocultures which we have now added to the discussion. We also briefly discuss transwell studies to assess migration as the reviewer requested.

      We also discuss in detail the two microfluidic coculture injury studies Babetto et al., 2020 and Vaqiue et al., 2019 extensively throughout the manuscript. We have added further discussion of the similarities and differences between theirs and our approach in the discussion.

      Introduction.

      Protocols exist where endogenous mouse SCs are used to myelinate dissociated or non-dissociated DRG explant cultures. (Shen et al, 2014; Harty et al, 2019; Sundaram et al, 2021; Stettner et al, 2013; Numata-Uematasu et al, 2023)__. Furthermore, another protocol seeded exogenous SCs onto non-dissociated DRG explant cultures (Päiväläinen et al, 2008)__. Other laboratories seed cultured rat SCs onto dissociated mouse DRG axons (Taveggia & Bolino, 2018)__. Use of dissociated or non-dissociated DRG explants cultures preclude many experimental uses, such as using SCs from different transgenic animals and separate transfection of SCs and neurons with viruses for live imaging or genetic manipulation, and easy use of microfluidic chambers to allow injury studies and separate drug treatments to neurons or SCs. The reason for this is that antimitotics cannot be used in dissociated or non-dissociated DRG explant cultures as this depletes SCs, and the culture quickly becomes contaminated with other non-neuronal cell types, such as satellite cells and fibroblasts migrating out of the DRG. Furthermore, use of exogenous SCs in a non-dissociated DRG explant culture risks, after a period of antimitotic exposure, which was developed by Päiväläinen et al., 2008 still risks potential contamination from endogenous SCs and satellite glia migrating out of the DRG explant over time. This occurs because antimitotic treatment is unlikely to fully penetrate the whole DRG without prior dissociation. Additionally, a compartmentalised culture system cannot be readily used with non-dissociated DRG explant cultures (Päiväläinen et al, 2008)__.

      Discussion.

      Furthermore, our protocol is complementary to the recently described 3D mouse myelinating SC-motor neuron coculture system using collagen hydrogels (Hyung et al, 2021; Park et al, 2021)__. It will be interesting in the future to up titrate the concentration of Matrigel®, which is similar to collagen hydrogels, in our cultures to see whether further increasing extracellular matrix viscosity and stiffness improves our myelination efficiency even further. While it is possible to study cell migration in microfluidic cell culture devices, transwell models offer significant advantages to study this cellular phenomenon (Negro et al, 2022)__. To date, there have been no published studies of successful myelination in human SC-neuron coculture systems. Despite this rat SCs have been shown to readily myelinate human-induced pluripotent stem cell (iPSC)-derived sensory neurons and an iPSC-derived peripheral nerve organoid system which does contain myelinating SCs has recently been described (Clark et al, 2017; Van Lent et al, 2022)__.

      These findings confirm the observations of both Babetto et al., 2020 and Vaquié et al., 2019 who used rat SCs in similar microfluidic culture systems (Vaquié et al, 2019; Babetto et al, 2020)__. We have shown that the axo-protective observation seen by Babetto et al., 2020 does not rely on myelination status, which was an outstanding question from that study (Babetto et al, 2020)__. Furthermore, in an advance from previous studies, we have visualised the axo-protective and axon clearance phenomena in the same culture and shown that they are temporally separated, with axon fragmentation and debris clearance by SCs occurring at much later timepoints after axotomy. Several different culture and experimental conditions preclude direct comparison of our study with both those of Vaquié et al., 2019 and Babetto et al., 2020. These include the use of rat SCs in both prior studies and that Babetto et al., 2020 mixed rat SC with mouse DRG axons; length of time in culture, time points quantified after injury and distance from injury and site of analysis. Babetto et al., 2020 performed axotomy on relatively short term cocultures (6 days in vitro) whereas Vaiquié et al., 2019 cultured for at least 4 weeks and in our case 6 weeks prior to axotomy. Vaiquié et al., 2019 removed nerve growth factor (NGF) prior to laser axotomy and quantified proximally (though also imaged distally) whereas both our study and Babetto et al., 2020 kept NGF in the medium, performed axotomy with a scalpel and quantified more distally and in our case extremely distal, where only individual neurites and no axon bundles were visible. Vaquié et al., 2019 had SCs on both sides of the barrier in the microfluidic chambers whereas we seeded SCs only in the axonal/bottom compartment. Additionally, our cocultures had both forskolin and b__NRG1 added to help induce myelination, whereas these factors are not required in rat myelinating cocultures. Finally, it is important to permeabilise myelinated cultures with acetone after fixation__, as we did, to visualise the entire axon through heavily myelinated segments otherwise axon integrity cannot be reliably assessed in a quantitative manner (Vaquié et al, 2019; Babetto et al, 2020)__.

      Because of the lack of key novel mechanisms, and lack of discussion on what this approach is superior to others in vitro approach limits the impact of the study and the excitement of the reader, even from the SC-axon community.

      Author response: We have developed the first compartmentalised fully dissociated mouse myelinating coculture system in over twenty years. Thanks to the reviewer’s suggestions, we have now shown that myelination is comparable to the original rat cocultures from the Bunge lab, which is the gold standard in the field, and superior to recently described compartmentalised rat coculture system by Vaquie et al., 2019. We have provided a detailed step by step protocol to allow other researchers to use our technique.

      Additionally, thanks to the reviewer, we have now described in detail exactly how our protocol differs from others and why we succeeded to get mouse SCs to myelinate so robustly in a fully dissociated coculture (see previous answers). This is an incremental but important advance given that studies currently use a coculture system using entirely cells from rat, or where rat Schwann cells are seeded on mouse axons, or dissociated or non-dissociated mouse explant cultures are used which abrogates using neurons and SCs from different transgenic mice.

      As this is a methods paper, we did not intend to describe novel molecular mechanisms though our method is already being used for such purposes by ourselves and a collaborator as outlined above in prior answers. We did not make this clear and we hope the extensive revision of the manuscript now addresses this point. Despite this, we have shown that the axo-protective effect of SCs is independent of myelination status, which is an advance on what is known in the literature.

      Finally, in addition to myelination, we have demonstrated that one can study all the key components of the SC and axonal response to injury in a quantifiable way in addition to demonstrating that these cocultures can be live imaged and used for drug studies. None of the prior mouse studies looked at injury responses of axons nor SCs. We believe this will be of use to the community.

      Reviewer #2 (Evidence, reproducibility and clarity):

      In the manuscript entitled "Distinct axo-protective and axo-destructive roles for Schwann cells after injury in a novel compartmentalised mouse myelinating coculture system", Arthur-Farraj and colleagues detail a method of dissociated coculture of mouse-DRG neurons and Schwann cells in microfluidic chambers. In this system, neurons and Schwann cells harvested from the same or from different animals are grown in different compartments that are connected by microgrooves, thereby allowing for spatial and diffusive separation. Neurons are shown to extent their axons across the microgroove barrier to the glial compartment where Schwann cells align with the axons and myelination can be induced. Detailed analysis of myelination and axon injury/degradation are presented as use cases, including the capability to genetically and pharmacologically manipulate neurons and Schwann cells independently, which also enabled fluorescent life cell imaging. The authors then examine the effect of immature/premyelinating and myelinating Schwann cells on the rate of axon degeneration. Upon axotomy Schwann cells significantly delayed degeneration, with no difference between non-myelinating and myelinating Schwann cells. Finally, live imaging during axon degeneration with fluorescent proteins separately expressed in neurons and Schwann cells demonstrated that Schwann cells ingest axonal fragments.

      Major comment:<br /> In establishing a much needed in-vitro system for PNS myelination and injury research the paper represents a valuable contribution to the PNS community. However, I find the presentation of aspects concerning a protective/destructive role of Schwann cells somewhat inconclusive. That these roles exist has been known, as the authors discuss. Then what does this study contribute concerning the open question that was raised by the discrepancies between Babetto et al., 2020 and Vaquié et al., 2019, i.e. how Schwann cells contribute to axon survival/regeneration after injury? Essentially, the only significant conclusion in this regard is that myelinating and non-myelinating mouse Schwann cells do not differ in their capability to protect axons from degeneration. The manuscript, including the title, would benefit from focusing more on this aspect. In particular, the discussion of the factors that lead to the still remaining apparent discrepancies between Babetto et al., 2020 and Vaquié et al., 2019 and this study should be revised. The authors state that "The study by Vaiquié et al., 2019 quantified axon fragmentation proximally in the microgrooves at timepoints starting from 12 hours after axotomy." (Discussion). While this observation is accurate, Jacob and colleagues also show accelerated, obviously distal axon degeneration in the presence of Schwann cells (Figure 3C in Vaquié et al., 2019). It is therefore unlikely that the discrepancies stem from analysis of more proximal vs. more distal axons, or the timepoints of analyses. In my opinion, a further study (using the coculture system presented in this manuscript) that compares the role of Schwann cells from rats and mice, and that includes analysis of more proximal and distal axon degeneration as well analysis of axon regeneration is needed. In a rework of the manuscript, the authors may therefore elaborate more on the shortcomings of the present study, or alternatively soften the aims of the study in the first place.

      Author response: We thank the reviewer for their comments, and we agree that the title and the aims of the manuscript were confusing. We didn’t make it explicit that this was a methods paper, and that we didn’t intend to show entirely novel findings but instead thoroughly characterise our mouse system so that it is comparable to what has been done for rat cocultures. We have now made this clearer by removing the word novel in the title and changing the title to the following:

      A method for mouse myelinating Schwann cell-DRG neuron compartmentalised cocultures: myelination status does not influence the axo-protective effect of Schwann cells after injury.

      Author response: We have decided to focus the manuscript more on the comparison of myelinating versus non-myelinating cocultures, given that we have shown that the axo-protective effect of SCs is independent of myelination status, which is an advance on what is known in the literature. We have added further characterisation of our aligned cocultures with p75NTR immuno and EM images (Fig.1D and E). We have added the following summary of how our study relates to findings of Babetto et al., 2020 and Vaquie et al., 2019 in the discussion, in line with the reviewer’s suggestions.

      Discussion.

      These findings confirm the observations of both Babetto et al., 2020 and Vaquié et al., 2019 who used rat SCs in similar microfluidic culture systems (Vaquié et al, 2019; Babetto et al, 2020)__. We have shown that the axo-protective observation seen by Babetto et al., 2020 does not rely on myelination status, which was an outstanding question from that study (Babetto et al, 2020)__. Furthermore, in an advance from previous studies, we have visualised the axo-protective and axon clearance phenomena in the same culture and shown that they are temporally separated, with axon fragmentation and debris clearance by SCs occurring at much later timepoints after axotomy.

      Author response: We have now added more in-depth discussion of the similarities and differences between Babetto et al., 2020 and Vaquie et al., 2019 and our approach in the discussion.

      Discussion.

      Several different culture and experimental conditions preclude direct comparison of our study with both those of Vaquié et al., 2019 and Babetto et al., 2020. These include the use of rat SCs in both prior studies and that Babetto et al., 2020 mixed rat SC with mouse DRG axons; length of time in culture, time points quantified after injury and distance from injury and site of analysis. Babetto et al., 2020 performed axotomy on relatively short term cocultures (6 days in vitro) whereas Vaiquié et al., 2019 cultured for at least 4 weeks and in our case 6 weeks prior to axotomy. Vaiquié et al., 2019 removed nerve growth factor (NGF) prior to laser axotomy and quantified proximally (though also imaged distally) whereas both our study and Babetto et al., 2020 kept NGF in the medium, performed axotomy with a scalpel and quantified more distally and in our case extremely distal, where only individual neurites and no axon bundles were visible. Vaquié et al., 2019 had SCs on both sides of the barrier in the microfluidic chambers whereas we seeded SCs only in the axonal/bottom compartment. Additionally, our cocultures had both forskolin and b__NRG1 added to help induce myelination, whereas these factors are not required in rat myelinating cocultures. Finally, it is important to permeabilise myelinated cultures with acetone after fixation__, as we did, to visualise the entire axon through heavily myelinated segments otherwise axon integrity cannot be reliably assessed in a quantitative manner (Vaquié et al, 2019; Babetto et al, 2020)__.

      Author response: We would like to add that we showed Claire Jacob, senior author of the Vaquie et al., 2019 study, our manuscript prior to peer review and she offered helpful comments, which we incorporated into the manuscript, which is why she is acknowledged. We have also discussed our findings with Elisabetta Babetto as well.

      While it would be a great future study to compare both axon degeneration rates in rat and mouse cocultures this was not the original intention of our study. We believe we have included enough detail of our experimental procedures, including the distance from the barrier we imaged axon degeneration, a crucial bit of information missing from the other studies, should others want to perform a comparative study between rats and mice.

      Methods

      Five images at a distance of between 1.2-1.4mm from the microgroove barrier (the most distal part of the culture that could be imaged) were quantified per culture, taken in comparable locations in each culture.

      Author response: I would add that one of our previous studies (Arthur-Farraj et al., 2012 Neuron, Fig5I) has already looked at axon outgrowth/regeneration in dissociated non-myelinating mouse SC-DRG co-cultures. We showed the presence of Schwann cells accelerates axon regeneration/outgrowth and this relies upon Schwann cell c-JUN.

      We have now added quantification of the extent of myelination in our cocultures and it is comparable to the original Bunge lab rat cocultures and more extensive than the Vaquie et al., 2019 coculture.

      Results

      Quantifying the number of PRX positive myelin segments we found that there were 325.33 ± 12.3 sheaths per mm2, comparable to what has been originally described in rat SC/DRG cocultures and two fold more extensive myelination than recently described compartmentalised rat cocultures models (n=3; Table 1; Eldridge et al, 1987; Vaquié et al, 2019)__. Furthermore, 25.47 ± 1% of Schwann cells were myelinating in our cultures (n=3; Table 1).

      Methods

      To quantify the number of myelin segments per area, we counted the number of myelin segments for five areas per culture for three cultures and normalised this per mm2. To quantify the percentage of Schwann cells in myelinating cocultures that are actively myelinating, we quantified the number of myelin segments and the number of DAPI-positive nuclei for five areas per culture for three cultures.

      Minor comments:<br /> - Figure 2D: From the electron micrograph there is no doubt that compact myelin is formed, however to me it seems the compaction is not complete. A rough estimation with the aid of the provided scale bar resulted in an interperiodic distance of about 17 nm, which contrasts with the remarkably well reproduced values reported in multiple reports using conventional specimen preparation (like in this study), of which I am citing just a few: about 13 nm in rat ex vivo nerve (Peterson and Pease, J. Ultrastruct Res 1972; Fledrich et al., Nat Commun 2018), 12 nm (Giese et al., Cell 1992), 12.2 nm (Perot et al., J Neurosci 2007), about 12 (Fernando et al., Nat Commun 2016), about 13 nm (García-Mateo et al., Glia 2017) or about 13 nm in mouse ex vivo (Boutary et al., Commun Biol 2021), which was also reproduced with about 13 nm in rat in vitro (Taveggia and Bolino, Methods Mol Biol 2018). The authors should acknowledge this deviation and might discuss possible reasons.

      Author response: We have now provided a more representative EM image of our myelination (Fig. 2D). Additionally, thanks to the reviewer’s comments we have now quantified the interperiodic distance and find it is comparable to the studies the reviewer suggested. We have added the data to the new Table 1 and added the references the reviewer advised. Please see the additions to the methods and the results section regarding this new data below.

      Results

      To confirm that cocultured myelinated SCs formed compact myelin we performed electron microscopy (EM), which revealed compact myelin formation with multiple myelin wraps and formation of readily visible major dense (MDL) and intraperiod lines (IPL; Fig. 2D). Additionally, we measured the periodicity, i.e., the distance between two adjacent major dense lines, to make sure myelin was compacted. Interperiodic distance was 12.16 ± 0.28 nm, in line with previous reports (n=3; Table 1; Boutary et al, 2021; Fernando et al, 2016; García-Mateo et al, 2018; Giese et al, 1992; Perrot et al, 2007)__.

      Methods

      To measure interperiodic distance, we measured at least 10 periods per myelinated fibre for at least three fibres per sample for three separate samples.

      • Figure 4A,B: The result of a slowed axon degeneration in coculture relies on the accurate assessment of continuity of the NFL staining. While the authors report that acetone permeabilization was necessary to afford complete penetration of the used antibodies in myelinating cultures, I cannot see why the authors have not used the same staining protocol for all cultures, as it is detailed in the method section. While I consider it unlikely that the staining conditions have led to an apparent delay of degeneration in coculture, experiments should generally be performed under identical conditions, unless there are good reasons not to do so. If this is not the case, it will be reassuring to see the same effect when identical staining conditions are employed. On the same note, do the compared cultures have the same age, i.e. have the neuron monocultures been in vitro for the same time as the cocultures?

      Author response: We thank the reviewer for picking up this inaccuracy in the manuscript. We can confirm that for the purposes of the axon degeneration experiment all cultures were stained using exactly the same staining protocol. Additionally, we were very careful to maintain all cultures for exactly the same time in culture – 6 weeks. Additionally, axon only cultures were maintained in myelination medium to make sure medium constituents were not responsible for the observed differences in degeneration rate. We have added the following elaboration to the methods section to clarify these points.

      Methods

      Axon only cultures related to Figure 1 were permeabilised in PBS + 0.5% Triton (Merck) + 5% HS + 5% donkey serum (DS, Merck - D9663) at RT for 1 hour. For the purposes of quantifying the rate of axon degeneration (Figure 4) both axon only cultures and cocultures with SCs were permeabilised in 50% Acetone for 2 minutes, 100% Acetone for 2 minutes, 50% Acetone for 2 minutes (all at RT), and then blocked in PBS + 0.5% Triton + 5% HS + 5% DS at RT for 1 hour.

      Methods

      All cultures (axon only, aligned SCs and myelinating SCs) were cultured for 6 weeks prior to axotomy experiments. To minimise the possibility that medium constituents were responsible for differences in axon degeneration rates, axonal compartments of axon only cultures were cultured in medium containing 10 ng ml-1 b__NRG1 and 10 m__M forskolin (axon only medium, extended methods section D6) once SCs were seeded on other cultures, and then switched into myelination medium (additional Matrigel__â and 50 m__g ml−1 L-Ascorbic Acid), 24 hours before axotomy. Bottom compartments of aligned SC cultures, 24 hours before axotomy, were switched into DRG/SC medium containing 10 ng ml-1 b__NRG-1, 10 m__M forskolin and 50 m__g ml−1 L-Ascorbic Acid, which is insufficient to induce myelination in mouse cultures. Bottom compartments of myelinating cocultures were medium changed into fresh myelination medium (Extended methods section D7) 24 hours prior to axotomy.

      • In several instances of the manuscript, the term "transfection" is used to refer to lentiviral gene transfer. I advise to use the more appropriate term "transduction" instead
      • I could not seem to find a meaningful reference to the microfluidic chambers that were used in the study. The protocol should contain details on the device and source of supply in order to enable potential readers to execute the protocol

      Author response: We thank the reviewer for this comment. We have replaced transfection with transduction throughout the manuscript. Please see the track changes manuscript for all instances.

      Reviewer #2 (Significance):

      The paper presents a convincing establishment of a dissociated coculture derived exclusively from mouse that leads to robust myelination. As the manuscript correctly states, Schwann cell culture and especially coculture with neurons has been experienced difficult in the field, and by providing a detailed protocol as well as demonstrating how the coculture system can be used to address important questions of PNS myelination and repair, the paper fills an important gap. However, the experiments directed to the role of Schwann cells in axon degeneration do not clarify much, which should be better addressed in the discussion and also by modifying the title accordingly.<br /> The paper will be of high value for basic researchers that are interested in performing studies addressing cellular and molecular mechanisms of myelination and repair in the PNS. Importantly, the paper can pave the way to usage of transgenic or knockout mouse models in coculture. Thereby it might spark interest also in those researchers that use transgenic and knockout mouse models and who have so far refrained from using coculture models.

      Field of expertise of the reviewer: Cellular and molecular mechanisms of myelination and growth signaling in the PNS; in-depth experience with DRG coculture models from rats and mice

      Author response: We thank the reviewer for their kind comments. We have now modified the title, aims and discussion of the manuscript in line with the reviewer’s suggestions.

      Reviewer #3 (Evidence, reproducibility and clarity):

      SUMMARY<br /> The authors present a detailed protocol for co-cultures of mouse DRGs with mouse SCs using microfluidics. In this model, cells of interest grow in different compartments while allowing for axons to grow in between, thereby making them accessible to injury induction. Using this experimental system, the authors show that myelination occurs, myelin gets compacted and acquires nodal organization. The authors then show that such a system allows for compartment-specific lentivirus transduction and live imaging. Next, they perform physical and chemical axonal injury and show that at early time point pos- injury the presence of SCs protects from axonal degeneration regardless of the myelination status, and helps with clearing of damaged axons at later time points.<br /> Major comments:

      The novelty of the study is questionable.<br /> While the model is well described and appears to be useful for the proposed applications (live imaging, transduction, injury model), the arguments provided regarding its novelty are not fully convincing. The main argument from the authors of this paper is that there are no established protocols describing the use of mouse SCs with mouse DRG neurons in dissociated myelinating cocultures. However, this appears to be inaccurate, as the model described in Stevens et al., 1998 (cited in the paper) uses mouse DRG neurons dissected at E13.5 with mouse SCs dissected at P3 to study myelination. Also, in Päiväläinen et al., 2008, mouse DRGs and SCs are cultured from transgenic mice at different developmental ages, thereby arguing that coculture models have been previously successfully implemented. The main difference appears to be rather the compartmentalization of SCs and DRGs which appears to be a mouse adaptation of the rat model described by Vaquie et al,2019. Based on the above, it seems imperative for the authors to tone down the novelty aspect and provide a more thorough discussion on how the current novel differs from protocols in published study, highlighting advantages and caveats for each.

      Author response: We agree with the reviewer that we did not make the case clear enough for how our coculture model adds to what is currently described in the literature. We have now changed the title and removed the word novel. New title:

      A method for mouse myelinating Schwann cell-DRG neuron compartmentalised cocultures: myelination status does not influence the axo-protective effect of Schwann cells after injury.

      Author response: We have now added the following paragraph to the introduction:

      Introduction.

      Indeed there has only ever been one laboratory detailing convincing myelin formation in dissociated mouse myelinating SC-DRG neuron cocultures, however this was never published as a step by step detailed protocol (Stevens et al, 1998; Stevens & Fields, 2000)__. In the last twenty years, there have been no published studies demonstrating myelination in fully dissociated mouse SC-mouse DRG cocultures. This has largely prevented the use of cells, particularly SCs, from transgenic mice in cocultures and thus restricted the ability to study SC-axon interactions in a system that can be readily manipulated and live imaged and results directly applied back to in vivo findings in the same species.

      Author response: Regarding the study by Päiväläinen et al, 2008_,_ they did not fully dissociate their DRGs (see Fig.1 which demonstrates a DRG explant) and thus it is a non-dissociated DRG explant model. While they demonstrated convincing myelination due to the use of Matrigel which we acknowledge them for, their model is not perfectly suited for the use of neurons and SCs from different transgenic animals as the use of a DRG explant, even with temporary use of an antimitotic, risks contamination by endogenous SCs and satellite glia over time, especially as their model is not compartmentalised. We discuss the caveats of their protocol and those using dissociated mouse explant cocultures in a revised paragraph in the introduction.

      Introduction.

      Protocols exist where endogenous mouse SCs are used to myelinate dissociated or non-dissociated DRG explant cultures. (Shen et al, 2014; Harty et al, 2019; Sundaram et al, 2021; Stettner et al, 2013; Numata-Uematasu et al, 2023)__. Furthermore, another protocol seeded exogenous SCs onto non-dissociated DRG explant cultures (Päiväläinen et al, 2008)__. Other laboratories seed cultured rat SCs onto dissociated mouse DRG axons (Taveggia & Bolino, 2018)__. Use of dissociated or non-dissociated DRG explants cultures precludes many experimental uses, such as using SCs from different transgenic animals and separate transfection of SCs and neurons with viruses for live imaging or genetic manipulation, and easy use of microfluidic chambers to allow injury studies and separate drug treatments to neurons or SCs. The reason for this is that antimitotics cannot be used in dissociated or non-dissociated DRG explant cultures as this depletes SCs, and the culture quickly becomes contaminated with other non-neuronal cell types, such as satellite cells and fibroblasts migrating out of the DRG. Furthermore, use of exogenous SCs in a non-dissociated DRG explant culture risks, after a period of antimitotic exposure, which was developed by Päiväläinen et al., 2008 still risks potential contamination from endogenous SCs and satellite glia migrating out of the DRG explant over time. This occurs because antimitotic treatment is unlikely to fully penetrate the whole DRG without prior dissociation. Additionally, a compartmentalised culture system cannot be readily used with non-dissociated DRG explant cultures (Päiväläinen et al, 2008)__.

      Author response: We have also added further discussion on how our protocol differs from the Stevens 1998 and other protocols in the discussion.

      Discussion

      Our protocol differs somewhat from the one used by Stevens et al., 1998 to induce myelination in dissociated mouse SC-DRG cocultures__, as they used ascorbic acid and 10% horse serum and presumably plated their cultures on laminin, though they do not explicitly detail this (Stevens et al, 1998)__. In our preliminary experiments we were unable to visualise much myelination with use of laminin, ascorbic acid or indeed if b__NRG1 and high concentration forskolin was added to the medium for up to four weeks. However, if we plated cocultures on Matrigel® and continuously added it to the myelination medium then we saw comparable levels of myelination in our mouse cocultures to that of rat cocultures (Eldridge et al, 1987)__. This approach of using Matrigel® to enhance myelination has previously been successfully employed in cultures of human iPSC sensory neurons with rat SCs and in in non-dissociated mouse DRG explant cultures (Clark et al, 2017; Päiväläinen et al, 2008)__. Importantly, we used growth factor depleted Matrigel® as standard Matrigel® preparations contain substantial amounts of Transforming growth factor b (TGF b__) which is a known inhibitor of myelination (Einheber et al, 1995)__. Additionally, the majority of rat and mouse coculture protocols plate cells on glass whereas we found cultures were healthier and myelinated better when cultured on plastic Alcar® coverslips.

      Author response: We have added further discussion of comparable models in the literature in the discussion.

      Discussion.

      Furthermore, our protocol is complementary to the recently described 3D mouse myelinating SC-motor neuron coculture system using collagen hydrogels (Hyung et al, 2021; Park et al, 2021)__. It will be interesting in the future to up titrate the concentration of Matrigel®, which is similar to collagen hydrogels, in our cultures to see whether further increasing extracellular matrix viscosity and stiffness improves our myelination efficiency even further. While it is possible to study cell migration in microfluidic cell culture devices, transwell models offer significant advantages to study this cellular phenomenon (Negro et al, 2022)__. To date, there have been no published studies of successful myelination in human SC-neuron coculture systems. Despite this rat SCs have been shown to readily myelinate human-induced pluripotent stem cell (iPSC)-derived sensory neurons and an iPSC-derived peripheral nerve organoid system which does contain myelinating SCs has recently been described (Clark et al, 2017; Van Lent et al, 2022)__.

      Next, the authors emphasize the conflicting results of two articles, Babetto et al., 2020 and Vaquie et al., 2019, as the basis to use their newly developed model in the same species and testing two ages corresponding to distinct myelination states. However, both studies reach the same conclusion as the current study, that SCs have a protective role, although at two different developmental time points. As such, it is likely that multiple mechanisms may account for the protective effect of SC on axonal damage, and therefore the different studies do not seem conflicting but rather complementary. Yet, it is interesting that this manuscript shows that the myelination status of SCs does not impact their ability to slow down degeneration and yet it confirms that different timing after injury elicits different behaviors in SCs, as suggested by the studies of Babetto et al., 2020 and Vaquie et al., 2019. In other words, a more accurate description of the results of these two studies is needed and a better explanation of what the authors consider to be conflicting and why (there could be more differences than species and myelination, for instance, such as the method used for axotomy - laser vs cut with scalpel which tear and pull membranes).

      Author response: We would like to humbly correct the reviewer that the studies by Babetto et al., 2020 and Vaquie et al., 2019 do not reach the same conclusion that Schwann cells have a protective role. Instead, they describe axon protection (Babetto et al., 2020) and axon fragmentation (Vaquie et al., 2019). Our studies now visualise both phenomena in the same culture system. We have now made this point more explicit as well as highlighted the one conceptual advance our methods paper makes on the current literature, which is that myelination status does not influence the SC axo-protection, as the reviewer suggested.

      Discussion.

      These findings confirm the observations of both Babetto et al., 2020 and Vaquié et al., 2019 who used rat SCs in similar microfluidic culture systems (Vaquié et al, 2019; Babetto et al, 2020)__. We have shown that the axo-protective observation seen by Babetto et al., 2020 does not rely on myelination status, which was an outstanding question from that study (Babetto et al, 2020)__. Furthermore, in an advance from previous studies, we have visualised the axo-protective and axon clearance phenomena in the same culture and shown that they are temporally separated, with axon fragmentation and debris clearance by SCs occurring at much later timepoints after axotomy.

      Author response: We have now added more in-depth discussion of the similarities and differences between Babetto et al., 2020 and Vaquie et al., 2019 and our approach in the discussion.

      Discussion.

      Several different culture and experimental conditions preclude direct comparison of our study with both those of Vaquié et al., 2019 and Babetto et al., 2020. These include the use of rat SCs in both prior studies and that Babetto et al., 2020 mixed rat SC with mouse DRG axons; length of time in culture, time points quantified after injury and distance from injury and site of analysis. Babetto et al., 2020 performed axotomy on relatively short term cocultures (6 days in vitro) whereas Vaiquié et al., 2019 cultured for at least 4 weeks and in our case 6 weeks prior to axotomy. Vaiquié et al., 2019 removed nerve growth factor (NGF) prior to laser axotomy and quantified proximally (though also imaged distally) whereas both our study and Babetto et al., 2020 kept NGF in the medium, performed axotomy with a scalpel and quantified more distally and in our case extremely distal, where only individual neurites and no axon bundles were visible. Vaquié et al., 2019 had SCs on both sides of the barrier in the microfluidic chambers whereas we seeded SCs only in the axonal/bottom compartment. Additionally, our cocultures had both forskolin and b__NRG1 added to help induce myelination, whereas these factors are not required in rat myelinating cocultures. Finally, it is important to permeabilise myelinated cultures with acetone after fixation__, as we did, to visualise the entire axon through heavily myelinated segments otherwise axon integrity cannot be reliably assessed in a quantitative manner (Vaquié et al, 2019; Babetto et al, 2020)__.

      Author response: We would like to add that we showed Claire Jacob, senior author of the Vaquie et al., 2019 study, our manuscript prior to peer review and she offered helpful comments, which we incorporated into the manuscript, which is why she is acknowledged. We have also discussed our findings with Elisabetta Babetto as well.

      Overall, the title does not appear to be the most appropriate because the content rather proposes a detailed protocol and gives examples of applications, rather than focusing on the protective versus destructive role of SCs on axons. It also appears to be misleading, as "axo-destructive" appears to suggest a negative role of Schwann cells on axons, whereas SC are rather helpful in clearing degenerative axons, a step which facilitates regeneration.

      Author response: We have now changed the title and the focus of the manuscript in line with the reviewer’s comments. New title:

      A method for mouse myelinating Schwann cell-DRG neuron compartmentalised cocultures: myelination status does not influence the axo-protective effect of Schwann cells after injury.

      Author response: We have removed the phrase axo-destructive throughout the manuscript and instead referred to axon fragmentation and axon debris clearance roles of SCs in line with the reviewer’s suggestion. Please see track changes manuscript for all instances where this was modified.

      The number of biological replicates for each experiment is not always indicated, and if the "n=" represent cultures prepared independently/passaged or wells/cell. It is essential to be rigorous and clearly indicate the number of technical replicates and biological samples throughout the manuscript and provide a thorough description of them. One example is Fig. 4 E were only 10 cells from a single culture appeared to have been imaged. Is this accurate? This aspect is essential to evaluate reproducibility, especially in view of the technical and biological variability.

      Author response: We have now added quantification of myelin segments per mm2, percentage of SCs that myelinate and quantification of the interperiodic distance of the myelin formed. This is all included in a new version of TABLE 1. We have discussed this data in the results section as follows.

      Quantifying the number of PRX positive myelin segments we found that there were 325.33 ± 12.3 sheaths per mm2, comparable to what has been originally described in rat SC/DRG cocultures and two fold more extensive myelination than recently described compartmentalised rat cocultures models (n=3; Table 1; Eldridge et al, 1987; Vaquié et al, 2019)__. Furthermore, 25.47 ± 1% of Schwann cells were myelinating in our cultures (n=3; Table 1). To confirm that cocultured myelinated SCs formed compact myelin we performed electron microscopy (EM), which revealed compact myelin formation with multiple myelin wraps and formation of readily visible major dense (MDL) and intraperiod lines (IPL; Fig. 2D). Additionally, we measured the periodicity, i.e., the distance between two adjacent major dense lines, to make sure myelin was compacted. Interperiodic distance was 12.16 ± 0.28 nm, in line with previous reports (n=3; Table 1; Boutary et al, 2021; Fernando et al, 2016; García-Mateo et al, 2018; Giese et al, 1992; Perrot et al, 2007)__.

      Author response: We have discussed the number of cultures used for each quantification in the methods section. See below.

      Quantification of Myelination in cocultures

      To quantify the number of myelin segments per area, we counted the number of myelin segments for five areas per culture for three cultures and normalised this per mm2. To quantify the percentage of Schwann cells in myelinating cocultures that are actively myelinating, we quantified the number of myelin segments and the number of DAPI-positive nuclei for five areas per culture for three independently prepared cultures. To measure interperiodic distance, we measured at least 10 periods per myelinated fibre for at least three fibres per sample for three separate samples.

      Quantification of Degeneration

      Five images at a distance of between 1.2-1.4mm from the microgroove barrier (the most distal part of the culture that could be imaged) were quantified per culture, taken in comparable locations in each culture. A line was drawn across each image, and each axon crossing this line was either scored as degenerated or intact. Images were blinded prior to quantification. A minimum of three independently prepared cultures were assessed per timepoint for each condition.

      Author response: We have removed Fig.4E and instead quoted the data in the results section as follows:

      When we quantified this phenomenon, we found that 97.84 ± 1.462% (n=2) of SCs in our cocultures contained mCherry-labelled axonal fragments.

      Author response: We apologise as the n number for this experiment was 2 (not 10), with cells in 10 areas quantified throughout all imaging timepoints from each independently prepared culture. We have included the following description in the methods section:

      To quantify number of SCs with fragments, each cell was defined as a region of interest and checked for the presence of mCherry positive fragments at all timepoints. Two separate independently prepared cultures and cells in 10 areas per culture were analysed.

      Author response: Additionally for Fig. 4B we have now included individual data points from independently prepared cultures.

      N numbers are included in all figure legends and always refers to independently prepare cultures/biological replicates.

      We have added to the relevant figure legends (Fig.3 and 4 and Table 1) the phrase:

      n number refers to independently prepared cultures from separate litters of mice.

      Minor comments:

      • Does myelination reach axons in the microgrooves (it seems to from 2C, but up to where)? Where is axotomy performed and are axons myelinated where the cut was performed?

      Author response: Myelination occasionally reaches the beginning of the microgrooves. We didn’t visualise myelination in the DRG cell body compartment. We have added the following detail to the methods section:

      Traumatic axotomies were carried out by carefully removing the microfluidic chamber (SND150 and RND150, Xona Microfluidics__Ò__) from the Aclar__â coverslip using sterile forceps and severing axons with a surgical blade under a light microscope. Axotomies were carried out at the level of the microgroove barrier. To confirm all axons were severed, a second higher cut was performed and axons between the cut sites removed using the surgical blade.

      Author response: Given this, we cannot exclude that the odd proximal myelin segment is cut, but the vast majority of axons are not myelinated at the site of cut (lower cut).

      • Since the model allows for comparison of aligned vs myelinating SCs, and that both aligned and myelinating SCs seem to slow down degeneration, and that c-JUN is upregulated after in vivo injury, have the authors measured if c-JUN levels increase similarly in both myelinating vs aligned SCs?

      Author response: We thank the reviewer for this suggestion. We have now quantified the JUN upregulation after injury in both myelinating and aligned cocultures as well as adding images of JUN upregulation in aligned cocultures. See Fig. 3B-E).

      Additionally, we noted a strong upregulation of JUN protein in SCs 12 hours after axotomy (Fig. 3B and C). We also saw significant JUN upregulation 12 hours after axotomy in cocultures with aligned SCs (Fig. 3D and E).

      Author response: We have decided to focus the manuscript more on the comparison of myelinating versus non-myelinating cocultures, given that we have shown that the axo-protective effect of SCs is independent of myelination status, which is an advance on what is known in the literature. In addition to changing the title, as we have mentioned previously, we have added further characterisation of our aligned cocultures with p75NTR immuno and EM images (Fig.1D and E).

      We have

      • On clarity:<br /> - In the step-by-step protocol, wording needs to be improved.

      Author response: We have substantially edited the step-by-step protocol. Please see track changes document for all specific changes in wording.

      • Temperatures for centrifugations are missing.

      Author response: We have added temperatures for all centrifugation steps. Please see track changes document

      • The MOI described for lentivirus is 2-10 in the protocol but 200 in the legend of Figure 3F.

      Author response: The MOI for DRGs was 2-10 and SCs was 200 in Figure 3F. This is described similarly in the extended methods section. DRGs are transduced much more easily than SCs.

      We have added the following sentence to the results section to emphasise this point:

      Importantly dissociated mouse SCs required a much higher multiplicity of infection (MOI) than dissociated mouse DRGs (see extended methods section).

      • Certain citations in the references list are incomplete (i.e. Babetto et al.; Catenaccio et al.,).

      Author response: We have updated the reference list.

      Reviewer #3 (Significance):

      The advance for the field proposed by this paper is mostly technical, as it details a new model to be used by the field, of mouse SCs-mouse DRGs in dissociating myelinating cultures. The tested applications allowed the authors to also confirm a protective role for SCs on axonal damage, which was independent from myelination status.

      Being a method paper, it is essential that the authors provide clear statements on the number of biological replicates, and technical repeats, as well as a very thorough and accurate description of the methodology.

      The model described has similarities with existing models in the field such as Stevens et al., 1998 and Vaquié et al., 2019. To place it in context in a more helpful way, the authors should emphasize on the novelty brought by their protocol compared to existing models. The authors compare their findings to results from Vaquié et al., 2019 and Babetto et al., 2020 that they describe as conflicting, when it seems they rather address different mechanisms of SCs in protection and repair, occurring at different time points.

      Audience might be interested in the detailed step by step protocol to use this in vitro model for the applications described, and investigate further why SCs myelination status does not influence their ability to protect from neurodegeneration early on or how to make use of this for neuroprotection studies.

      Author response: We have now rephrased the description of Vaquié et al., 2019 and Babetto et al., 2020 studies in line with the reviewer’s suggestions. We have now added further discussion of our model in the context of all other models in the field as we have outlined in detail in above responses.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary

      The authors present a detailed protocol for co-cultures of mouse DRGs with mouse SCs using microfluidics. In this model, cells of interest grow in different compartments while allowing for axons to grow in between, thereby making them accessible to injury induction. Using this experimental system, the authors show that myelination occurs, myelin gets compacted and acquires nodal organization. The authors then show that such a system allows for compartment-specific lentivirus transduction and live imaging. Next, they perform physical and chemical axonal injury and show that at early time point pos- injury the presence of SCs protects from axonal degeneration regardless of the myelination status, and helps with clearing of damaged axons at later time points.

      Major comments:

      The novelty of the study is questionable.<br /> While the model is well described and appears to be useful for the proposed applications (live imaging, transduction, injury model), the arguments provided regarding its novelty are not fully convincing. The main argument from the authors of this paper is that there are no established protocols describing the use of mouse SCs with mouse DRG neurons in dissociated myelinating cocultures. However, this appears to be inaccurate, as the model described in Stevens et al., 1998 (cited in the paper) uses mouse DRG neurons dissected at E13.5 with mouse SCs dissected at P3 to study myelination. Also, in Päiväläinen et al., 2008, mouse DRGs and SCs are cultured from transgenic mice at different developmental ages, thereby arguing that coculture models have been previously successfully implemented. The main difference appears to be rather the compartmentalization of SCs and DRGs which appears to be a mouse adaptation of the rat model described by Vaquie et al,2019. Based on the above, it seems imperative for the authors to tone down the novelty aspect and provide a more thorough discussion on how the current novel differs from protocols in published study, highlighting advantages and caveats for each.

      Next, the authors emphasize the conflicting results of two articles, Babetto et al., 2020 and Vaquie et al., 2019, as the basis to use their newly developed model in the same species and testing two ages corresponding to distinct myelination states. However, both studies reach the same conclusion as the current study, that SCs have a protective role, although at two different developmental time points. As such, it is likely that multiple mechanisms may account for the protective effect of SC on axonal damage, and therefore the different studies do not seem conflicting but rather complementary. Yet, it is interesting that this manuscript shows that the myelination status of SCs does not impact their ability to slow down degeneration and yet it confirms that different timing after injury elicits different behaviors in SCs, as suggested by the studies of Babetto et al., 2020 and Vaquie et al., 2019. In other words, a more accurate description of the results of these two studies is needed and a better explanation of what the authors consider to be conflicting and why (there could be more differences than species and myelination, for instance, such as the method used for axotomy - laser vs cut with scalpel which tear and pull membranes).

      Overall, the title does not appear to be the most appropriate because the content rather proposes a detailed protocol and gives examples of applications, rather than focusing on the protective versus destructive role of SCs on axons. It also appears to be misleading, as "axo-destructive" appears to suggest a negative role of Schwann cells on axons, whereas SC are rather helpful in clearing degenerative axons, a step which facilitates regeneration.

      The number of biological replicates for each experiment is not always indicated, and if the "n=" represent cultures prepared independently/passaged or wells/cell. It is essential to be rigorous and clearly indicate the number of technical replicates and biological samples throughout the manuscript and provide a thorough description of them. One example is Fig. 4 E were only 10 cells from a single culture appeared to have been imaged. Is this accurate? This aspect is essential to evaluate reproducibility, especially in view of the technical and biological variability.

      Minor comments:

      • Does myelination reach axons in the microgrooves (it seems to from 2C, but up to where)? Where is axotomy performed and are axons myelinated where the cut was performed?
      • Since the model allows for comparison of aligned vs myelinating SCs, and that both aligned and myelinating SCs seem to slow down degeneration, and that c-JUN is upregulated after in vivo injury, have the authors measured if c-JUN levels increase similarly in both myelinating vs aligned SCs?
      • On clarity:
      • In the step-by-step protocol, wording needs to be improved.
      • Temperatures for centrifugations are missing.
      • The MOI described for lentivirus is 2-10 in the protocol but 200 in the legend of Figure 3F.
      • Certain citations in the references list are incomplete (i.e. Babetto et al.; Catenaccio et al.,).

      Significance

      The advance for the field proposed by this paper is mostly technical, as it details a new model to be used by the field, of mouse SCs-mouse DRGs in dissociating myelinating cultures. The tested applications allowed the authors to also confirm a protective role for SCs on axonal damage, which was independent from myelination status.

      Being a method paper, it is essential that the authors provide clear statements on the number of biological replicates, and technical repeats, as well as a very thorough and accurate description of the methodology.

      The model described has similarities with existing models in the field such as Stevens et al., 1998 and Vaquié et al., 2019. To place it in context in a more helpful way, the authors should emphasize on the novelty brought by their protocol compared to existing models. The authors compare their findings to results from Vaquié et al., 2019 and Babetto et al., 2020 that they describe as conflicting, when it seems they rather address different mechanisms of SCs in protection and repair, occurring at different time points.

      Audience might be interested in the detailed step by step protocol to use this in vitro model for the applications described, and investigate further why SCs myelination status does not influence their ability to protect from neurodegeneration early on or how to make use of this for neuroprotection studies.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      In the manuscript entitled "Distinct axo-protective and axo-destructive roles for Schwann cells after injury in a novel compartmentalised mouse myelinating coculture system", Arthur-Farraj and colleagues detail a method of dissociated coculture of mouse-DRG neurons and Schwann cells in microfluidic chambers. In this system, neurons and Schwann cells harvested from the same or from different animals are grown in different compartments that are connected by microgrooves, thereby allowing for spatial and diffusive separation. Neurons are shown to extent their axons across the microgroove barrier to the glial compartment where Schwann cells align with the axons and myelination can be induced. Detailed analysis of myelination and axon injury/degradation are presented as use cases, including the capability to genetically and pharmacologically manipulate neurons and Schwann cells independently, which also enabled fluorescent life cell imaging. The authors then examine the effect of immature/premyelinating and myelinating Schwann cells on the rate of axon degeneration. Upon axotomy Schwann cells significantly delayed degeneration, with no difference between non-myelinating and myelinating Schwann cells. Finally, live imaging during axon degeneration with fluorescent proteins separately expressed in neurons and Schwann cells demonstrated that Schwann cells ingest axonal fragments.

      Major comment:

      In establishing a much needed in-vitro system for PNS myelination and injury research the paper represents a valuable contribution to the PNS community. However, I find the presentation of aspects concerning a protective/destructive role of Schwann cells somewhat inconclusive. That these roles exist has been known, as the authors discuss. Then what does this study contribute concerning the open question that was raised by the discrepancies between Babetto et al., 2020 and Vaquié et al., 2019, i.e. how Schwann cells contribute to axon survival/regeneration after injury? Essentially, the only significant conclusion in this regard is that myelinating and non-myelinating mouse Schwann cells do not differ in their capability to protect axons from degeneration. The manuscript, including the title, would benefit from focusing more on this aspect. In particular, the discussion of the factors that lead to the still remaining apparent discrepancies between Babetto et al., 2020 and Vaquié et al., 2019 and this study should be revised. The authors state that "The study by Vaiquié et al., 2019 quantified axon fragmentation proximally in the microgrooves at timepoints starting from 12 hours after axotomy." (Discussion). While this observation is accurate, Jacob and colleagues also show accelerated, obviously distal axon degeneration in the presence of Schwann cells (Figure 3C in Vaquié et al., 2019). It is therefore unlikely that the discrepancies stem from analysis of more proximal vs. more distal axons, or the timepoints of analyses. In my opinion, a further study (using the coculture system presented in this manuscript) that compares the role of Schwann cells from rats and mice, and that includes analysis of more proximal and distal axon degeneration as well analysis of axon regeneration is needed. In a rework of the manuscript, the authors may therefore elaborate more on the shortcomings of the present study, or alternatively soften the aims of the study in the first place.

      Minor comments:

      • Figure 2D: From the electron micrograph there is no doubt that compact myelin is formed, however to me it seems the compaction is not complete. A rough estimation with the aid of the provided scale bar resulted in an interperiodic distance of about 17 nm, which contrasts with the remarkably well reproduced values reported in multiple reports using conventional specimen preparation (like in this study), of which I am citing just a few: about 13 nm in rat ex vivo nerve (Peterson and Pease, J. Ultrastruct Res 1972; Fledrich et al., Nat Commun 2018), 12 nm (Giese et al., Cell 1992), 12.2 nm (Perot et al., J Neurosci 2007), about 12 (Fernando et al., Nat Commun 2016), about 13 nm (García-Mateo et al., Glia 2017) or about 13 nm in mouse ex vivo (Boutary et al., Commun Biol 2021), which was also reproduced with about 13 nm in rat in vitro (Taveggia and Bolino, Methods Mol Biol 2018). The authors should acknowledge this deviation and might discuss possible reasons.
      • Figure 4A,B: The result of a slowed axon degeneration in coculture relies on the accurate assessment of continuity of the NFL staining. While the authors report that acetone permeabilization was necessary to afford complete penetration of the used antibodies in myelinating cultures, I cannot see why the authors have not used the same staining protocol for all cultures, as it is detailed in the method section. While I consider it unlikely that the staining conditions have led to an apparent delay of degeneration in coculture, experiments should generally be performed under identical conditions, unless there are good reasons not to do so. If this is not the case, it will be reassuring to see the same effect when identical staining conditions are employed. On the same note, do the compared cultures have the same age, i.e. have the neuron monocultures been in vitro for the same time as the cocultures?
      • In several instances of the manuscript, the term "transfection" is used to refer to lentiviral gene transfer. I advise to use the more appropriate term "transduction" instead
      • I could not seem to find a meaningful reference to the microfluidic chambers that were used in the study. The protocol should contain details on the device and source of supply in order to enable potential readers to execute the protocol

      Significance

      The paper presents a convincing establishment of a dissociated coculture derived exclusively from mouse that leads to robust myelination. As the manuscript correctly states, Schwann cell culture and especially coculture with neurons has been experienced difficult in the field, and by providing a detailed protocol as well as demonstrating how the coculture system can be used to address important questions of PNS myelination and repair, the paper fills an important gap. However, the experiments directed to the role of Schwann cells in axon degeneration do not clarify much, which should be better addressed in the discussion and also by modifying the title accordingly.

      The paper will be of high value for basic researchers that are interested in performing studies addressing cellular and molecular mechanisms of myelination and repair in the PNS. Importantly, the paper can pave the way to usage of transgenic or knockout mouse models in coculture. Thereby it might spark interest also in those researchers that use transgenic and knockout mouse models and who have so far refrained from using coculture models.

      Field of expertise of the reviewer: Cellular and molecular mechanisms of myelination and growth signaling in the PNS; in-depth experience with DRG coculture models from rats and mice

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary:

      In this original article from Mutscher et al., the authors developed a compartmentalized dissociated mouse myelinating SC-DRG coculture system to investigate the distinct roles of Schwann cells in axon protection and degeneration after injury. The innovation of this approach relies on (i) the use of mouse SCs and mouse DRGS neurons instead of rat cells; (ii) the use of microfluidic chambers, seeded by axons and SCs in different compartment; (iii) the possibility to perform a traumatic injury in vitro. While this novel approach offers new ways to study peripheral nerve regeneration and SC-axon interaction, and technical the study is robust, the paper is currently limited by the exploration of their model.

      Major points:

      It is unclear is this approach will ever lead to the identification of key mechanism or key candidates. This is a major miss in the current manuscript form. In short: the authors should demonstrate that their in vitro system can lead to significant leap in our understanding of peripheral nerve regeneration by identifying novel targets/pathways or mechanisms.

      The use of embryonic DRG neurons or SC isolated from P2 animals are arguably physiologically not the same cells that are affected by traumatic nerve injury, which happen most often than not in adult. This is a problem in the long-term reliance on this approach to study axotomy peripheral nerve regeneration.

      Moderate points:

      "there are no established protocols in the field describing the use of mouse SCs with mouse DRG neurons in dissociated myelinating cocultures". The use of mouse cells is laudable, but it is not necessarily a technical innovation, or at least the current manuscript does not explain why their approach particularly suitable to mouse Schwann cells.

      The figures in the paper are largely descriptive. They are very little quantitative measurement. Thus, the readers will have a hard to determine, if they replicate the proposed approach, whether their efficient is on par with the current authors.

      Minor points:

      Fig.4B and E should show individual data points.

      Significance

      In addition, to the demonstration of feasibility of this in vitro approach, the main finding by the authors is that SCs have a role in neuronal protection and support is key for peripheral nerve regeneration. Thus while in vitro approach does not add key information that do not already exists in the field, it somewhat confirms that the effect is SC autonomous. Overall the approach is interesting and has potential, but the study currently lack a demonstration of its usefulness to the community.

      It would have been interesting to have the authors discuss the advantages of their approach in comparison to other innovative approaches to study SC-axon interactions that have been developed in the last decade (i.e., 3D environment, microfluidic approach, transwell systems). There is also a lack of citations about similar studies in the field.

      Because of the lack of key novel mechanisms, and lack of discussion on what this approach is superior to others in vitro approach limits the impact of the study and the excitement of the reader, even from the SC-axon community.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity):

      In this manuscript the author is presenting a deep-learning model used to predict the development stage of zebrafish embryo. A robust method that can accurately classify a zebrafish into different development stages is highly relevant for many researchers working with zebrafish and hence the importance in developing methods like this is high.

      The manuscript is overall ok. However, more data is needed to convince the reader that the method is robust enough to work with other samples in other labs. This would greatly improve the impact of the publication.

      We agree with the reviewer and have included in our revised manuscripts additional test data that was acquired at a different laboratory to the training data (Figures 5 - 7).

      Page 6.<br /> - How is the data acquired?

      Images used to do initial model training are the same as those used in a previous study - the details of image acquisition are contained in the relevant publication (doi: 10.12688/wellcomeopenres.18313.1). However, we have now added “Zebrafish Husbandry” and “Live Imaging” for newly-acquired images. We have added a table (Table 1) listing details of all image data used in the study.

      Page 8.<br /> "This indicates that whileKimmelNet can be used successfully with noisier test data than that on which it was trained,there is an upper limit to how noisy the data can be."<br /> - This is an obvious statement there will always be an upper limit on noise.

      We agree with the reviewer that this statement is not terribly informative. This section (“KimmelNet’s prediction accuracy is not significantly impacted by moderate levels of additive noise”) has been removed from the revised manuscript in favour of incorporating a section detailing testing of the model on newly-acquired images (“KimmelNet can generalise to previously unseen data”).

      Page 9.<br /> - Are only wildtype embryos used? How would this work on different mutants. To evaluate the robustness of the method this it would be valuable to test on some mutant line with known developmental difference from the wild type.

      We agree with the reviewer that testing on a mutant line would lend more weight to our findings. For example, the p53-/- zebrafish has a reported, published developmental delay, but we did not have access to that line. However, the developmental delay reported for the p53-/- mutant is virtually indistinguishable from that effected by a temperature change. We therefore focussed our efforts on developmental delay affected by altering incubation temperature only.

      Image data.<br /> - I would strongly suggest that the author should include a description of the data in the manuscript. A description of how the data is acquired, microscope, different batches, type of embryos used.

      The image data used in the first draft of the manuscript is the same as that used in a previous publication (Jones et al. 2022), which contains all the relevant details the reviewer seeks. However, we have now added the relevant information for the newly-acquired image data.

      "Random160translation in the y-direction was avoided due to the aspect ratio of the images (width>161height) - any artifacts introduced by translation in the x-direction would be removed by the162centre crop layer, but this would not be the case for translation in the y-direction."<br /> - Could this be solved by using some border method reflection, repetition or fixed value?

      The reviewer is correct that some form of image reflection or repetition could be utilised. However, given the nature of our images, if an embryo is located close to the image boundary, reflection/repetition can result in some odd artefacts, so we minimised the use of such fill methods (also used by the random zoom augmentation layer). We could instead use an arbitrary fixed value, as the reviewer suggested, but finding a value suitable for all images is difficult.

      Page 10.<br /> Addition of Noise to Image Data<br /> - This should be added in the training phase. This would probably improve the robustness of the network and also improve the results on the test data.

      We agree with the reviewer and have now added a random Gaussian noise layer for data augmentation purposes during model training (see Figure 1).

      • Supplementary 3 images with high noise. It is worrying that the network is not able to handle the noise in this figure. Looks like the features that is used to distinguish the developmental stage of the embryo is still clearly seen with this high noise level? Retrain the model with noise as an augmentation to improve this.

      As the reviewer suggested, addition of random noise is now incorporated into model training. The new version of the manuscript does not include the same supplemental figures, but instead includes additional testing on newly-acquired data instead.

      Reviewer #1 (Significance):

      The development of methods like this is highly relevant in the zebrafish community. Staging and evaluating the developmental stage for zebrafish is common and is of interest to the broad community. A lot of this work today is done manually, limiting the throughput, and adding human bias.

      The limit of this study is the dataset used for training and evaluation. Firstly, it is not clear about the structure of the data and how it is acquired, different types of fish or imaging setup etc. For a method to be useful to the community it needs to be robust enough to handle different types of fish (transgenic lines). The manuscript would be greatly improved by adding this to the training and evaluation.

      We have now added additional datasets for the purposes of evaluating the model.

      My expertise is image analysis and machine learning for quantification of biological samples, with focus on zebrafish screening.

      Reviewer #2 (Evidence, reproducibility and clarity):

      Summary<br /> The paper "Automated staging of zebrafish embryos with KimmelNet" by Barry et al., presents a method to automatically stage developmental timepoints of zebrafish embryos based on convolutional neural networks (CNN). The authors show that a CNN trained on ~20k images can determine time post fertilization on test-image sets with an accuracy on the range of a few hours. This technique undoubtedly has the potential to become very useful for any zebrafish researchers interested in developmental timing as it eases analysis and removes potential subjective bias.

      Major comments<br /> In its current form the paper lacks sufficient graph annotations and method descriptions. This makes it hard in places to judge the validity of the claims. I do believe that the presented method can be useful and is likely valid but to be convincing, the authors need to spend more time expanding the methods, justifying their choices of analysis and clarifying figure annotations.

      We believe that we have addressed the reviewer’s concerns in this revised manuscript, as detailed in response to the specific points below.

      Specific points:<br /> 1) The annotation of the training data is not described and specifically it is unclear how valid the staging of the training data itself is. The authors state in the introduction "the hours post fertilization (hpf) [...] provides only and approximation of the actual developmental stage". It is therefore critical to know how this was accounted for in the annotation of the training data. Since the quality of the training data will ultimately limit the best-case quality of Kimmel Net. The authors need to go into some detail here even though the training data appears to be from another published dataset.

      The reviewer raises a valid point – two individual zebrafish embryos that are x hours post-fertilisation are not necessarily at the same developmental stage. However, we believe it is reasonable to assume that two populations of embryos x hours post-fertilisation are, on average, at the same developmental stage and it is this assumption that forms the basis for our approach. Given the inherent variability the reviewer refers to, we are not suggesting that our model would be particularly accurate for staging individual embryos. However, we are very confident (and we believe the data in the manuscript supports this) that given a population of embryos, our model will provide an accurate rate of development. We have added a paragraph (lines 131-141) to address this point.

      2) Why were "test predictions fit to a straight line through the origin". On the one hand this makes sense (since the slope would indicate the correspondence) but it should be clarified why an intercept was omitted in the fit. After all it is unclear if Kimmel net correctly identifies 0Hpf embryos.

      The reviewer makes a valid point – we do not know what predictions KimmelNet would produce for images of embryos closer to 0 hpf. However, we felt an equation of the form y=mx was a reasonable choice for two reasons. First of all, it matches the form of the Kimmel equation, which, despite its flaws, we are using as a benchmark of sorts – the absence of a y intercept makes comparisons with the Kimmel equation straightforward. Secondly, a “perfect” model would produce a straight line fit with y=x, so the lack of a y intercept seemed a reasonable constraint to impose. We have added some brief text (lines 103-105) to clarify this choice.

      3) The methods do not list how the mean of the absolute error was calculated from 3B/C. I think this should be the mean of the absolute error (not the mean of the error) but in that case the numbers listed in the text appear rather small given the histograms in 3 B/C. A clear statement in the methods would clarify this issue.

      We have now added a “Statistical Analysis” section under Materials & Methods to detail exactly what was used to calculate the values given for error analysis. We have calculated the mean of the error, not the mean of the absolute error, as we wish to illustrate that the mean is close to zero. We have used the standard deviation of the errors to illustrate that there is a significant spread in the error values, as depicted in Figure 3C and D.

      Minor comments<br /> 1) The Y-axis in Figure 2B is simply labeled "Loss" - what is the unit of this loss? HPF? Or HPF^2? This is important for judging the quality of the fit

      We thank the reviewer for drawing our attention to this omission. The loss is hpf2 (mean squared error) and we have updated the plot to reflect this.

      2) Figure 3 B. I would suggest changing the labels of the confidence intervals in the legend. "Inner and outer" is already clear from the figure itself, so labels that state that these are derived from n=100 vs. n=20 test image sized samples would be more useful to the reader

      We thank the reviewer for this suggestion – we have updated the figure legend accordingly.

      Referees cross-commenting

      I concur with comments issued by the other reviewers. I think it will be especially important to address the comments related to testing the method on mutants (Reviewer #1) and training the model in the presence of noise to increase its robustness (Reviewers #1 and #3) as well as addressing the overall annotation/generation of the training data (Reviewers #1 and #2).

      We believe we have now addressed all of these concerns. The model has been retrained with additional data augmentation incorporating random noise, tested on newly-acquired data and we have added tables summarising the details of all image data used in this study.

      I think these points will be critical to make the paper useful by increasing transparency and ensuring reproducibility in other labs with different imaging conditions, strains, mutants, etc.

      Reviewer #2 (Significance):

      Developmental delay is a common occurrence that can be caused by genetic and environmental background effects. It is therefore highly desirable to properly quantify this variable. The work presented here makes an important step in this direction, by allowing to quantify developmental timepoints independent of subjective staging. This speeds up analysis, increases reproducibility and enhances rigor. However, as my comments above indicate, the significance also depends on the ability of potential users to judge the quality of the work. Once those issues have been addressed, I think the work will be of broad interest to the developmental biology community, first and foremost labs utilizing the zebrafish model. However, as the authors state, the presented model architecture could be trained with the data from other species as well.

      Expertise: Zebrafish, quantitative analysis, behavior, neuroscience

      We thank the reviewer for their positive feedback.

      Reviewer #3 (Evidence, reproducibility and clarity):

      Summary:

      Properly staging embryos of zebrafish embryos is important, yet provides challenging since it can depend on many factors, such as temperature, water quality, fish population density, etc. Here, the authors provide a deep-learning-based model called KimmelNet that allows the prediction of the age of zebrafish embryos, using 2D brightfield images. The technique is robust to weak measurement noise and can also be used to identify developmental delays from a very small number of experimental data.

      The code is accessible to the reader, open-source and should be useable by experimentalists without huge effort.

      Major comments:

      I suggest retraining the model and application of the model to additional data for the following reasons:<br /> • Why did the authors not train for (high) measurement noise and heterogeneous background illumination? Would that not make the model more robust? In principle, creating training should not be considerably harder than before, since the manipulation of the images has been already shown in the manuscript and the authors just need to run it again on the HPC cluster. If there are no technical or administrative constraints (access to the cluster, computational effort, high costs, etc.), the authors should retrain their model.

      We thank the reviewer for this suggestion. As detailed in Figure 1, with a view to making the model more robust, we have now added several more layers of data augmentation, including the addition of random noise, and retrained our model.

      • For Fig. S2 and S3 it is not clear if there is such a strong deviation from the Kimmel equation due to measurement noise or due to the background illumination. The saliency maps appear as if they are mainly affected by the illumination, and maybe less by the noise. Would it be possible to apply the model to a case without artificial noise, but with heterogeneous background illumination to identify what has a bigger impact?

      We thank the reviewer for this suggestion. We have now replaced the “artificial” examples used in the previous version of the manuscript with newly-acquired data (Figure 5), which exhibits different characteristics to that used for training.

      Additionally, the authors need to clarify what exactly they are comparing in this manuscript and rework their interpretation of their findings:<br /> • When comparing the predictions between KimmelNet and the Kimmel equation, the authors use an equation of the form y=mx. Could the authors please elaborate on why they introduce the constraint of y(0)=0? It might be naturally given by the so-called Kimmel equation, but by looking at Fig 3a, it seems like an equation of the form y=mx+a would be more appropriate and it appears like KimmelNet introduces an offset of around a=2h for 25 Celsius. The authors need to discuss this.

      The main rationale for using an equation of the form y=mx is to be consistent with the Kimmel equation (see lines 103-105). The reviewer is correct that an equation of the form y=mx+c may well produce a better fit to the data, but omitting a y intercept makes comparison with the Kimmel Equation trivial.

      • In lines 5-8 the authors say that developmental stage progression does not only depend on temperature, but also on population density, water quality etc. and they explain that usually not only hpf, but also staging guides based on morphological criteria are used! If that is true, how good is their training data set that only uses hpf and not the other important guides? How did the authors test that these factors have no impact on their training data?

      We have now added a paragraph (lines 131-141) to address this point.

      Since this tool has the potential to have a big impact on zebrafish research, it would be nice to provide some examples of how exactly this could be achieved:<br /> • Could the authors discuss how exactly their tool is useful to experimentalists? Is it the idea that if an experimentalist wants to investigate an embryo of a particular stage, they apply KimmelNet to images of embryos to identify the stage of the embryo and only then undertake their planned experiment? Is that a realistic undertaking?

      As is evidenced by the errors presented in Figure 3C & D, testing KimmelNet on individual images of embryos may well result in a large error in the predicted hpf. As such, it is not appropriate to use the tool in such a manner. However, to modify the example provided by the reviewer, should an experimentalist have a population of embryos they wished to stage, then yes, KimmelNet would certainly be an appropriate tool for doing so.

      • Would it be possible to provide a tutorial (or even video tutorial if such skills are available in the group of authors) that provides real examples of how to apply the technique? This would make it easier for people without advanced Python/Deep-Learning skills to use the tool, hence improving the impact of KimmelNet.

      A lack of user-friendly interfaces for applying deep learning methods in biology is well-documented – basic knowledge of python and tools like jupyter notebooks are often necessary (https://doi.org/10.1038/s41592-023-01900-4). However, we have endeavoured to make the running of KimmelNet as easy as possible for new users. A jupyter notebook instance can be run on Binder with absolutely no set-up required. To run KimmelNet on their own data, biologists just need to download the Git repo and replace the test images with their own data. We have updated the landing page on the GitHub repo to provide more specific step-by-step instructions for each of these tasks. We will also endeavour to upload our model to the BioImage Model Zoo (https://bioimage.io/#/) to further increase accessibility.

      I am very critical towards equation 1. Please note that I don't think this has any impact on the quality of the technique provided in this manuscript and the significant flaws can already be found in Kimmel 1995 (which is not under review here). That is why I suggest rewriting of this manuscript to not support an over-interpretation of this equation.<br /> • I do not think that the Kimmel equation is an established term. At least a Google Scholar Search for "Kimmel equation" only gives one result: the preprint of this manuscript.<br /> • The equation has no mathematical meaning regarding its units (subtracting temperature and a unitless value). I also very rarely see equations with Degrees Celsius and not Kelvin.<br /> • Additionally, the equation provides a linear relationship between the development time and temperature h(T) and in Kimmel et al, it is shown that this is only true for 25-33 Celsius. Such a linearisation is not very surprising for a small temperature range, but I am not sure how true it is for higher temperature differences. Hence, I feel that it is very bold to give a specific name to such an equation, giving it an importance that it does not deserve.

      We appreciate the reviewer’s concerns and have removed explicit references to “The Kimmel Equation”, without substantively changing the content of the manuscript.

      Minor comments:

      • For the measurement noise cases it would be nice to have some example images of fish with the specific noise levels in Fig S1 and Fig S2.

      We have now removed the “synthetic” additive noise test data, previously depicted in Figures S1-3, in favour of newly-acquired images in Figures 5-7.

      • Could the authors hypothesize why they predict a slower dynamic for 25 Celsius than predicted by the Kimmel equation?

      Referring to Figure 2 in Kimmel et al (1995), it is apparent that the straight lines are by no means perfect fits to the datapoints. In Fig 2A in particular, some datapoints for the 25C data fall well below the line fit. While the published equation suggests a rate of development 80.5% of the rate at 28.5C, according to Fig 2A, an alternative line fit could give a developmental rate as low as 70-75%, which would be in agreement with our data.

      Reviewer #3 (Significance):

      Strengths of the study:

      An easy-to-use method to automatically stage zebrafish embryos and identify differences in the developmental stage is very important for the zebrafish community and the technique in this manuscript definitely novel. The tool is can be used without large effort and the authors suggest that it can also find applications beyond zebrafish embryos. Hence, it is not only interesting to the zebrafish community, but to a broader developmental biology audience.

      Weakness of the study:<br /> The main weakness of the manuscript is in the training data used for the deep-learning model and the apparent large impact of heterogeneous background illumination. If that is not solved, it is unclear if this technique will find an application in the zebrafish community.

      We believe this weakness has now been addressed by incorporating additional data augmentation measures in the training process and testing the model on newly-acquired data.

      Field of expertise of the reviewer: Image Analysis, Mathematical Modelling, Biological Physics. While I have limited experience in deep learning, I cannot evaluate the specific details of the network architecture. I also have no experience in zebrafish research.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary:

      Properly staging embryos of zebrafish embryos is important, yet provides challenging since it can depend on many factors, such as temperature, water quality, fish population density, etc. Here, the authors provide a deep-learning-based model called KimmelNet that allows the prediction of the age of zebrafish embryos, using 2D brightfield images. The technique is robust to weak measurement noise and can also be used to identify developmental delays from a very small number of experimental data.

      The code is accessible to the reader, open-source and should be useable by experimentalists without huge effort.

      Major comments:

      I suggest retraining the model and application of the model to additional data for the following reasons:<br /> - Why did the authors not train for (high) measurement noise and heterogeneous background illumination? Would that not make the model more robust? In principle, creating training should not be considerably harder than before, since the manipulation of the images has been already shown in the manuscript and the authors just need to run it again on the HPC cluster. If there are no technical or administrative constraints (access to the cluster, computational effort, high costs, etc.), the authors should retrain their model.<br /> - For Fig. S2 and S3 it is not clear if there is such a strong deviation from the Kimmel equation due to measurement noise or due to the background illumination. The saliency maps appear as if they are mainly affected by the illumination, and maybe less by the noise. Would it be possible to apply the model to a case without artificial noise, but with heterogeneous background illumination to identify what has a bigger impact?

      Additionally, the authors need to clarify what exactly they are comparing in this manuscript and rework their interpretation of their findings:<br /> - When comparing the predictions between KimmelNet and the Kimmel equation, the authors use an equation of the form y=mx. Could the authors please elaborate on why they introduce the constraint of y(0)=0? It might be naturally given by the so-called Kimmel equation, but by looking at Fig 3a, it seems like an equation of the form y=mx+a would be more appropriate and it appears like KimmelNet introduces an offset of around a=2h for 25 Celsius. The authors need to discuss this.<br /> - In lines 5-8 the authors say that developmental stage progression does not only depend on temperature, but also on population density, water quality etc. and they explain that usually not only hpf, but also staging guides based on morphological criteria are used! If that is true, how good is their training data set that only uses hpf and not the other important guides? How did the authors test that these factors have no impact on their training data?

      Since this tool has the potential to have a big impact on zebrafish research, it would be nice to provide some examples of how exactly this could be achieved:<br /> - Could the authors discuss how exactly their tool is useful to experimentalists? Is it the idea that if an experimentalist wants to investigate an embryo of a particular stage, they apply KimmelNet to images of embryos to identify the stage of the embryo and only then undertake their planned experiment? Is that a realistic undertaking?<br /> - Would it be possible to provide a tutorial (or even video tutorial if such skills are available in the group of authors) that provides real examples of how to apply the technique? This would make it easier for people without advanced Python/Deep-Learning skills to use the tool, hence improving the impact of KimmelNet.

      I am very critical towards equation 1. Please note that I don't think this has any impact on the quality of the technique provided in this manuscript and the significant flaws can already be found in Kimmel 1995 (which is not under review here). That is why I suggest rewriting of this manuscript to not support an over-interpretation of this equation.<br /> - I do not think that the Kimmel equation is an established term. At least a Google Scholar Search for "Kimmel equation" only gives one result: the preprint of this manuscript.<br /> - The equation has no mathematical meaning regarding its units (subtracting temperature and a unitless value). I also very rarely see equations with Degrees Celsius and not Kelvin.<br /> - Additionally, the equation provides a linear relationship between the development time and temperature h(T) and in Kimmel et al, it is shown that this is only true for 25-33 Celsius. Such a linearisation is not very surprising for a small temperature range, but I am not sure how true it is for higher temperature differences. Hence, I feel that it is very bold to give a specific name to such an equation, giving it an importance that it does not deserve.

      Minor comments:

      • For the measurement noise cases it would be nice to have some example images of fish with the specific noise levels in Fig S1 and Fig S2.
      • Could the authors hypothesize why they predict a slower dynamic for 25 Celsius than predicted by the Kimmel equation?

      Significance

      Strengths of the study:

      An easy-to-use method to automatically stage zebrafish embryos and identify differences in the developmental stage is very important for the zebrafish community and the technique in this manuscript definitely novel. The tool is can be used without large effort and the authors suggest that it can also find applications beyond zebrafish embryos. Hence, it is not only interesting to the zebrafish community, but to a broader developmental biology audience.

      Weakness of the study:

      The main weakness of the manuscript is in the training data used for the deep-learning model and the apparent large impact of heterogeneous background illumination. If that is not solved, it is unclear if this technique will find an application in the zebrafish community.

      Field of expertise of the reviewer:

      Image Analysis, Mathematical Modelling, Biological Physics. While I have limited experience in deep learning, I cannot evaluate the specific details of the network architecture. I also have no experience in zebrafish research.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary

      The paper "Automated staging of zebrafish embryos with KimmelNet" by Barry et al., presents a method to automatically stage developmental timepoints of zebrafish embryos based on convolutional neural networks (CNN). The authors show that a CNN trained on ~20k images can determine time post fertilization on test-image sets with an accuracy on the range of a few hours. This technique undoubtedly has the potential to become very useful for any zebrafish researchers interested in developmental timing as it eases analysis and removes potential subjective bias.

      Major comments

      In its current form the paper lacks sufficient graph annotations and method descriptions. This makes it hard in places to judge the validity of the claims. I do believe that the presented method can be useful and is likely valid but to be convincing, the authors need to spend more time expanding the methods, justifying their choices of analysis and clarifying figure annotations.

      Specific points:

      1. The annotation of the training data is not described and specifically it is unclear how valid the staging of the training data itself is. The authors state in the introduction "the hours post fertilization (hpf) [...] provides only and approximation of the actual developmental stage". It is therefore critical to know how this was accounted for in the annotation of the training data. Since the quality of the training data will ultimately limit the best-case quality of Kimmel Net. The authors need to go into some detail here even though the training data appears to be from another published dataset.
      2. Why were "test predictions fit to a straight line through the origin". On the one hand this makes sense (since the slope would indicate the correspondence) but it should be clarified why an intercept was omitted in the fit. After all it is unclear if Kimmel net correctly identifies 0Hpf embryos.
      3. The methods do not list how the mean of the absolute error was calculated from 3B/C. I think this should be the mean of the absolute error (not the mean of the error) but in that case the numbers listed in the text appear rather small given the histograms in 3 B/C. A clear statement in the methods would clarify this issue.

      Minor comments

      1. The Y-axis in Figure 2B is simply labeled "Loss" - what is the unit of this loss? HPF? Or HPF^2? This is important for judging the quality of the fit
      2. Figure 3 B. I would suggest changing the labels of the confidence intervals in the legend. "Inner and outer" is already clear from the figure itself, so labels that state that these are derived from n=100 vs. n=20 test image sized samples would be more useful to the reader

      Referees cross-commenting

      I concur with comments issued by the other reviewers. I think it will be especially important to address the comments related to testing the method on mutants (Reviewer #1) and training the model in the presence of noise to increase its robustness (Reviewers #1 and #3) as well as addressing the overall annotation/generation of the training data (Reviewers #1 and #2).

      I think these points will be critical to make the paper useful by increasing transparency and ensuring reproducibility in other labs with different imaging conditions, strains, mutants, etc.

      Significance

      Developmental delay is a common occurrence that can be caused by genetic and environmental background effects. It is therefore highly desirable to properly quantify this variable. The work presented here makes an important step in this direction, by allowing to quantify developmental timepoints independent of subjective staging. This speeds up analysis, increases reproducibility and enhances rigor. However, as my comments above indicate, the significance also depends on the ability of potential users to judge the quality of the work. Once those issues have been addressed, I think the work will be of broad interest to the developmental biology community, first and foremost labs utilizing the zebrafish model. However, as the authors state, the presented model architecture could be trained with the data from other species as well.

      Expertise: Zebrafish, quantitative analysis, behavior, neuroscience

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      In this manuscript the author is presenting a deep-learning model used to predict the development stage of zebrafish embryo. A robust method that can accurately classify a zebrafish into different development stages is highly relevant for many researchers working with zebrafish and hence the importance in developing methods like this is high.

      The manuscript is overall ok. However, more data is needed to convince the reader that the method is robust enough to work with other samples in other labs. This would greatly improve the impact of the publication.

      Page 6.<br /> - How is the data acquired?

      Page 8.<br /> "This indicates that whileKimmelNet can be used successfully with noisier test data than that on which it was trained,there is an upper limit to how noisy the data can be."<br /> - This is an obvious statement there will always be an upper limit on noise.

      Page 9.<br /> - Are only wildtype embryos used? How would this work on different mutants. To evaluate the robustness of the method this it would be valuable to test on some mutant line with known developmental difference from the wild type.

      Image data.<br /> - I would strongly suggest that the author should include a description of the data in the manuscript. A description of how the data is acquired, microscope, different batches, type of embryos used.

      "Random160translation in the y-direction was avoided due to the aspect ratio of the images (width>161height) - any artifacts introduced by translation in the x-direction would be removed by the162centre crop layer, but this would not be the case for translation in the y-direction."<br /> - Could this be solved by using some border method reflection, repetition or fixed value?

      Page 10.<br /> Addition of Noise to Image Data<br /> - This should be added in the training phase. This would probably improve the robustness of the network and also improve the results on the test data.

      • Supplementary 3 images with high noise. It is worrying that the network is not able to handle the noise in this figure. Looks like the features that is used to distinguish the developmental stage of the embryo is still clearly seen with this high noise level? Retrain the model with noise as an augmentation to improve this.

      Significance

      The development of methods like this is highly relevant in the zebrafish community. Staging and evaluating the developmental stage for zebrafish is common and is of interest to the broad community. A lot of this work today is done manually, limiting the throughput, and adding human bias.

      The limit of this study is the dataset used for training and evaluation. Firstly, it is not clear about the structure of the data and how it is acquired, different types of fish or imaging setup etc. For a method to be useful to the community it needs to be robust enough to handle different types of fish (transgenic lines). The manuscript would be greatly improved by adding this to the training and evaluation.

      My expertise is image analysis and machine learning for quantification of biological samples, with focus on zebrafish screening.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      1. General Statements

      The authors greatly appreciate Review Commons’ innovative approach to scientific review and publishing. We thank the reviewers for their kind words regarding the manuscript overall quality and for highlighting the quantitative approach and reproducibility of this work. We further thank the concerns raised and suggestions made that have contributed to improving the manuscript. Below is a point-by-point response to the reviewers, organized into sections that discriminate the alterations already made and plans for further experiments and revisions. We hope that they appropriately address the reviewers' concerns.

      2. Description of the planned revisions

      Reviewer #2: Figure 6 in particular, the number of analyzed embryos is small, given the fact that there is a lot of inter-individual heterogeneity in this process it could well be that the authors got, by chance, two embryos out of three having the same pattern of Hairy1 expression.

      R: The authors appreciate the concern raised by Reviewer#2. This experiment is very time consuming and difficult to execute, which is why the number of samples is limited. Overall, we analyzed 7 embryos and 5 recapitulated the pattern of gene expression. We were particularly interested in the occipital somites and, in this time window, 3 out of 4 showed the same expression pattern. Nevertheless, further experiments will be performed to increase the number of analyzed individuals. We are confident this will contribute to strengthening the conclusions of our work.

      Reviewer #2: I believe that an additional shorter time point (+15 or 30 min) with a different pattern of the oscillatory gene would also add to the characterization of the dynamics (same for Fig 5c). This is particularly true given that the domain of expression of Hairy 1 analyzed in Figure 6 is localized quite rostral which might be interpreted as a phase 1 or a phase 2 as well (as initially described in Palmeirim et al 1997).

      R: We thank Reviewer#2 for this suggestion. We have preliminary data (20-40 min) evidencing different patterns of expression and will perform more experiments to complement these results. Figures will be modified to include samples incubated for shorter time intervals, to evidence different expression patterns obtained in these conditions.

      3. Description of the revisions that have already been incorporated in the transferred manuscript

      Reviewer #1 | Major points:

      1. In the RESULTS session, "Occipital somites are formed faster than cervical and trunk somites," the authors argue that the occipital somites form with greater temporal variability than the neck and trunk somites. Judging from Figures 3C and 3D, I feel it is the case. However, the authors should demonstrate it through statistical analysis.

      2. In the RESULTS session, "Anterior-posterior length of rostral somites," the authors concluded that "the large variability in measurements of somites 17-20 most probably results from the rotation of the embryo body in these developmental stages." Probably they mentioned data of the length of #17-#20 somites in Table1. They should demonstrate it through statistical analysis to show the large variability in the specific area. I understand that embryo rotation could be a reason for the variability. The authors should show evidence. Or they should discuss various possibilities from a broad perspective.

      R: The authors thank Reviewer#1 for suggesting a statistical approach to better characterize the data variability obtained. We performed a Brown-Forsythe test and found that, indeed, there is a statistically significant difference between the temporal variability (period) of somites 1-7 and 8-20 (P-value = 0.02319). Application of the Brown-Forsythe test also found a non-equal variability in the length of the somites #17-20 (P-value = 0.005403). A new Supplementary Figure 4 was added, displaying these results.

      The Brown-Forsythe test is a statistical method used to assess the equality of variances in a dataset across different groups, in this case, the period or the length of early and late somites. It is a robust alternative to the traditional Levene's test, particularly useful when the assumption of homogeneity of variances is not met, such as when the data distributions are skewed or contain outliers (which is the case with our data). It calculates the absolute deviations of individual observations from their respective group medians, which makes it less sensitive than the Levene's test to extreme values. By comparing these deviations between groups, the Brown-Forsythe test helps determine whether the variance differs significantly across the groups. We are confident that this result confers robustness to our claims, and hope that it appropriately addresses the reviewer's concerns.

      Regarding the reasons underlying the variability in the length of #17-#20 somites, we believe it is mainly due to technical constraints. In early developmental stages the chicken embryo is flat, and measurements are easily performed along the anterior-posterior (A-P) somite axis. When somites 17-20 are formed, the embryonic axis starts undergoing rotation, meaning that in some cases we may be measuring along a rotated somite axis. In our work somite length is determined as soon as the posterior intersomitic cleft is formed, so an alternative explanation could be that each somite is formed with a variable length, that is soon after consolidated, resulting in the characteristic consistent metameric organization of somites along the embryo body axis. This is highly unlikely because the length of somites 17-20 long after they are formed (Herrmann et al., 1951) is within the same value range we observed.

      The manuscript has been altered to include the above-mentioned information, as follows:

      • Methods section, under Statistical analysis (Line 153): “To assess the homogeneity of variances between early and late somites, we applied the Brown-Forsythe test on both the period and length measurements. This method involves computing the absolute deviations of individual observations from their respective group medians, rendering it less sensitive to extreme values (outliers). Through the comparison of these deviations across groups, the Brown-Forsythe test aids in determining the statistical significance of variance disparities.”
      • Results section, under Anterior-posterior length of rostral somites (Line 201): “A larger variability was obtained for measurements of somites 17-20 (Supplementary Figure 4A), although this most probably results from the rotation of the embryo body in these developmental stages, hindering precise length measurements of the somite A-P axis.”
      • Results section, under Occipital somites are formed faster than cervical and trunk somites (Line 216): “Remarkably, there is substantial variability in the time of formation of the early-most somites (Figure 3C; Supplementary Figures 3), which gradually stabilizes until somite 8 onwards, where both somite formation time and variability is equivalent to that observed for somites 15-20 (Figure 3C, D; Table 1; Supplementary Figure 4B).”
      1. The authors do not describe the expression patterns of hairy1 in the PSM in the manuscript, but they merely judged whether they are different or the same (recapitulate). The description of the expression pattern needs to be revised totally. The authors should describe the expression patterns of hairy1 in the PSM of each sample carefully and in detail. Fortunately, the previous report (Pourquie and Tam, Developmental Cell, 1, 619-620, 2001) categorized the expression patterns of the EC genes into three phases. The authors should at least categorize each sample according to the criterion by Pourque and Tam. If arrows of brackets indicate the area of expression, it is reader-friendly.

      R: A thorough characterization of segmentation clock gene expression (including hairy1) in the PSM of early somitogenesis chick embryos has been previously described (Rodrigues et al, 2006). For this reason, the authors focused mainly on a comparative analysis of the hairy1 expression patterns obtained in explants incubated for different periods of time. The authors acknowledge, however, the need for further description of the expression patterns obtained in early gastrulation stages, which haven’t been previously documented. Overall, the following alterations were made to the manuscript text:

      Line 231: “The regions with greater variability of hairy1 expression included the neural plate, anterior to the node, the epiblast posterior to the node encompassing the precursors of the paraxial mesoderm (Psychoyos & Stern, 1996) and the caudal-most epiblast. hairy2 expression was also very dynamic along the embryo A-P axis (n=20) (Figure 4B), evidencing chevron-like expression domains, that appear at different levels of the primitive streak, as previously described by Jouve and collaborators (Jouve et al., 2002).”

      Line 245: “As somitogenesis takes place, hairy1 and hairy2 expression patterns retain their dynamic properties in the PSM (Figure 5A, B), as previously described (Rodrigues et al., 2006).”

      The authors further thank Reviewer#1’s suggestion to include brackets indicating the areas of hairy1 expression. Figures 4, 5 and 6 have been altered accordingly, which, indeed, makes figure interpretation more reader-friendly. The gene expression phases presented in Palmeirim et al (1997) and then by Pourquié and Tam (2001) mean to summarize a dynamic expression, with continuous intermediate phases, making it difficult to clearly categorize each pattern obtained. Since our purpose was to evaluate if the entire expression pattern was recapitulated (irrespective of the specific phase), we believe that categorizing each sample in phases is not paramount for result interpretation.

      Reviewer #1 | Minor points.

      1. In the RESULTS session, "Anterior-posterior length of rostral somites," the authors described that the average length of somites ranges from 118 - 191 μ__m. But according to Table-1, the lowest length of somite is 115.92 μ__m (__∼__116 μ__m). So, the lower limit should be corrected here.

      R: The authors thank the reviewer for pointing this out. The text has been appropriately modified in Line 199 of the revised manuscript.

      1. In the DISCUSSION section, the authors mentioned that they presented a thorough characterization of the size and time of formation of the first ten somites in the chicken embryo. But based on the tables and figures, it will only be the first nine somites, not the ten.

      R: The authors agree with the reviewer’s comment. The text has been appropriately modified in Lines 128, 176 and 279 of the revised manuscript.

      1. In the GRAPHICAL ABSTRACT, if the color of the oscillation line is the same as the corresponding somites, it is intuitive.

      R: The authors thank the reviewer for this suggestion and have modified the graphical abstract accordingly. We employed multiple shades of the same color (corresponding to different positions along the body axis) to represent different embryos.

      1. It would be helpful if the manuscript contained both page and line numbers.

      R: Page and line numbers were added to the revised manuscript.

      Reviewer #2

      Minor comment: In 5 c similar patterns and newly formed somites should be pointed out by arrows on the figure to help the readers.

      R: We thank Reviewer#2 for this suggestion. Arrows and brackets have been added to the figure to highlight the newly formed somites and gene expression domains, respectively.

      Minor comment: To be more specific the term segmentation clock should be used instead of embryonic clock as I believe there are other embryonic clocks (cell cycle, circadian, etc..)

      R: The authors appreciate the suggestion of Reviewer#2 regarding the term used to identify the molecular oscillator in our work. The term “segmentation clock” or “somitogenesis clock” is commonly used to refer to oscillations in hairy1/2 gene expression because their discovery and subsequent study has mainly focused on the somitogenesis process. Oscillations of hairy1/2 expression (Hes1/7 in mouse), however, have also been described in cells and developmental stages that are not associated with somite formation, and herein we describe dynamic expression in epiblast regions containing precursors that don’t give rise to segmented structures. As discussed in our recent paper (Carraco et al., Front. Cell Dev. Biol, 2022), the broader term Embryo Clock may be used to refer to molecular oscillations in embryonic cells, controlled by negative feedback regulation, that play a role in temporally controlled morphogenetic processes and/or cell fate specification.

      In the beginning of our manuscript (Line 75), we clearly state that we are referring to the embryo clock operating during somitogenesis: “(…) somitogenesis embryo clock (EC), comprising genes with cell-autonomous oscillatory expression in the PSM driven by negative feedback loops (reviewed in Carraco et al., 2022)”, so we believe that the term used will be clearly perceived by the reader. For further clarification, however, the subtitle The Embryo Clock in early somitogenesis in the Discussion section has been modified to (Line 321): “The Embryo Segmentation Clock in early somitogenesis”

      Reviewer #3 | Minor comments:

      The authors did not consider the fact that the first formed somite is the second somite. After the formation of the second somites, the real first somite forms anterior to the second somite. Furthermore, the real first and the third somite seems to be formed simultaneously. It is worthy for the authors to re-examine the data, whether the real first somite and the third somite are formed at the same time. And to check whether the first somite was counted to the segmented region. And this point should be at least discussed.

      R: The authors thank Reviewer#3 for the opportunity to clarify this important issue in our manuscript. It was previously described that the first morphological somite formed is, in fact, the second somite, while the “real” first somite is formed later, anteriorly to this one (Hamburger and Hamilton, 1951). This rostral-most somite-like structure is not anteriorly delimited by a fissure and has thus been termed an “incomplete” or “rudimentary” somite (Hinsch & Hamilton, 1956). Since the methodology used in our work relies on measuring the length between the rostral-most and the posterior-most intersomitic clefts, the “rudimentary” somite is not included in our data, and we considered somite #1 as the first somite delimited both anteriorly and posteriorly by intersomitic clefts. This was stated in the Methods section, under Embryo measurements, and has now also been made explicit under Somite nomenclature (Line 120): “Only structures delimited both anteriorly and posteriorly by intersomitic clefts were counted as somites.”

      We, indeed, observe the formation of the “rudimentary” somite anteriorly to somite #1, when somites 3-4 and formed. This information was included in the Discussion section, under Spatio-temporal properties of the rostral somite segmentation (Line 311): “Note that our analysis did not consider the “rudimentary somite”, as defined by Hinsch and Hamilton (Hinsch & Hamilton, 1956) since it does not possess an anterior somitic cleft. We found that this structure becomes clearly visible, rostrally to somite 1, as somites 3-4 are formed.”

      4. Description of analyses that authors prefer not to carry out

      Reviewer #1 (Significance (Required)):

      General assessment: The results are not conceptually new or surprising. However, their careful quantitative analysis is informative and worthwhile to be published because it has yet to be done in the early somite formation. In this manuscript, the authors focused on the early stage of somitogenesis. It will be more informative if they complete this analysis for the whole somite area including the end of somitogenesis.

      Reviewer #3 | Referees cross commenting****

      I find all criticisms are justified. The most advance, as stated by other both reviewers, is the quantitative assay of the somite formation, since this is no yet done previously. As suggested by the first reviewer, it will be more informative if the authors complete this analysis for all regions. For all somites would be too much work, but they can select some representative somites of each region in addition to occipital region, such as 3 somites for one region, including the cervical, thoracic, lumbal, sacral and caudal region. Thus, the dynamic of the temporal somite formation of the whole embryo can be analysed using the same method. This will provide much more impact for this work.

      R: The authors thank the Reviewers #1 and #3 for the kind words and for highlighting the quantitative approach taken. Regarding completing the analysis for the whole somite area including the end of somitogenesis, the authors agree that this would be interesting for the community. The focus of this work, however, was a detailed understanding of early somite segmentation, where measurements of somites 14-20 were performed for validation purposes alone of the technical approach developed, since their time of formation has been previously well established. Characterization of somite formation dynamics along the entire embryonic axis, while informative, would entail significant technical challenges, which are beyond the focus of this work. Briefly, we performed live imaging using the EC culture system (Chapman et al, 2001). This appropriately reproduces in ovo development of early embryos but imposes significant constraints on embryo development in older developmental stages, including the ones corresponding to the formation of the last somites. A possible alternative to perform these measurements would be to apply the tissue explant culture system developed by Palmeirim et al., 1997 to different portions of the embryo body, and optimize it for real-time imaging. However, we believe that the time and effort required are beyond the scope of this work and would not significantly contribute to elucidating the main questions addressed in this manuscript.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Somites form consecutively along the anterior to posterior (AP) axis. The time of the formation of a somite is controlled by the segmentation clock, oscillation of cyclic genes in the presomitic mesoderm. The length of an oscillation cycle differs between species and should also differ between the axial levels. In chicken embryos, one cycle for a trunk somite requires 90 minutes, while it is much slower (150 minutes) for a posterior-most somite. Is this quicker or slower for an anterior-most somite? Andrade' group addressed this question and measured the time of the formation of each occipital somites (somite 1-5). They found that the formation of an occipital somite requires only 75 minutes, while somites from somite 6 onwards takes as long as the trunk somites (about 90 minutes). The faster formation of occipital somites is correlated with the time of the cyclic expression of hairy1 and hairy2.

      Major comments:

      The conclusion is well supported by the data. The measurement of the length increments of the segmented region and then assay using algorithm are well established. Thus, the data are well reproducible.

      Minor comments:

      The authors did not consider the fact that the first formed somite is the second somite. After the formation of the second somites, the real first somite forms anterior to the second somite. Furthermore, the real first and the third somite seems to be formed simultaneously. It is worthy for the authors to re-examine the data, whether the real first somite and the third somite are formed at the same time. And to check whether the first somite was counted to the segmented region. And this point should be at least discussed.

      Referees cross commenting

      I find all criticisms are justified. The most advance, as stated by other both reviewers, is the quantitative assay of the somite formation, since this is no yet done previously. As suggested by the first reviewer, it will be more informative if the authors complete this analysis for all regions. For all somites would be too much work, but they can select some representative somites of each region in addition to occipital region, such as 3 somites for one region, including the cervical, thoracic, lumbal, sacral and caudal region. Thus, the dynamic of the temporal somite formation of the whole embryo can be analysed using the same method. This will provide much more impact for this work.

      Significance

      Significance: The measurement of the length increments of the segmented region and then assay using algorithm are the novelty and strengths of this study. So, the data are reproducible and objective.

      The results of this study extend our understanding about the dynamic process of the somitogenesis. Especially, the most interesting point is that based on this result, we can see that the segmentation clock runs faster in the head region, and then slow down gradually along the AP axis.

      Audience: specialized, basic research<br /> The developmental biologist will be interested in this topic.

      My expertise is the somite development, somite differentiation, mesoderm development.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      In this study, Maia-Fernandez et al used time-lapse imaging of chicken embryos to analyze the formation of the first formed somite, by doing in situ hybridization, they checked the dynamic nature of well-known segmentation clock genes during this process. They found that the segmentation clock period is faster for the formation of the first five somites (most anterior) and that this process is underlain by dynamic/cyclic expression of Hairy 1 and Hairy 2 as it has been described for more posterior somites.

      • I believe that there few issues that should be addressed to strengthen the conclusions of the manuscript:<br /> Figure 6 in particular, the number of analyzed embryos is small, given the fact that there is a lot of inter-individual heterogeneity in this process it could well be that the authors got, by chance, two embryos out of three having the same pattern of Hairy1 expression.
      • I believe that an additional shorter time point (+15 or 30 min) with a different pattern of the oscillatory gene would also add to the characterization of the dynamics (same for Fig 5c). This is particularly true given that the domain of expression of Hairy 1 analyzed in Figure 6 is localized quite rostral which might be interpreted as a phase 1 or a phase 2 as well (as initially described in Palmeirim et al 1997).

      Minor comments:

      In 5 c similar patterns and newly formed somites should be pointed out by arrows on the figure to help the readers.<br /> To be more specific the term segmentation clock should be used instead of embryonic clock as I believe there are other embryonic clocks (cell cycle, circadian, etc..)

      Significance

      In this study, the authors address the question of the formation of the first-formed somites using bird a model system; this is a conceptual advance in the sense that our knowledge of the dynamics of these critical morphological events is minimal. The technical advances (time-lapse, image analysis, dissection) made by the authors are quite remarkable and allow for filling the gap of knowledge the community has in this particular domain. The article is well written, data are well presented and it is interesting for a large community of developmental biologists.

      My expertise is in cell and tissue morphogenesis of amniote embryos

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Maia-Fernandes et al. investigated somite formation dynamics in the chick embryo's early stage in this manuscript. They found that the cranial most somites (1-5) form faster than the trunk. They also show that the oscillatory expression pattern of hairy1, regarded as the somitogenesis embryo clock (EC), is coupled to the somite segmentation in the occipital somites. The results are not conceptually new or surprising; they merely show what has been widely believed. However, their careful quantitative analysis is informative and worthwhile to be published because it has yet to be done in the early somite formation. To improve the manuscript, I have several concerns to be addressed.

      Major points.

      1. In the RESULTS session, "Occipital somites are formed faster than cervical and trunk somites," the authors argue that the occipital somites form with greater temporal variability than the neck and trunk somites. Judging from Figures 3C and 3D, I feel it is the case. However, the authors should demonstrate it through statistical analysis.
      2. The authors do not describe the expression patterns of hairy1 in the PSM in the manuscript, but they merely judged whether they are different or the same (recapitulate). The description of the expression pattern needs to be revised totally. The authors should describe the expression patterns of hairy1 in the PSM of each sample carefully and in detail. Fortunately, the previous report (Pourquie and Tam, Developmental Cell, 1, 619-620, 2001) categorized the expression patterns of the EC genes into three phases. The authors should at least categorize each sample according to the criterion by Pourque and Tam. If arrows of brackets indicate the area of expression, it is reader-friendly.
      3. In the RESULTS session, "Anterior-posterior length of rostral somites," the authors concluded that "the large variability in measurements of somites 17-20 most probably results from the rotation of the embryo body in these developmental stages." Probably they mentioned data of the length of #17-#20 somites in Table1. They should demonstrate it through statistical analysis to show the large variability in the specific area. I understand that embryo rotation could be a reason for the variability. The authors should show evidence. Or they should discuss various possibilities from a broad perspective.

      Minor points.

      1. In the RESULTS session, "Anterior-posterior length of rostral somites," the authors described that the average length of somites ranges from 118 - 191 μm. But according to Table-1, the lowest length of somite is 115.92 μm (∼116 μm). So, the lower limit should be corrected here.
      2. In the DISCUSSION section, the authors mentioned that they presented a thorough characterization of the size and time of formation of the first ten somites in the chicken embryo. But based on the tables and figures, it will only be the first nine somites, not the ten.
      3. In the GRAPHICAL ABSTRACT, if the color of the oscillation line is the same as the corresponding somites, it is intuitive.
      4. It would be helpful if the manuscript contained both page and line numbers.

      Significance

      General assessment: The results are not conceptually new or surprising. However, their careful quantitative analysis is informative and worthwhile to be published because it has yet to be done in the early somite formation. In this manuscript, the authors focused on the early stage of somitogenesis. It will be more informative if they complete this analysis for the whole somite area including the end of somitogenesis.

      Advance: Previously no one provided quantitative data for somite formation. In this viewpoint, this manuscript has an advantage.

      Audience: Their data could be helpful in the generation of mathematical models.

      My field: Developmental Biology

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1:

      1. The authors claim PA-induced mTORC1 activation is dependent on acetyl-CoA derived from mitochondrial fatty acid oxidation. Using isotope tracing to determine the contribution of PA to acetyl-CoA, would improve.

      Response: We thank the reviewer for this valuable comment. As suggested, we measured the contribution of PA to the total cellular acetyl-CoA pool using metabolic flux assays. The results showed that the incorporation of 13C from [U-13C16]-palmitate into acetyl-CoA was exceeding 60% (Page 40, Figure 4D in the revised version), indicating that exceeding 60% of the acetyl-CoA pool was PA derived. Likewise, we also found that the incorporations of 13C from [U-13C16]-palmitate into 6:0-CoA, 8:0-CoA, 10:0-CoA and 12:0-CoA were all exceeding 50% (Page 40, Figure 4D in the revised version). Thus, these results suggested that a large portion of the PA was used for fatty acid oxidation upon entering the cell. Moreover, we found that fatty acid β oxidation blocked by perhexiline maleate inhibited PA-induced increase of acetyl-CoA, suggesting that the induction of acetyl-CoA content was largely dependent on the fatty acid oxidation of PA. Furthermore, we also demonstrated that inhibition of mitochondrial fatty acid β oxidation by pharmacological inhibitor or genetic knockdown abrogated PA-induced activation of mTORC1. However, using sodium acetate treatment to elevate cellular acetyl-CoA levels rescued impaired mTORC1 activity induced by the inhibition of fatty acid β oxidation under PA condition. Together, these results revealed that PA-induced mTORC1 activation is dependent on acetyl-CoA derived from mitochondrial fatty acid oxidation.

      1. They showed PA increases fatty acid oxidation related gene expression and acetyl-CoA level, while OA and LA could not. why only PA could increases fatty acid oxidation and acetyl-CoA level, considering both of these lipids could be oxidation in mitochondria? Is there any differences in mitochondria among treatment of PA, OA and LA? It is better to monitor fatty acid oxidation in real time using seahorse. And add discussions.

      Response: We thank the reviewer for this constructive question. As suggested, we performed the seahorse real-time cell metabolic analysis and the results showed that PA treatment enhanced mitochondrial OCR and elevated maximal oxygen consumption rates compared with OA or LA treatment in fish myocytes (Page 37, Figure 3B in the revised version). Likewise, we also found that PA-induced increase of fatty acid oxidation-related gene expressions was more robust than OA or LA in vivo and in vitro. Thus, these results indicated that the induction of mitochondrial fatty acid oxidation by pa treatment was stronger than OA or LA treatment.

      In this study, using LC–MS, we showed that PA treatment increased the contents of short/medium-chain acyl-CoA and acylcarnitine in comparison with OA or LA treatment. Thus, these results suggested that although all three fatty acids can be oxidation in mitochondria, PA may be preferred to enter the mitochondria for fatty acid β oxidation, compared with OA or LA. Previous studies have found that OA is more inclined to synthesize triglycerides to induce the formation of lipid droplets than PA (Chen et al., 2023; Plötz et al., 2016). Likewise, we also found that OA significantly increased the contents of 18:1-CoA in comparison with PA. Thus, we speculate that, after entering the cell, OA is more preferentially synthesized to triglyceride for storage than fatty acid oxidation. Moreover, LA is considered to be a precursor of arachidonic acid, and can be converted to a myriad of bioactive compounds called eicosanoids (Whelan & Fritsche, 2013). Similarly, we found that LA markedly elevated the contents of 18:2-CoA/18:3-CoA. Thus, we conjecture that LA preferentially synthesizes functional lipids compared to entering mitochondria for fatty acid oxidation. Together, differences in the levels of acetyl-CoA produced by these three fatty acids may be related to their metabolic pathway preferences.

      There may be two reasons for why PA prefers to enter mitochondrial for fatty acid oxidation. On one hand, due to differences in the structure of PA, OA and LA, the substrate affinity of CPT1B to these fatty acyl-CoAs may be different, that may contribute to the different rates of fatty acid to enter into mitochondria. On the other hand, in contrast to the β-oxidation of SFAs, the β-oxidation of UFAs requires the involvement of 2,4-dienoyl-CoA reductase (You et al., 1989), and thus the β-oxidation of SFAs may be more efficient.

      At present, the understanding of differences in fatty acid oxidation between SFAs and UFAs is insufficient, so more studies are needed in the future to further explore the underling mechanisms behind these differences. The reviewers have raised a very important direction for research, and so we will continue to address this issue in future.

      We have expanded this section of the Discussion (Page 14, line 388-412 in the revised version).

      1. The authors present lots of western blot images, suggest to provide quantification data of these blots.

      Response: We thank the reviewer for their careful assessment of our study. We apologize for not providing quantification data for western blot images in our initial manuscript. To support our conclusions, we have now added a densitomentric and statistical analysis of all western blots in the revised version (Page 33, Figure 1 in the revised version; Page 35, Figure 2 in the revised version; Page 37, Figure 3 in the revised version; Page 40, Figure 4 in the revised version; Page 43, Figure 5 in the revised version; Page 45, Figure 6 in the revised version; Page 47, Figure 7 in the revised version; Page 51, Figure S1 in the revised version; Page 53, Figure S2 in the revised version; Page 54, Figure S3 in the revised version; Page 56, Figure S4 in the revised version; Page 57, Figure S5 in the revised version).

      1. It is better to discuss the relationship between fatty acid oxidation and mTOR signaling.

      Response: The reviewer’s comments are very valuable. We apologize for not discussing enough for the relationship between fatty acid oxidation and mTOR signaling in our initial manuscript. We have now expanded this section of the Discussion (Page 13, line 366-387 in the revised version, see below).

      Growing lines of evidence suggested a strong link between mitochondrial fatty acid oxidation and mTORC1 signaling (Ricoult & Manning, 2013). As a central regulator of anabolism, mTORC1 is considered to inhibit fatty acid β oxidation pathway for energy storage or ketogenesis (Aguilar et al., 2007). Several studies revealed that restrained mTORC1 by rapamycin induced fatty acid β oxidation in rat hepatocytes through increasing expression of fatty acid β oxidation related enzymes (Brown et al., 2007; Peng et al., 2002). Likewise, mice with whole-body knockout of S6K1 showed enhanced fatty acid β oxidation and increased expression levels of CPT1 in isolated adipocytes (Um et al., 2004), and S6K1/S6K2 double-knockout mice also exhibited elevated fatty acid β oxidation of fatty acids in isolated myoblasts by activating AMPK (Aguilar et al., 2007). Furthermore, a recent study has established that FOXK1 can mediate the inhibition of fatty acid β oxidation by mTORC1 (Fujinuma et al., 2023). Thus, these collective data revealed that fatty acid β oxidation was restrained by mTORC1. However, conversely, the role of mitochondrial fatty acid oxidation in the regulation of mTORC1 is still controversy. A study in prostate cancer cells suggested that inhibited fatty acid β oxidation by etomoxir reduced mTORC1 activity (Schlaepfer et al., 2014), and another study found that deleting CPT1B specifically in skeletal muscle of mice suppressed mTORC1 by provoking AMPK activation (Vandanmagsar et al., 2016). Consistent with these studies, our results showed that acetyl-CoA derived from mitochondrial fatty acid β oxidation induced mTORC1 activation under PA treatment, indicating that acetyl-CoA may be a novel insight linking fatty acid β oxidation and mTORC1 signaling. Paradoxically, unlike other studies, a recent study found that mice with heart-specific CPT2-deficient exhibited induction of mTORC1 pathway. Thus, the effects of fatty acid β oxidation on mTORC1 pathway are complicated and may differ under variable physiological and pathological conditions. Further studies are needed to determine the sophisticated mechanisms underlying the regulation of fatty acid β oxidation on mTORC1 signaling.

      Reviewer #2:

      Major comments:

      Initial experiment: Among several fatty acid-rich diets, fish were fed a palmitic acid (PA) rich (PO) diet for 10 weeks, and the PO diet significantly raised fasting blood glucose levels compared to control diet (fish oil of equal lipid content). The PO diet also impaired the fish's glucose and insulin tolerance. The PO diet also led to decreased phosphorylation levels of AKT, which regulates glucose metabolism. Therefore, the researchers initially concluded that a palmitic acid-rich diet leads to systemic insulin resistance in fish.

      1. I have a couple of questions on this initial experiment on which all the subsequent studies are based. In Figure 1A, the body weight was identical in control and PO group. Don't you expect PO feeding lead to obesity in fish, as HFD induces obesity in mice?

      Response: We thank the reviewer for this constructive question. In our study, we found that dietary PO diet for 10 weeks failed to affect the body weight of fish, compared with CON diet. Similar to our results, a study in human also found that there was no significant differences in the body weight and body mass index (BMI) between saturated fat diet and monounsaturated fat diet (Vessby et al., 2001). Unlike high-fat diet, the lipid content level of PO diet was not elevated, but only the fatty acid composition was altered, with palmitic acid composition being significantly increased in comparison with CON diet (Page 60, Table S1). Thus, this may be the reason of why the PO diet did not induce weight gain.

      Although accumulating evidence showed that the onset of insulin resistance was often accompanied by weight gain and obesity (Kahn & Flier, 2000; Shoelson et al., 2007), some studies also found that insulin resistance occurred without obesity. A recent study found that mice with liver knockout of Lpcat3 exhibited improved insulin sensitivity without a change in the body weight (Tian et al., 2023). Moreover, another study in mice showed that dietary phenylalanine-rich diet induced insulin resistance, but had no effects on the body weight (Zhou et al., 2022). Likewise, our study also found that dietary PO diet provoked systemic insulin resistance, while did not affect the body weight in fish. Thus, these studies indicated that the development of insulin resistance may not always be entirely accompanied by obesity.

      1. Figure 1G and 1H show glucose and insulin tolerance after PO feeding for 10 weeks. The area under curve (AUC) should be compared to determine if GTT and ITT were statistically different. The ITT curve is particularly interesting as the control fish did not seem to respond to insulin, while the PO-fed fish responded more robustly. The only difference is the initial glucose level. Are the GTT and ITT done after fasting? How long is the fasting? The curves suggest that even though PO increased (fasting) blood glucose levels, it improved insulin sensitivity - therefore the premise that PO induces insulin resistance is not supported here. The lack of insulin induced response in the control group is worrisome. I suggest that the measures should be retaken, and AUC should be used to support if there are any differences in GTT and ITT.

      Response: We thank the reviewer for this valuable comment. We apologize for making this confusion in the initial manuscript and we thank the reviewer for providing this opportunity to correct our manuscript. As suggested, to further investigate whether dietary PO could cause impairment of insulin sensitivity, we have re-performed the GTT and ITT assays. Considering that fish have a poor capacity to utilize glucose, we extended the assay time to 8 h. To make the results more accurate, we also added the biological replicates. Moreover, before injection of glucose or insulin, fish were fasted for 24 h. Furthermore, we added area under curve (AUC) of GTT and ITT, and performed statistical analyses of the AUC data.

      Our results showed that dietary PO diet reduced glucose tolerance and insulin tolerance in fish (Page 33, Figure 1G and 1H in the revised version). Moreover, compared with CON diet, the AUC of GTT and ITT were significantly increased in PO diet (Page 33, Figure 1G and 1H in the revised version). Similarly, we found that dietary PO diet elevated fasting blood glucose levels and plasma insulin concentrations. Furthermore, we showed that dietary PO diet decreased the phosphorylation levels of AKT in the liver and skeletal muscle. In addition, we also demonstrated that PA treatment could induce cellular insulin resistance in fish myocytes and C2C12 myotubes. Thus, in our opinion, the above results could indicate that dietary PO induced insulin resistance in fish.

      1. Based on the assumption that PO induces IR (which needs to be confirmed based on the previous comments), the researchers attempted to understand how PA triggers IR through a series of experiments, predominantly western blot analysis. All the Western blots should be quantified. The model is that PA activates FAO in mitochondrial that elevates cytosolic acetyl-coA, which acetylates Rheh to activate mTORC1. mTORC1 on one hand alters IRS1 phosphorylation and on the other hand inhibits transcriptional activity of TFEB to reduce Irs1 mRNA level. Together reduces IRS1 leads to Insulin Resistance.

      Response: The reviewer’s comments were very important to verify the validity of our findings. We have now added a densitomentric and statistical analysis of all western blots in the revised version (Page 33, Figure 1 in the revised version; Page 35, Figure 2 in the revised version; Page 37, Figure 3 in the revised version; Page 40, Figure 4 in the revised version; Page 43, Figure 5 in the revised version; Page 45, Figure 6 in the revised version; Page 47, Figure 7 in the revised version; Page 51, Figure S1 in the revised version; Page 53, Figure S2 in the revised version; Page 54, Figure S3 in the revised version; Page 56, Figure S4 in the revised version; Page 57, Figure S5 in the revised version).

      1. Figure 1. PA reduces basal and insulin stimulated AKT phosphorylation in fish liver and muscle, as well as in culture fish and murine myocytes (Fig. 1I-M). The results appear to be solid but need to be quantified.

      Response: We thank the reviewer for this kind suggestion. We have now added a densitomentric and statistical analysis of all western blots in Figure 1 (Page 33, Figure 1 in the revised version).

      1. Figure 2 shows that PA provoked hyperactivation of mTORC1 (indicated by elevated phosphorylated S6K levels. This effect was abolished by Rapamycin treatment (an mTORC1 inhibitor) and also abolished by insulin stimulation (2F). Again, the western blots should be quantified.

      Response: We thank the reviewer for this excellent suggestion. We have now added a densitomentric and statistical analysis of all western blots in Figure 2 (Page 35, Figure 2 in the revised version).

      1. Figure 6: the researchers measured the effect of PA treatment on IRS1 phosphorylation in order to understand the mechanism of insulin resistance induced by mTORC1 activation under PA treatment. A PO diet intensified S636/S639 phosphorylation in fish muscle. In fish myocytes and C2C12 myotubes, PA treatment elevated S636/S639 phosphorylation but decreased the Y612 phosphorylation of IRS1 in a dose-dependent manner. Treatment of fish myocytes and C2C12 myotubes with an mTOR inhibitor blocked increased IRS1 S636/S639 phosphorylation levels under PA treatment. Also, PA specifically reduced mRNA levels of Irs1. This indicates that PA-induced, mTOR-dependent alteration of IRS1 phosphorylation and transcription may have contributed to insulin resistance. It is unclear how mTORC induces either increase or decrease in IRS1 phosphorylation depending on the residuals.

      Response: We appreciate the reviewers for this important question. In fact, previous studies have clearly explored how mTORC1 pathway affects S636/S639 phosphorylation of IRS1. On one hand, as a kinase complex, mTORC1 could directly induce S636/S639 phosphorylation of IRS1 in vitro (Ozes et al., 2001). On the other hand, mTORC1 could activate S6K to promote S636/S639 phosphorylation of IRS1 (Shah & Hunter, 2006; Um et al., 2004). In addition, considering that the serine/threonine phosphorylation status of IRS has been shown to affect its tyrosine phosphorylation and protein degradation (Copps & White, 2012), we speculate that the decrease of Y612 phosphorylation of IRS1 is dependent on the induction of IRS1 S636/S639 phosphorylation.

      In this study, we found that PA could induce S636/S639 phosphorylation of IRS1 in a mTORC1-dependent manner. Considering that previous studies have explored the mechanism by which mTORC1 induced IRS1 S636/S639 phosphorylation, we did not conduct further studies on this issue. Notably, we found that mTORC1 could also regulate the transcription of IRS1, so we subsequently investigated the mechanism by which mTORC1 inhibited IRS1 transcription.

      1. Figure 7 shows that PA inhibits nuclear translocation of TFEB to suppress IRS1 transcription. The EMSA in 7D is not convincing.

      Response: We thank the reviewer for this valuable comment and we apologize for providing unclear blots in the initial manuscript. To support our conclusions, we have now re-performed the EMSA assays. The results suggested that TFEB can directly bind to the IRS1 promoter at these two sites (Page 47, Figure 7D in the revised version).

      Minor comments:

      1. Some data appears to weaken the results and/or contradictory. For example, the paper initially showed reduced AKT phosphorylation to support PA induced IR, but shouldn't a lower level of pAKT reduces mTORC activation? But then the rest of the manuscript explores how PA activates mTOR. Part of the IR is manifested by impaired mTORC1 activation, yet the PA activates mTORC1. The authors should present the rationale and flow of the ideas in a better way.

      Response: We thank the reviewer for this excellent suggestion. We appreciate the points that in some insulin resistance conditions, as a downstream of the insulin pathway, mTORC1 activity is manifested to be inhibited. However, mTORC1 activity showed different under other insulin resistance conditions.

      In fact, multiple negative feedback signals exist in cells to maintain cellular homeostasis under diverse environmental challenges and stimulations (Kearney et al., 2021). However, aberrant of negative feedback can lead to impaired intracellular signaling pathway and induce a variety of diseases (Nguyen & Kholodenko, 2016). Similarly, numerous negative feedback mechanisms also exist in insulin signaling to prevent the development of cancers that may be induced by hyperactivation of insulin pathway. The negative feedback of insulin pathway is mainly mediated by mTORC1, which has been found to inhibit insulin signaling transduction by directly or indirectly affecting IRS1 phosphorylation (Copps & White, 2012; Shah & Hunter, 2006; Um et al., 2004). However, under some pathological or stress conditions, mTORC1 is over-activated, resulting in the amplification of the negative feedback of insulin pathway and the development of insulin resistance. A recent study found that imidazole propionate, a metabolite produced by the gut microbiota, provoked insulin resistance through inducing mTORC1 activation and phosphorylation of IRS1 (Koh et al., 2018). Other studies also showed that elevated abundance of branched-chain amino acids (BCAAs) or branched-chain α-keto acid (BCKA) could cause insulin resistance by boosting mTORC1 pathway (Zhou et al., 2019). Thus, mTORC1 activation induced-negative feedback inhibition of insulin pathway may be a critical factor in the development of insulin resistance.

      Consistently, our study found that PA could activate mTORC1 in an acetylation modification-dependent manner. Moreover, activation of mTORC1 inhibited the phosphorylation of AKT and caused insulin resistance by affecting the phosphorylation and transcription of IRS1.Indeed, AKT is considered to activate mTORC1 in multiple manners, and inhibition of AKT results in the reduction of mTORC1 activity. However, mTORC1 activity is not only affected by AKT, but is also regulated by a diverse set of upstream signals (Saxton & Sabatini, 2017). Thus, we considered that the activating effect of PA on mTORC1 activity is higher than the negative effect of mTORC1 activity produced by AKT inhibition. This also led to the fact that mTORC1 remained in an activated state despite the inhibition of AKT in PA condition.

      1. There are also many run-on sentences and grammar issues, making it very hard to read. The writing can be improved.

      Response: We thank the reviewer for this valuable comment and we apologize for these grammar mistakes in the initial manuscript. Following the reviewer’s suggestion, we have invited native speaker to guide the English writing and carefully corrected these run-on sentences and grammar issues. We thank the reviewer for this careful evaluation of our manuscript.

      Reviewer #3:

      Major issues affecting the conclusions:

      1. The conclusions are supported by the data. However, I suggest to perform a densitomentric and statistical analysis of western blots, especially when the authors report a representative blot, showing samples loaded in single.

      Response: We thank the reviewer for this excellent suggestion. We agree that it would be important to verify the validity of our findings and we apologize for not providing quantification data for western blot images in our initial manuscript. We have now added a densitomentric and statistical analysis of all western blots in the revised version (Page 33, Figure 1 in the revised version; Page 35, Figure 2 in the revised version; Page 37, Figure 3 in the revised version; Page 40, Figure 4 in the revised version; Page 43, Figure 5 in the revised version; Page 45, Figure 6 in the revised version; Page 47, Figure 7 in the revised version; Page 51, Figure S1 in the revised version; Page 53, Figure S2 in the revised version; Page 54, Figure S3 in the revised version; Page 56, Figure S4 in the revised version; Page 57, Figure S5 in the revised version).

      1. The methods are clear and reproducible. The authors should better explain how they have dissolved all the powders (i.e. fatty acids) to obtain the stock solutions next diluted (from what concentration?) in the media

      Response: We thank the reviewer for this valuable comment. We apologize for not explaining how to dissolved all the powders in the media. As suggested, we have now provided a detailed explanation of how to dissolved all the powders to obtain the stock solutions in the Methods (Page 19, line 537-586 in the revised version, see below).

      For PA, OA or LA in vitro treatment, fatty acid free BSA (Equitech-Bio, USA) was dissolved in FBS-free DMEM at room temperature according the ratio 1:100 (1 g fatty-acid free BSA: 100 ml FBS-free DMEM). 500 mg PA (Merck, Cat#P0500), OA (Merck, Cat#O1008) or LA (Merck, Cat#L1376) was dissolved in 10 ml ethanol to obtain PA, OA or LA stock solution respectively. Then PA, OA or LA stock solution was blow-drying with nitrogen gas and was dissolved in 0.1 M NaOH and warming at 75°C until clear to obtain 100 mM PA, OA or LA solution. Subsequently, 100 mM PA, OA or LA solution was added to 1% BSA solution according the ratio 1:100 (100 mM PA:1% BSA, v/v) at 50°C. Finally, the mixture was filtered using a 0.45 μM filter and stored at -20°C. For insulin in vitro treatment, insulin powder (Merck, USA) was dissolved in hydrochloric acid (pH=2) to obtain 1 mg/ml stock solution. For rapamycin or Torin1 in vitro treatment, rapamycin (Med Chem Express, #HY-10219, USA) or Torin1 (Med Chem Express, #HY-13003, USA) was dissolved in dimethyl sulfoxide (DMSO, Solarbio, China) to obtain 1 mM stock solution respectively. For MHY1485 in vitro treatment, MHY1485 (Med Chem Express, #HY-B0795, USA) was dissolved in DMSO (Solarbio, China) to obtain 10 mM stock solutions.

      For etomoxir or perhexiline maleate in vitro treatments, etomoxir (Med Chem Express, #HY-50202, USA) or perhexiline maleate (Med Chem Express, #HY-B1334A, USA) was dissolved in DMSO (Solarbio, China) to obtain 50 mM stock solution respectively. For BMS-303141 treatment, BMS-303141 (Med Chem Express, #HY-16107, USA) was dissolved in DMSO (Solarbio, China) to obtain 25 mM stock solutions. For sodium acetate treatment, sodium acetate (Merck, #S2889, USA) was dissolved in ultrapure water from a Milli-Q water system to obtain 5M stock solution. For C646, spermidine or MB-3 treatment, C646 (Med Chem Express, #HY-13823, USA), spermidine (Med Chem Express, #HY-B1776, USA) or MB-3 (Merck, #M2449, USA) was dissolved in DMSO (Solarbio, China) to obtain 50 mM stock solution respectively. For MG149 treatment, MG149 (Med Chem Express, #HY-15887, USA) was dissolved in DMSO (Solarbio, China) to obtain 150 mM stock solution. For TFEB activator 1 treatment, TFEB activator 1 (Med Chem Express, #HY-135825) was dissolved in DMSO (Solarbio, China) to obtain 10 mM stock solution.

      1. Anova analysis should be performed to analyze western blot densitometries.

      Response: The reviewer raises an important point and we appreciate this comment. As suggested, we have now added statistical analyses of all western blot densitometries in the revised version. The data are presented as the means ± SEM and were analyzed using independent t-tests for two groups and one-way ANOVA with Tukey’s test for multiple groups (Page 33, Figure 1 in the revised version; Page 35, Figure 2 in the revised version; Page 37, Figure 3 in the revised version; Page 40, Figure 4 in the revised version; Page 43, Figure 5 in the revised version; Page 45, Figure 6 in the revised version; Page 47, Figure 7 in the revised version; Page 51, Figure S1 in the revised version; Page 53, Figure S2 in the revised version; Page 54, Figure S3 in the revised version; Page 56, Figure S4 in the revised version; Page 57, Figure S5 in the revised version).

      Minor comments:

      Prior studies are referenced appropriately, text and figures are clear. I suggest to add in the abstract all the model systems used. HEK293 also should be inserted in the description of the results. Please add the reference to figure 8 in the text. Please, describe cell origin.

      Response: We thank the reviewer for this careful assessment of our study. We apologize for not making this clearer in our initial manuscript. We have now added all the model systems used in the Abstract (Page 2, line 24-28 in the revised version, see below).

      Here, using a croaker model, we report that dietary palmitic acid (PA), but not oleic acid or linoleic acid, leads to dysregulation of mTORC1 signaling which provokes systemic insulin resistance and glucose intolerance. Mechanistically, using croaker primary myocytes, mouse C2C12 myotubes and HEK293T cells, we show that PA-induced mTORC1 activation is dependent on mitochondrial fatty acid β oxidation.

      Moreover, we have now added the description of HEK293T cells in the Results (Page 10, line 261-265 in the revised version; Page 11-12, line 309-315 in the revised version, see below).

      To further investigate whether the regulation of mTORC1 by Tip60 is dependent on the acetylation of Rheb, the interaction between Tip60 and Rheb was analyzed via co-immunoprecipitation assays, and the results showed that Tip60 can interact with Rheb in HEK293T cells (Figure 5E). Moreover, overexpressed Tip60 reinforced the acetylation of Rheb and phosphorylation levels of S6K in HEK293T cells (Figure 5F).

      Dual luciferase experiments in HEK293T cells showed that TFEB had the strongest ability to elevate the luciferase activity of the IRS1 promoter among the crucial downstream transcription factors of mTORC1 (Figure 7A). Moreover, TFEB enhanced the promoter activity of IRS1 in a dose-dependent manner (Figure 7B) and mutations of the predicted TFEB binding site 4 and site 6 in the IRS1 promoter significantly reduced the promoter activity of IRS1 in HEK293T cells (Figure 7C). Furthermore, ChIP and EMSA experiments in HEK293T cells verified that TFEB can directly bind to the IRS1 promoter at site 4 and site 6 (Figures 7D and 7E).

      As suggested, we have added the reference to figure 8 in the Discussion (Page 17, line 483-487 in the revised version, see below).

      In summary, our work unveils an evolutionarily conserved mechanism by which mitochondrial fatty acid β oxidation flux of acetyl-CoA induces mTORC1 activation through enhancing Tip60-mediated Rheb acetylation under PA condition. Subsequently, hyperactivation of mTORC1 boosted serine phosphorylation of IRS1 and inhibited TFEB-mediated transcription of IRS1, leading to insulin resistance (Figure 8).

      As suggested, we have added the description of cell origin in the Methods (Page 19, line 526-527 in the revised version; Page 19, line 533-534 in the revised version, see below).

      Mouse C2C12 myoblast cells were obtained from the Cell Bank of the Chinese Academy of Sciences (Shanghai, China).

      HEK293T cells were obtained from the Cell Bank of the Chinese Academy of Sciences (Shanghai, China).

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      The manuscript by Zhao et al aimed to elucidate the mechanisms by which palmitic acids drives insulin resistance. The authors performed experiments in croaker, fish myocytes and mouse differentiated C2C12. They manage in vivo metabolic assays, enzymatic assays, immunoblot procedures, double luciferase assays, RNA analysis, pharmacological inhibitions and genetic knockdown. By using these model systems and procedures, the authors demonstrate that palmitic acid, but not oleic and linoleic acids, induces systemic and cellular insulin resistance through the hyper activation of mTORC1. They show that palmitic acid stimulates the mitochondrial fatty acid β oxidation, increasing the acetyl-CoA levels which enhances the acetylation of Rheb, a well known activator of mTORC1, by Tip60. Moreover, the authors show that mTORC1, beside reinforcing IRS1 phosphorylation, inhibits nuclear translocation of TFEB, thus preventing IRS1 transcription.

      Major issues affecting the conclusions:

      The conclusions are supported by the data. However, I suggest to perform a densitomentric and statistical analysis of western blots, especially when the authors report a representative blot, showing samples loaded in single.<br /> The methods are clear and reproducible. The authors should better explain how they have dissolved all the powders (i.e. fatty acids) to obtain the stock solutions next diluted (from what concentration?) in the media<br /> Anova analysis should be performed to analyze western blot densitometries.<br /> Minor comments:<br /> Prior studies are referenced appropriately, text and figures are clear. I suggest to add in the abstract all the model systems used. HEK293 also should be inserted in the description of the results. Please add the reference to figure 8 in the text. Please, describe cell origin.

      Referee Cross-commenting

      All reviewers have requested densitomentric (and statistical) analysis of western blot to prove the strength of the results. This is the major point to be addressed. Other points should also be only discussed.

      Significance

      The study extends the knowledge in the field of fatty acids-induced insulin resistance, that is a field studied by many researchers from many years, but with a lot of unclear mechanisms yet. Thus, the nature of the advance is conceptual and mechanistic. The only limitation is the lack of evidence in human samples/cells.

      Basic researchers and experts in translational medicine will be interested by this research.<br /> This is the point of view of a basic researcher, mainly interested in the molecular mechanisms underlining type 2 diabetes/obesity and cancer.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary:

      The goal of this study was to investigate how saturated fat induce insulin resistance through activating mTOR. More specifically, the researchers show that palmitate (saturated fat) activates mTORC1 to induce insulin via transcriptional and posttranslational suppression of IRS1. Specifically, the researchers show that PA stimulates FAO to raise cytosolic acetyl-coA levels, promoting Tip60-mediated acetylation of Rheb to activate mTORC1 activity. To study the relationship between mTORC1 activity, fatty acid stimulation and insulin resistance, the researchers decided to conduct their study in fish, as fish are known to be glucose intolerant by nature, making them an appropriate model organism for studying insulin resistance. Mouse C2C12 cell line and fish primary myocytes were also used to validate key results.

      Major comments:

      Initial experiment: Among several fatty acid-rich diets, fish were fed a palmitic acid (PA) rich (PO) diet for 10 weeks, and the PO diet significantly raised fasting blood glucose levels compared to control diet (fish oil of equal lipid content). The PO diet also impaired the fish's glucose and insulin tolerance. The PO diet also led to decreased phosphorylation levels of AKT, which regulates glucose metabolism. Therefore, the researchers initially concluded that a palmitic acid-rich diet leads to systemic insulin resistance in fish.

      I have a couple of questions on this initial experiment on which all the subsequent studies are based. In Figure 1A, the body weight was identical in control and PO group. Don't you expect PO feeding lead to obesity in fish, as HFD induces obesity in mice?<br /> Figure 1G and 1H show glucose and insulin tolerance after PO feeding for 10 weeks. The area under curve (AUC) should be compared to determine if GTT and ITT were statistically different. The ITT curve is particularly interesting as the control fish did not seem to respond to insulin, while the PO-fed fish responded more robustly. The only difference is the initial glucose level. Are the GTT and ITT done after fasting? How long is the fasting? The curves suggest that even though PO increased (fasting) blood glucose levels, it improved insulin sensitivity - therefore the premise that PO induces insulin resistance is not supported here. The lack of insulin induced response in the control group is worrisome. I suggest that the measures should be retaken, and AUC should be used to support if there are any differences in GTT and ITT.

      Based on the assumption that PO induces IR (which needs to be confirmed based on the previous comments), the researchers attempted to understand how PA triggers IR through a series of experiments, predominantly western blot analysis. All the Western blots should be quantified. The model is that PA activates FAO in mitochondrial that elevates cytosolic acetyl-coA, which acetylates Rheh to activate mTORC1. mTORC1 on one hand alters IRS1 phosphorylation and on the other hand inhibits transcriptional activity of TFEB to reduce Irs1 mRNA level. Together reduces IRS1 leads to Insulin Resistance.

      Figure 1. PA reduces basal and insulin stimulated AKT phosphorylation in fish liver and muscle, as well as in culture fish and murine myocytes (Fig. 1I-M). The results appear to be solid but need to be quantified.

      Figure 2 shows that PA provoked hyperactivation of mTORC1 (indicated by elevated phosphorylated S6K levels. This effect was abolished by Rapamycin treatment (an mTORC1 inhibitor) and also abolished by insulin stimulation (2F). Again, the western blots should be quantified.

      Figure 3: PA treatment increases mRNA expression levels of fatty acid beta oxidation genes in fish myocytes and C2C12 myotubes, and subsequent suppression of CPT1B and CPT2 (rate-limiting enzymes of FAO) inhibited mTORC1 activity and signaling in muscle, C2C12 myotubes and fish myocytes under PA treatment. This suggests PA-induced mTORC1 activation is dependent on mitochondrial FAO. Inhibition of CPT1 improved suppression of insulin stimulated phosphorylation of AKT under PA treatment in fish myocytes and C2C12 myotubes, indicating that mitochondrial FAO is heavily involved in PA-induced mTORC1 activation that contributes to insulin resistance.

      Figure 4: PO diet increases acetyl-CoA levels in muscle and PA treatment increases intracellular acetyl-CoA in a dose-dependent manner in fish myocytes. Inhibition of FAO by perhexiline maleate diminished induction of acetyl-CoA under PA treatment. In vivo dsRNA knockdown of ATP citrate lyase (ACLY, catalyze acetyl-CoA synthesis from mitochondrial citrate) decreased mTORC1 activity in muscle, and inhibition of ACLY in fish myocytes and C2C12 myotubes decreased induction of mTORC1 activity under PA treatment. This indicates that palmitic acid promotes mTORC1 activation through acetyl-CoA that is derived from mitochondrial FAO. PA treatment elevates acetylation of Rheb in a dose-dependent manner. Inhibition of FAO by perhexiline maleate attenuated PA-stimulated Rheb acetylation, while fish myocytes and C2C12 myotubes treated with sodium acetate (which can enhance acetyl-CoA) exhibited enhanced Rheb acetylation. The data indicate that acetyl-CoA produced by FAO activates mTORC1 signaling through increased Rheb acetylation. Phosphorylation of AKT were enhanced in muscle with dsACLY knockdown injection, and sodium acetate addition blocked recovery of insulin-stimulated glucose uptake and phosphorylation levels of AKT by perhexiline maleate under PA treatment. ACLY inhibition promoted insulin stimulated phosphorylation of AKT under PA treatment. So, in terms of acetyl-CoA's role in PA-induced insulin resistance, the data suggest that acetyl-CoA derived from FAO mediates PA-induced mTORC1 activation and insulin resistance.<br /> Figure 5: Acetyl-CoA can activate lysine acetyltransferases, and the researchers found that mRNA expression of tip60 was elevated in fish myocytes and C2C12 myotubes under PA treatment. Cultured fish myocytes and C2C12 myotubes treated with a Tip60 inhibitor prevented the induction of mTORC1 activity under PA treatment. Tip60 knockdown also blocked PA-induced mTORC1 activation in C2C12 myotubes. The researchers then determined that Tip60 regulation of mTORC1 is dependent on the acetylation of Rheb by studying the interaction between the two via a CoIP assay, which indeed indicated that Tip60 and Rheb interact. Additionally, Tip60 knockdown impaired PA-induced acetylation of Rheb, supporting the notion that Tip60 mediates the acetylation of Rheb under PA treatment. Inhibition of Tip60 attenuated PA-induced suppression of insulin-stimulated glucose uptake in C2C12 myotubes, and inhibition of Tip60 also restored insulin-stimulated phosphorylation of AKT under PA treatment. These data suggest that Tip60 mediates the regulation of Rheb acetylation under PA treatment and may be a novel therapeutic target for insulin resistance.

      Figure 6: the researchers measured the effect of PA treatment on IRS1 phosphorylation in order to understand the mechanism of insulin resistance induced by mTORC1 activation under PA treatment. A PO diet intensified S636/S639 phosphorylation in fish muscle. In fish myocytes and C2C12 myotubes, PA treatment elevated S636/S639 phosphorylation but decreased the Y612 phosphorylation of IRS1 in a dose-dependent manner. Treatment of fish myocytes and C2C12 myotubes with an mTOR inhibitor blocked increased IRS1 S636/S639 phosphorylation levels under PA treatment. Also, PA specifically reduced mRNA levels of Irs1. This indicates that PA-induced, mTOR-dependent alteration of IRS1 phosphorylation and transcription may have contributed to insulin resistance. It is unclear how mTORC induces either increase or decrease in IRS1 phosphorylation depending on the residuals.

      Figure 7 shows that PA inhibits nuclear translocation of TFEB to suppress IRS1 transcription. The EMSA in 7D is not convincing.

      Minor comments:

      Some data appears to weaken the results and/or contradictory. For example, the paper initially showed reduced AKT phosphorylation to support PA induced IR, but shouldn't a lower level of pAKT reduces mTORC activation? But then the rest of the manuscript explores how PA activates mTOR. Part of the IR is manifested by impaired mTORC1 activation, yet the PA activates mTORC1. The authors should present the rationale and flow of the ideas in a better way.

      There are also many run-on sentences and grammar issues, making it very hard to read. The writing can be improved.

      Referee cross-commenting

      Other than lacking quantification of western blots, my major concern is the ITT curve in Figure 1G, which does not support the conclusion that PO induces insulin resistance and therefore the rest of the study is based on a faulty premise. The curve shows that insulin reduced blood glucose much more robustly in the PO group than in the control group, suggesting PO increased insulin sensitivity. Area above curve should be calculated to quantify the difference.

      Significance

      Overall, this research was trying to show that PA-induced IR is dependent on hyper activation of mTORC1. More specifically, acetyl-CoA induces mTORC1 activation under a palmitic acid diet, and this is achieved through Tip60-mediated Rheb acetylation, which ultimately leads to insulin resistance through IRS1 suppression. The study is very mechanistic and important for understanding IR that is associated with diets high in saturated fatty acids and could potentially leads to therapeutic targets for combating insulin resistance and glucose intolerance.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      In this work, the authors showed PA induces hyperactivation of mTORC1 and insulin resistance. They found acetyl-CoA derived from mitochondrial fatty acid oxidation is required for PA-induced mTORC1 activation and insulin resistance by increasing TIP60-mediated Rheb acetylation. They also showed PA induced mTORC1 activation enhances IRS1 phosphorylation and inhibits transcription of IRS1 by impeding TFEB nuclear translocation. Overall, the authors did a lot of experiments to prove that PA causes mTORC1 activation and insulin resistance. The results are generally convincing, and the finding is novel and instructive.

      1. The authors claim PA-induced mTORC1 activation is dependent on acetyl-CoA derived from mitochondrial fatty acid oxidation. Using isotope tracing to determine the contribution of PA to acetyl-CoA, would improve.
      2. They showed PA increases fatty acid oxidation related gene expression and acetyl-CoA level, while OA and LA could not. why only PA could increases fatty acid oxidation and acetyl-CoA level, considering both of these lipids could be oxidation in mitochondria? Is there any differences in mitochondria among treatment of PA, OA and LA? It is better to monitor fatty acid oxidation in real time using seahorse. And add discussions.
      3. The authors present lots of western blot images, suggest to provide quantification data of these blots.
      4. It is better to discuss the relationship between fatty acid oxidation and mTOR signaling.

      Significance

      The finding is interesting and significant.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      The authors do not wish to provide a response at this time.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary:

      The present manuscript presents a thorough description of the relative localization (in space and time) of a number of proteins of the early secretory pathway. To that aim, the authors used by their custom-made 3D live cell super-resolution microscope (SCLIM) and the yeast S. cerevisiae as a model system. The main claim of these data is that the early secretory pathway in S. cerevisiae is organized by maturation from a newly proposed yeast ERGIC compartment all the way to the trans-Golgi network (TGN).

      Major comments:

      I have two major comments regarding this manuscript:

      1. It is not clear to me how the presented data shows the existence of an ERGIC in the yeast S. cerevisiae. I understand, and appreciate from this text too, that a clear definition of ERGIC, even in a mammalian system, is unclear. For this reason, I would first suggest that the authors provide a clear definition of what ERGIC means to them. Next, the experiments herein presented are all based on a very careful, thorough and nicely organized spatio-temporal mapping of a large number of early secretory pathway proteins, including the ERGIC53 yeast "counterpart" Emp46 (it would help to add, even as a supplementary figure, an alignment/sequence comparison between the human ERGIC53 and the S. cerevisiae emp46). However, the data presented here does not clearly indicate to me that there is a bona fide ERGIC in yeast. Couldn't it just be that what the authors call ERGIC is a cis-Golgi cisterna? I understand that the BFA experiments show a different behavior for some proteins, which fits with what the authors previously names GECCO in plants, so why not calling this GECCO? Again, it will be important to provide definitions of these compartments for the audience. Next, my main concern here is that this is all based on SCLIM, which is a very nice technique, but the resolution is limited in both space and time (by the way, it would be nice to explicitly measure of quantify the spatial resolution in x-y-z). Hence, it is not possible to discern whether an "independent" ERGIC is formed as compared to cis-Golgi cisterna. Electron microscopy (possibly CLEM) could help somehow resolve that and massively increase the strength of the claims, but I do understand this might be difficult for this group and very time consuming, so it might be important to clearly state the limitation of the herein presented data. A possible alternative to test if protein that are seen segregated are within the same membrane (as claimed here) would be to do trapping experiments where a reagent induces dimerization between the two proteins (when tagged with specific tags, such as FKBP/FRB).
      2. I could not find any details (maybe I have missed them) about how many times experiments were replicated and the statistical significance of the findings herein reported. In most figures, examples of microscopy images/videos are shown, and selected lines profiles are presented. However, it is not clear how robust these experiments are. Some ideas:

      2.1.) The major source of quantification is the peak-to-peak time distance between two proteins. In Table S1 some stdev is presented, but not clear how it is find (it is the sted of all n number of puncta? or of the mean duration per cell? or of the mean duration per experiment? I would suggest that the authors provide the results shown in Table S1 plotted as a histogram or superplot (see e.g. https://rupress.org/jcb/article/219/6/e202001064/151717/SuperPlots-Communicating-reproducibility-and) and clearly explain how statistics is performed.

      2.2) Also, the time-lapse movies are acquired with a 5s gap between time points. How is this included in the incertainty of the peak-to-peak duration in Table S1?

      2.3) In pg. 7 the authors write "Although experimental variation was high, the two zones appeared to be spatially segregated". Can the authors provide quantitative and statistical support of this claim?

      2.4) It is not clear to me how the puncta for analysis are selected. For example, in Fig. 1C, the punctum shown already shows some initial co-localization (it could be e.g. that a peak value was prior or after the duration of the time lapse movie, thereby biassing the computation of the peak-to-peak duration). So, if one would consider those spots e.g., positive for Emp46 that do not contain Mnn9 signal, how often do you see conversion (that is, appearance of Mnn9 signal)? Along the same lines, in pg. 8 the authors write "... signal appeared first and then mnn9-mCherry came up". Details on how this quantification is done and statistical analysis would be needed, to my opinion, to support the claim.

      Minor comments:

      1. The color code for the 3 color microscopy images is nice, however, the use of green and red for the 2 color images is a bit unfortunate for some people (like myself) who suffer from color blindness. I'd suggest to use green and magenta instead.
      2. Pg. 8: have the authors tested Rer1 vs Emp46?
      3. Pg. 8: I was of the impression that GRASP65 (GORASP1) is considered to be a cis-Golgi protein (see e.g., Tie et al eLife 2018). Then, what the authors call "ERGIC" couldn't it simply be a cis-Golgi cisterna?
      4. pg. 13: "propose to define Grh1, Rer1, and Sed5 as yeast ERGIC/GECCO...". What about Emp46?
      5. The first part of the manuscript (up to mid page 13) is clearly focused on defining ERGIC in yeast, then the paper appears as a set of experiments aimed at adding more components in their spatio-temporal mapping. This is ok, but is should be clearly motivated and explained in the Title, abstract and intro.
      6. The visualization of colocalization according to the opacity (as said in the methods) is somehow confusing to me. Are the 3D images projections or 3D renderings (no axes are seen)? In e.g. Fig. 6G or 8L, regions where green and magenta (or green and red) are colocalized do not appear white (or yellow), which visually suggests to the inattentive reader that there is no colocalization, when there is.
      7. I have not understood what this sentence in pg. 18 means: Similar segregation patterns are also observed during the Golgi-TGN maturation process (Tojima et al., 2019). "We propose that the ERGIC, Golgi, and TGN can coexist as structurally and functionally distinct zones within a single, maturing cisterna." Are they referring to ERGIC, Golgi, and TGN steady state components (proteins) or the structures themselves?
      8. The introduction of new data (mammalian data) in the discussion is odd. It might be ok, but I would frame it within a results section and use it later in the discussion.
      9. Fig.9: the arrows should go from protein to protein (some seem to go from in between proteins, such as the bottom-most arrow with 87.8 s time duration. Also in panel B, bottom part, some proteins are missing (Erd2 ad Chs5).
      10. Fig. 1 and many other: in the line profiles the distance in the x axis has no units or labels. Please add this and the direction of the line profile (an arrowhead would suffice).

      Significance

      General assessment:

      The experiments herein presented are based on a very careful, thorough and nicely organized spatio-temporal mapping of a large number of early secretory pathway proteins, including the ERGIC53 yeast "counterpart" Emp46. However, the data presented here does not convincingly show that there is a bona fide ERGIC in yeast. A major limitation is that the experiments are all based on a state-of-the-art, but still with a limited resolution, fluorescence microscopy technique. Ultrastructural data (e.g., CLEM) would massively help support or revisit the claims presented in this manuscript regarding the existence of an ERGIC compartment in yeast. Also, adding the information about the number of biological replicates and proper statistical analyses on the presented results would be needed to further support the claims.

      Advance:

      This manuscript builds on the authors' custom build 4D super-resolution microscope (SCLIM) and on previous results (e.g., Tojima et al., J. Cell Sci. 2019). The main novelty is in the study of a number of new early secretory pathway proteins and in the proposal of the existence of a non-stable, maturing ERGIC compartment in S. cerevisiae.

      Audience:

      This paper might be attractive for a broad audience of cell biologists, especially those interested in membrane biology, cell compartmentalization, and intracellular trafficking and secretion.

      Describe your expertise:

      I am an expert in membrane trafficking.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary

      Tojima and colleagues present a very exciting 4D SCLIM analysis of 20 key-proteins controlling or occupying different stages of Golgi-mediated protein trafficking, taking the reader on a trip from the ER export sites to the ERGIC/GECCO, cis-, medial- and trans-Golgi towards the TGN/recycling endosome. The choice of molecular markers to be patiently time-resolved in 3D allowed the authors to assemble a temporal roadmap for the molecular players studied. This is most impressive.

      Major comments

      There is an enormous amount of patient systematic analysis packed in this paper, following a punctate organelle as it emerges from the dark, evolves over time (following a combination of 2 markers) with fluorescence peaks at specific time points, after which the signal disappears again. I am certain other cell biologists will be impressed, as I was, viewing the individual images and graphs presented, culminating ultimately in figure 9 that could go straight into a textbook to form a starting point for anybody who wishes to study a particular protein of interest and chose the most appropriate markers to compare it with. The authors propose that the ERGIC/GECCO/Golgi-remnants compartment is an evolutionary conserved structure even though it has a different subcellular distribution/morphology in different classes of eukaryotes. The data presented here and in earlier work seem to support this notion. In particular, the authors demonstrate that the yeast GRASP 65 homologue Grh1 is the earliest to appear closely followed by Ypt1 and Emp46. The fact that RER1 and ERD2 come slightly later is in line with a proposed gate-keeper function, because if they were instead to recycle continuously they should appear first in line. I agree with the authors that the simple model of ER-derived COPII vesicles fusing with each other and thus creating an ERGIC/GECCO de novo is probably too simplistic. The idea of a more permanent structure, pulsating between cargo-loading and cargo-releasing events, possibly associated with creating zones/subdomains within a single cisterna seems very attractive given the data shown here. This work is descriptive, but it is of very high importance to anybody engaged with experimental approaches to study protein sorting from the Golgi-apparatus back to the ER, or on to the plasma membrane or the lytic compartments. The 5 functional stages proposed for Golgi-maturation is an attractive starting point for future research, and I very much like the notion that ERGIC and cis-Golgi cisternae may start as zones/subdomains within a single cisternae, possibly formed via phase separations involving both protein-protein and protein-lipid interactions.

      Minor comments

      The title strongly focusses on the ERGIC and therefore the earliest sorting steps in the ER-Golgi system, but this manuscripts offers so much more. I was fascinated to learn that Ypt1 appears twice during cisternal maturation in yeast. This may be a yeast-specific phenomenon but it is very interesting. The same can be said about the proposal that Gea1 and Gea2 have different roles in the Golgi and act in different cisternae, and the localisation of AP-3 at the trans-Golgi rather than the TGN. The functional distinction between trans-Golgi and TGN and the differences in their origin are important points and it will be a shame if readers don't realise that this manuscript offers further insight into later steps in Golgi-mediated transport. There may be a case to add something to the abstract and/or modify the title accordingly, but then I also feel that long titles are not ideal and the ERGIC/GECCO portion is the more important take-home message. This is a case for the editorial team and the authors to make the most of the findings.

      Given the importance of ERD2 in sorting soluble proteins to be returned back to the ER, the authors may consider using the biological active XFP-TM-ERD2 fusion instead of ERD2-GFP, but this may be kept for future work. In plants, ERD2-GFP is mainly at the Golgi when overexpressed, its erroneous leakage to the ER is only observed at low expression or when K/HDEL proteins are co-expressed. The XFP-TM-ERD2 construct may be better confined to the ERGIC-GECCO and may have a different temporal pattern.<br /> The authors may consider citing Stornaiuolo et al., Mol Biol Cell 2003 Mar;14(3):889-902 who compared the trafficking of KDEL and KKXX pathways and concluded that KDEL proteins are retrieved prior to KKXX proteins....as this fits nicely into the current findings showing that ERD2 and RER1 appear sooner than COPI markers.

      Significance

      The significance of the work is high because it basically allows to add facts to models. We use models to explain an elusive process because it escapes direct observation. Once we can observe directly, a model can become fact. In simple terms, the authors allow us to see things that we could only speculate about in the past. Therefore, the results present a very significant advance and will be highly relevant to the entire cell biology community. This paper is an important landmark and will help the field to formulate new experimental approaches and models to understand the origins of the Golgi apparatus, the core of the secretory pathway that defines being a eukaryote.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary:

      It is generally believed that budding yeast does not have the ER-Golgi intermediate compartment (ERGIC). In this study, the authors attempt to prove the existence of the ERGIC. They measured the kinetics of a few Golgi markers involved in the early secretory pathway in live cell imaging and observed the recruitment of Grh1 precedes that of Mnn9, suggesting the presence of pre-early cisternae. The authors propose that the Grh1-positive cisternae, assembling at the ER exit site (ERES) and progressing to become the early Golgi cisternae, represent the equivalent of the mammalian ERGIC in yeast.

      Major comments:

      The concept of the ERGIC in mammalian cells was initially proposed based on the protein ERGIC-53 in the 1990s. However, recent nanoscopy imaging data from the Lippincott-Schwartz lab challenges the conventional view of ERGIC by revealing the "ERGIC" is a membrane domain of the ERES (Wegel et al., Cell, 2021; PMID: 33852913), suggesting it might not be appropriate to adopt this concept.

      The authors' observations could be interpreted differently. Since the ERGIC is not molecularly defined in their study, the authors cannot prove its existence in yeast unequivocally. Their data indicate the presence of Golgi cisternae, characterized by Grh1, that precede the earliest known cisternae. Although the authors refer to these Grh1-positive cisternae as the "ERGIC", they are essentially "pre-early" cisternae that progress to become the early Golgi cisternae. Nevertheless, their findings could extend the budding yeast Golgi cisternal progression unit further upstream to include the ERES as the starting point for Golgi cisternal maturation. To further explore this, it would be interesting to investigate the kinetics of COPII subunits in cisternal progression along with Grh1 or Mnn9 and to plot COPII components in the Figure 9 map.

      The second half of the manuscript appears to deviate from the main focus on identifying the ERGIC. This section primarily presents the Golgi localization of four Golgi proteins (Ypt1, Gea1, Gea2, and Alp6) deduced from kinetics. However, it lacks functional studies to substantiate the authors' claims on their cellular functions. As a result, this part of the study remains purely speculative and might not support the authors' claims. Given that Figure 9 provides a highly informative summary of all kinetics and localization data, I recommend the authors keep but significantly abridge this section.

      The manuscript also has a few major concerns.

      1. The analysis of only one fluorescent particle or Golgi cisternal punctate structure is insufficient for a Golgi marker, considering the substantial variation of Golgi cisternae. To improve statistical robustness, the authors should select multiple fluorescent particles from multiple cells, displaying plots with averaged intensities, error bars, and sample sizes (n).
      2. In Fig. 5A, the BFA-induced lumps positive for Grh1, Rer1, and Sed5 may potentially represent the ERES, as observed in mammalian cells (Ward et al., JCB, 2002; PMID: 11706049). To verify this, the authors should co-label these lumps with COPII subunits.
      3. The authors previously reported the "hug-and-kiss" model for cargo transport from the ERES to the early Golgi cisternae. As the current study is highly relevant to the "hug-and-kiss" model, it is disappointing that the authors did not provide further data and comment on it. The "hug-and-kiss" and "ERGIC" transport modes are two distinct ways for secretory cargo transport from the ERES to the early Golgi cisternae. The authors should verify the "hug-and-kiss" transport and report the relative frequency of the two transport modes.
      4. The current version has minimal background knowledge of ERGIC in mammalian and yeast cells. Therefore, the authors should provide a comprehensive introduction to ERGIC.

      Significance

      General assessment:

      The data presented in the manuscript is novel and appears to be convincing. However, one of the major concerns is the lack of statistical robustness, which requires to be addressed. Furthermore, the manuscript's data could be interpreted differently, as elaborated in the major comments.

      Advance:

      While the interpretation of the manuscript's data requires reconsideration, it can contribute to our understanding of secretory trafficking at the Golgi. The manuscript could fill a crucial gap in our knowledge in this field by addressing the major comments.

      Audience:

      Cell biologists in the membrane trafficking field, particularly those working on the Golgi, would find the manuscript interesting.

      My expertise:

      My research focuses on membrane trafficking at the ER, Golgi, and endosome.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity):

      Unfortunately, this paper adds only a little to our understanding of uptake in to the flagellar pocket of trypanosomes. It tends to add only detail to information that has been well characterised elsewhere and indeed, as the authors themselves point out, (lines 92-98) it is rather incremental.

      We were disappointed that the reviewer was so unsupportive of the work presented here. It seems possible that the reviewer is partly objecting that the title - which emphasised the main finding of the paper - does not fully capture the content of the paper. We have therefore modified the title to emphasise that the paper is principally a characterisation of TbSmee1 rather than an investigation of the flagellar pocket, with the insight into cargo entry being the most notable finding.

      Not only has Tbsmee1 been studied before but this data in bloodstream forms is not particularly novel since it gives much the same information as the canonical hook protein TbMORN. This work follows the pattern of conclusions made previously with the protein TbMORN. It focusses on the protein TbSmee where RNAi mutants are interpreted to show flagellar pocket enlargement and impaired access by surface bound cargo. Unfortunately, there is little mechanistic or functional conclusion to the study in terms of how TbSmee operates naturally in the cell.

      This is deliberately downplaying the value of the work. TbSmee1 has not previously been characterised in bloodstream form cells, and neither TbMORN1 nor the hook complex are as well-characterised as other cytoskeletal components such as the flagellum and basal body. To criticise the paper for not providing a molecular mechanism of TbSmee1's function is unreasonable given the volume of work provided and the fact that this is a first characterisation of the protein in this life cycle stage. Expectation of a complete molecular mechanism is setting a very high bar for a first characterisation.

      It is also possible that the reviewer has not grasped the main thrust of the argument - when TbMORN1 was characterised it was the first protein shown to have this cargo entry defect. We show here that not only does TbSmee1 share this defect, but that it is in fact a previously-unacknowledged feature of all phenotypes of this type, exemplified by clathrin. We have modified the text to make this finding more clearly emphasised (see for example lines 654-661 in the tracked-changes version of the manuscript).

      There are other possible explanations for the phenotype. That would need to be studied. This large flagellar pocket phenotype is seen with RNAi mutants of many different types of proteins in the trypanosome and so pleiotropic effects are highly likely. Also, there are a good number of alternative possibilities to account for reduced access to the pocket in these mutants and this data could be usefully added.

      This is another statement that seems intended primarily to disparage the paper rather than attempt to improve it. It would have been extremely helpful if the reviewer highlighted what these other possible explanations are instead of making vague allusions. The widespread prevalence of this kind of phenotype means that our insight into restricted cargo access to the flagellar pocket is of general relevance in the trypanosome field.

      Specific points<br /> 1. The transient location for the TbSmee at the FAZ tip - or in this case the groove region - was seen in procyclics (Perry, 2018) so this bloodstream indication merely confirms that concept.

      The reviewer is again downplaying the value of the work rather than providing constructive criticism. While FLAM3 has been shown to be at the tip of the new flagellum in bloodstream form cells (Sunter et al., 2015), at the time of the preprint being published Smee1 was actually the first protein (besides the DOT1 antigen) shown to localise to the groove region in bloodstream form cells. It is also worth noting that procyclic form cells and bloodstream form cells are fairly different in this regard - in procyclic cells, there is an entire flagellar connector structure that is not present in bloodstream form cells, and so demonstrating that Smee1 was present in the groove region was an important experiment. Since this preprint was published, Smithson et al. have identified 13 additional proteins localising to the groove (Smithson et al., 2022) - we have modified the text to include these points (see lines 542-545 of the tracked-changes manuscript).

      1. The C terminal region required for targeting is a reasonable deletion analysis of regions of the protein. But can this data (line 228) be said to "mediate targeting" - or is it just required. For instance, targeting might be OK but it might be needed for stable association, etc etc.

      We have changed the text to say "required" for targeting instead of "mediating" targeting (line 312 of tracked-changes manuscript).

      1. This protein has already been shown to be phosphorylated and the sites and cell cycle possibilities have been mapped by Urbaniak. So that section adds little. https://doi.org/10.1371/journal.ppat.1008129

      The reviewer is again disparaging the significance of the work rather than critiquing it. This is after all only a single panel of a figure and ~15 lines of text, and therefore a minor but still noteworthy element of the manuscript. This also misunderstands what the Urbaniak study does and does not show - while that work showed that Smee1 is phosphorylated, it remained possible that other post-translational modifications were occurring. This experiment shows that the "fuzzy" appearance (variable electrophoretic migration) of TbSmee1 in gels can be solely attributed to phosphorylation as opposed to other post-translational modification. We contacted Dr. Urbaniak to confirm this - his answer is below.

      "__I think your approach to look at the fuzzy banding is actually rather elegant; our data shows that phosphorylation occurs but we did not look for any other PTMs that could influence migration on a gel and probably wouldn't see them without a different enrichment and analysis method. We often see a fuzzy pattern with glycosylation due to the heterogeneity, and I suspect other modifications will also results in a smear. Given that the band collapses to a single band after phosphatase treatment and not with an inhibitor present it is fair to conclude that phosphorylation is responsible for the fuzzy band, not other undefined PTMs like glycosylation.__"

      1. Essentiality in BS forms and pocket enlargement. This is not surprising. A very large number of cytoskeletal proteins show this in RNAi knockdown. Flagella mutants (extensive publications from many groups (Hill, Bastin, Gull, etc) over last 15 years show this very well and so this protein is just one more example.

      This appears to be another comment aimed at downplaying the value of the manuscript rather than providing constructive feedback. The fact that we have demonstrated something previously unobserved in a common phenotype makes the data of general interest to the community, we feel.

      1. I didn't find that the explanations for flagella pocket enlargement are soundly based. The experiments focus on endocytosis and uptake and ignore other plausible reasons and some evidence in literature.

      Again, the reviewer's feedback would be considerably more constructive if they had taken the time to specifically cite the evidence in the literature that they are alluding to, and present some of the "other plausible reasons" they are aware of. We have consulted widely in the community and have not been able to find anybody who knew what work the reviewer is referring to here.

      Lines 84/85. Enlarged pockets may be indicative of endocytosis failure. Presumably the rationale is that endocytosis fails, but exocytosis still occurs and the pocket membrane enlarges. What evidence is there that exocytosis of membrane still occurs? This simple concept might indeed operate in a clathrin mutant but is surface membrane/content exocytosis is maintained in these cytoskeleton mutants? There is good evidence for glycoconjugates within the flagellar pocket. Are these depleted or present still?

      The reviewer is correct that we have not specifically assayed for exocytosis, but the fact that we are able to make the same observations in both the clathrin RNAi (where exocytosis has been assayed - Allen et al., 2003) and the Smee1 RNAi means that this is not a problematic omission. The effect of the enlarged flagellar pocket phenotype on the glycoconjugates in the flagellar pocket is an interesting question but far outside the current focus of the paper.

      1. There are also a number of other publications indicating that clathrin pits are still present on the enlarged pockets of various mutants when viewed by EM. The authors have looked at the flagellar pockets by EM but the EM methods described have extensive washings and centrifugations before fixation. This is a very poor approach and will mean that endo and exocytic traffic is disturbed (extensive references in literature in other systems? This is not a useful approach for exo/endocytosos studies where flux of traffic demands fast chemical or freezing fix in media.

      The reviewer has misunderstood the aim of the experiments described in Figure 5D, which was to observe the morphological changes caused by depletion of TbSmee1. As the reviewer is no doubt aware, high-pressure freezing of trypanosomes gives much better morphological preservation than chemical fixing in media, so the choice of method is not "very poor" but tailored to the experimental aims. We have modified the text to make this point more clearly (lines 355-358 of tracked-changes version). Once again, the referee offers no citation to back up their assertion that endo- and exocytic traffic is disturbed by wash steps, either in trypanosomes or elsewhere.

      1. The EMs and Light microscopy does show that the mutant pockets are substantially abnormal in their cytoskeletal arrangement. They have multiple flagella profiles, flagella structures have not connected with the membrane and are sometimes in the cytoplasm (see a glance of the paraflagellar rod in the cytoplasm in FigS5C and internalised FAZ attachment plaques in Fig 4 D bottom right cell). Given these extensive (and expected) cytoskeletal abnormalities it is highly likely that these pocket abnormalities are a result of motility, cell division/developmental issues and the differential uptake phenotypes merely consequential.

      This is another misinformed argument that is seeking to disparage the data. The reviewer has apparently overlooked the fact that the same phenotype is seen in clathrin RNAi, when flagellar pocket enlargement precedes any downstream effects on cell division cycle progression. We have gone to great lengths (Fig 6) to demonstrate that the enlargement of the flagellar pocket almost certainly precedes the onset of the growth defect in the TbSmee1 RNAi, and it is therefore likely to precede the cytoskeletal abnormalities that the reviewer has highlighted. An effect on cellular motility is possible and would be interesting to investigate in future work.

      1. The authors speak about early phenotypes , but these are often at 15-24 hours. That is probably a couple of cell cycles and so not early.

      To be informative, the analyses of RNAi phenotypes have to be done as soon as possible after the onset of the growth defect, and we have gone to great lengths (Figure 5) to define this point as being at 21 hours. This is already difficult as the number of phenotypic cells at the onset of the growth defect will not be high. We have clarified the text to emphasise that "early" refers to soon after the onset of the phenotype (lines 388-389 of tracked-changes version).

      In relation to the above question of comparison to the same morphology produced by flagella mutants it would be good to know if these hook mutants produce motility phenotypes and whether these are manifest before the uptake phenotypes. There is evidence (cited here) that forward motility of the trypanosome directs material on surface into the pocket. If these cells have motility defects (primary or via failed division) then surely that would provide an alternative simple explanation for uptake differences.

      The reviewer is overlooking the observation that the surface-bound endocytic cargoes (ConA, BSA) are still being sorted/directed as far as the entrance to the flagellar pocket - what is interesting is that the cargo is apparently unable to enter the flagellar pocket. As noted above, it would certainly be interesting to look at motility effects in follow-up work.

      1. There is a general point that if studies are to have real relevance to uptake in the trypanosome then they need to deal with uptake of natural ligands rather than artificial surrogates such as dextran. Such tracers were used historically, but in the last decade a series of receptors and ligands for fluid phase and particularly membrane mediated endocytosis have been discovered. With the investment of a little time these important ligand / receptors such as haptoglobin, transferrin, etc would be much more relevant.

      Dextran is still state-of-the-art as it is an inert fluid phase marker. We are not aware - and have asked widely - of any readily-available alternative to dextran as a fluid phase marker, especially seeing as we have demonstrated in this study that BSA does not behave as a fluid phase marker in the experimental conditions used. The reviewer is also being disingenuous in suggesting that there is a panel of validated physiological reporters for trypanosomes that are readily available commercially - this is not the case. Transferrin is probably the only example, but the transferrin receptor is confined to the flagellar pocket and therefore not relevant to the question of how surface-bound material enters the flagellar pocket in the first place. As suggested by Reviewer 3 and endorsed by Reviewer 2, we have looked at the uptake of anti-VSG antibodies (which are a physiological cargo) in additional experiments and obtained evidence that the same effects are seen (Figure 9).

      **Referees cross-commenting

      this session includes comments from Reviewer 1 and Reviewer 2.<br /> *

      Reviewer 2<br /> <br /> Dear Reviewers 1 and 3:<br /> I agree with many of the points with Reviewer 1 and our divergence is partly a matter of degree. While it is true that this manuscript is incremental in its contribution to our understanding of TbSmee1, it nonetheless adds to our understanding of the role of this protein in the bloodstream life stage and because of that I find value in the work. The fact that it mirrors what was seem in other protein knockdown studies (e.g. TbMORN) doesn't negate its contribution for me. Reviewer 1 makes an important point, however, when stating that this work does not add a mechanistic or functional conclusion as to how TbSmee1 operates and for me that is the biggest shortcoming of the work. Offering mechanistic insight is a high bar and while it would make for a much more exciting story it does not discount the value of the work as presented. What I do appreciate is the speculation about this observation that endocytosis is required for entrance of surface bound material into the pocket and although they are unable to show that this is not a side affect of other processes being disrupted it is and intriguing point. These observation have the potential of stimulating further investigations into crosstalk between the entrance to the pocket and endocytosis. I also agree that the use of ligands for known receptors like transferrin would be far more informative. While I assumed the transferrin receptor was in the pocket itself it would be interesting to see if the ESAG6/7 is also located outside the pocket and transiently binds cargo before being brought inside for endocytosis.<br /> I think that Reviewer 3 brings up a great point with the focus on VSG's. I think that examining VSG turnover in these mutants can add value to the analysis and inform our view of how affecting the hook complex alters VSG endocytosis.

      We appreciate Reviewer 2 taking the time to defend the value of the work, and we concur with Reviewer 2's assessment. Reviewer 2 is also correct that the transferrin receptor appears to be primarily or wholly confined to the flagellar pocket interior, making this likely less informative in this context. Concerning the uptake of anti-VSG antibodies highlighted by Reviewer 3 and endorsed by Reviewer 2, we have carried out these experiments and obtained similar results to those published in the first version of the preprint (Figure 9).

      Reviewer 1<br /> <br /> some fair comment and agreement. This is being sent to general cell biology journals.<br /> when one looks at this area in the round it is it is nearly 50 years (1975) since Langreth and Balber published their seminal work on protein uptake and digestion in bloodstream and culture forms of T. brucei. There has been 50 years intense study and the genome has been around for nearly 20 years as well. So, put simply - for both a general science audience and the wider parasite community - if this is a paper about one protein, TbSmee1,then it has surely has to say something functional about that protein. If it is a paper about uptake in trypanosomes (where mutants are one means of interrogation) then it surely has to say something about mechanisms of uptake of physiological relevant ligands. The days of dextran etc are past.

      Hence, my comment that this does neither and so is very incremental to what is known already. It is 2022 not 1975. Langreth and Baber published their seminal work in J Protozoology for very good reasons no doubt.

      It is striking that Reviewer 1 here extends their aggressive and uncivil approach to attack Reviewer 2's assessment, again substituting forceful wording for informed argument. Reviewer 1 again inexplicably and mistakenly criticises the use of dextran when no state-of-the-art alternative exists. They then go on to needlessly disparage the work done by Langreth & Balber when this work was produced in a totally different publishing landscape. They also appear to fundamentally misunderstand the Review Commons concept, which is to provide journal-independent preprint peer review; it is also worth noting that there are specialist journals such as PLoS Pathogens in the RevComm affiliates as well as general cell biology journals. Given that the mechanism of variant surface glycoprotein (VSG) switching has not yet been fully articulated despite the efforts of multiple labs and many projects over a decades-long time period, it seems extremely unreasonable to be making such demands of this paper.

      Reviewer 2<br /> Thank you for replying and I agree with the spirit of your critique. My only comment, which could result from my own naivete, is to say that despite the incredible work that has been done in dissecting endocytosis in T. brucei over these past 50 years, it appears that we still do not understand how many fundamental of aspects of this activity works in this parasite. Even basic questions regarding how cargo, e.g. transferrin, binding to surface receptors is sensed by the parasite remains unknown and the identity of the specific signaling components which transmit this information internally to initiate endocytosis have not been characterized. In many ways it seems that we don't even understand how the parasite partitions the end/exocytic pathways in the pocket and maintains membrane homeostasis. While we know that some kinases and traditional signaling components must be involved, a high resolution understanding of this process in T. brucei seems lacking. I only say all this to suggest that the field maybe isn't yet that advanced to reject work of this type as so many mechanistic unknowns still remain to be uncovered and maybe incremental advances and phenomenology still can add value to the field. However, I respect your opinion on the matter and my perspective could be due to a lack of a full appreciation of the literature on the subject.

      We completely agree with Reviewer 2's assessment here, which neatly summarises our rationale for the present work. Reviewer 2 is, if anything, being overly accommodating by suggesting that their perspective may be due to a lack of a full appreciation of the literature - on the contrary, Reviewer 2 appears to have a very sound grasp of the topic.

      Reviewer #1 (Significance):

      Unfortunately, I did not find tis to be very significant. It covers old ground in terms of the phenotype described. Many groups have shown the differences between procyclic and bloodstream phenotypes in this enlarged pocket phenomenon. The work is rather incremental from these and other author's work on these hook proteins.<br /> There are alternative explanations for understanding the effect of flagella pocket structure and uptake of ligands into the pocket and trypanosome cell. These would need to be studied before one could see a functional, mechanistic link established.<br /> Other parts of this are of nicely done but do not move on our understanding (eg targeting/phosphorylation) from what has been done previously.

      As noted repeatedly, it appears that Reviewer 1's priority is disparaging the value of the work here and downplaying its significance rather than providing constructive feedback. The reviewer repeatedly makes unrealistic demands (a mechanistic model, use of non-standard reagents), misunderstands the aim of experiments (use of high-pressure freezing), makes vague allusions to other work in the literature but without citing anything specific to support their case, and makes strong and assertive statements that are factually incorrect (design of RNAi experiments, use of dextran). We find this approach unhelpful, uncivil, and unprofessional. It is desperately disappointing that we should have to spend the majority of our response rebutting Reviewer 1's comments rather than implementing constructive criticisms that would strengthen the manuscript.

      Reviewer #2 (Evidence, reproducibility and clarity):

      Summary:<br /> In this manuscript the authors have advanced our understanding of the hook complex component TbSmee1 through a detailed analysis of this protein's role in the endocytosis of surface bound proteins via the flagellar pocket in bloodstream form Trypanosoma brucei. The TbSmee1 protein, previously identified using proximity labeling using TbMORN1 and TbPLK, and characterized in procyclic T. brucei, was confirmed to target to both the shank portion of the hook complex as well as the growing end of the new FAZ in replicating cells. The protein was also shown to likely be phosphorylated as had been suggested previously due to its association with the kinase TbPLK. A domain deletion analysis demonstrated that domains 2 and 3 are important for TbSmee1's proper localization to the hook complex. Loss of TbSmee1 using RNAi based knockdown resulted in a quick cessation of growth in the bloodstream form within 24 hours in contrast to what was seen previously in procyclic cells which had only a decreased growth rate. Loss of TbSmee1 also resulted in an enlargement of the flagellar pocket and in many ways mirrored the phenotype observed with knockdown of TbMORN1. Although prior work on TbSmee1 in procyclic T. brucei demonstrated that loss of this protein altered the morphology of TbMORN1, no such change was seen in bloodstream form cells and only an alteration in the morphology of TbLRRP1 was observed. In characterizing the effect of TbSmee1 depletion on endocytosis the authors showed that the fluid phase marker Dextran could enter into the flagellar pocket of TbSmee1 depleted parasites while the surface bound ConA and BSA remained outside of the flagellar pocket suggesting that TbSmee1 may play a role in allowing larger protein components into the pocket regions. Similar observations were also previously seen with TbMORN1 depletion. Importantly, a knockdown of clathrin recapitulated the TbSmee1 knockdown phenotype suggesting that endocytosis itself was required to allow material bound at the surface to enter into the flagellar pocket. In addition to adding to our understanding of hook complex components, this work raises some interesting questions regarding the role of the hook complex in facilitating endocytosis in this important human pathogen.

      Thank you for the positive assessment.

      Major Critiques:<br /> This is a superbly written manuscript with robust high-quality data that strongly support the major conclusions made by the authors. The flow the article is logical and easy to follow making it accessible to a wide array of readers.

      We are glad that the Reviewer appreciated the effort that went into writing the paper.

      Although I appreciate the brevity of the introduction and how the article gets straight to the point, additional background information on the components and function of the flagellar pocket collar protein could help contextualize the goals of the project. The way in which the flagellar collar structures are introduced to the reader is quite abrupt (beginning on line 75) and simply states the names of TbBILBO1, the centrin arm and hook complex as simple facts without much discussion about the background of these components/regions. A graphical representation of the centrin arm or hook complexes relative to other components like the pocket itself, FAZ or axoneme could make following the story much easier. An expansion of this background could also go a long way to convince readers of the importance of this region in the basic biology and virulence of T. brucei.

      Implemented. We have added more background details on the hook complex, flagellar pocket collar, and centrin arm and added a new schematic image to Figure 1 showing these structures as well as the FAZ (Figure 1A).

      On lines 84-86 the authors cite the way in which 'small' vs 'large' macromolecules enter into the pocket without defining what exactly is meant by these terms as they are relative in nature. Setting some boundaries of size could provide some context to the reader.

      Implemented. We have provided more detail on the approximate sizes in nm (lines 110-113 of tracked-changes manuscript).

      In the domain localization analysis beginning in Figure 4 there is a missed opportunity to also assess which portions of the TbSmee1 protein are important for overall function as well. By either an examination of dominant negative phenotypes resulting from overexpression of the truncated mutant or the expression of the truncated forms designed to be RNAi resistant in the TbSmee1 knockdown cell line, one could also assess which portions of this protein are essential for endocytic function in addition to targeting. Is there a reason this was not performed?

      This is a good point; we did actually investigate overexpression of the TbSmee1(161-766) construct which can target correctly but is missing the first folded domain, but did not observe any phenotypic effects. We have added this point to the results (lines 301-302 of tracked-changes version). We agree that it would be interesting to express the truncations in a TbSmee1 RNAi background in order to simultaneously assay for targeting and function, but this was (unfortunately, perhaps) not part of the original experimental design. To do so now would require generating a completely new panel of truncation constructs with recoded DNA (in order to make them RNAi-resistant) and then generating a new panel of cell lines. While this would be informative, we feel that it would be impractical at present.

      In the analysis of viability changes due to TbSmee1 depletion (lines 237) the authors state that at "72 h post-induction showed widespread lysis, ..." This phenotype seems inconsistent with other related endocytic defect mutants. There is no further mention of this lysis phenomenon here or in the discussion and considering how unique this seems it deserves either additional data to demonstrate or further discussion as to the basis of the phenotype. It seems, at least from this study of TbStarkey1 and prior studies which result in the enlarged flagellar pocket phenotype, that having an enlarged pocket is not the cause of lysis and doesn't even naturally lead to a growth defect.

      Widespread lysis is the usual outcome of bloodstream form cells with strong endocytic defects - we have observed this directly for the clathrin, TbMORN1, and TbSmee1 RNAi cell lines, and it has been documented in a number of other publications (see for example Natesan et al., 2010, Manna et al., 2017). We have clarified this point in the text (see for examples lines 359-341, 474-478 of tracked-changes manuscript).

      The authors do not comment on what is the source for the cessation in growth following TbSmee1 knockdown. Is it nutrient depravation like in other endocytic defect mutants?

      Implemented (see for example lines 359-361, 605-610 of the tracked-changes manuscript). The source of the growth defect is likely to be due to impaired cell division cycle progression due to the gross enlargement of the flagellar pocket and subsequent steric hindrance and imbalance of membrane homeostasis.

      In the end, one of the most interesting observations made by the authors is that loss of TbSmee1 inhibits endocytosis and this has the appearance of not allowing large molecule substrates like ConA and BSA to enter into the flagellar pocket. This appeared to have nothing to do with a gatekeeping type function of the hook complex/flagellar collar and instead, as shown through clathrin knockdown, was related to the ability of the parasite to endocytose. There are a lot of potential interpretations of this phenomenon with one being a simple perturbation of the normal membrane trafficking to and from the flagellar pocket being involved. An analysis of knockdown of exocytic components might reveal whether or not this inability to enter into the pocket is also seen when exocyst proteins are also depleted. It may be impossible to tease apart these two interrelated activities but it might eliminate one side of the equation if these proteins can still enter the flagellar pocket when exocytosis if perturbed although this reviewer understands that that dimension of T. brucei membrane trafficking is poorly understood relative to endocytosis.

      This is an interesting point, and the reviewer is also correct in highlighting that exocytosis is far less characterised than endocytosis in Trypanosoma brucei. The exocyst has been characterised in bloodstream form T. brucei (Boehm et al., 2017) and shown to also have a role in endocytosis, so teasing out the relative contributions of these pathways would undoubtedly be challenging. We would prefer not to go in this direction in this present study, but it is an obvious avenue for future work.

      An intriguing possibility that the authors allude to and which if answered would make this manuscript have a far broader appeal is to determine if loss of TbSmee1 alters the lipid kinase distribution and if this is the source of the negative impact on endocytosis. One important dimension of endocytosis in T. brucei which remains poorly understood is the role of signaling machinery in triggering endocytic events. It is possible that the hook complex serves as the gatekeeping or signaling platform that recruits signaling components (like lipid kinases) that identify and/or modify the membrane lipid phosphatidylinositols harboring cargo laden receptors thus marking them for endocytosis within the pocket. It still seems unclear when in the process of endocytosis is the decision made to pull things into the pocket but it seems that the assumption is that this occurs deep within the pocket. This data suggests that there is possibly another decision point prior to being allowed entrance into the pocket. It may be that this isn't a gatekeeping decision but rather a stop vs. go activity where once cargo laden membrane reaches the collar a choice is made to pull this material in or not there and not after material is already in the pocket.

      These are all really interesting ideas and would be fascinating topics for future work.

      This obvious enigma based on the observation that loss of hook complex components affect the spatially separated site of endocytosis support the idea that the actual endocytic signaling platforms are located at the hook complex and that this area may make the membrane modifications that mark membrane as being ready to be endocytosed via clathin coated vesicles at the bottom of the pocket. This would still allow for fluid phase small molecule entrance which does not require binding to surface proteins. The obvious problems of having both endo/exocytosis occurring in the same close proximity makes the dissection of this phenomenon difficult but it is worth potentially expounding on further in the discussion as this idea is very appealing and adds an important dimension to our understanding of endocytosis in this organism.

      Implemented (lines 722-727 of the tracked-changes manuscript). We have added some more detail to these points in the Discussion. We agree with the reviewer that there are some profoundly interesting questions concerning membrane identify and membrane protein uptake here.

      Minor Critiques:<br /> The authors commit significant time to the analysis of the phosphorylation of TbSmee1, but there is little stated about the role of TbPLK in this activity or the potential connection of TbSmee1 phosphorylation to the cell cycle. Would a knockdown of TbPLK using RNAi potentially demonstrate an altered migration of TbSmee1 due to a lack of phosphorylation? An analysis of radiolabeled TbSmee1 using p32 in vivo would likely support this claim as well. Has mass spectrometry identified potential phosphorylation sites to examine? Additionally, the loss of TbSmee1 has been shown to disrupt localization of TbPLK in procyclic cells and so why this was not also assessed in bloodstream form cells subjected to RNAi was not clear.

      Partly implemented. We have added some discussion of the possible role of TbSmee1 phosphorylation in the cell cycle to the Discussion (lines 562-565 of tracked-changes manuscript), and emphasised the identification of phosphorylation sites in previous phosphoproteomics work (citations of Nett et al., 2009, Urbaniak et al., 2013). Given that the strongest and earliest effect of TbSmee1 depletion was on endocytosis and cargo uptake, we chose to focus on this angle rather than exploring its contribution to the biogenesis of cytoskeleton-associated structures and its interaction with TbPLK. For that reason we would prefer not to carry out the experiments looking at the effects of TbSmee1 depletion on TbPLK or vice versa.

      In the results section (lines 104-108) a model of the protein structure as predicted for example by AlphaFold might be informative and complement the domain analysis work depending on the quality of the prediction.

      Implemented. The AlphaFold prediction is consistent with the predictions made by the other structural analyses, and we have noted this in the text (lines 145-148 and 551 of the tracked-changes version).

      There is an arrow in the Figure 1B Western blot but I can find no mention of what it is trying to highlight in the text.

      Corrected.

      For Figure 1D there is no loading control or control for the distribution of the soluble fraction to validate the separation of the two compartments.

      Implemented. We have carried out additional experiments to show the partitioning of a cytoplasmic protein (the endoplasmic reticulum chaperone BiP) into the detergent-soluble fraction. These results are now displayed in the updated Figure 1.

      The authors fail to comment on the lack of changes in hook complex components they see to that observed by Perry et. al. 2018. This difference merits some minor comment or speculation.

      Implemented. We have added this commentary to the Discussion (lines 592-600 of the tracked-changes version).

      Line 228: domain should be capitalized.

      Implemented.

      Line 230: FigS5C should have a space and period after Fig. and S5C.

      Implemented.

      Line 244: "on" should be inserted in the sentence "...TbSmee1 protein depletion ON either side of the onset..."

      Implemented.

      Line 400: the '...20/21 h post-induction...' is slightly confusing and may read better as 20-21 h.

      Implemented.

      Line 463: a space is needed between '...2009).The...'.

      Implemented.

      Reviewer #2 (Significance):

      This manuscript advances our current conception of endocytosis in T. brucei. Although this model kinetoplastid parasite has been extensively studied with respect to endocytosis there is still a great deal we do not yet understand regarding how this process is regulated at a mechanistic level. This work has begun to connect previously unappreciated aspects of endocytosis in T. brucei by highlighting a potentially novel connection between the flagellar collar/hook complex and the physically separated endocytic events within the flagellar pocket itself. It may be that what appears as regulated entrance into the pocket is in fact the source of signaling that triggers the endocytic events carried out by clathrin. This is an interesting notion that no doubt requires further investigation which lies outside of the scope of this report. While this work appeals primarily to those studying kinetoplastids parasites it has the potential to provide insight into basic protozoan biology as well. Due to my related interest in kinetoplastid endocytosis, I find this work to be of high quality, conceptually interesting and employs many of the cutting-edge techniques currently available in the study of T. brucei.

      We are very happy that the Reviewer formed a favourable impression of the work.

      Reviewer #3 (Evidence, reproducibility and clarity):

      This manuscript begins to dissect the function of the hook complex protein SMEE1 in the mammalian infective form of T. brucei. The hook complex is a cytoskeletal structure associated with the flagellar pocket, the only site of endo/exocytosis in these cells. The authors demonstrate that SMEE1 is required for endocytosis in these cells and that this can occur with minimal change to the molecular make-up of the hook complex. The authors show that endocytosis is important for the access of large molecules e.g. ConA into the flagellar pocket.

      Major comments

      The key conclusion of this study are convincing and the data is generally well presented and clear. The interpretation of the figures matches well with the data presented - there are a few minor issues though that I have highlighted below in minor comments. The authors use a range of molecular cell biology approaches to define the role of SMEE1 and these are appropriate and are well controlled.

      Thank you.

      My major comment focuses on the use of different tracers to study endocytosis but the elephant in the room is what is happening to VSG as this is the surface protein that needs to rapidly removed from the cell surface and cleaned. Given the importance of removal of antibodies bound to the VSG - have the authors looked at this in the SMEE1 depleted cells? Do VSG-antibody complexes accumulate in this region? This is an important experiment as this would give key physiologically relevant data to this study. All the material should be readily available for this as there are a number of VSG antibodies.

      We agree with the Reviewer that the behaviour of these VSG-bound antibodies is a key test of the physiological relevance of the observations we have made using ConA and BSA, and have implemented this request - the results are in the new Figure 9. Although they sound simple, these assays turned out to be far from trivial and much more technically challenging than the other uptake assays, owing to the extremely fast kinetics (seconds) of anti-VSG uptake (Engstler et al., 2007) and the unexpectedly and incredibly high losses of bound antibodies during the assay. This might be due to shedding, as noted in the Discussion.

      Minor comments<br /> Perhaps I have been overthinking this but is surface-bound the right way to describe the cargo, as it clearly goes in both directions onto and off the surface and in fact the experiments in this manuscript are focussing on the removal of this material from the surface so is not surface-bound.

      We have clarified that "surface-bound" refers to material that binds to the surface glycoprotein coat of the trypanosomes and which is subsequently internalised, not material that is bound for (i.e being directed to) the cell surface (lines 77-78 of tracked-changes version). We hope this addresses the Reviewer's point?

      Have the authors investigated the structure of the protein using alphafold and if so how does that compare to the domain structure that was presented in this manuscript?

      Implemented (lines 145-148, 551 of tracked-changes version). We have checked the AlphaFold prediction of the three-dimensional structure of TbSmee1 and noted it in the Results; the prediction is consistent with the earlier bioinformatic analyses.

      The authors raised a number of antibodies to TbSMEE1 and TbSTARKEY1 but it was not clear in the figures which antibody was ultimately used for analysis by western and IF - could the authors clarify, as some looked to have a higher background than others. Line 150 states the same localisation was seen for all three antibodies and references S3C but I couldn't see that data presented.

      Implemented - the 304 antisera was used for most subsequent experiments and we have noted this in the M&M (lines 793-798 of tracked-changes version). Figure S3C shows that the Ty1-TbSmee1 recapitulates the localisation of the antibodies against the endogenous protein - we have clarified this point as well (lines 206-207 of tracked-changes version).

      Line 169 - can the authors provide more detail about the global correlation methodology as I was unable to follow the details in the methods? Is this a pixel per pixel correlation over the image or on a selected region over the area of potential signal overlap? In figure 2E it appears that BILBO1 signal correlates more closely with the SMEE1 signal than MORN and LRRP1 and from the images that would not seem to be the case. Have I interpreted this figure incorrectly?

      Implemented. The original analysis was a global correlation analysis that was determining whether the signals were correlated with each other regardless of spatial overlap, and we agree with the reviewer that these outputs were non-intuitive to interpret. In the revision, we have carried out a new analysis (and updated the accompanying text and M&M section), measuring the degree of spatial correlation between each pair of signals on a pixel-by-pixel basis over the area of each cell, with a total of 30 cells analysed in each pairing. We believe that this addresses the reviewer's point. See lines 223-243, 963-974 of the tracked-changes version).

      The authors have generated a number of different clones and performed experiments on these clones generally more than twice, which is clearly explained in the figure legends but in places the data is then put together and it is difficult to know which experiments/clones it comes from - for example 7C/7F what do those percentages represent? Is this the sum of all experiments? A representative experiment? How many cells per experiment were analysed?

      Implemented. We have double-checked all the figure legends and clarified this point where necessary. Quantifications were always made by compiling data from multiple independent experiments using multiple separate clones - see in particular lines 1323-1324, 1363-1365, 1380-1382 of the tracked-changes version.

      Line 200 - From the image it is not convincing that SMEE1 is slightly behind DOT1 - I agree it looks enveloped but would appear level with the distal end of the DOT1 signal.

      Implemented. We have adopted the Reviewer's wording for this text (line 271 of tracked-changes version).

      For the truncation experiments the authors should explain that these are performed with cells in which the endogenous SMEE1 will be expressed and this may influence the localisation of the truncations, especially as there is no information about whether SMEE1 forms complexes with itself or other proteins.

      Implemented (lines 296-298 of tracked-changes version).

      Figure 4D - should be 1 not T-

      We have relabelled this as "TbSmee1". The values in this column are the immunoblot signal intensities obtained for the endogenous TbSmee1 protein in the -Tet condition. We have also clarified this in the figure legend.

      Line 223 - given the low expression of constructs 2 and 9 I'm not sure it is possible to infer anything from the lack of localisation of these constructs as they appear unstable and would be unlikely to localise to a specific location.

      We have added this caveat to the text (lines 558-562 of tracked-changes version).

      Figure S7 - The images presented were not convincing that there was a reduction in the localisation of LRRP1 to the hook complex on depletion of either SMEE1 or MORN1. The difference looks particularly minor if present at all.

      Agreed, there was some debate in the group about these results. We have changed to text to fit the Reviewer's interpretation (lines 347-348 of the tracked-changes version).

      Line 264 - "implied that the lethal phenotype might be due to a loss of function" - this seems an odd thing to say as it doesn't provide any insight as of course the phenotype is due to a loss of function.

      We have clarified this point (lines 350-353 of the tracked-changes version). We would however disagree with the reviewer that RNAi phenotypes are exclusively due to a loss of individual protein's function(s) - when proteins are present in multiprotein complexes (as is often the case with cytoskeleton-associated proteins), then destabilisation of the complex due to loss of the entire protein can cause the observed phenotype, rather than the loss of the function performed by the individual protein within the complex (this may be a semantic point, however). A very good example of this is with the outer arm dynein complex component LC1 (Ralston et al., 2011) - RNAi against LC1 is lethal because the entire outer arm dynein complex is destabilised, whereas expression of non-functional mutants of LC1 produces viable cells with motility defects due to the specific loss of LC1 function.

      Line 412 - can the authors clarify what they mean by geometric problems?

      Implemented (lines 605-610 of tracked-changes version). We were referring to the fact that enlargement of the flagellar pocket will probably create difficulties for the progression of the cell division cycle.

      Throughout the manuscript can you use log scale for the growth curves.

      Implemented.

      Line 756 - add citation

      Whoops! Implemented (line 1058 of tracked-changes version).

      Line 465/66 - the authors states that the ability of the fluid phase cargo being still able to enter the pocket is evidence that the channel lumen is still open; however, I would think that despite the close apposition of the cell membrane to the flagellar membrane in the flagellar pocket neck region this would be unlikely to impede fluid/soluble material from entering the pocket, as presumably VSG protein can move through this region. This does not alter the ultimate conclusion the authors are drawing but without microscopy evidence for the state of the channel lumen it is difficult to be sure of its status.

      Fair point. We have modified this statement (line 701 in tracked-changes version).

      Reviewer #3 (Significance):

      The flagellar pocket is the key portal into and out of the trypanosome cell and as such has a vital role to play in host-parasite interactions. The flagellar pocket is supported by a number of cytoskeletal structures including the hook complex and the role of these structures in flagellar pocket function are poorly understood. The flagellar pocket is particularly important in the bloodstream form of the trypanosome parasite which infects the mammalian host as it is the route for the surface protein VSG to get onto and off the surface. The VSG is required for antigenic variation and the removal of VSG-antibody complexes helps 'clean' the surface of the parasite. SMEE1 is a component of the hook complex and the manuscript here dissects its role in the mammalian infective parasite and shows that it is vital for the endocytosis of material off the surface. Intriguingly, a block in endocytosis causes a blockage of material outside of the pocket, suggesting a multi-step process in the regulation of uptake of material from the parasite's surface.<br /> This manuscript will be of specific interest to those researchers investigating the long-term persistence of these parasites in the mammalian host. There are potentially some insights into the control of membrane domains for endocytosis that are of interest to more general cell biologists as well.

      We are very grateful to the reviewer for the supportive comments and the constructive evaluation. Many thanks!

      Expert in molecular cell biology of trypanosomes and Leishmania.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      This manuscript begins to dissect the function of the hook complex protein SMEE1 in the mammalian infective form of T. brucei. The hook complex is a cytoskeletal structure associated with the flagellar pocket, the only site of endo/exocytosis in these cells. The authors demonstrate that SMEE1 is required for endocytosis in these cells and that this can occur with minimal change to the molecular make-up of the hook complex. The authors show that endocytosis is important for the access of large molecules e.g. ConA into the flagellar pocket.

      Major comments

      The key conclusion of this study are convincing and the data is generally well presented and clear. The interpretation of the figures matches well with the data presented - there are a few minor issues though that I have highlighted below in minor comments. The authors use a range of molecular cell biology approaches to define the role of SMEE1 and these are appropriate and are well controlled.

      My major comment focuses on the use of different tracers to study endocytosis but the elephant in the room is what is happening to VSG as this is the surface protein that needs to rapidly removed from the cell surface and cleaned. Given the importance of removal of antibodies bound to the VSG - have the authors looked at this in the SMEE1 depleted cells? Do VSG-antibody complexes accumulate in this region? This is an important experiment as this would give key physiologically relevant data to this study. All the material should be readily available for this as there are a number of VSG antibodies.

      Minor comments

      Perhaps I have been overthinking this but is surface-bound the right way to describe the cargo, as it clearly goes in both directions onto and off the surface and in fact the experiments in this manuscript are focussing on the removal of this material from the surface so is not surface-bound.

      Have the authors investigated the structure of the protein using alphafold and if so how does that compare to the domain structure that was presented in this manuscript?

      The authors raised a number of antibodies to TbSMEE1 and TbSTARKEY1 but it was not clear in the figures which antibody was ultimately used for analysis by western and IF - could the authors clarify, as some looked to have a higher background than others. Line 150 states the same localisation was seen for all three antibodies and references S3C but I couldn't see that data presented.

      Line 169 - can the authors provide more detail about the global correlation methodology as I was unable to follow the details in the methods? Is this a pixel per pixel correlation over the image or on a selected region over the area of potential signal overlap? In figure 2E it appears that BILBO1 signal correlates more closely with the SMEE1 signal than MORN and LRRP1 and from the images that would not seem to be the case. Have I interpreted this figure incorrectly?

      The authors have generated a number of different clones and performed experiments on these clones generally more than twice, which is clearly explained in the figure legends but in places the data is then put together and it is difficult to know which experiments/clones it comes from - for example 7C/7F what do those percentages represent? Is this the sum of all experiments? A representative experiment? How many cells per experiment were analysed?

      Line 200 - From the image it is not convincing that SMEE1 is slightly behind DOT1 - I agree it looks enveloped but would appear level with the distal end of the DOT1 signal.

      For the truncation experiments the authors should explain that these are performed with cells in which the endogenous SMEE1 will be expressed and this may influence the localisation of the truncations, especially as there is no information about whether SMEE1 forms complexes with itself or other proteins.

      Figure 4D - should be 1 not T-

      Line 223 - given the low expression of constructs 2 and 9 I'm not sure it is possible to infer anything from the lack of localisation of these constructs as they appear unstable and would be unlikely to localise to a specific location.

      Figure S7 - The images presented were not convincing that there was a reduction in the localisation of LRRP1 to the hook complex on depletion of either SMEE1 or MORN1. The difference looks particularly minor if present at all.

      Line 264 - "implied that the lethal phenotype might be due to a loss of function" - this seems an odd thing to say as it doesn't provide any insight as of course the phenotype is due to a loss of function.

      Line 412 - can the authors clarify what they mean by geometric problems?

      Throughout the manuscript can you use log scale for the growth curves.

      Line 756 - add citation

      Line 465/66 - the authors states that the ability of the fluid phase cargo being still able to enter the pocket is evidence that the channel lumen is still open; however, I would think that despite the close apposition of the cell membrane to the flagellar membrane in the flagellar pocket neck region this would be unlikely to impede fluid/soluble material from entering the pocket, as presumably VSG protein can move through this region. This does not alter the ultimate conclusion the authors are drawing but without microscopy evidence for the state of the channel lumen it is difficult to be sure of its status.

      Significance

      The flagellar pocket is the key portal into and out of the trypanosome cell and as such has a vital role to play in host-parasite interactions. The flagellar pocket is supported by a number of cytoskeletal structures including the hook complex and the role of these structures in flagellar pocket function are poorly understood. The flagellar pocket is particularly important in the bloodstream form of the trypanosome parasite which infects the mammalian host as it is the route for the surface protein VSG to get onto and off the surface. The VSG is required for antigenic variation and the removal of VSG-antibody complexes helps 'clean' the surface of the parasite. SMEE1 is a component of the hook complex and the manuscript here dissects its role in the mammalian infective parasite and shows that it is vital for the endocytosis of material off the surface. Intriguingly, a block in endocytosis causes a blockage of material outside of the pocket, suggesting a multi-step process in the regulation of uptake of material from the parasite's surface.<br /> This manuscript will be of specific interest to those researchers investigating the long-term persistence of these parasites in the mammalian host. There are potentially some insights into the control of membrane domains for endocytosis that are of interest to more general cell biologists as well.

      Expert in molecular cell biology of trypanosomes and Leishmania.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary:

      In this manuscript the authors have advanced our understanding of the hook complex component TbSmee1 through a detailed analysis of this protein's role in the endocytosis of surface bound proteins via the flagellar pocket in bloodstream form Trypanosoma brucei. The TbSmee1 protein, previously identified using proximity labeling using TbMORN1 and TbPLK, and characterized in procyclic T. brucei, was confirmed to target to both the shank portion of the hook complex as well as the growing end of the new FAZ in replicating cells. The protein was also shown to likely be phosphorylated as had been suggested previously due to its association with the kinase TbPLK. A domain deletion analysis demonstrated that domains 2 and 3 are important for TbSmee1's proper localization to the hook complex. Loss of TbSmee1 using RNAi based knockdown resulted in a quick cessation of growth in the bloodstream form within 24 hours in contrast to what was seen previously in procyclic cells which had only a decreased growth rate. Loss of TbSmee1 also resulted in an enlargement of the flagellar pocket and in many ways mirrored the phenotype observed with knockdown of TbMORN1. Although prior work on TbSmee1 in procyclic T. brucei demonstrated that loss of this protein altered the morphology of TbMORN1, no such change was seen in bloodstream form cells and only an alteration in the morphology of TbLRRP1 was observed. In characterizing the effect of TbSmee1 depletion on endocytosis the authors showed that the fluid phase marker Dextran could enter into the flagellar pocket of TbSmee1 depleted parasites while the surface bound ConA and BSA remained outside of the flagellar pocket suggesting that TbSmee1 may play a role in allowing larger protein components into the pocket regions. Similar observations were also previously seen with TbMORN1 depletion. Importantly, a knockdown of clathrin recapitulated the TbSmee1 knockdown phenotype suggesting that endocytosis itself was required to allow material bound at the surface to enter into the flagellar pocket. In addition to adding to our understanding of hook complex components, this work raises some interesting questions regarding the role of the hook complex in facilitating endocytosis in this important human pathogen.

      Major Critiques:

      This is a superbly written manuscript with robust high-quality data that strongly support the major conclusions made by the authors. The flow the article is logical and easy to follow making it accessible to a wide array of readers. Although I appreciate the brevity of the introduction and how the article gets straight to the point, additional background information on the components and function of the flagellar pocket collar protein could help contextualize the goals of the project. The way in which the flagellar collar structures are introduced to the reader is quite abrupt (beginning on line 75) and simply states the names of TbBILBO1, the centrin arm and hook complex as simple facts without much discussion about the background of these components/regions. A graphical representation of the centrin arm or hook complexes relative to other components like the pocket itself, FAZ or axoneme could make following the story much easier. An expansion of this background could also go a long way to convince readers of the importance of this region in the basic biology and virulence of T. brucei.

      On lines 84-86 the authors cite the way in which 'small' vs 'large' macromolecules enter into the pocket without defining what exactly is meant by these terms as they are relative in nature. Setting some boundaries of size could provide some context to the reader.

      In the domain localization analysis beginning in Figure 4 there is a missed opportunity to also assess which portions of the TbSmee1 protein are important for overall function as well. By either an examination of dominant negative phenotypes resulting from overexpression of the truncated mutant or the expression of the truncated forms designed to be RNAi resistant in the TbSmee1 knockdown cell line, one could also assess which portions of this protein are essential for endocytic function in addition to targeting. Is there a reason this was not performed?

      In the analysis of viability changes due to TbSmee1 depletion (lines 237) the authors state that at "72 h post-induction showed widespread lysis, ..." This phenotype seems inconsistent with other related endocytic defect mutants. There is no further mention of this lysis phenomenon here or in the discussion and considering how unique this seems it deserves either additional data to demonstrate or further discussion as to the basis of the phenotype. It seems, at least from this study of TbStarkey1 and prior studies which result in the enlarged flagellar pocket phenotype, that having an enlarged pocket is not the cause of lysis and doesn't even naturally lead to a growth defect.

      The authors do not comment on what is the source for the cessation in growth following TbSmee1 knockdown. Is it nutrient depravation like in other endocytic defect mutants?

      In the end, one of the most interesting observations made by the authors is that loss of TbSmee1 inhibits endocytosis and this has the appearance of not allowing large molecule substrates like ConA and BSA to enter into the flagellar pocket. This appeared to have nothing to do with a gatekeeping type function of the hook complex/flagellar collar and instead, as shown through clathrin knockdown, was related to the ability of the parasite to endocytose. There are a lot of potential interpretations of this phenomenon with one being a simple perturbation of the normal membrane trafficking to and from the flagellar pocket being involved. An analysis of knockdown of exocytic components might reveal whether or not this inability to enter into the pocket is also seen when exocyst proteins are also depleted. It may be impossible to tease apart these two interrelated activities but it might eliminate one side of the equation if these proteins can still enter the flagellar pocket when exocytosis if perturbed although this reviewer understands that that dimension of T. brucei membrane trafficking is poorly understood relative to endocytosis.

      An intriguing possibility that the authors allude to and which if answered would make this manuscript have a far broader appeal is to determine if loss of TbSmee1 alters the lipid kinase distribution and if this is the source of the negative impact on endocytosis. One important dimension of endocytosis in T. brucei which remains poorly understood is the role of signaling machinery in triggering endocytic events. It is possible that the hook complex serves as the gatekeeping or signaling platform that recruits signaling components (like lipid kinases) that identify and/or modify the membrane lipid phosphatidylinositols harboring cargo laden receptors thus marking them for endocytosis within the pocket. It still seems unclear when in the process of endocytosis is the decision made to pull things into the pocket but it seems that the assumption is that this occurs deep within the pocket. This data suggests that there is possibly another decision point prior to being allowed entrance into the pocket. It may be that this isn't a gatekeeping decision but rather a stop vs. go activity where once cargo laden membrane reaches the collar a choice is made to pull this material in or not there and not after material is already in the pocket.

      This obvious enigma based on the observation that loss of hook complex components affect the spatially separated site of endocytosis support the idea that the actual endocytic signaling platforms are located at the hook complex and that this area may make the membrane modifications that mark membrane as being ready to be endocytosed via clathin coated vesicles at the bottom of the pocket. This would still allow for fluid phase small molecule entrance which does not require binding to surface proteins. The obvious problems of having both endo/exocytosis occurring in the same close proximity makes the dissection of this phenomenon difficult but it is worth potentially expounding on further in the discussion as this idea is very appealing and adds an important dimension to our understanding of endocytosis in this organism.

      Minor Critiques:

      The authors commit significant time to the analysis of the phosphorylation of TbSmee1, but there is little stated about the role of TbPLK in this activity or the potential connection of TbSmee1 phosphorylation to the cell cycle. Would a knockdown of TbPLK using RNAi potentially demonstrate an altered migration of TbSmee1 due to a lack of phosphorylation? An analysis of radiolabeled TbSmee1 using p32 in vivo would likely support this claim as well. Has mass spectrometry identified potential phosphorylation sites to examine? Additionally, the loss of TbSmee1 has been shown to disrupt localization of TbPLK in procyclic cells and so why this was not also assessed in bloodstream form cells subjected to RNAi was not clear.

      In the results section (lines 104-108) a model of the protein structure as predicted for example by AlphaFold might be informative and complement the domain analysis work depending on the quality of the prediction.

      There is an arrow in the Figure 1B Western blot but I can find no mention of what it is trying to highlight in the text.

      For Figure 1D there is no loading control or control for the distribution of the soluble fraction to validate the separation of the two compartments.

      The authors fail to comment on the lack of changes in hook complex components they see to that observed by Perry et. al. 2018. This difference merits some minor comment or speculation.

      Line 228: domain should be capitalized.

      Line 230: FigS5C should have a space and period after Fig. and S5C.

      Line 244: "on" should be inserted in the sentence "...TbSmee1 protein depletion ON either side of the onset..."

      Line 400: the '...20/21 h post-induction...' is slightly confusing and may read better as 20-21 h.

      Line 463: a space is needed between '...2009).The...'.

      Significance

      This manuscript advances our current conception of endocytosis in T. brucei. Although this model kinetoplastid parasite has been extensively studied with respect to endocytosis there is still a great deal we do not yet understand regarding how this process is regulated at a mechanistic level. This work has begun to connect previously unappreciated aspects of endocytosis in T. brucei by highlighting a potentially novel connection between the flagellar collar/hook complex and the physically separated endocytic events within the flagellar pocket itself. It may be that what appears as regulated entrance into the pocket is in fact the source of signaling that triggers the endocytic events carried out by clathrin. This is an interesting notion that no doubt requires further investigation which lies outside of the scope of this report. While this work appeals primarily to those studying kinetoplastids parasites it has the potential to provide insight into basic protozoan biology as well. Due to my related interest in kinetoplastid endocytosis, I find this work to be of high quality, conceptually interesting and employs many of the cutting-edge techniques currently available in the study of T. brucei.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Unfortunately, this paper adds only a little to our understanding of uptake in to the flagellar pocket of trypanosomes. It tends to add only detail to information that has been well characterised elsewhere and indeed, as the authors themselves point out, (lines 92-98) it is rather incremental. Not only has Tbsmee1 been studied before but this data in bloodstream forms is not particularly novel since it gives much the same information as the canonical hook protein TbMORN.

      This work follows the pattern of conclusions made previously with the protein TbMORN. It focusses on the protein TbSmee where RNAi mutants are interpreted to show flagellar pocket enlargement and impaired access by surface bound cargo. Unfortunately, there is little mechanistic or functional conclusion to the study in terms of how TbSmee operates naturally in the cell. There are other possible explanations for the phenotype. That would need to be studied. This large flagellar pocket phenotype is seen with RNAi mutants of many different types of proteins in the trypanosome and so pleiotropic effects are highly likely.

      Also, there are a good number of alternative possibilities to account for reduced access to the pocket in these mutants and this data could be usefully added.

      Specific points

      1. The transient location for the TbSmee at the FAZ tip - or in this case the groove region - was seen in procyclics (Perry, 2018) so this bloodstream indication merely confirms that concept.
      2. The C terminal region required for targeting is a reasonable deletion analysis of regions of the protein. But can this data (line 228) be said to "mediate targeting" - or is it just required. For instance, targeting might be OK but it might be needed for stable association, etc etc.
      3. This protein has already been shown to be phosphorylated and the sites and cell cycle possibilities have been mapped by Urbaniak. So that section adds little. https://doi.org/10.1371/journal.ppat.1008129
      4. Essentiality in BS forms and pocket enlargement. This is not surprising. A very large number of cytoskeletal proteins show this in RNAi knockdown. Flagella mutants (extensive publications from many groups (Hill, Bastin, Gull, etc) over last 15 years show this very well and so this protein is just one more example.
      5. I didn't find that the explanations for flagella pocket enlargement are soundly based. The experiments focus on endocytosis and uptake and ignore other plausible reasons and some evidence in literature.<br /> Lines 84/85. Enlarged pockets may be indicative of endocytosis failure. Presumably the rationale is that endocytosis fails, but exocytosis still occurs and the pocket membrane enlarges. What evidence is there that exocytosis of membrane still occurs? This simple concept might indeed operate in a clathrin mutant but is surface membrane/content exocytosis is maintained in these cytoskeleton mutants? There is good evidence for glycoconjugates within the flagellar pocket. Are these depleted or present still?
      6. There are also a number of other publications indicating that clathrin pits are still present on the enlarged pockets of various mutants when viewed by EM. The authors have looked at the flagellar pockets by EM but the EM methods described have extensive washings and centrifugations before fixation. This is a very poor approach and will mean that endo and exocytic traffic is disturbed (extensive references in literature in other systems? This is not a useful approach for exo/endocytosos studies where flux of traffic demands fast chemical or freezing fix in media.
      7. The EMs and Light microscopy does show that the mutant pockets are substantially abnormal in their cytoskeletal arrangement. They have multiple flagella profiles, flagella structures have not connected with the membrane and are sometimes in the cytoplasm (see a glance of the paraflagellar rod in the cytoplasm in FigS5C and internalised FAZ attachment plaques in Fig 4 D bottom right cell). Given these extensive (and expected) cytoskeletal abnormalities it is highly likely that these pocket abnormalities are a result of motility, cell division/developmental issues and the differential uptake phenotypes merely consequential.
      8. The authors speak about early phenotypes , but these are often at 15-24 hours. That is probably a couple of cell cycles and so not early. In relation to the above question of comparison to the same morphology produced by flagella mutants it would be good to know if these hook mutants produce motility phenotypes and whether these are manifest before the uptake phenotypes. There is evidence (cited here) that forward motility of the trypanosome directs material on surface into the pocket. If these cells have motility defects (primary or via failed division) then surely that would provide an alternative simple explanation for uptake differences.
      9. There is a general point that if studies are to have real relevance to uptake in the trypanosome then they need to deal with uptake of natural ligands rather than artificial surrogates such as dextran. Such tracers were used historically, but in the last decade a series of receptors and ligands for fluid phase and particularly membrane mediated endocytosis have been discovered. With the investment of a little time these important ligand / receptors such as haptoglobin, transferrin, etc would be much more relevant.

      Referees cross-commenting

      This session includes comments from Reviewer 1 and Reviewer 2.

      Reviewer 2

      Dear Reviewers 1 and 3:<br /> I agree with many of the points with Reviewer 1 and our divergence is partly a matter of degree. While it is true that this manuscript is incremental in its contribution to our understanding of TbSmee1, it nonetheless adds to our understanding of the role of this protein in the bloodstream life stage and because of that I find value in the work. The fact that it mirrors what was seem in other protein knockdown studies (e.g. TbMORN) doesn't negate its contribution for me. Reviewer 1 makes an important point, however, when stating that this work does not add a mechanistic or functional conclusion as to how TbSmee1 operates and for me that is the biggest shortcoming of the work. Offering mechanistic insight is a high bar and while it would make for a much more exciting story it does not discount the value of the work as presented. What I do appreciate is the speculation about this observation that endocytosis is required for entrance of surface bound material into the pocket and although they are unable to show that this is not a side affect of other processes being disrupted it is and intriguing point. These observation have the potential of stimulating further investigations into crosstalk between the entrance to the pocket and endocytosis. I also agree that the use of ligands for known receptors like transferrin would be far more informative. While I assumed the transferrin receptor was in the pocket itself it would be interesting to see if the ESAG6/7 is also located outside the pocket and transiently binds cargo before being brought inside for endocytosis.<br /> I think that Reviewer 3 brings up a great point with the focus on VSG's. I think that examining VSG turnover in these mutants can add value to the analysis and inform our view of how affecting the hook complex alters VSG endocytosis.

      Reviewer 1

      some fair comment and agreement. This is being sent to general cell biology journals.<br /> when one looks at this area in the round it is nearly 50 years (1975) since Langreth and Balber published their seminal work on protein uptake and digestion in bloodstream and culture forms of T. brucei. There has been 50 years intense study and the genome has been around for nearly 20 years as well. So, put simply - for both a general science audience and the wider parasite community - if this is a paper about one protein, TbSmee1,then it has surely has to say something functional about that protein. If it is a paper about uptake in trypanosomes (where mutants are one means of interrogation) then it surely has to say something about mechanisms of uptake of physiological relevant ligands. The days of dextran etc are past. Hence, my comment that this does neither and so is very incremental to what is known already. It is 2022 not 1975. Langreth and Baber published their seminal work in J Protozoology for very good reasons no doubt.

      Reviewer 2<br /> Thank you for replying and I agree with the spirit of your critique. My only comment, which could result from my own naivete, is to say that despite the incredible work that has been done in dissecting endocytosis in T. brucei over these past 50 years, it appears that we still do not understand how many fundamental of aspects of this activity works in this parasite. Even basic questions regarding how cargo, e.g. transferrin, binding to surface receptors is sensed by the parasite remains unknown and the identity of the specific signaling components which transmit this information internally to initiate endocytosis have not been characterized. In many ways it seems that we don't even understand how the parasite partitions the end/exocytic pathways in the pocket and maintains membrane homeostasis. While we know that some kinases and traditional signaling components must be involved, a high resolution understanding of this process in T. brucei seems lacking. I only say all this to suggest that the field maybe isn't yet that advanced to reject work of this type as so many mechanistic unknowns still remain to be uncovered and maybe incremental advances and phenomenology still can add value to the field. However, I respect your opinion on the matter and my perspective could be due to a lack of a full appreciation of the literature on the subject.

      Significance

      Unfortunately, I did not find tis to be very significant. It covers old ground in terms of the phenotype described. Many groups have shown the differences between pro cyclic and bloodstream phenotypes in this enlarged pocket phenomenon. The work is rather incremental from these and other author's work on these hook proteins.

      There are alternative explanations for understanding the effect of flagella pocket structure and uptake of ligands into the pocket and trypanosome cell. These would need to be studied before one could see a functional, mechanistic link established.

      Other parts of this are of nicely done but do not move on our understanding (eg targeting/phosphorylation) from what has been done previously.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity (Required)):

      In this manuscript, Kagermeier et al. present a novel and interesting study that attempts to model a severe neurodevelopmental disorder, pontocerebellar hypoplasia type 2a, using neocortical and cerebellar organoids. Brain organoids are an appropriate and promising approach to elucidate disease mechanisms in neurodevelopmental diseases. The authors show a reduction in the size of the organoids which is more pronounced in the cerebellar compared to neocortical organoids. While this finding is interesting and reminiscent of the clinical PCH2a phenotype, i.e., cerebellar hypoplasia, the study is very preliminary and the conclusions of the manuscript are not supported by the data. Additional information and further experiments are necessary to support the claims made.

      Major concerns:

      1. hiPSC lines show considerable inter- and intra-individual variability and therefore the size differences observed between these control and patient-derived organoids may arise from differences in the hiPSC lines used. While the data sufficiently demonstrates the pluripotency of the multiple novel hiPSC lines, major concerns remain as to the appropriateness of the control hiPSC lines. The manuscript should include a table describing the age and sex matching as well as mode of reprogramming for all control and patient lines. Patient and control lines should be matched as closely as possible. Furthermore, figure legends should clearly indicate which clones and lines are shown in the various figure panels.

      We agree with the reviewer that hiPSC variability is an important concern in the field. In order to minimize such effects, all iPSCs lines used in this study were generated following the same protocol in the same lab. All cell lines are derived from male donors, thus, eliminating sex-based variability. Further, there is no report of sex-based variance in the clinical phenotype of PCH2a children and this finding is further corroborated by a currently on-going natural history study in our research team. While it would be ideal to also have age-matched controls, this is not possible for ethical reasons as skin biopsies from healthy children cannot easily be obtained to match the pediatric PCH2a cases. However, based on the literature, we believe that epigenetic age is erased upon reprogramming (Strassler et al 2018, Studer et al 2015). Following the reviewer’s recommendation, we provide a table that clearly indicates the origin of all six cell lines used (see Methods section) and information of respective lines was added to the figure legends as suggested by the reviewer.

      As the hiPSC lines used are not isogenic, it is important that the authors characterise these lines further. This should include a quantification of the rates proliferation and apoptosis in all used hiPSC lines, as these might impact the growth rate of the embryoid bodies / organoids.

      We thank the reviewer for raising this concern. To address the variability of hiPSC lines, we performed an extensive characterization of pluripotency, proliferation and cell cycle dynamics of all six hiPSC lines through immunocytochemistry against pluripotency marker OCT4, proliferation marker Ki-67 and EdU incorporation experiments. We further assessed the apoptosis rate of hiPSCs by staining against apoptotic marker cCas3. These experiments were carried out in three consecutive passages of all iPSC lines providing statistical power to the analyses. All experiments did not result in significant differences between PCH2a and control iPSC lines (see Figure 2).

      The authors state that the hiPSC lines have been characterised by SNP arrays to show that no genomic / chromosomal aberrations have been accrued due to reprogramming. The manuscript should include information as to when the SNP array was performed (i.e., immediately after reprogramming, after initial passaging, etc) and also include the results of the SNP array as additional information. What passage were the hiPSC when the presented experiments were carried out?

      In agreement with this comment, we provide data of SNP arrays that were performed to ensure the chromosomal integrity of all cell lines (see supplement). Further, we added details on passages of the cell lines in the respective figure legends as suggested by the reviewer. In brief, all cell lines were kept below passage 20 and were subjected to pluripotency testing before differentiations were started.

      Given that TSNE54 is broadly and strongly expressed in the developing nervous system, the very limited staining of the organoids for TSNE54 in Figure 2 is surprising. Can the authors provide an explanation for the fact that TSNE54 is only expressed in a small subset of cells? Which cell types are these? Moreover, high-magnification images should be shown to demonstrate subcellular staining pattern of TSNE54. Quantification of TSNE54 protein levels by immunoblotting would also be beneficial.

      Related to this observation, it is puzzling that the large size differences that the authors observe in their organoids would be driven by such a small number of TSNE54-expressing cells. How do the authors explain this discrepancy?

      We thank the reviewer for this comment. We have carefully assessed human cerebellar development transcriptomic datasets which demonstrate that TSEN54 is in fact not strongly but moderately expressed in the human developing nervous system. Additionally, TSEN54 expression is expressed in various different cell types (not limited to a subset of cell types) (Aldinger et al 2021, Sepp et al 2021). We agree with this reviewer and reviewer 3 that Western Blotting or other types of quantification would be informative as well as investigation of the subcellular localization of the protein. However, these questions go beyond the scope of the current manuscript, which aims to present a disease model. We have therefore decided to remove the characterization of TSEN54 expression in organoids from our revised manuscript.

      The generated organoids need to be better characterised with a broader range of markers using both qPCR and immunostaining. At the moment, their identity as "cortical" and "cerebellar" organoids remain unconvincing. This is particularly true for cerebellar organoids, which are challenging to generate and are not widely used. The authors should include additional markers (for example, see PMIDs 25640179, 29397531, 32117945) and immunostaining should clearly show expected staining patterns.

      In Figure 5, it appears that some markers (e.g., SATB2) are expressed differently between control and patient lines, yet this is not commented on by the authors who conclude that control and patient lines show differentiation into organoids.

      We thank the reviewer for this suggestion. We performed further immunostainings using the markers that were used in other cerebellar organoid papers (Muguruma et al 2015, Silva et al 2020, Watson et al 2018) as the reviewer suggested. In detail, we added immunohistochemistry experiments on Day 30 and Day 50 of differentiation for early Purkinje cell markers OLIG2 and SKOR2. We also included ATOH1 as a marker for rhombic lip-derived granule cells. For the neocortical organoids, we believe that the performed characterization is sufficient since the protocol we used is well-established and widely used as also indicated by the reviewer. We agree that the cellular composition of the organoids should be investigated in detail (for instance using single-cell transcriptomics). However, we believe this is out of the scope of this manuscript, which describes the establishment of a brain-region specific model platform.

      The authors attempt to look into a potential mechanism for the size differences observed between control and patient organoids. However, only cleaved caspase-3 is used as a marker for apoptosis and no differences were observed. The authors should include further markers for potential cell death. In addition, immunostaining for proliferation markers (i.e., KI67) should be performed to evaluate whether the difference in organoid size could stem from decreased proliferation rather than increased cell death.

      We agree with the reviewer and included a quantification of the proliferation marker Ki-67 within the SOX2 positive population of cerebellar and neocortical organoids as well as the quantification of SOX2 positive areas within the organoids (Figure 6). We observed significant differences in proliferation between PCH2a and control cerebellar organoids. Moreover, we also analyzed the morphology of organoids and quantified the thickness and number of rosettes and find significant differences between control and PCH2a cerebellar organoids corroborating the notion that proliferation is altered in cerebellar organoids. Neocortical organoids do not show any significant differences in proliferation and Sox2+ structures. Only the thickness of the Sox2+ areas is slightly decreased in neocortical PCH2a organoids compared to controls. In order to deepen our analysis of a possible increased apoptosis in PCH2a organoids, we also quantified cCas3 in Sox2+ structures (Figure 5) as also suggested by Reviewer 2. These analyses did not show any significant differences between PCH2a and control organoids. We therefore suggest that at the early stages of differentiation studied here, proliferative differences are the main reason for the size differences between PCH2a and control organoids.

      Reviewer #1 (Significance (Required)):

      The authors present an innovative approach to study neurodevelopmental disorders using brain organoids and should be of interest to researchers and clinicians working on neurodevelopmental diseases. However, the data presented are too limited to support any conclusions about the phenotype observed. Furthermore, questions remain about the used methodology and more work is needed to demonstrate the successful generation of both cortical and cerebellar organoids.

      Reviewer #2 (Evidence, reproducibility and clarity (Required)):

      Please find enclosed my recommendation for the paper submitted by Kagermeier et al entitled' Human organoid model of PCH2a recapitulates brain region-specific pathology'. It describes the development of a human model for PCH2a and its characterization. My overall assessment of the paper is 'Major revision' which is explained below.

      Although the paper is very well written and clearly interesting in that it describes the generation and initial analyses of a human organoid model for PCH2a it should be revised such that it will proof the points it is trying to make. The authors are meticulous in their studies combining cellular characterization and a thorough initial screen of organoid (both cerebellar as well as cortical) integrity, yet hardly any mechanistic data is provided. Nevertheless, if the authors are able to add additional experiments and are able to address the points raised, the reviewer may be willing to consider a more positive outcome.

      Major concerns

      1) The overall quality of the figures is poor. There is a lot of overexposure such that often cellular or tissue structures are blended. It starts with Figure 1 G and H but can be observed throughout the manuscript. Deconvolution would greatly enhance their results.

      We are thankful for this comment and we have improved the quality of all microscopy images.

      2) Especially figure 4 and 5 could have been complemented with quantitative data. It furthermore seems more supplemental figure as these are just proof-of-principle stainings. No conclusions can be drawn from the panels except that all markers are there in the various conditions. And while they are showing a neural rosette in Fig 4A, just tiny ones can be observed in 4B. It is also not clear what the whole mount IHC ads in comparison to the IHC on sections. It is also strange that there is still a lot of SOX2 in the CALB/MAP2-positive area, but again with this magnification hard to appreciate.

      We agree with the reviewer that so far we presented qualitative proof-of-principle stainings that demonstrate cerebellar and neocortical differentiation, respectively. In order to address the comment of the reviewer, we improved the quality of the images and also provided higher magnification and enhanced resolution. Additionally, we now provide detailed quantifications of SOX2+ and Ki67+ neural progenitor cells and show that differences observed between PCH2a and control cerebellar organoids may explain the size differences observed between organoids (Figure 6). Our study provides the basis for more in-depth analysis of differences in differentiation and cell type composition between PCH2a and control organoids in the future, for example through single-cell RNAseq.

      3) If the authors would like to proof the point that cerebellar/cortical development is hampered, more functional assays could have been done. Nothing is analyses on the fraction of progenitor cells present (such as the percentage of Tbr2+ IPC in VZ/CP). Furthermore, if there is a suspicion that the number of cells is affected (which is also not shown), proliferation/cell cycle exit experiments using BrdU/EdU should have been performed. Early cell cycle exit still cannot be rules out and should have been tested by the combination of Ki67-/EdU+ percentage of a certain faction of progenitor cells (eg PAX6+ pool).

      We thank the reviewer for this valuable suggestion and agree that it would be interesting to carry out respective experiments. In this study, we show the establishment of a brain-region-specific organoid platform as a disease model for PCH2a and are only at the beginning of deciphering the underlying mechanism. In the revised manuscript, we quantified Ki-67+/Sox2+ cells in proliferative zones in the organoids. We believe that future studies including BrdU / EdU incorporation assays as well as scRNA-seq will answer the questions raised here and decipher the disease-causing mechanism on both cellular and molecular levels but are beyond the scope of this manuscript.

      4) Instead the author chose to only perform a cCas3 staining. From the panels in Figure 6 it is hard to appreciate which cells are actually cCas3+. Also the analyses were performed on the total pool of cell while it might have been more interesting to look for cell death of the various progenitor pools (eg the SOX2+ pool).

      We agree with the reviewer that a more in-depth analysis of apoptotic cell populations is interesting and performed cCas3/Sox2+ quantification for cerebellar and neocortical organoids. We did not observe significant differences of cCas3 expression within the SOX2+ cell population. (Figure 5)

      Minor concerns

      1) It would greatly enhance the review process if line numbers are added

      We have added line numbers to the manuscript.

      2) On general concepts (such as the generation of organoids in the context of disease) more references could have been added

      We have added more references and discussed the topic of brain organoids as disease models as suggested by this reviewer (Eichmüller & Knoblich 2022, Khakipoor et al 2020, Velasco et al 2020).

      Figures

      Fig. 1: In A, the square is clearly visible and not similar to B. An annotation of which is the control and which is the patient is missing in the figure. The arrows are hardly visibly, would make them slightly bigger and remove the black outer lining. Figure 1C can easily go to the Supplemental material. Fig 1 D is hard to appreciate the staining, a close-up with bright field microscope will help. E-I Most of the panels but especially G and H are overexposed. In J, it is hard to appreciate the TSEN54 staining. Maybe separate channels and a merge?

      We thank the reviewer for bringing these details to our attention. We have changed the arrows in the figure to enhance their visibility. Further we have adjusted the quality of the images overall. Lastly, we have made a comment in the figure legend clearly stating which scan came from which child. The described square was added to hide facial features of the imaged individuals hence they are not identical.

      Fig. 3: Usually go into the supplementals.

      Since organoid size is a major first readout when modeling a disorder that is characterized by a reduction of the volume of specific brain regions, we decided to keep this readout in the main text.

      Fig 4/5: Lack of quantitative data and poor quality of figures (overexposure).

      Fig 6: Many of the SOX2 panels are overexposed

      We thank the reviewer for the suggestions on the figures and addressed the concerns in the revised manuscript.

      CROSS-CONSULTATION COMMENTS

      I completely agree with reviewers #1 and #3. It is good to notice that we are overall on the same page.

      Reviewer #2 (Significance (Required)):

      The authors definitely made an excellent start to model PCH2a. Three controls and three patient lines are good to begin with but isogenic controls using one parental line and a patient line where the mutation is fixed would have been ideal. It is interesting that there seem to be a brain area specific pathology of the phenotype. Yet, more thorough analyses could have been performed such as proliferation and differentiation and cell cycle exit experiments. As for now the mostly descriptive data are only scratching the surface and little can be concluded on the molecular framework they are trying to solve.

      Reviewer #3 (Evidence, reproducibility and clarity (Required)):

      Summary:

      In this study Kagermeier et al. use human cerebellar and neocortical organoids to investigate the effects of the PCH2a-causing homozygous TSEN54c.919G>T variant on the neurodevelopment of different brain regions. They reveal a substantial growth defect in both neocortical and cerebellar regions with a more profound phenotype in the cerebellum. They continue to investigate major cell types of neurodevelopment in both regions and briefly potential mechanisms underlying the phenotypes. The study is well conceived and addresses the current gap of disease-modeling in cerebellar organoids; nevertheless, some major claims are not sufficiently substantiated in the current version. Below, I provide suggestions on how to improve the manuscript with some additional minor comments that might help with readability and accessibility of the work.

      Major comments:

      1. TSEN54 expression levels: The authors compare RNA and protein expression levels for TSEN54 to investigate the mutation's effect. For this the authors use qPCR on iPSCs and organoids of different age and immunostainings and conclude "we did not find differences in expression between cell and tissue types". There are some issues with this analysis as explained below:

      -The qPCR data (Fig. 2B) is first normalized to a housekeeping gene (GAPDH), however, then all organoid data are additionally normalized to the respective iPSC line. Thus, in case there is already a difference on iPSC level, this normalization might mask any difference in the organoids. It is unclear why this approach was chosen, and it seems more appropriate to show the data just normalized to GAPDH than additionally normalizing to the iPSCs, or at least to show first that iPSCs do not have differences in TSEN54 expression. Furthermore, even though apparently not statistically significant there seems to be a strong trend of lower TSEN54 levels in PCH2a in neocortical organoids, but even more so in cerebellar organoids. In my view this would fit very well with the study and should be further explored before concluding there is no statistical difference. Considering the high error bars of the cerebellar organoid samples, a higher N-number might be necessary to reach statistical significance in the difference in expression. Most importantly, it would be appropriate to show single data points where possible and to mark the different cell lines (as done in other figures), as otherwise it is not possible to judge whether there is a cell line bias in the data.

      -The evidence for protein expression of TSEN54 is immunofluorescence stainings for all conditions. As there is no quantification, the authors should not conclude differences, or the lack thereof, based on this qualitative data. Furthermore, in fact in the on example shown the PCH2a cerebellar condition (Fig 2D) seems to show lower expression levels compared with other conditions. This could be due to the selected image, as all other examples include large neural rosettes with strong staining in the center of the rosettes. Furthermore, it is unclear what cell line these stainings come from, even whether the PCH2a cerebellar and neocortical stainings come from the same cell line. Thus, the authors should select comparable examples for all conditions, and ideally provide staining examples (e.g., as supplementary data) for the other replicates to ensure expression in all replicates. If the authors want to comment on differences in protein expression, maybe a quantitative approach (e.g., quantitative western blot) would be more appropriate. Otherwise, the statements should be adjusted to not conclude whether TSEN54 protein levels differ or not.

      -Irrespective of the above comments the conclusion of the section "TSEN54 expression in cerebellar and neocortical organoids", that currently reads "we did not find differences in expression between cell and tissue types" should be changed, as the authors did not investigate whether there are cell type-specific differences of TSEN54 expression.

      We thank the reviewer for this comment. We agree that the provided data is not suitable for quantitative analysis of TSEN54 expression. Please also see our related response to the similar concern raised by reviewer 1. Thanks to these suggestions, we have decided to exclude the TSEN54 expression data from the current manuscript as a detailed analysis should be part of an extensive future study.

      Organoid growth analysis:

      The organoid growth analysis in Figure 3 and supplementary Figure 2 shows the main phenotype of the study that seems to be very strong. The authors use unpaired t-tests to compare within the different timepoints. Unfortunately, I think this approach might not be appropriate as even though the Welch correction does not rely on similar SDs in the compared groups (Control vs. PCH2a), it still assumes that all data points within each group share the same variance. However, this is not the case, as e.g., the control condition includes three groups (Control-1 to -3), that between groups might have different variance as such not all datapoints are independent from each other. Potentially ANOVA analyses controlling for cell line and timepoint might be more appropriate. Or additionally, the authors could consider using the linear regression analysis in Supplementary Figure 2 to further investigate the difference in organoid growth by e.g., comparing the slope of the regression lines. This might be more appropriately reflecting the growth deficit over time than simply comparing each timepoint individually. Expanding on this analysis the regression analysis requires some more information on the fit (intercept, slope, R-squared of the model), which would help clarifying the growth dynamics in the different systems and conditions.

      We thank the reviewer for the suggestions on statistical analysis and adjusted our approach accordingly. Briefly we performed 3-way-ANOVA analysis for the growth curves which revealed no significant differences between the different lines within the groups (Control or PCH2a) at different time points. Additionally, we added the linear regression model to the results (See Figure 3 and supplementary table 2, with the information on the curve fit).

      The growth ratio analysis (Figure 3D) is essential to the major claim of the paper that the organoids replicate the region-specific differences. As the authors performed all experiments with matching cell lines this could additionally strengthen the argument by generating the ratio of size differences for each cell line separately (instead of just for all PCH2a lines together). This would allow comparison of the same genetic background in both cerebellar and neocortical condition and further corroborate the region-specific severity of the phenotype. Potentially, this would also enable to test these differences statistically.

      We appreciate the suggestion to compare the differentiation protocols by line. Below we display the line-by-line analysis between the two differentiation protocols at D30 (A), D50 (B), and D90 (C). In order to visualize the differences in size between the two protocols more clearly, we have generated ratios of the average organoid sizes between neocortical and cerebellar organoids (D). The analysis corroborates our previous visualizations and statistics (3-way ANOVA) by showing that PCH lines produce neocortical and cerebellar organoids that differ in size more than those of control lines. The differences are most pronounced at D30 and D90. However, we believe that this analysis does not add additional value to our manuscript and have therefore decided not to include it in the revised version.

      Additionally, all growth analyses for the neocortical organoids (Figure 3C, Supplementary Figure 2B and C) seem to lack the PCH-1 cell line and only contain PCH-2 and PCH-3. This cell line should be added or commented on why it was excluded from the analyses.

      We agree with the reviewer. Unfortunately, we experienced contamination in that specific differentiation and therefore cannot provide the data. We have made a related comment in the manuscript. Since all differentiations were performed in parallel, adding this line at a later time point would add additional confounders and is therefore undesirable.

      Potential mechanism of the phenotype (apoptosis analysis):

      In Figure 6 the authors investigate the hypothesis that increased apoptosis contributes to the phenotypes. In the cleaved Caspase 3 staining there appear to be no differences. Unfortunately, the analysis apparently only includes one replicate (one organoid?) per cell line and condition. Considering the variability in the data shown this seems inappropriately low and should ideally contain ~3 replicates per cell line condition to judge technical and biological variability if the authors want to make the point that there is no "significant difference between PCH2a and control organoids at any time point in both cerebellar and neocortical organoids". Otherwise, this claim does not seem to be substantiated enough by the data.

      Finally, due to the absence of a phenotype related to apoptosis the authors conclude that the phenotypes may be due to "deficits in the proliferation of progenitor cells". Although this is mentioned in the introduction and the discussion, there is no evidence in the current study that supports this interesting idea. By adding relatively straight forward co-staining experiments for e.g., SOX2 (progenitors) and Ki67 (proliferating cells), the authors could provide further evidence for this hypothesis using existing organoid sections. This would support this speculative idea and could add a more mechanistic insight to the study, thereby making it more exciting.

      To address this concern, we have now added a table to the supplement that described in detail which organoids / batches / cell lines were used for which experiment (Supplementary table 3). In addition to our previous cCas3 quantifications, we performed the quantification of cCas3 within the population of SOX2-positive cells, which was suggested by Reviewer 2 (Figure 5).

      To assess the alternative hypothesis, that proliferation deficits account for the size differences observed between organoids, we also performed quantifications of SOX2-positive zones in the organoids at D30 and D50 of differentiation as well as quantifications of Ki-67 positive cells within the SOX2-positive population. For cerebellar organoids we found significant differences in these experiments (Figure 6). We believe that this data supports the hypothesis of aberrant proliferation in PCH2a cerebellar organoids explaining the size differences.

      Minor comments:

      • Cell line and quality control: The authors recruit three male patients with PCH2a and reprogram iPSCs. These cell lines are subjected to a well performed extensive quality control. However, it is unclear what cell lines the stainings (e.g., Fig. 1D to I) originate from. Furthermore, the supplementary qPCR analysis (Supplementary Figure 1) includes only the PCH-1 line, and additionally two cell lines that are not explained (F-CO and hESC-I3). It is unclear what the relevance of showing the qPCR of these cell lines is. To ensure proper QC for all used cell lines the authors should provide data for all cell lines (PCH-1 to -3 and control-1 to -3), or at least summarize (e.g., in a table) what QC metrics were applied to which cell line. Most importantly, this information is completely lacking for the control cell lines and the QC is just mentioned in the text. Unfortunately, it is unclear where the control cell lines originate from, and some basic information would be required to judge whether they are appropriate controls: are they iPSC or ESC, were they reprogrammed with a similar paradigm as the PCH2a cells, what is the gender of the control cell lines (all PCH2a cell lines are apparently male)?

      In line with a similar comment from reviewer 1, we have included a table that provides information on the origin of all six cell lines used in the revised manuscript (methods section). Further we provide SNP-Array data on all cell lines as supplementary material. We also performed detailed characterization of pluripotency, proliferation and cell cycle dynamics of all six hiPSC lines through immunocytochemistry against pluripotency marker OCT4, proliferation marker Ki-67 and EdU incorporation experiments (Figure 2). We further assessed the apoptosis rate of hiPSCs by staining against apoptotic marker cCas3. All experiments did not result in significant differences between PCH2a and control iPSC lines (see Figure 2). In line with the suggestion of this reviewer, we removed the qPCR analysis of iPSCs from the manuscript.

      • To make the study more approachable for a medical audience and to judge the variability in phenotype presentation among the recruited patients it would be appreciated if more information on the patients would be provided. The authors write: "We identified three individuals that display the genetic, clinical and brain imaging features previously described for PCH2a.". This information including age/date of birth, as well as other medically relevant information could be provided in the supplementary figure (e.g., is there a difference in disease burden among the different patients?). This would allow judging the recruited cohort better.

      We thank the reviewer for this insightful comment. We provided a table with detailed clinical information (supplementary table 1).

      • According to the method section the cerebellar and neocortical organoids were cultured in very different medium especially at later timepoints. While neocortical organoids were kept in a neural maintenance medium based on Neurobasal-A, cerebellar organoids were kept in a medium based on BrainPhys. These media contain very different levels of nutrients, especially of glucose (25mM vs 2.5mM, Bardy et al. 2015). This can have a strong phenotype on proliferation of progenitors and proliferative phenotypes (e.g., see Eichmüller et al. 2022). Especially as the authors claim that there is a difference in the PCH2a phenotypes between brain regions, it should be excluded that this is due to medium differences at later timepoints. When investigating the growth curves of Figure 3B and C it seems like the major difference in growth speed seems to be that neocortical organoids grow faster in early timepoints (We agree that media composition can greatly influence growth dynamics of cells in 2D and 3D. However, in this study we assess the differences between two groups: the PCH2a and control iPSC-derived organoids. The differences we describe are in relation to the respective control group and iPSCs were generated following the same protocol in the same lab. We believe that by following two protocols and comparing the three PCH2a to the three control lines within each protocol predominantly, we account for different media composition possibly changing growth dynamics.

      • Staining examples shown and presentation: In several figures the authors could improve the presentation of the staining examples with some changes:

      o Cell line information for images: as the authors only ever note the condition (PCH2a or Control) but not the cell line it is unclear if the stainings all come from one cell line or from multiple different cell lines. This prevents comparing the different differentiation conditions. Additionally, for major conclusions the authors should consider including supplemental stainings or further information on how reproducible the results shown are (how many cell lines and batches were used?).

      We thank the reviewer for these suggestions. We added information on cell lines and passages for all experiments shown in this study in the figure legends. Moreover, we also added a table providing information on n-numbers for all experiments (supplementary table 3).

      o Selection of examples: in several cases (Fig 2C/D, 4A, 6A/B) the selected images depict very different regions, e.g., one condition shows a large rosette, while in the other condition no rosette can be seen. It would be more appropriate to show matching examples where possible.

      We agree with the reviewer and have chosen matched regions of interest in the figure panels in the revised version of the manuscript. Please note that for cerebellar organoids we observed a significant difference in the timepoint of appearance of these rosette-like structures. Therefore, an exact matching of regions of interest was not possible due to biological differences between the samples, which we have also quantified (Figure 6).

      o Color code of stainings: Colors do not match throughout the manuscript in immunofluorescence images. E.g., Fig. 4 uses blue, green, red, magenta and Fig. 5 uses blue, green, magenta, cyan. It would be preferable to adhere to one color code. Considering significant fraction of the population is having red-green blindness, the latter color code seems more appropriate as it should ensure readability also for color-blind audiences.

      We are thankful for this comment. We changed the color code to make figures more widely accessible.

      • Small typos:

      o Figure 1 legend: last sentence "The" instead of "Th"

      o Supplementary Figure 1B: PCH-2 is named "PCH-22"

      o Supplementary Figure 2: As in the main figure for neocortical organoids the PCH-1 condition is missing (see comment on organoid growth curves). Additionally, the color/shape code of the plots in B does not always match the legend (e.g., size in left plot is different and color of PCH-3 in middle and left plot differs from legend and right plot).

      o It is unclear why the cortical organoids are referred to as "neocortical organoids" in the figures and the text. The methods and the reference in the methods as well as all major papers rather use the word "cortical".

      We addressed these suggestions and thank the reviewer for bringing these to our attention. Unfortunately, we could not include data on PCH-01 in neocortical differentiation due to a contamination in this batch. We made sure to run all the batches presented here in parallel so that all conditions are equivalent, preventing us from including a different batch at a later time point.

      We believe that in the context of our study, it is important to highlight cortical organoids as neocortical organoids, because we are also showing cerebellar organoids and there is also a cerebellar cortex.

      References:

      Bardy, C. et al. Neuronal medium that supports basic synaptic functions and activity of human neurons in vitro. Proc National Acad Sci 112, E3312 (2015).

      Eichmüller, O. L. et al. Amplification of human interneuron progenitors promotes brain tumors and neurological defects. Science 375, (2022).

      CROSS-CONSULTATION COMMENTS

      I agree with the comments of the other reviewers and as they are mostly matching, this reinforces the importance to improve certain aspects of the manuscript. As there are no deviating issues I do not comment specifically on any reviewer comments.

      Reviewer #3 (Significance (Required)):

      This work is using organoid technology to shed light on brain region-specific phenotypes in PCH2a. Brain organoids have drastically changed the way we study human neurological diseases (Eichmüller and Knoblich 2022), however, most brain organoid research has focused on cortical organoids. Cerebellar organoid protocols exist for some time (Muguruma et al. 2015, Silva et al. 2020, Nayler et al. 2021) but were not yet applied to uncover new disease biology. Especially considering the important role of human-specific cerebellar processes in specific developmental disorders (Haldipur et al. 2021) and cancer (Hendrikse et al. 2022, Smith et al. 2022), disease modeling in human cerebellar organoids holds great potential for understanding disease biology. The work by Kagermeier et al. demonstrates that human cerebellar organoids are recapitulating brain region-specific growth deficits and thus is an important step forward for disease modeling. Therefore, this work will be interesting to researchers working on brain development and disease modeling, especially in in-vitro systems. Nevertheless, the mechanistic insight of the study is limited, as is the insight into how human-specific processes might be involved in the pathogenesis of PCH2a. Therefore, it will be interesting how this disease model will be used in future to investigate the cell types and mechanisms involved in the PCH2a phenotype.

      Personal field of expertise: Brain organoids and disease modeling in organoids especially of neurodevelopmental diseases. Analysis of organoids with stainings, as well as sequencing techniques, and bioinformatics.

      References:

      Eichmüller, O. L. & Knoblich, J. A. Human cerebral organoids - a new tool for clinical neurology research. Nat Rev Neurol 1-20 (2022) doi:10.1038/s41582-022-00723-9.

      Haldipur, P. et al. Evidence of disrupted rhombic lip development in the pathogenesis of Dandy-Walker malformation. Acta Neuropathol 142, 761-776 (2021).

      Hendrikse, L. D. et al. Failure of human rhombic lip differentiation underlies medulloblastoma formation. Nature 609, 1021-1028 (2022).

      Muguruma, K., Nishiyama, A., Kawakami, H., Hashimoto, K. & Sasai, Y. Self-Organization of Polarized Cerebellar Tissue in 3D Culture of Human Pluripotent Stem Cells. Cell Reports 10, 537-550 (2015).

      Nayler, S., Agarwal, D., Curion, F., Bowden, R. & Becker, E. B. E. High-resolution transcriptional landscape of xeno-free human induced pluripotent stem cell-derived cerebellar organoids. Sci Rep-uk 11, 12959 (2021).

      Silva, T. P. et al. Scalable Generation of Mature Cerebellar Organoids from Human Pluripotent Stem Cells and Characterization by Immunostaining. J Vis Exp (2020) doi:10.3791/61143.

      Smith, K. S. et al. Unified rhombic lip origins of group 3 and group 4 medulloblastoma. Nature 609, 1012-1020 (2022).

      References by the authors

      Aldinger KA, Thomson Z, Phelps IG, Haldipur P, Deng M, et al. 2021. Spatial and cell type transcriptional landscape of human cerebellar development. Nat Neurosci 24: 1163-75

      Eichmüller OL, Knoblich JA. 2022. Human cerebral organoids — a new tool for clinical neurology research. Nature Reviews Neurology 18: 661-80

      Khakipoor S, Crouch EE, Mayer S. 2020. Human organoids to model the developing human neocortex in health and disease. Brain Res 1742: 146803

      Muguruma K, Nishiyama A, Kawakami H, Hashimoto K, Sasai Y. 2015. Self-organization of polarized cerebellar tissue in 3D culture of human pluripotent stem cells. Cell Rep 10: 537-50

      Sepp M, Leiss K, Sarropoulos I, Murat F, Okonechnikov K, et al. 2021.

      Silva TP, Fernandes TG, Nogueira DES, Rodrigues CAV, Bekman EP, et al. 2020. Scalable Generation of Mature Cerebellar Organoids from Human Pluripotent Stem Cells and Characterization by Immunostaining. J Vis Exp

      Strassler ET, Aalto-Setala K, Kiamehr M, Landmesser U, Krankel N. 2018. Age Is Relative-Impact of Donor Age on Induced Pluripotent Stem Cell-Derived Cell Functionality. Front Cardiovasc Med 5: 4

      Studer L, Vera E, Cornacchia D. 2015. Programming and Reprogramming Cellular Age in the Era of Induced Pluripotency. Cell Stem Cell 16: 591-600

      Velasco S, Paulsen B, Arlotta P. 2020. 3D Brain Organoids: Studying Brain Development and Disease Outside the Embryo. Annu Rev Neurosci 43: 375-89

      Watson LM, Wong MMK, Vowles J, Cowley SA, Becker EBE. 2018. A Simplified Method for Generating Purkinje Cells from Human-Induced Pluripotent Stem Cells. Cerebellum 17: 419-27

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary: In this study Kagermeier et al. use human cerebellar and neocortical organoids to investigate the effects of the PCH2a-causing homozygous TSEN54c.919G>T variant on the neurodevelopment of different brain regions. They reveal a substantial growth defect in both neocortical and cerebellar regions with a more profound phenotype in the cerebellum. They continue to investigate major cell types of neurodevelopment in both regions and briefly potential mechanisms underlying the phenotypes. The study is well conceived and addresses the current gap of disease-modeling in cerebellar organoids; nevertheless, some major claims are not sufficiently substantiated in the current version. Below, I provide suggestions on how to improve the manuscript with some additional minor comments that might help with readability and accessibility of the work.

      Major comments: 1. TSEN54 expression levels: The authors compare RNA and protein expression levels for TSEN54 to investigate the mutation's effect. For this the authors use qPCR on iPSCs and organoids of different age and immunostainings and conclude "we did not find differences in expression between cell and tissue types". There are some issues with this analysis as explained below: -The qPCR data (Fig. 2B) is first normalized to a housekeeping gene (GAPDH), however, then all organoid data are additionally normalized to the respective iPSC line. Thus, in case there is already a difference on iPSC level, this normalization might mask any difference in the organoids. It is unclear why this approach was chosen, and it seems more appropriate to show the data just normalized to GAPDH than additionally normalizing to the iPSCs, or at least to show first that iPSCs do not have differences in TSEN54 expression. Furthermore, even though apparently not statistically significant there seems to be a strong trend of lower TSEN54 levels in PCH2a in neocortical organoids, but even more so in cerebellar organoids. In my view this would fit very well with the study and should be further explored before concluding there is no statistical difference. Considering the high error bars of the cerebellar organoid samples, a higher N-number might be necessary to reach statistical significance in the difference in expression. Most importantly, it would be appropriate to show single data points where possible and to mark the different cell lines (as done in other figures), as otherwise it is not possible to judge whether there is a cell line bias in the data. -The evidence for protein expression of TSEN54 is immunofluorescence stainings for all conditions. As there is no quantification, the authors should not conclude differences, or the lack thereof, based on this qualitative data. Furthermore, in fact in the on example shown the PCH2a cerebellar condition (Fig 2D) seems to show lower expression levels compared with other conditions. This could be due to the selected image, as all other examples include large neural rosettes with strong staining in the center of the rosettes. Furthermore, it is unclear what cell line these stainings come from, even whether the PCH2a cerebellar and neocortical stainings come from the same cell line. Thus, the authors should select comparable examples for all conditions, and ideally provide staining examples (e.g., as supplementary data) for the other replicates to ensure expression in all replicates. If the authors want to comment on differences in protein expression, maybe a quantitative approach (e.g., quantitative western blot) would be more appropriate. Otherwise, the statements should be adjusted to not conclude whether TSEN54 protein levels differ or not. -Irrespective of the above comments the conclusion of the section "TSEN54 expression in cerebellar and neocortical organoids", that currently reads "we did not find differences in expression between cell and tissue types" should be changed, as the authors did not investigate whether there are cell type-specific differences of TSEN54 expression.

      1. Organoid growth analysis: The organoid growth analysis in Figure 3 and supplementary Figure 2 shows the main phenotype of the study that seems to be very strong. The authors use unpaired t-tests to compare within the different timepoints. Unfortunately, I think this approach might not be appropriate as even though the Welch correction does not rely on similar SDs in the compared groups (Control vs. PCH2a), it still assumes that all data points within each group share the same variance. However, this is not the case, as e.g., the control condition includes three groups (Control-1 to -3), that between groups might have different variance as such not all datapoints are independent from each other. Potentially ANOVA analyses controlling for cell line and timepoint might be more appropriate. Or additionally, the authors could consider using the linear regression analysis in Supplementary Figure 2 to further investigate the difference in organoid growth by e.g., comparing the slope of the regression lines. This might be more appropriately reflecting the growth deficit over time than simply comparing each timepoint individually. Expanding on this analysis the regression analysis requires some more information on the fit (intercept, slope, R-squared of the model), which would help clarifying the growth dynamics in the different systems and conditions. The growth ratio analysis (Figure 3D) is essential to the major claim of the paper that the organoids replicate the region-specific differences. As the authors performed all experiments with matching cell lines this could additionally strengthen the argument by generating the ratio of size differences for each cell line separately (instead of just for all PCH2a lines together). This would allow comparison of the same genetic background in both cerebellar and neocortical condition and further corroborate the region-specific severity of the phenotype. Potentially, this would also enable to test these differences statistically. Additionally, all growth analyses for the neocortical organoids (Figure 3C, Supplementary Figure 2B and C) seem to lack the PCH-1 cell line and only contain PCH-2 and PCH-3. This cell line should be added or commented on why it was excluded from the analyses.

      2. Potential mechanism of the phenotype (apoptosis analysis): In Figure 6 the authors investigate the hypothesis that increased apoptosis contributes to the phenotypes. In the cleaved Caspase 3 staining there appear to be no differences. Unfortunately, the analysis apparently only includes one replicate (one organoid?) per cell line and condition. Considering the variability in the data shown this seems inappropriately low and should ideally contain ~3 replicates per cell line condition to judge technical and biological variability if the authors want to make the point that there is no "significant difference between PCH2a and control organoids at any time point in both cerebellar and neocortical organoids". Otherwise, this claim does not seem to be substantiated enough by the data. Finally, due to the absence of a phenotype related to apoptosis the authors conclude that the phenotypes may be due to "deficits in the proliferation of progenitor cells". Although this is mentioned in the introduction and the discussion, there is no evidence in the current study that supports this interesting idea. By adding relatively straight forward co-staining experiments for e.g., SOX2 (progenitors) and Ki67 (proliferating cells), the authors could provide further evidence for this hypothesis using existing organoid sections. This would support this speculative idea and could add a more mechanistic insight to the study, thereby making it more exciting.

      Minor comments: - Cell line and quality control: The authors recruit three male patients with PCH2a and reprogram iPSCs. These cell lines are subjected to a well performed extensive quality control. However, it is unclear what cell lines the stainings (e.g., Fig. 1D to I) originate from. Furthermore, the supplementary qPCR analysis (Supplementary Figure 1) includes only the PCH-1 line, and additionally two cell lines that are not explained (F-CO and hESC-I3). It is unclear what the relevance of showing the qPCR of these cell lines is. To ensure proper QC for all used cell lines the authors should provide data for all cell lines (PCH-1 to -3 and control-1 to -3), or at least summarize (e.g., in a table) what QC metrics were applied to which cell line. Most importantly, this information is completely lacking for the control cell lines and the QC is just mentioned in the text. Unfortunately, it is unclear where the control cell lines originate from, and some basic information would be required to judge whether they are appropriate controls: are they iPSC or ESC, were they reprogrammed with a similar paradigm as the PCH2a cells, what is the gender of the control cell lines (all PCH2a cell lines are apparently male)?

      • To make the study more approachable for a medical audience and to judge the variability in phenotype presentation among the recruited patients it would be appreciated if more information on the patients would be provided. The authors write: "We identified three individuals that display the genetic, clinical and brain imaging features previously described for PCH2a.". This information including age/date of birth, as well as other medically relevant information could be provided in the supplementary figure (e.g., is there a difference in disease burden among the different patients?). This would allow judging the recruited cohort better.

      • According to the method section the cerebellar and neocortical organoids were cultured in very different medium especially at later timepoints. While neocortical organoids were kept in a neural maintenance medium based on Neurobasal-A, cerebellar organoids were kept in a medium based on BrainPhys. These media contain very different levels of nutrients, especially of glucose (25mM vs 2.5mM, Bardy et al. 2015). This can have a strong phenotype on proliferation of progenitors and proliferative phenotypes (e.g., see Eichmüller et al. 2022). Especially as the authors claim that there is a difference in the PCH2a phenotypes between brain regions, it should be excluded that this is due to medium differences at later timepoints. When investigating the growth curves of Figure 3B and C it seems like the major difference in growth speed seems to be that neocortical organoids grow faster in early timepoints (<d30), but similar at later timepoints, which would exclude effects of the media at late timepoints. Nevertheless, considering the strong effect media glucose concentration can have the authors should investigate whether there is an effect at growth speed at later timepoints by comparing control organoids. This could also strengthen the region-specific phenotype due to PCH2a.

      • Staining examples shown and presentation: In several figures the authors could improve the presentation of the staining examples with some changes: o Cell line information for images: as the authors only ever note the condition (PCH2a or Control) but not the cell line it is unclear if the stainings all come from one cell line or from multiple different cell lines. This prevents comparing the different differentiation conditions. Additionally, for major conclusions the authors should consider including supplemental stainings or further information on how reproducible the results shown are (how many cell lines and batches were used?). o Selection of examples: in several cases (Fig 2C/D, 4A, 6A/B) the selected images depict very different regions, e.g., one condition shows a large rosette, while in the other condition no rosette can be seen. It would be more appropriate to show matching examples where possible. o Color code of stainings: Colors do not match throughout the manuscript in immunofluorescence images. E.g., Fig. 4 uses blue, green, red, magenta and Fig. 5 uses blue, green, magenta, cyan. It would be preferable to adhere to one color code. Considering significant fraction of the population is having red-green blindness, the latter color code seems more appropriate as it should ensure readability also for color-blind audiences.

      • Small typos: o Figure 1 legend: last sentence "The" instead of "Th" o Supplementary Figure 1B: PCH-2 is named "PCH-22" o Supplementary Figure 2: As in the main figure for neocortical organoids the PCH-1 condition is missing (see comment on organoid growth curves). Additionally, the color/shape code of the plots in B does not always match the legend (e.g., size in left plot is different and color of PCH-3 in middle and left plot differs from legend and right plot). o It is unclear why the cortical organoids are referred to as "neocortical organoids" in the figures and the text. The methods and the reference in the methods as well as all major papers rather use the word "cortical".

      References: Bardy, C. et al. Neuronal medium that supports basic synaptic functions and activity of human neurons in vitro. Proc National Acad Sci 112, E3312 (2015). Eichmüller, O. L. et al. Amplification of human interneuron progenitors promotes brain tumors and neurological defects. Science 375, (2022).

      CROSS-CONSULTATION COMMENTS I agree with the comments of the other reviewers and as they are mostly matching, this reinforces the importance to improve certain aspects of the manuscript. As there are no deviating issues I do not comment specifically on any reviewer comments.

      Significance

      This work is using organoid technology to shed light on brain region-specific phenotypes in PCH2a. Brain organoids have drastically changed the way we study human neurological diseases (Eichmüller and Knoblich 2022), however, most brain organoid research has focused on cortical organoids. Cerebellar organoid protocols exist for some time (Muguruma et al. 2015, Silva et al. 2020, Nayler et al. 2021) but were not yet applied to uncover new disease biology. Especially considering the important role of human-specific cerebellar processes in specific developmental disorders (Haldipur et al. 2021) and cancer (Hendrikse et al. 2022, Smith et al. 2022), disease modeling in human cerebellar organoids holds great potential for understanding disease biology. The work by Kagermeier et al. demonstrates that human cerebellar organoids are recapitulating brain region-specific growth deficits and thus is an important step forward for disease modeling. Therefore, this work will be interesting to researchers working on brain development and disease modeling, especially in in-vitro systems. Nevertheless, the mechanistic insight of the study is limited, as is the insight into how human-specific processes might be involved in the pathogenesis of PCH2a. Therefore, it will be interesting how this disease model will be used in future to investigate the cell types and mechanisms involved in the PCH2a phenotype.

      Personal field of expertise: Brain organoids and disease modeling in organoids especially of neurodevelopmental diseases. Analysis of organoids with stainings, as well as sequencing techniques, and bioinformatics.

      References:

      Eichmüller, O. L. & Knoblich, J. A. Human cerebral organoids - a new tool for clinical neurology research. Nat Rev Neurol 1-20 (2022) doi:10.1038/s41582-022-00723-9.

      Haldipur, P. et al. Evidence of disrupted rhombic lip development in the pathogenesis of Dandy-Walker malformation. Acta Neuropathol 142, 761-776 (2021).

      Hendrikse, L. D. et al. Failure of human rhombic lip differentiation underlies medulloblastoma formation. Nature 609, 1021-1028 (2022).

      Muguruma, K., Nishiyama, A., Kawakami, H., Hashimoto, K. & Sasai, Y. Self-Organization of Polarized Cerebellar Tissue in 3D Culture of Human Pluripotent Stem Cells. Cell Reports 10, 537-550 (2015).

      Nayler, S., Agarwal, D., Curion, F., Bowden, R. & Becker, E. B. E. High-resolution transcriptional landscape of xeno-free human induced pluripotent stem cell-derived cerebellar organoids. Sci Rep-uk 11, 12959 (2021).

      Silva, T. P. et al. Scalable Generation of Mature Cerebellar Organoids from Human Pluripotent Stem Cells and Characterization by Immunostaining. J Vis Exp (2020) doi:10.3791/61143.

      Smith, K. S. et al. Unified rhombic lip origins of group 3 and group 4 medulloblastoma. Nature 609, 1012-1020 (2022).

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Please find enclosed my recommendation for the paper submitted by Kagermeier et al entitled' Human organoid model of PCH2a recapitulates brain region-specific pathology'. It describes the development of a human model for PCH2a and its characterization. My overall assessment of the paper is 'Major revision' which is explained below.

      Although the paper is very well written and clearly interesting in that it describes the generation and initial analyses of a human organoid model for PCH2a it should be revised such that it will proof the points it is trying to make. The authors are meticulous in their studies combining cellular characterization and a thorough initial screen of organoid (both cerebellar as well as cortical) integrity, yet hardly any mechanistic data is provided. Nevertheless, if the authors are able to add additional experiments and are able to address the points raised, the reviewer may be willing to consider a more positive outcome.

      Major concerns

      1. The overall quality of the figures is poor. There is a lot of overexposure such that often cellular or tissue structures are blended. It starts with Figure 1 G and H but can be observed throughout the manuscript. Deconvolution would greatly enhance their results.
      2. Especially figure 4 and 5 could have been complemented with quantitative data. It furthermore seems more supplemental figure as these are just proof-of-principle stainings. No conclusions can be drawn from the panels except that all markers are there in the various conditions. And while they are showing a neural rosette in Fig 4A, just tiny ones can be observed in 4B. It is also not clear what the whole mount IHC ads in comparison to the IHC on sections. It is also strange that there is still a lot of SOX2 in the CALB/MAP2-positive area, but again with this magnification hard to appreciate.
      3. If the authors would like to proof the point that cerebellar/cortical development is hampered, more functional assays could have been done. Nothing is analyses on the fraction of progenitor cells present (such as the percentage of Tbr2+ IPC in VZ/CP). Furthermore, if there is a suspicion that the number of cells is affected (which is also not shown), proliferation/cell cycle exit experiments using BrdU/EdU should have been performed. Early cell cycle exit still cannot be rules out and should have been tested by the combination of Ki67-/EdU+ percentage of a certain faction of progenitor cells (eg PAX6+ pool).
      4. Instead the author chose to only perform a cCas3 staining. From the panels in Figure 6 it is hard to appreciate which cells are actually cCas3+. Also the analyses were performed on the total pool of cell while it might have been more interesting to look for cell death of the various progenitor pools (eg the SOX2+ pool).

      Minor concerns

      1. It would greatly enhance the review process if line numbers are added
      2. On general concepts (such as the generation of organoids in the context of disease) more references could have been added

      Figures

      Fig. 1: In A, the square is clearly visible and not similar to B. An annotation of which is the control and which is the patient is missing in the figure. The arrows are hardly visibly, would make them slightly bigger and remove the black outer lining. Figure 1C can easily go to the Supplemental material. Fig 1 D is hard to appreciate the staining, a close-up with bright field microscope will help. E-I Most of the panels but especially G and H are overexposed. In J, it is hard to appreciate the TSEN54 staining. Maybe separate channels and a merge?

      Fig. 3: Usually go into the supplementals

      Fig 4/5: Lack of quantitative data and poor quality of figures (overexposure).

      Fig 6: Many of the SOX2 panels are overexposed

      Referees cross-commenting

      I completely agree with reviewers #1 and #3. It is good to notice that we are overall on the same page.

      Significance

      The authors definitely made an excellent start to model PCH2a. Three controls and three patient lines are good to begin with but isogenic controls using one parental line and a patient line where the mutation is fixed would have been ideal. It is interesting that there seem to be a brain area specific pathology of the phenotype. Yet, more thorough analyses could have been performed such as proliferation and differentiation and cell cycle exit experiments. As for now the mostly descriptive data are only scratching the surface and little can be concluded on the molecular framework they are trying to solve.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      In this manuscript, Kagermeier et al. present a novel and interesting study that attempts to model a severe neurodevelopmental disorder, pontocerebellar hypoplasia type 2a, using neocortical and cerebellar organoids. Brain organoids are an appropriate and promising approach to elucidate disease mechanisms in neurodevelopmental diseases. The authors show a reduction in the size of the organoids which is more pronounced in the cerebellar compared to neocortical organoids. While this finding is interesting and reminiscent of the clinical PCH2a phenotype, i.e., cerebellar hypoplasia, the study is very preliminary and the conclusions of the manuscript are not supported by the data. Additional information and further experiments are necessary to support the claims made.

      Major concerns:

      1. hiPSC lines show considerable inter- and intra-individual variability and therefore the size differences observed between these control and patient-derived organoids may arise from differences in the hiPSC lines used. While the data sufficiently demonstrates the pluripotency of the multiple novel hiPSC lines, major concerns remain as to the appropriateness of the control hiPSC lines. The manuscript should include a table describing the age and sex matching as well as mode of reprogramming for all control and patient lines. Patient and control lines should be matched as closely as possible. Furthermore, figure legends should clearly indicate which clones and lines are shown in the various figure panels.
      2. As the hiPSC lines used are not isogenic, it is important that the authors characterise these lines further. This should include a quantification of the rates proliferation and apoptosis in all used hiPSC lines, as these might impact the growth rate of the embryoid bodies / organoids.
      3. The authors state that the hiPSC lines have been characterised by SNP arrays to show that no genomic / chromosomal aberrations have been accrued due to reprogramming. The manuscript should include information as to when the SNP array was performed (i.e., immediately after reprogramming, after initial passaging, etc) and also include the results of the SNP array as additional information. What passage were the hiPSC when the presented experiments were carried out?
      4. Given that TSNE54 is broadly and strongly expressed in the developing nervous system, the very limited staining of the organoids for TSNE54 in Figure 2 is surprising. Can the authors provide an explanation for the fact that TSNE54 is only expressed in a small subset of cells? Which cell types are these? Moreover, high-magnification images should be shown to demonstrate subcellular staining pattern of TSNE54. Quantification of TSNE54 protein levels by immunoblotting would also be beneficial. Related to this observation, it is puzzling that the large size differences that the authors observe in their organoids would be driven by such a small number of TSNE54-expressing cells. How do the authors explain this discrepancy?
      5. The generated organoids need to be better characterised with a broader range of markers using both qPCR and immunostaining. At the moment, their identity as "cortical" and "cerebellar" organoids remain unconvincing. This is particularly true for cerebellar organoids, which are challenging to generate and are not widely used. The authors should include additional markers (for example, see PMIDs 25640179, 29397531, 32117945) and immunostaining should clearly show expected staining patterns. In Figure 5, it appears that some markers (e.g., SATB2) are expressed differently between control and patient lines, yet this is not commented on by the authors who conclude that control and patient lines show differentiation into organoids.
      6. The authors attempt to look into a potential mechanism for the size differences observed between control and patient organoids. However, only cleaved caspase-3 is used as a marker for apoptosis and no differences were observed. The authors should include further markers for potential cell death. In addition, immunostaining for proliferation markers (i.e., KI67) should be performed to evaluate whether the difference in organoid size could stem from decreased proliferation rather than increased cell death.

      Significance

      The authors present an innovative approach to study neurodevelopmental disorders using brain organoids and should be of interest to researchers and clinicians working on neurodevelopmental diseases. However, the data presented are too limited to support any conclusions about the phenotype observed. Furthermore, questions remain about the used methodology and more work is needed to demonstrate the successful generation of both cortical and cerebellar organoids.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      We thank all three Reviewers for their thorough assessment of our manuscript and their constructive comments and suggestions.

      Reviewer #1 (Evidence, reproducibility and clarity (Required)):

      In this study, the authors generate several variants of actin that are internally tagged with short peptide tags. They identify one particular position that is able to tolerate various tags of 5-10 amino acids and still shows largely unaltered behavior in cells. They study incorporation of their tagged actins into filaments, characterize the interactions of G-actin variants with different associated proteins and show that retrograde actin flow in lamellipodia and the wound healing response of epithelial cells is not affected by the tagged variants. They then apply the tagged actin to study subcellular distribution of different actin isoforms in mammalian and yeast cells.

      The identification of a specific site in the actin protein that tolerates variable peptide insertions is very exciting and of fundamental interest for all research fields that deal with cytoskeletal rearrangements and cellular morphogenesis. The result demonstrating the functionality of actin variants with peptides inserted between aa 229 and 230 are generally convincing and well done. In particular, the generation of CRISPR/Cas9 genome edited versions of beta- and gamma actin are impressive. I therefore generally support publication of this study. There are however several technical and conceptual issues that should be addressed to improve quality and scope of the study. I listed some specific comments below:

      We thank the Reviewer for their constructive comments and general support for publication of our study.

      Major points

      - The biggest issue I have is the last section on the application of tagged actins to study isoform functions. In principle the application is very clear as there are simply no alternative ways to study isoform distribution in live cells. However, the experimental data are simply not convincing. What the authors define as "cortex" in Fig. 5A seems to rather represent cytosolic background mixed with radial fibers. I am not convinced that even the antibody staining with a relatively clear differential distribution of beta and gamma really shows a genuine accumulation of one isoform on stress fibers. It seems to me that the beta-actin staining has as higher cytosolic background and is generally weaker (gamma nicely labels transverse arcs), which reduces signal/noise and therefore yields a relatively increased level in areas with less-bundled actin. My suggestion is to select more clearly defined actin structures and to use micro-patterned cells to normalize the otherwise obstructing variability in actin organization. Possible structures would be cortical arcs in bow-shaped cells, lamellipodial edges (HT1080 seem to make very nice and large lamellipodia) or cell-cell contacts (confluent monolayer, provided cells don´t grow on top of each other). Stress fibers are possible but need to be segmented very precisely and I did not see any details on this in the methods section. For Fig. 5D: I assume cells were used where only one isoform was tagged? This is technical weak and the double-normalization is probably blurring any difference that might be occurring. Why not use a double-tagging strategy with ALFA/FLAG or ALFA/AU5 tags to exploit the constructs introduced in the previous figures? Also, the unique selling point of the strategy is the possibility of actual live imaging of specific isoforms. Cells that have stably integrated double tags and then transiently express nanobodies for ALFA and either AU5 or FLAG (or other if those don't exist) would make this possible. Considering the work already done in this manuscript, such an approach should actually be possible - did the authors attempt this or is there are reason it is not discussed? If double tagged cells are not possible for some reason it should at the very least be possible to combine ALFA-detection with the specific antibody against the other isoform and get rid of the double normalization.

      We thank the Reviewer for the various suggestions regarding the comparison between the localization of the tagged and native isoforms. In our reply below, we will separately discuss the possibilities and our considerations for fixed samples and live cell imaging. We apologize for the lengthy response but for transparency reasons, we would like to give a thorough overview of our efforts for isoform-specific localization in cells, something for which we have limited space in the manuscript.

      Fixed samples:

      It was a significant experimental challenge to comparing the labeling of the β- and γ-actin specific antibodies with our internally tagged actin system (Fig. 5A-D). The reason for this is that the labeling of the samples with the β- and γ-actin specific antibodies requires treatment with methanol (Dugina et al., J Cell Sci, 2009), most likely to disturb the interaction of actin with actin-binding proteins that prevent the binding of the antibodies due to steric hindrance. Methanol treatment, however, precludes the co-labeling with phalloidin, likely due to changes in the tertiary/quaternary protein structure of F-actin. Initially, we have put a lot of effort in trying to simultaneously label phalloidin with the actin specific antibodies but even very brief methanol treatment (seconds), before or after phalloidin labeling, completely prevents/reverses the binding of phalloidin. Importantly, also the ALFA tag labeling was suboptimal after methanol treatment.

      The fact that we could not perform these double labelings led us to perform different ratio calculations for the β- and γ-actin antibody and the ALFA tag labeling. In the case of the antibody immunofluorescence labeling, we simply divided the signal of the β-actin and γ-actin since we could simultaneously label the isoforms in the same cell. In the case of the ALFA tag labeling, we used phalloidin for independent signal normalization and then performed a second normalization. Although this complicates the normalization procedure (ALFA tag signal of β- and γ-actin is first normalized to total F-actin and then a ratio is calculated) and understandably leads to some confusion, this was the only way forward to obtain the results presented in the manuscript.

      The Reviewer points out that “What the authors define as "cortex" in Fig. 5A seems to rather represent cytosolic background mixed with radial fibers.”. In our images, we observe very little cytosolic background from both antibody stainings. More importantly, for the quantitative analysis, the fluorescence intensity values were corrected for the background values observed in cytosolic areas so even if the signal is present, it should not affect our analysis. We do admit though that we could have been more careful with the term “cortex” since the observed signal could indeed be a mix of radial fibers and the actin cortex. The reviewer further states that “I am not convinced that even the antibody staining with a relatively clear differential distribution of beta and gamma really shows a genuine accumulation of one isoform on stress fibers.” Although the differences are small, we consistently observe a differential fluorescence intensity of β- and γ-actin in actin-based structures with a relatively stronger signal of γ-actin in stress fibers (Fig. 5C). Since we always normalize the fluorescent signal intensity per cell, this strongly indicates a genuine accumulation of one isoform over the other in specific actin-based structures. This observation is very consistent in our experiments and also aligns with many published studies where differences in the localization of β- and γ-actin are reported in various cell types (Pasquier et al., Vasc Cell, 2015; van den Dries et al., Nat Comms, 2019; Malek et al., Int J Mol Sci, 2020). As for the segmentation, we mentioned in the Methods section that we selected small regions (0.5x0.5mm) that exclusively contain stress fiber or “cortex” regions. The regions shown in Fig. 5B are therefore larger than the analyzed regions, something which we will better indicate in the revised manuscript.

      Planned revision: We will provide a more detailed explanation of our quantitative analysis in the Methods section such that it is more clear how our normalization procedure was performed. Furthermore, we will adapt Fig. 5A-B such that it better visualizes how we defined the regions for quantification. As per the Reviewer’s suggestion, we will also apply a different experimental method to show that the tagged isoforms properly localize to actin-based structures. For this, we will attempt to use micropatterned cells to induce clearly define actin-bases structures (the crossbows as suggested by the Reviewer) and also explore the possibilities of investigating the differential localization in double-tagged cells. We will also reconsider the use of the term “cortex” for the region that is pointed out in Fig. 5A-B.

      Live cell imaging:

      We agree with the Reviewer that it would be very valuable to attempt simultaneous live cell imaging of two isoforms. Yet, for this, we would need two tag/fluorophore systems that allow the visualization of internally tagged isoforms in living cells. As presented in our original manuscript, we have successfully inserted many different epitope tags (FLAG/AU1/AU5/ALFA) in the T229/A230 position to demonstrate the versatility of our tagging approach. Yet, despite significant efforts to identify a second tag/fluorophore system that would allow isoform-specific live cell imaging, we only succeeded in designing one strategy to perform live cell imaging, i.e. with the ALFA tag (Götzke, Nat Comms, 2019). Part of the reason for this is that so far, no high affinity nanobodies have been generated against the classical epitope tags (FLAG, AU5 etc.). This is an established challenge since classical epitope tags are typically linear/unstructured while nanobodies require folded secondary structures for epitope recognition such as alpha helices (the ALFA tag was specifically designed as such).

      Besides the successful ALFA tag approach we have tried the following additional approaches for live cell imaging: 1) __full-length GFP, 2) full-length GFP with linker, 3) GFP11 (to complement with GFP1-10 (Cabantous et al., Nat Biotech, 2005) 4) GFP11 with linker 5) FLAG Frankenbodies (Zhao et al., Nat Comms, 2019; Liu et al., Genes Cells, 2021) in FLAG IntAct cells and 6) __Tetracysteine/FlAsH labeling. Importantly, each of these additional internally tagged actins, except for those that contained full-length GFP, showed a high colocalization with the cytoskeleton, again demonstrating the versatility of the T229/A230 position to tag actin. Unfortunately, none of these approaches satisfactorily visualized the actin isoforms in living cells. We will therefore briefly summarize our findings here.

      (1-2, integration of full-length GFP and GFP with linker) Probably not surprisingly, but integrating the entire coding sequence of GFP or GFP flanked by linkers (each 5AA in length) within the T229/A230 position did not results in a proper localization of actin.

      (3-4, integration of GFP11 and GFP11 with linker) Next, we assessed the localization of the GFP11 tagged actin versions (GFP11: 16AA, GFP11+linker: 26AA). Because GFP11 is not visible without GFP1-10 complementation, we also tagged actin at the N-terminus simply for proof of concept where the internally tagged actins would end up. Interestingly, both GFP11-actin and GFP11+linker-actin properly integrated within the cytoskeleton as demonstrated by the FLAG staining. This again demonstrates the versatility of the T229/A230 position and strongly suggests that even the integration of 26AA within this position does only minimally affect the polymerization of actin into the cytoskeleton.

      (3-4) After confirmation of the proper integration of GP11-actin and GFP+linker-actin we continue to express the GFP1-10 in these cells. Unfortunately, this resulted in no or only very minimal localization of the actin to the cytoskeleton, demonstrating that GFP-complementation hampers the integration into the cytoskeleton.

      (5, use of FLAG Frankenbodies) We also expressed FLAG Frankenbodies into our FLAG IntAct cells in an attempt to visualize the isoforms in living cells. FLAG Frankenbodies are single chain antibodies fused to GFP and can be expressed in cells to visualize FLAG-tagged proteins (Liu et al., Genes Cells, 2021). Although a cytoskeletal labeling was indeed discernable in some cells, the FLAG Frankenbody signal overlapped much less with the total actin signal as compared to the FLAG immunofluorescence labeling, indicating that the incorporation of the FLAG-tagged actin was much less in the presence of the FLAG Frankenbody. Also, a significant fraction of the cells demonstrated a homogenous cytosolic signal.

      (6, Use of tetracysteine/FlAsH) Although the tetracysteine tag/FlAsH system is widely known to induce artefacts, we still aimed to evaluate if for live cell imaging of IntAct actins. Similar to GFP11, we first determined the integration of tetracysteine-actin into the cytoskeleton with the use of an additional N-terminal FLAG tag and demonstrate that it was properly integrated into the actin cytoskeleton. Unfortunately, after brief incubation with FlAsH-EDT2, we noted 1) a significant amount of background fluorescence, preventing proper actin visualization and 2) that the cell became static indicating toxicity of the FlAsH-EDT2 compound. Titrating down the amount of FlAsH-EDT2 did not alleviate these drawbacks and only resulted in less fluorescence.

      Overall, based on these experiments, we concluded that the T229/A230 position itself is very versatile, as demonstrated by the proper localization of the GFP11-actin variants and the TetraCys-actin. At the same time, none of these tag/fluorophore systems properly visualized actin in living cells. Although we are unsure what the reason is for this, it is easily imaginable that the on/off kinetics of the split GFP system and the FLAG Frankenbodies are suboptimal to allow for the rapid and continuous integration of actin monomers into the F-actin cytoskeleton. We therefore also concluded that currently, the ALFA tag/nanobody system is apparently unique in its ability to visualize epitope tagged actin in living cells (as shown in the manuscript). For simultaneous visualization of multiple isoforms, we rely on progress on the development of novel nanobody-based tags, something we hope the Reviewer will agree is outside the scope of the current work.

      *- The authors make a point of comparing the internally tagged actin to N-terminal tags that are mostly functional but have been shown to affect translational efficiency. I would strongly suggest to include N-terminally tagged actin as control for all assays in this study. Also for the physiological assays (retrograde flow, wound healing), a positive control is missing that shows some effect. Previous studies showed defects with transiently expressed actin with an N-terminal GFP. As retrograde flow measurements are very sensitive to the exact position of the kymographs and wound healing assays is a very crude and indirect readout, such a positive control is essential. *

      We acknowledge that N-terminally tagged actin has been used extensively for actin research (especially before the introduction of Lifeact). For our studies, however, we were specifically interested in whether the internally tagged actins show similar characteristics as compared to wildtype actin. We have not included N-terminally tagged actin in all of our experiments, since this would not affect our conclusions with respect to the functionality of our internally tagged actins. We expect that for future investigations to for example further establish the importance of actin N-terminal modifications in the differential regulation of actin isoforms, the comparison between internally and N-terminally tagged actins could be very instrumental. Yet, we consider this comparison outside the scope of the current manuscript. For now, the results in the manuscript provide evidence that our approach is unique with respect to the fact that it allows isoform-specific tagging without manipulating the N-terminus. As such, our internal tagging system complements the already existing repertoire of actin reporting methods (N-terminal fusion, Lifeact, F-Tractin, actin nanobodies) and allows researchers to study so far unknown properties of actin variants.

      *- Expression of tagged actins in yeast is a very nice idea but it would be far more informative to express the tagged forms as the only copy of actin. This can either be done by directly replacing endogenous actin gene in S. cerevisiae, or (if the tagged versions are not viable) - using the established plasmid shuffle system (express actin on counter-selectable plasmid, then knock out endogenous copy and introduce additional plasmid with tagged actin, then force original plasmid out). In the presence of endogenous S. cerevisiae actin the shown effects are very hard to interpret as nothing is known about relative protein levels (endogenous vs. introduced). Also, if constitutive expression of the ALFA nanobody is harmful for integration into cables, why not perform inducible expression of the nanobody and observe labeling after induction. For the live imaging a robust cable marker is needed, like Abp140-GFP. Finally, indicate the sequence differences between the used actin forms in yeast (supplementary figure with sequence alignment and clear indication of all variations) *

      We thank the reviewer for their positive comments and feedback regarding expression of IntAct variants in yeast. Currently, we have expressed IntAct as an extra copy in the presence of native Act1 of S. cerevisiae. All the IntAct variants have been expressed under a commonly used constitutive TEF1 promoter. We agree with the Reviewer that it would be valuable to attempt to express the tagged forms as the only copy of actin.

      Planned revisions:

      1) As per the Reviewer’s suggestion, we will attempt to make yeast strains with IntAct as the sole expressing actin copy by using the well-established 5-FOA-based plasmid shuffle system in yeast. We will use a ∆act1 strain containing wildtype act1 in a centromeric ura-plasmid described in Harrer et. al, 2007 (generously shared by Prof. Jessica and Prof. Amberg at Upstate Medical University of New York, USA) and express IntAct exogenously via additional plasmids. Shuffling of these strains on 5-FOA will cause the loss of ura-plasmid containing the wildtype act1 copy and will determine whether yeast cells will be able to survive with IntAct as the sole source of actin. If the cells do survive with IntAct as a sole copy, we will perform subsequent analysis for assessing actin cytoskeleton organization under these conditions.

      2) As the reviewer has mentioned, expression of NbALFA during live-cell imaging experiments hindered incorporation of IntAct into linear actin cables in yeast (Suppl. Fig. S13). As per the reviewer’s suggestion, we will now try to create an inducible-expression system for the NbALFA-mNG and observe its effects on incorporation into formin-made actin cables after induction. We have already created NbALFA-mNG constructs under galactose-inducible GALS and GAL1 promoters and are currently constructing yeast strains for these experiments.

      __3) __We will add an extra supplementary Figure to indicate the sequence differences of the various actin variants that we have expressed in yeast.

      - As the authors clearly show good integration of several tagged actins into filaments I would expand the structural characterization: perform alpha fold predictions of actin monomer structures including the various tags to show the expected orientation. It is striking that the only integration site that seems to work well is at the last position of a short helix, indicating that the orientation of the integrated peptide might be fixed in space and be optimal to minimize interference. Also, a docking of the tag onto the recently published cryoEM structures of the actin filament should be shown to indicate where it resides compared to tropomyosin or the major groove where most side binding proteins seem to bind.

      We already performed AlphaFold predictions of the tagged actin monomers, but we have decided to not include these predictions in the manuscript because of two reasons. First and foremost, while the prediction confidence of the non-tagged region is very high (pLDDT > 90), the prediction confidence of the tagged region is very low (pLDDT https://alphafold.ebi.ac.uk/faq), pLDDT values below 70 should be treated with caution and values below 50 should not be interpreted. Intriguingly, the low confidence aligns with the fact that for both tags, the prediction does not match with known features of the tag. The FLAG tag should be a linear/unstructured region in order to be recognized by the antibody and the ALFA tag should organize into an alpha helix (Götzke et al., Nat Comms, 2019). Yet, in the prediction, the FLAG tag partially continues as an alpha helix and the ALFA tag is only a small helix with part of the tag being unstructured. Second, more minor, reason for not including the predictions is that AlphaFold does not predict to what extend the tag is flexible, which means that even if the tagged region is predicted correctly, it is difficult to say whether the regions will interfere with binding of proteins.

      Despite the low prediction confidence, we used the published actin-tropomyosin cryoEM structure (von der Ecken et al., Nature, 2015) to replace WT actin with ALFA tag actin and the results are shown below. Again, although results should be interpreted with caution, the tag does not seem to obstruct monomer-monomer interactions within an F-actin filament and also the tropomyosin binding surface is relatively distant from the tag region, suggesting that these interactions are likely not disturbed by introducing the tag.

      - For any claims regarding usability of tagged variants for isoform research it would be very important to characterize the known posttranslational modifications of tagged actin variants - are the differences between beta and gamma maintained on this level as well?

      Planned revision: Following the Reviewer’s suggestion, we will perform a western blot analysis to compare posttranslational modification (arginylation) of tagged and wildtype actins.

      Technical issues

      - There is no scale for the color coding in Fig. 5A, B

      We deliberately did not add a numerical scale because the images are normalized which means that presenting the actual numbers might be misleading. The numbers could be interpreted as if they actually present the amount of β-actin relative to γ-actin which is not the case due to staining differences and the normalization procedure.

      - The y-scales for Fig. 5C and D need to be identical to allow direct comparison

      Planned revision: We will adapt the scale of Fig. 5D to make it identical to Fig. 5C. Following the other suggestions of the reviewer, we will also critically evaluate our normalization procedure and present those numbers in Fig. 5C-D if the values turn out to be different.

      - Pearson coefficient should not be normalized to a control value as its already a dimensionless parameter. Always report actual R-value - also remove R2 values for Pearson as this makes no sense in this context (not sure if it was a typo or intended).

      We normalized the Pearson coefficient values for visual representation of the results. The majority of the raw coefficient values (more than 80%) are between 0.20 and 0.75 (see raw values in the associated excel file). Theoretically, Pearson coefficient values are possible between 1 (or-1 for negative correlations) and 0. The much smaller window in our values as compared to the theoretical window (0.55 vs 1) led us to normalize the values such that they can be presented on a scale from “maximum expected colocalization” to “minimum expected colocalization”. In this way, the differences between the various tagged actins are much better appreciated in the Figure. As to reporting the R2, the Reviewer is correct. Reporting the R2 is an inadvertent mistake from our side and we will correct it.

      Planned revision: We will change the R2 in the text to PCC or Pearson Correlation Coefficient.

      *- All values on subcellular regions (like stress fiber or cortex) dependet critically on the way thesese regions were thresholded or identified. Provide all details on how this was done in the methods section and ensure that adequate background subtraction and normalization is applied. Optimally, an unbiased (AI or automated) approach based on simple image statistics is used for this to avoid personal bias. *

      Planned revision: As also indicated above, we will add new experiments to better compare the localization of the isoforms in tagged and parental cells. These new experiments will also be accompanied by a more detailed explanation of how the regions were selected and quantified.

      - In Fig. 2A only heterozygous FLAG-actin cells are used. Why not use a homozygous line (for both beta and gamma actin)? The nice band shift of the FLAG version would allow the precise quantification of the fraction of total actin covered by beta and gamma actin, which then could provide some additional info for the apparently weaker beta staining in Fig. 5 (if beta expression is simply weaker). This would be a very simple and useful advantage of the internal tags that could be widely applied.

      In Fig. 2A, we used the heterozygous FLAG-actin cells to directly compare the production of β-actin from the knock-in allele and the wildtype allele in the same cells. The fact that the two bands observed in this western blot analysis (upper and lower) are almost the same (with the FLAG band being a bit more intense) provides the strongest indication that the tag does not interfere with the expression of actin. In Suppl. Fig. 5D, we show that the expression of β-actin is also unaffected in the hemizygous FLAG actin cells, which exclusively express tagged actin.

      Planned revision: As per the Reviewer’s suggestion, we will also add a western blot analysis on the expression of both actin isoforms and total actin in hemizygous cells.

      *- Fig. 3: control with N-terminal tag is missing. Also, why is it not possible to assay filament binding factors like Myosin, Filamin or alpha actinin - instead of co-IP a simple co-sedimentation assay with cell extracts in F-buffer should pick up any major difference in decoration of filaments containing the ALFA tag. Using two speeds for centrifugation it might even be possible to observe effects on filament bundling. The best approach for this would of course be to purify tagged actins and perform in vitro assays but this is clearly beyond the scope of what the authors intended here. I personally think that a broad acceptance of the marker will only come once the biochemistry has been sufficiently characterized so this is a future direction I would strongly encourage. *

      We kindly refer to our response on Page 5/6 for why we have not included the N-terminal control.

      Planned revision: The co-sedimentation assay is an excellent suggestion by the reviewer. Following the Reviewer’s suggestion, we will perform F/G-actin fractionation and assess the presence of several F-actin associated proteins in the F-actin fraction.

      - Fig. 2A has no loading control

      We show this western blot to indicate that the WT actin and tagged actin are expressed at similar levels in the heterozygous knock-in cells. For this, no loading control is needed because we only compare the intensity of the upper band (tagged actin) with the lower band (WT actin).

      - The RPE-1 data are confusing as several constructs show very different localization (completely cytosolic) to HT1080 cells and there is no possible explanation given for this. Maybe simply remove this data set?

      We agree with the reviewer that the differences in the localization between some of the internally tagged actins between the HT1080 and RPE1 cells might be confusing, especially for the A230-A231 variant for example. Yet, the fact that also in these cells, the T229-A230 variant performs equally well as compared to N-terminally tagged actin is an important confirmation that this variant is properly integrated into actin-based structures, independent of cell type. This makes the support for choosing this variant to continue with our studies stronger. A possible explanation for the differences is that RPE1 cells in general tend to form more stress fibers as compared to the HT1080. Since the localization to stress fibers is different between the internally tagged actins, this may explain the differences observed in colocalization.

      __Planned revision: __We will add a short text, in the Results or the Discussion, on the differences between the colocalization values between HT1080 and RPE1 cells.

      *- The angel measurements for lamellipodial actin is not very meaningful: the angel is determined for the radial bundles, which do not correspond to the Arp2/3 angel of single filaments and is likely the results of different nucleation factors, I would suggest to remove this. If angel measurement are really intended, cryoEM needs to be performed. *

      We apologize for this misapprehension from our side which is also noted by the other two reviewers. In the treadmilling videos of the lamellipodia in HT1080 cells, which were obtained using Airyscan super-resolution microscopy, we clearly observe a consistent filament formation at a constant angle, something which we interpreted as the angle between the mother filament and the daughter filament. After consulting the literature, we indeed have to admit that this cannot be interpreted as such and we will remove these datasets.

      Planned revision: We will remove the datasets with the angle measurements (Suppl. Fig. 7A-B) from our manuscript.

      - Replace all SEM with SD values - use at least 3 biological replicates (4D SEM of n=2)

      Planned revision: We will carefully check our statistics and revise where appropriate.

      Minor points

      - Intro: after listing all the details already understood on actin isoforms it is not very convincing to simply state the molecular principles remain largely unclear (l 34) - maybe better "there is no way to study actin dynamics due to current limitations of specific antibodies to fixed samples. Interesting option would be actually to develop nanobodies that are isoform specific.

      We will rephrase the text in the introduction. Regarding the development isoform-specific nanobodies. Although this sounds like a promising way forward, this would likely not result in isoform-specific targeting in living cells. Similar to the antibodies, isoform-specific nanobodies would have to be generated against the N-terminus which, under native conditions, is likely not available due to the occupation with actin-binding protein. Also, since the N-terminus is not structured, it may be extremely challenging to generate nanobodies against these epitopes.

      *- L 71: "involved" in the kinetics is not a good term - maybe affects or regulates.... *

      We will rephrase the text.

      - L148: "suspect" instead of "expect" - this clonal variation is actually a big danger of the employed approach as possible defects in actin organization could be masked by compensatory changes - it would generally be good to show critical data for at least 3 independent clones to rule out dominant selection effects.

      We will rephrase. We agree that clonal variation could be a danger if actin levels are to be investigated. For future follow-up studies, we plan to make additional cell lines to avoid clone-specific conclusions.

      ***Referees cross-commenting** *

      *I completely agree with the comments by reviewer 2 on the various missing controls - adding several or all of those will make the results much more convincing. The key for the adaptation of any new actin probe will be the level of confidence researchers have on the doumented effects. Even some negative effects on actin behavior (I am sure there will be some) should not prevent usage of the strategy as long as there is robust and convincing documentation of those effects. I also agree that including some basic in vitro characterization will go a long way to convince people dierectly working on actin (there is a very high level of biochemical understanding in that field). *

      Planned revision: We will perform the essential controls as suggested by Reviewer 2. Furthermore, for future experiments, we do envisage the production and purification of internally tagged actins and investigate their binding properties in in vitro reconstitution assays. We have already started with optimizing these approaches through our ongoing collaboration (KD, SP).

      Reviewer #1 (Significance (Required)):

      *Significance: Very useful finding that can be applied to any question related to actin-dependent cellular processes (morphogenesis, cell division, cell polarization, cell migration etc.) *

      *Strength: main finding convincing, strong genome edited cell lines *

      *Limitations: application to study of isoforms very limited and data not convincing, statistics and image quantifications need improvement *

      *Advance: identify new location for integral tagging of actin, which was not really possible before. The main relevance is for fundamental cell biology but the approach can also be applied to the study of disease variants in actin. *

      Audience: general cell biology - very broad interest

      __Reviewer #2 (Evidence, reproducibility and clarity (Required)): __

      Actin is highly sensitive to modifications, and tagging it with fluorescent proteins or even smaller motifs can affect its function. The most well-known example of this is that fission yeast where actin has been replaced with GFP-actin are inviable (Wu and Pollard, Science 2005) because the labeled actin cannot incorporate into the formin-dependent filaments that make up the cytokinetic ring. Subsequent experiments revealed that formins filter out GFP-actin monomers, as well as monomers that are labeled with smaller fluorescent motifs (Chen et al, J. Structural Biology 2012). Further, attempts to make mammalian cells lines where GFP-beta-actin was knocked into one allele resulted in extreme down-regulation of the GFP-labeled actin, indicating that there is some implicit toxicity with the labeled version. To my knowledge, all attempts at making homozygous GFP-actin knock-ins have been unsuccessful. Therefore, while GFP-actin or other labeled variants can be over-expressed in many different cell types with some success, there is always the question of how faithful the labeled actin represents bona fide actin localization and dynamics.

      To address this van Zwam et al. have developed a clever strategy of screening actin for internal motifs that can tolerate incorporation of a tag without affecting its function. They appear to have found a good candidate, named IntAct, and provide evidence that this tagging position allows the actin to be functional in both human and yeast cells. The work is very promising, and many of the assays performed satisfy the criteria of rigor and reproducibility. Importantly, the authors have created knock-in human cell lines where the tagged actin is expressed at normal levels, including a double allele knock-in that is viable and has normal proliferation and motility. Additionally, the authors show that labeled S. cerevisiae actin can incorporate into actin cables, which are formin dependent. IntAct constructs were shown to interact with several well-known actin binding proteins and localized well to many different actin structures. There was also interesting data obtained from tagging both beta and gamma actin in human cells. However, as an actin scientist eager for new probes to visualize actin in cells, there are still questions about the functionality of these probes. Addressing these issues, listed below, would alleviate the concerns I still have about IntActs after going through the manuscript. IntActs have the potential to have a large impact on cytoskeletal research if it can be rigorously documented that they are functionally as close to unlabeled actin as possible.

      We thank the Reviewer for their constructive comments and general positive evaluation of our study.

      *Reviewer #2 (Significance (Required)): *

      Concerns:

      1. There are no negative controls performed for either the fixed or live-cell imaging of IntAct. Since the fixed cell data is heavily reliant on the presence of flag-labeled puncta at actin filaments, it is important to show that the immunocytochemistry protocol doesn't produce anything that would mimic the localization of actin. For the live cell data, there has been no effort made to show that the binding of the nanobody to the ALFA tag on InAct is specific.

      Planned revision: __We will add the following controls to exclude that any of the labeling procedures produces anything that would mimic the localization of actin: 1) Immunofluorescence staining of the used tags (FLAG/ALFA) in cells that do not have tagged actins 2) Expression of ALFA-Nb-GFP and ALFA-Nb-mScarlet in cells that do not have tagged actins 3)__ Expression of free GFP in cells that have tagged actins. We will co-stain these cells with phalloidin to visualize F-actin and determine if any signal is specifically localized to the actin cytoskeleton.

      2. The homozygous ALFA-tagged IntAct cells have a 50% reduction in the amount of actin expression (Fig. 2D). What is the F:G ratio in these cells? The F:G measurement is only shown for the FLAG-tagged heterozygous IntAct cells, which have the worst co-localization with phalloidin (Fig. 2F) and were not used for subsequent figures. I appreciate that motility and proliferation were measured and shown to not be affected (Fig. 4D,E) , but in our lab reducing the amount of polymerized actin by 50% (which may be more in ALFA-tagged IntAct cells if the F:G changes) has catastrophic effects on other cytoskeletal and organelle systems. Since the homozygous ALFA IntAct cells are the main ones used in the manuscript, they should be the ones that are fully characterized.

      We would like to point out that the reduction is only 20-25 percent depending on the specific western blot analysis and the loading control. Still, the Reviewer is correct about the necessity of the F:G actin measurements of the ALFA-tagged IntAct cells and we therefore included those as Suppl. Fig. 9 in the original manuscript (text on page 9). The quantification of these assays clearly demonstrated that the F-G actin ratio in the ALFA-tagged IntAct cells is the same as in parental cells.

      3. It is not addressed if expressing the ALFA-Nb-GFP construct in ALFA-IntAct cells alter actin properties? This is essential information for live cell imaging experiments.

      Planned revision: We have already performed proliferation and migration experiments in cells that stably express the ALFA-Nb-GFP. These data indicated that proliferation and migration are not affected by the presence of the nanobody and these data will be included in the revised manuscript. To note, in the original manuscript, we already showed that treadmilling of actin at the lamellipodia is not affected by the presence of the ALFA-Nb-GFP.

      4. It is not addressed how much of the ALFA-IntAct gets labeled with ALFA-Nb-GFP and how uniform the labelling.

      We do not understand this specific request of the Reviewer. To our knowledge, it is not possible to assess how much of a probe (in this case the ALFA-Nb-GFP) binds the target (in this case the ALFA-IntAct actins) in living cells. This is not only the case for the ALFA-Nb-GFP but also for any other probe. As an example, when expressing Lifeact, we also do not know how much of the actin molecules within F-actin get labeled with Lifeact and how uniform the labeling is. From the results of the live-cell imaging we can only conclude that the binding is at least so effective that we can readily observe and discern all the actin-based structures that are also observed by Lifeact (see Suppl. Fig. 8 for Lifeact-GFP/ALFA-Nb-mScarlet cotransfection). Whether the regions that do not have F-actin only contain ALFA-Nb-GFP that is bound to actin monomers or also contains a significant fraction of free ALFA-Nb-GFP seems an issue that cannot be addressed.

      5. To assess lamellapodia architecture, "branched actin angle" is measured using AiryScan imaging of actin filaments. This type of microscopy does not offer the ability to image individual actin filaments; what is actually being measured is the orientation of actin bundles to each other. It should be impossible to image the orientation of actin filaments in Arp2/3 dendritic networks and it is surprising that the measurements average to 70 degrees. A suitable substitute for this would be to measure the size and amount of F-actin in phalloidin-stained lamellipodia using kymograph analysis.

      We apologize for this misapprehension from our side which is also noted by the other two reviewers. In the treadmilling videos of the lamellipodia in HT1080 cells, which were obtained using Airyscan super-resolution microscopy, we clearly observe a consistent filament formation at a constant angle, something which we interpreted as the angle between the mother filament and the daughter filament. After consulting the literature, we indeed have to admit that this cannot be interpreted as such and we will remove these datasets.

      Planned revision: We will remove the datasets with the angle measurements (Suppl. Fig. 7A-B) from our manuscript.

      6. Was it possible to make an IntAct gene substitution in yeast?

      Planned revision: We thank the reviewer for this interesting question and as also suggested by Reviewer 1, we are now constructing yeast strains with IntAct as the sole expressing actin copy by using the well-established plasmid shuffle system in yeast. The results of these experiments will determine the ability of IntAct to completely substitute actin function in yeast.

      Also, while this is not necessary for this manuscript, making a fission yeast strain where actin has been substituted with IntAct and demonstrating that IntAct gets incorporated into the cytoplasmic ring and into Cdc12p-polymerized filaments would alleviate MANY potential concerns people would have about these probes by directly assessing situations were other labeled actins have been documented to fail. Along the same lines, it would have been nice to see a comparison in some of the assays of ALFA-IntAct and GFP-actin or another labeled actin variant.

      We appreciate the reviewer for their constructive feedback and completely agree that it is important to document how IntAct behaves in scenarios where other labelled actins have failed. As a proof of principle, IntAct incorporates into both formin- and Arp2/3- made linear and branched actin filaments in yeast (Fig.5E, Suppl. Fig. 14) and this data shows that IntAct labelling strategy is the first to achieve good integration into both these structures as previous efforts with labelled actin such as GFP-Actin fail to incorporate into formin-made actin filaments (Doyle et al., PNAS, 1996). Thus, we believe that IntAct does perform better than other labelled actins in yeast, although, further optimizations are required to overcome limitations regarding incorporation into actin cables in the presence of the ALFA nanobody.

      Planned revision: We have already extended applicability of IntAct to another well-known fungal model system, the fission yeast Schizosaccharomyces pombe (S. pombe). We expressed IntAct variants of human β- and γ- actin, budding yeast actin (Sc-IntAct) and fission yeast actin (Sp-IntAct) from an exogenous plasmid under the native S. pombe actin promoter in an S. pombe strain that constitutively expresses the Nb-ALFA-mNG. Live-cell microscopy of S. pombe cells expressing these proteins revealed that all IntAct variants localize to actin patch-like structures located at the cell poles and cell division site (during cytokinesis). These structures show similar dynamics as reported for actin patches of S. pombe previously (Pelham et al., Nat Cell Biol, 2001). These preliminary results suggest that IntAct proteins show a similar localization pattern to only branched actin networks found in the actin patches of S. pombe like we had previously observed for the budding yeast, S. cerevisiae (Fig. S13 in manuscript). The underlying mechanism for this exclusion from linear actin cable network from both budding and fission yeast remain unknown and may represent an inherent specificity and sensitivity of yeast formins. Our current and future experiments will express IntAct variants in absence of the ALFA nanobody and determine the level of incorporation into actin cables, patches, and actomyosin ring.

      Planned revision: We have also already performed a quantitative analysis to ascertain the effect of Sc-IntAct expression of cortical actin patch dynamics which represent sites of endocytosis in yeast (Young et al., J Cell Biol, 2004; Winter et al., Curr Biol, 1997). We compared actin cortical patch lifetimes between wildtype cells and cells expressing Sc-Act1 or Sc-IntAct as an extra copy. We used Abp1-3xmcherry as a marker for actin patches and quantified the time window between the appearance and disappearance of a patch (actin patch lifetime) from time-lapse microscopy experiments. Our preliminary results indicate that actin patch lifetimes are unaffected by exogenous expression of both Sc-Act1 or Sc-IntAct suggesting that IntAct does not negatively influence or alter actin patch dynamics. These observations suggest its applicability as a direct visualization strategy for actin at the cortical patches in budding yeast alongside existing surrogate markers like Abp1, Arc15, etc (Goode et al., Genetics, 2015; Wirshing et al., J Cell Biol, 2023).

      __Reviewer #3 (Evidence, reproducibility and clarity (Required)): __

      *Summary: *

      This paper tackles a new strategy to tag actin in cells, by identifying that incorporation of a tag of moderate size in subdomain 4 of actin minimally affects actin dynamics in cells, and does not perturb its interaction with known partners, as observed in pull-down assays.

      *Major comments: *

      The paper is interesting and experiments are convincing.

      *My main concerns are the following : *

      - Varland et al, is reporting a phosphorylation on Thr229 : I think the authors should mention and discuss this potential PTM that could be affected in IntAct.

      We thank the Reviewer for pointing this out. We are aware of this review that includes phosphorylation on Thr229 as a possible PTM. Yet, this PTM is only reported in one of the Tables of the Review and not further discussed in the text. It is also unclear how the authors determined that Thr229 is a possible phosphorylation site except for the notion that this residue is a threonine and exposed at the surface of the actin molecule. Together with the fact that there is no evidence from primary studies that Thr229 is phosphorylated, we therefore decided to not include it in our discussion.

      - The sequence in subdomain 4 (the alpha helix containing T229A230) is extremely conserved in animals, as well as in between the 6 human actin isoforms. This usually indicates a strong selection pressure on the residues. I think the authors should discuss how surprising it is that the T229A230 position can accomodate various tags while it is probably the place of interaction with other proteins and is playing an important role in the mechanical structural integrity of the actin itself.

      We thank the Reviewer for bringing up this important point. To a certain extent, the conservation argument is true for all of the residues/domains in actin. Any manipulation will change a conserved part of the actin molecule in one way or another and thereby potentially modify its function. This is also evident from the fact that for most of the internally tagged actins, we observed a very poor colocalization with the actin cytoskeleton (Fig. 1). While for the T229/A230, we have not observed any major effects yet, this certainly does not mean that no further changes or defects will be uncovered in future experiments. Nonetheless, since our approach is unique with respect to the fact that it allows isoform-specific tagging without manipulating the N-terminus, our internal tagging system complements the already existing repertoire of actin reporting methods (N-terminal fusion, Lifeact, F-Tractin, actin nanobodies) and allows researchers to study so far unknown properties of actin variants. We have already included in the discussion that, at this point, we can only speculate as to why this variant performs much better than the others (Page 16 of the manuscript) and that possible explanations are the location at the inner domain and the higher structural plasticity of this region as compared to the rest of the molecule, as found during an alanine mutagenesis screen (Rommelaere et al., Structure, 2003).

      - It is now well established that actin plays active and important roles in the nucleus : is ALFA-actin correctly translocated to the nucleus ?

      Planned revision: This is an interesting suggestion. We will perform nuclear-cytosol fractionation experiments and determine whether ALFA-actin is still correctly translocated to the nucleus.

      *- OPTIONAL: one may regret that there is no classical in vitro assays, such as pyrene assays to assess some kinetcis parameters on epitope-tagged actins. I guess this would make the paper a bit too large. Although, it will prove useful to better understand how much formin activity is affected (see below) *

      For further biochemical characterization and a detailed investigation of the precise assembly kinetics of the tagged actins, we (KD, SP) are already working together to set up in vitro reconstitution experiments. Yet, as also indicated by the Reviewer, we consider these experiments outside of the scope of the current work.

      *Minor comments: *

      Below are points that could be addressed by the authors to improve the manuscript readability and highlight some important points that are sometimes missing or are not properly discussed:

      -line 40 "...but the distinct N-terminal epitope is not available under native conditions preventing" is a bit too obscure. Can the authors say clearly what is meant by 'native conditions'?

      In our understanding, the term ‘native’ is generally used when referring to conditions in which proteins are in their natural state, without alterations due to heat or denaturants, and possibly also still interacting with their binding partners. We will rephrase to better indicate that in this specific case, we mean that the region that harbors the N-terminus is usually occupied by actin-binding proteins, preventing the binding of the antibody due to steric hindrance.

      - figure 1A : make a clearer correspondance between the number shown in panel A and the amino acid numbers displayed in panel C and G.

      Planned revision: This is a good point, we will add extra annotation in the graph to better link the panels with each other. We will also add additional annotation in Fig. 1D-F for the same purpose.

      - figure 1A : it could be informative to indicate subdomains in this panel.

      Planned revision: We will add the numbers for the subdomains in Fig. 1A.

      - figure 1C : normalized correlation cell : I am not sure I understand how the normalization of the Pearson coefficient is done. It is therefore not clear how can it >1 or >-1 ? This should be clearly explained in the method section of the paper.

      __Planned revision: __We will better explain the normalization procedure in the Methods section.

      - figure S4 : comes a bit too early when ALFA-actin has not been yet introduced in the main text. Please, reposition this part or provide data with the FLAG-tag version.

      Planned revision: This is a good point and completely overlooked by us. We will introduce this Figure later such that the ALFA tag is already introduced.

      - section starting line 121 : this section should be better motivated = Why are different tags being tested ? This comes later in the discussion, but the reader fails at following the reasoning/motivation here.

      Planned revision: We will add extra motivation for why we added multiple tags.

      - figure 2D, line 145 "We also evaluated actin protein expression in the homozygous ALFA-β-actin cells and this showed that the total amount of β-actin was slightly lower in the ALFA-β-actin cells compared to parental HT1080 cells (Fig. 2C-D)." 'Slightly' is not a very quantitative nor accurate term. please rephrase. Besides, a statistical test for the paired data would also be informative. Besides, data in figure S6B-D indeed show a correlated increase in the expression of Gamma-actin that compensate for the decrease in the Beta-actin level in ALFA-Beta-actin. Can the authors explain why they conclude otherwise?

      Planned revision: This indeed is an important point and we will change the phrasing of this section to provide a more quantitative and accurate description of the western blot quantifications.

      - figure S7B: I am not ure anyone has ever reported measurement of angle of branched actin filament using epifluorescence microscopy. I would remove this panel, or the authors should explain how this measurement can be done objectively.

      We apologize for this misapprehension from our side which is also noted by the other two reviewers. In the treadmilling videos of the lamellipodia in HT1080 cells, which were obtained using Airyscan super-resolution microscopy, we clearly observe a consistent filament formation at a constant angle, something which we interpreted as the angle between the mother filament and the daughter filament. After consulting the literature, we indeed have to admit that this cannot be interpreted as such and we will remove these datasets.

      Planned revision: We will remove the datasets with the angle measurements (Suppl. Fig. 7A-B) from our manuscript.

      *- Figure 2F : can the authors comment on the (significant ?) lower value for FLAG-tag actin ? *

      The lower value for FLAG-tag actin has likely to do with the properties of the antibody and suitability for immunofluorescence. For reason that we do not know, we usually detect more background for the FLAG tag antibody as compared to the other antibodies/ALFA tag nanobody. Since the Pearson correlation coefficient quickly decreases with suboptimal labeling, this is likely the reason that the values for FLAG-actin are lower as compared to the other tagged actins. Importantly, in our biochemistry experiments (F/G-actin), we detect no difference between FLAG-actin and ALFA-actin indicating that it is rather the immunofluorescence and sensitive Pearson correlation analysis than the integration of actin that causes this difference.

      - line 205 "The results from these experiments show that both DIAPH1 and FMNL2 associate with ALFA-β-actin (Fig. 3D),". It is not so obvious that these formins directly interact with monomeric actin via their FH2 domains in co-immunoprecipitation assays. It might very well be mediated by the interaction with profilin, that in turn bind to the FH1 domain of formins. For me, this assay does not make a correct proof that epitope-labelled actin do not interfere with formin activity.

      Planned revision: The point that the co-immunoprecipitation does not demonstrate direct interactions between formins and actin is well taken. We, however, do not claim that this assay proofs that formin activity, or formin-based integration of actin monomers, is similar with tagged actin as compared to wildtype actin. Nonetheless, we will critically re-evaluate the relevant passages and rephrase the text to avoid any confusion.

      - figure 5C&D : both graph should use the same scale for the y-axis for easier comparison.

      Planned revision: We will adapt the scale of Fig. 5D to make it identical to Fig. 5C. Following the other suggestions of the Reviewer (and of Reviewer #1), we will also critically evaluate our normalization procedure and present those numbers in the Figures if the values turn out to be different.

      - figure 5D: I think the way the ratio is performed is misleading. Why not look at the Beta/Gamma ratio using the isoform specific antibodies used in parental cells, and show the results for ALFA-Beta-actin and for ALFA-Gamma-actin separately ?

      We kindly refer to our answer to Reviewer #1 on Page 2 for a detailed explanation on the experimental challenge of comparing the localization of wildtype and tagged actin isoforms.

      Planned revision: We will critically evaluate our normalization procedure and present those numbers in the Figures if the values turn out to be different. Furthermore, we will add a different experimental method to show that the tagged isoforms properly localize to actin-based structures. For this, we will attempt to use micropatterned cells to induce clearly define actin-bases structures and also explore the possibilities of investigating the differential localization in double-tagged cells.

      *- The limitation observed for unbranched cables in yeast that nanobody-tagged ALFA-actin does not incorporate correctly should be discussed and stressed further in the discussion, as it might prove to be a strong limitation for live-cell imaging to reliably study any type of actin networks. *

      We acknowledge the reviewer’s concern regarding the inability of ALFA-tagged actin to incorporate into yeast actin cables when NbALFA is co-expressed and will discuss this point further in the revised manuscript. We have now observed the same limitation for fission yeast actin cables as well and combined, these observations may represent a tighter control and sensitivity of yeast formins towards any perturbations in actin size (since NbALFA binds to ALFA tag with picomolar affinity). To address this issue and as also suggested by Reviewer 1, we are now creating yeast strains with inducible control of NbALFA expression under GALS/GAL1 promoters and observe the labelling of actin structures after this approach. Additionally, expression of variants of NbALFA with high dissociation rates may also allow labelling of actin cables and would be certainly worth a try in the future. A structural comparison between mammalian and yeast formins may be required to shed some light on the molecular basis of this fundamental difference.

      However, since in the absence of the nanobody, this limitation is overcome (Fig. 5E, Suppl. Fig. 14), we believe that with additional modifications and fast developments in imaging technologies, this limitation can be overcome in the future. Thus, IntAct as a labeling strategy represents an advancement over existing labelled actins with the most important aspect being the identification of the T229/A230 residue pair to be permissive for integration of various tags even as large as GFP11 fragment including a linker (26AA) (Reviewer Fig. 2). Importantly, the T229/A230 site is conserved across many organisms (such as Chlamydomonas reinhardatii, Cryptococcus neoformans, etc) and may act as a framework to study the actin cytoskeleton especially in organisms where known surrogate markers like phalloidin and Lifeact may not work or work only sub optimally.

      *Reviewer #3 (Significance (Required)): *

      *General assessment: *

      *This paper provides a new tagging strategy to monitor actin activity in cells, by specifically inserting the tag along the amino acid sequence. *

      *Advance: *

      *This is a very useful tool, as most existing available probes bind to actin in regions that are common to many other actin binding proteins. The authors provide extensive experiments to validate that tagged-actin are functional and do not perturb the actin expression level, actin network architecture nor dynamics. *

      *Audience: *

      *This research paper will be of interest to a rather broad audience (many cell biologists) that are either sutyding actin dynamics or know that actin is involved in the cell functions they study. *

      *Expertise: *

      *My expertise is in vitro actin biochemistry. *

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary:

      This paper tackles a new strategy to tag actin in cells, by identifying that incorporation of a tag of moderate size in subdomain 4 of actin minimally affects actin dynamics in cells, and does not perturb its interaction with known partners, as observed in pull-down assays.

      Major comments:

      The paper is interesting and experiments are convincing.

      My main concerns are the following :

      • Varland et al, is reporting a phosphorylation on Thr229 : I think the authors should mention and discuss this potential PTM that could be affected in IntAct.
      • The sequence in subdomain 4 (the alpha helix containing T229A230) is extremely conserved in animals, as well as in between the 6 human actin isoforms. This usually indicates a strong selection pressure on the residues. I think the authors should discuss how surprising it is that the T229A230 position can accomodate various tags while it is probably the place of interaction with other proteins and is playing an important role in the mechanical structural integrity of the actin itself.
      • It is now well established that actin plays active and important roles in the nucleus : is ALFA-actin correctly translocated to the nucleus ?
      • OPTIONAL: one may regret that there is no classical in vitro assays, such as pyrene assays to assess some kinetcis parameters on epitope-tagged actins. I guess this would make the paper a bit too large. Although, it will prove useful to better understand how much formin activity is affected (see below)

      Minor comments:

      Below are points that could be addressed by the authors to improve the manuscript readability and highlight some important points that are sometimes missing or are not properly discussed :

      • line 40 "...but the distinct N-terminal epitope is not available under native conditions preventing" is a bit too obscure. Can the authors say clearly what is meant by 'native conditions' ?
      • figure 1A : make a clearer correspondance between the number shown in panel A and the amino acid numbers displayed in panel C and G.
      • figure 1A : it could be informative to indicate subdomains in this panel.
      • figure 1C : normalized correlation cell : I am not sure I understand how the normalization of the Pearson coefficient is done. It is therefore not clear how can it >1 or >-1 ? This should be clearly explained in the method section of the paper.
      • figure S4 : comes a bit too early when ALFA-actin has not been yet introduced in the main text. Please, reposition this part or provide data with the FLAG-tag version.
      • section starting line 121 : this section should be better motivated = Why are different tags being tested ? This comes later in the discussion, but the reader fails at following the reasoning/motivation here.
      • figure 2D, line 145 "We also evaluated actin protein expression in the homozygous ALFA-β-actin cells and this showed that the total amount of β-actin was slightly lower in the ALFA-β-actin cells compared to parental HT1080 cells (Fig. 2C-D)." 'Slightly' is not a very quantitative nor accurate term. please rephrase. Besides, a statistical test for the paired data would also be informative. Besides, data in figure S6B-D indeed show a correlated increase in the expression of Gamma-actin that compensate for the decrease in the Beta-actin level in ALFA-Beta-actin. Can the authors explain why they conclude otherwise ?
      • figure S7B: I am not ure anyone has ever reported measurement of angle of branched actin filament using epifluorescence microscopy. I would remove this panel, or the authors should explain how this measurement can be done objectively.
      • Figure 2F : can the authors comment on the (significant ?) lower value for FLAG-tag actin ?
      • line 205 "The results from these experimentsshow that both DIAPH1 and FMNL2 associate with ALFA-β-actin (Fig. 3D),". It is not so obvious that these formins directly interact with monomeric actin via their FH2 domains in co-immunoprecipitation assays. It might very well be mediated by the interaction with profilin, that in turn bind to the FH1 domain of formins. For me, this assay does not make a correct proof that epitope-labelled actin do not interfere with formin activity.
      • figure 5C&D : both graph should use the same scale for the y-axis for easier comparison.
      • figure 5D: I think the way the ratio is performed is misleading. Why not look at the Beta/Gamma ratio using the isoform specific antibodies used in parental cells, and show the results for ALFA-Beta-actin and for ALFA-Gamma-actin separately ?
      • The limitation observed for unbranched cables in yeast that nanobody-tagged ALFA-actin does not incorporate correctly should be discussed and stressed further in the discussion, as it might prove to be a strong limitation for live-cell imaging to reliably study any type of actin networks.

      Significance

      General assessment:

      This paper provides a new tagging strategy to monitor actin activity in cells, by specifically inserting the tag along the amino acid sequence.

      Advance:

      This is a very useful tool, as most existing available probes bind to actin in regions that are common to many other actin binding proteins. The authors provide extensive experiments to validate that tagged-actin are functional and do not perturb the actin expression level, actin network architecture nor dynamics.

      Audience:

      This research paper will be of interest to a rather broad audience (many cell biologists) that are either sutyding actin dynamics or know that actin is involved in the cell functions they study.

      Expertise:

      My expertise is in vitro actin biochemistry.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Actin is highly sensitive to modifications, and tagging it with fluorescent proteins or even smaller motifs can affect its function. The most well-known example of this is that fission yeast where actin has been replaced with GFP-actin are inviable (Wu and Pollard, Science 2005) because the labeled actin cannot incorporate into the formin-dependent filaments that make up the cytokinetic ring. Subsequent experiments revealed that formins filter out GFP-actin monomers, as well as monomers that are labeled with smaller fluorescent motifs (Chen et al, J. Structural Biology 2012). Further, attempts to make mammalian cells lines where GFP-beta-actin was knocked into one allele resulted in extreme down-regulation of the GFP-labeled actin, indicating that there is some implicit toxicity with the labeled version. To my knowledge, all attempts at making homozygous GFP-actin knock-ins have been unsuccessful. Therefore, while GFP-actin or other labeled variants can be over-expressed in many different cell types with some success, there is always the question of how faithful the labeled actin represents bona fide actin localization and dynamics.

      To address this van Zwam et al. have developed a clever strategy of screening actin for internal motifs that can tolerate incorporation of a tag without affecting its function. They appear to have found a good candidate, named IntAct, and provide evidence that this tagging position allows the actin to be functional in both human and yeast cells. The work is very promising, and many of the assays performed satisfy the criteria of rigor and reproducibility. Importantly, the authors have created knock-in human cell lines where the tagged actin is expressed at normal levels, including a double allele knock-in that is viable and has normal proliferation and motility. Additionally, the authors show that labeled S. cerevisiae actin can incorporate into actin cables, which are formin dependent. IntAct constructs were shown to interact with several well-known actin binding proteins and localized well to many different actin structures. There was also interesting data obtained from tagging both beta and gamma actin in human cells. However, as an actin scientist eager for new probes to visualize actin in cells, there are still questions about the functionality of these probes. Addressing these issues, listed below, would alleviate the concerns I still have about IntActs after going through the manuscript. IntActs have the potential to have a large impact on cytoskeletal research if it can be rigorously documented that they are functionally as close to unlabeled actin as possible.

      Significance

      Concerns:

      1. There are no negative controls performed for either the fixed or live-cell imaging of IntAct. Since the fixed cell data is heavily reliant on the presence of flag-labeled puncta at actin filaments, it is important to show that the immunocytochemistry protocol doesn't produce anything that would mimic the localization of actin. For the live cell data, there has been no effort made to show that the binding of the nanobody to the ALFA tag on InAct is specific.
      2. The homozygous ALFA-tagged IntAct cells have a 50% reduction in the amount of actin expression (Fig. 2D). What is the F:G ratio in these cells? The F:G measurement is only shown for the FLAG-tagged heterozygous IntAct cells, which have the worst co-localization with phalloidin (Fig. 2F) and were not used for subsequent figures. I appreciate that motility and proliferation were measured and shown to not be affected (Fig. 4D,E) , but in our lab reducing the amount of polymerized actin by 50% (which may be more in ALFA-tagged IntAct cells if the F:G changes) has catastrophic effects on other cytoskeletal and organelle systems. Since the homozygous ALFA IntAct cells are the main ones used in the manuscript, they should be the ones that are fully characterized.
      3. It is not addressed if expressing the ALFA-Nb-GFP construct in ALFA-IntAct cells alter actin properties? This is essential information for live cell imaging experiments.
      4. It is not addressed how much of the ALFA-IntAct gets labeled with ALFA-Nb-GFP and how uniform the labelling.
      5. To assess lamellapodia architecture, "branched actin angle" is measured using AiryScan imaging of actin filaments. This type of microscopy does not offer the ability to image individual actin filaments; what is actually being measured is the orientation of actin bundles to each other. It should be impossible to image the orientation of actin filaments in Arp2/3 dendritic networks and it is surprising that the measurements average to 70 degrees. A suitable substitute for this would be to measure the size and amount of F-actin in phalloidin-stained lamellipodia using kymograph analysis.
      6. Was it possible to make an IntAct gene substitution in yeast?

      Also, while this is not necessary for this manuscript, making a fission yeast strain where actin has been substituted with IntAct and demonstrating that IntAct gets incorporated into the cytoplasmic ring and into Cdc12p-polymerized filaments would alleviate MANY potential concerns people would have about these probes by directly assessing situations were other labeled actins have been documented to fail. Along the same lines, it would have been nice to see a comparison in some of the assays of ALFA-IntAct and GFP-actin or another labeled actin variant.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      In this study, the authors generate several variants of actin that are internally tagged with short peptide tags. They identify one particular position that is able to tolerate various tags of 5-10 amino acids and still shows largely unaltered behavior in cells. They study incorporation of their tagged actins into filaments, characterize the interactions of G-actin variants with different associated proteins and show that retrograde actin flow in lamellipodia and the wound healing response of epithelial cells is not affected by the tagged variants. They then apply the tagged actin to study subcellular distribution of different actin isoforms in mammalian and yeast cells.

      The identification of a specific site in the actin protein that tolerates variable peptide insertions is very exciting and of fundamental interest for all research fields that deal with cytoskeletal rearrangements and cellular morphogenesis. The result demonstrating the functionality of actin variants with peptides inserted between aa 229 and 230 are generally convincing and well done. In particular, the generation of CRISPR/Cas9 genome edited versions of beta- and gamma actin are impressive. I therefore generally support publication of this study. There are however several technical and conceptual issues that should be addressed to improve quality and scope of the study. I listed some specific comments below:

      Major points

      • The biggest issue I have is the last section on the application of tagged actins to study isoform functions. In principle the application is very clear as there are simply no alternative ways to study isoform distribution in live cells. However, the experimental data are simply not convincing. What the authors define as "cortex" in Fig. 5A seems to rather represent cytosolic background mixed with radial fibers. I am not convinced that even the antibody staining with a relatively clear differential distribution of beta and gamma really shows a genuine accumulation of one isoform on stress fibers. It seems to me that the beta-actin staining has as higher cytosolic background and is generally weaker (gamma nicely labels transverse arcs), which reduces signal/noise and therefore yields a relatively increased level in areas with less-bundled actin. My suggestion is to select more clearly defined actin structures and to use micro-patterned cells to normalize the otherwise obstructing variability in actin organization. Possible structures would be cortical arcs in bow-shaped cells, lamellipodial edges (HT1080 seem to make very nice and large lamellipodia) or cell-cell contacts (confluent monolayer, provided cells don´t grow on top of each other). Stress fibers are possible but need to be segmented very precisely and I did not see any details on this in the methods section. For Fig. 5D: I assume cells were used where only one isoform was tagged? This is technical weak and the double-normalization is probably blurring any difference that might be occurring. Why not use a double-tagging strategy with ALFA/FLAG or ALFA/AU5 tags to exploit the constructs introduced in the previous figures? Also, the unique selling point of the strategy is the possibility of actual live imaging of specific isoforms. Cells that have stably integrated double tags and then transiently express nanobodies for ALFA and either AU5 or FLAG (or other if those don't exist) would make this possible. Considering the work already done in this manuscript, such an approach should actually be possible - did the authors attempt this or is there are reason it is not discussed? If double tagged cells are not possible for some reason it should at the very least be possible to combine ALFA-detection with the specific antibody against the other isoform and get rid of the double normalization.
      • The authors make a point of comparing the internally tagged actin to N-terminal tags that are mostly functional but have been shown to affect translational efficiency. I would strongly suggest to include N-terminally tagged actin as control for all assays in this study. Also for the physiological assays (retrograde flow, wound healing), a positive control is missing that shows some effect. Previous studies showed defects with transiently expressed actin with an N-terminal GFP. As retrograde flow measurements are very sensitive to the exact position of the kymographs and wound healing assays is a very crude and indirect readout, such a positive control is essential.
      • Expression of tagged actins in yeast is a very nice idea but it would be far more informative to express the tagged forms as the only copy of actin. This can either be done by directly replacing endogenous actin gene in S. cerevisiae, or (if the tagged versions are not viable) - using the established plasmid shuffle system (express actin on counter-selectable plasmid, then knock out endogenous copy and introduce additional plasmid with tagged actin, then force original plasmid out). In the presence of endogenous S. cerevisiae actin the shown effects are very hard to interpret as nothing is known about relative protein levels (endogenous vs. introduced). Also, if constitutive expression of the ALFA nanobody is harmful for integration into cables, why not perform inducible expression of the nanobody and observe labeling after induction. For the live imaging a robust cable marker is needed, like Abp140-GFP. Finally, indicate the sequence differences between the used actin forms in yeast (supplementary figure with sequence alignment and clear indication of all variations)
      • As the authors clearly show good integration of several tagged actins into filaments I would expand the structural characterization: perform alpha fold predictions of actin monomer structures including the various tags to show the expected orientation. It is striking that the only integration site that seems to work well is at the last position of a short helix, indicating that the orientation of the integrated peptide might be fixed in space and be optimal to minimize interference. Also, a docking of the tag onto the recently published cryoEM structures of the actin filament should be shown to indicate where it resides compared to tropomyosin or the major groove where most side binding proteins seem to bind.
      • For any claims regarding usability of tagged variants for isoform research it would be very important to characterize the known posttranslational modifications of tagged actin variants - are the differences between beta and gamma maintained on this level as well?

      Technical issues

      • There is no scale for the color coding in Fig. 5A, B
      • The y-scales for Fig. 5C and D need to be identical to allow direct comparison
      • Pearson coefficient should not be normalized to a control value as its already a dimensionless parameter. Always report actual R-value - also remove R2 values for Pearson as this makes no sense in this context (not sure if it was a typo or intended).
      • All values on subcellular regions (like stress fiber or cortex) dependet critically on the way thesese regions were thresholded or identified. Provide all details on how this was done in the methods section and ensure that adequate background subtraction and normalization is applied. Optimally, an unbiased (AI or automated) approach based on simple image statistics is used for this to avoid personal bias.
      • In Fig. 2A only heterozygous FLAG-actin cells are used. Why not use a homozygous line (for both beta and gamma actin)? The nice band shift of the FLAG version would allow the precise quantification of the fraction of total actin covered by beta and gamma actin, which then could provide some additional info for the apparently weaker beta staining in Fig. 5 (if beta expression is simply weaker). This would be a very simple and useful advantage of the internal tags that could be widely applied.
      • Fig. 3: control with N-terminal tag is missing. Also, why is it not possible to assay filament binding factors like Myosin, Filamin or alpha actinin - instead of co-IP a simple co-sedimentation assay with cell extracts in F-buffer should pick up any major difference in decoration of filaments containing the ALFA tag. Using two speeds for centrifugation it might even be possible to observe effects on filament bundling. The best approach for this would of course be to purify tagged actins and perform in vitro assays but this is clearly beyond the scope of what the authors intended here. I personally think that a broad acceptance of the marker will only come once the biochemistry has been sufficiently characterized so this is a future direction I would strongly encourage.
      • Fig. 2A has no loading control -
      • The RPE-1 data are confusing as several constructs show very different localization (completely cytosolic) to HT1080 cells and there is no possible explanation given for this. Maybe simply remove this data set?
      • The angel measurements for lamellipodial actin is not very meaningful: the angel is determined for the radial bundles, which do not correspond to the Arp2/3 angel of single filaments and is likely the results of different nucleation factors, I would suggest to remove this. If angel measurement are really intended, cryoEM needs to be performed.
      • Replace all SEM with SD values - use at least 3 biological replicates (4D SEM of n=2)

      Minor points

      • Intro: after listing all the details already understood on actin isoforms it is not very convincing to simply state the molecular principles remain largely unclear (l 34) - maybe better "there is no way to study actin dynamics due to current limitations of specific antibodies to fixed samples. Interesting option would be actually to develop nanobodies that are isoform specific 
      • L 71: "involved" in the kinetics is not a good term - maybe affects or regulates....
      • L148: "suspect" instead of "expect" - this clonal variation is actually a big danger of the employed approach as possible defects in actin organization could be masked by compensatory changes - it would generally be good to show critical data for at least 3 independent clones to rule out dominant selection effects.

      Referees cross-commenting

      I completely agree with the comments by reviewer 2 on the various missing controls - adding several or all of those will make the results much more convincing. The key for the adaptation of any new actin probe will be the level of confidence researchers have on the doumented effects. Even some negative effects on actin behavior (I am sure there will be some) should not prevent usage of the strategy as long as there is robust and convincing documentation of those effects. I also agree that including some basic in vitro characterization will go a long way to convince people dierectly working on actin (there is a very high level of biochemical understanding in that field).

      Significance

      Significance: Very useful finding that can be applied to any question related to actin-dependent cellular processes (morphogenesis, cell division, cell polarization, cell migration etc.)

      Strength: main finding convincing, strong genome edited cell lines

      Limitations: application to study of isoforms very limited and data not convincing, statistics and image quantifications need improvement

      Advance: identify new location for integral tagging of actin, which was not really possible before. The main relevance is for fundamental cell biology but the approach can also be applied to the study of disease variants in actin.

      Audience: general cell biology - very broad interest

  2. Aug 2023
    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      We thank the reviewers for their time, the positive reviews and the useful comments. We answer below and explain the changes made to the manuscript. The comments of the reviewers are in italics.

      Reviewer #1

      1. 'For GWAS, the strains that were fertile after 20 generations were considered non-Mrt.' One aspect of Fig 1D that could be clarified are the dots at generation 21. If these represent strains that were always fertile at generation 21, then perhaps give these a different color to indicate that sterility was never observed?

      Response: This is a good idea. We added colors in Figure 1, which makes it clearer.

      We also provide a different color for surviving replicates in all relevant figures.

      1. 'The mean Mrt values of strains ranged from sterile at 3 generations to fertile after 20 generations at 25°C, with a skewed distribution toward high values (Figure 1B).' Based on Table S2, part of the explanation for this skewed distribution in later generations is that some strains became sterile rapidly for some blocks, whereas the same strain did not become sterile in other blocks. For example, JU1200, JU360, PB303. I suggest providing a second color for Fig. 1D for strains that sometimes displayed sterility and sometimes did not.

      __Response: __We now colored the isolates that never became sterile, with the same color code as in panel B. Because we stopped the scoring at G20 and code fertility at G20 as '21', those with a mean below 21 show some sterility in at least one case.

      Because the number of generations at which we stopped the phenotyping (20) is arbitrary, the fact a line stayed fertile at 20 generations in one replicate is not very meaningful, especially considering that the number of replicates is not the same for all strains. The key point of the variance graph is to show that the strains with the most variance are those with high but

      For those that were sometimes fertile and sometimes sterile, I suggest creating a graph in Figure 1 that shows generations at sterility or lack of sterility, color coded by block. This will allow the significance of strains with high generation Mrt values to be better appreciated for readers who do not look at the supplementary table.

      __Response: __Yes, we added this graph in Figure S1. This is indeed useful.

      1. The GWAS section could benefit from a simple explanation of the premise of GWAS for non-specialist readers.

      __Response: __Yes, we added: "A genome-wide association study (GWAS) is a genetic mapping that uses the natural diversity of a panel of organisms of a given species to test for statistical independence between the allelic state of polymorphic markers and the phenotype of interest (Andersen and Rockman 2022). A statistical association between the marker and the phenotype indicates that a polymorphism tightly linked to the marker in the data (i.e. in linkage disequilibrium with it) causes the variation in phenotype. For statistical reasons, GWAS can only detect polymorphisms that are at intermediate frequencies in the panel, i.e. cases where both alleles occur at frequencies higher than 5%. We only used such polymorphisms in the GWAS (see Methods)."

      And further down:

      "To diminish the multiple testing burden, the initial analysis in Figure 1E used a restricted set of markers, after pruning those that were in high linkage to each other."

      1. One problem might be that the Mrt phenotype is widespread among wild strains. To the authors' credit, they consider results observed in different laboratories as valid, even when the results do not agree. If the Mrt phenotype is influenced by the environment, then some laboratory environments might result in 'false negative' Mrt results that could be ignored in favor of positive results from another lab that appear strong. Might focusing on strains with a set of strong positive results from one lab allow the authors to draw stronger GWAS conclusions?

      2. The authors' perform GWAS based on the variance of the Mrt phenotype data. Would the GWAS data be more illuminating if the authors only considered strains that become sterile fairly rapidly, within 10 generations. The authors might then have a second category that included strains that become sterile from generation 11-20. If the genetic basis for the Mrt phenotypes is the same, then GWAS of strains that become sterile in less than 10 generations might yield similar peaks as GWAS for strains that become sterile between generations 11-20.

      __Response: __These two comments are strongly related so we answer them together. Note that the GWAS is not mapping the variance values but the Mrt values themselves.

      We actually initially only used block 1 (a single replicate, all strains performed in parallel in our laboratory) and also detected the chromosome III association using a categorical variable (threshold at 11), but decided to show the results with all data to maximize power, taking into account the generation value and block effects.

      We investigated other ways to code the data (e.g. categorically) and removing the strains of the most variable middle category, as proposed by the reviewer. This changed the p values and the rank of the markers on chromosome III but not the overall result.

      In summary, we did a variety of tests, which pointed to chromosome III, a region that was validated using crosses (Figure 2).

      Note that in the revision, we updated the GWAS plot and fine mapping table as we noticed a few problems in our previous mapping. 1) We removed 3 isolates that were classified in Lee et al. 2021 as divergent. 2) We included strains that had been lost in the pipeline because their names did not match CeNDR isotypes. This increased the significance of the chromosome III peak.

      __Response: __There was no comment 6.

      1. 'We did not investigate whether a second locus present in JU775 on the right arm of Chr III might have a lesser effect.'

      __Response: __We are not sure what the reviewer meant. Considering the difficulties with the stronger effect locus, we did not try to study loci with a weaker effect.

      1. It might be interesting to test the memory of growth on beneficial bacteria on JU4134, which had a Mrt phenotype that was strongly suppressed by the beneficial bacteria.

      __Response: __We agree that testing other strains would be useful but given the duration of such experiments (30 generations and two weeks of preparation before), we respectfully decline to perform this experiment that does not seem strictly necessary.

      1. The Mrt phenotype of mutants in small RNA inheritance and histone modifying enzymes 'appears however distinct from that of the prg-1/piwi mutant (for which the cause of sterility is debated), especially the latter does not show temperature dependence and is suppressed by starvation.' While it is true that the cause of sterility is debated for the prg-1/piwi mutant, this mutant is defective for small RNA silencing and likely has parallels with some defects in histone modifying enzymes. Anecdotal reports suggest that starvation might affect the Mrt phenotype or longevity of histone modifying enzyme mutants. Moreover, the cause of sterility is not clear for small RNA inheritance and histone modifying enzyme mutants. It is fair to say that the distinction between temperature-sensitivity or lack of temperature sensitivity of small RNA mutants is not understood. Could the authors please comment here about whether any of the wild strains display sterility at 20°C.

      __Response: __The temperature-dependence of the wild isolates is progressive between 20-25°C. We previously showed that strains with a very strong Mrt phenotype, such as QX1211, can display sterility at 20°C (Figure 1B in Frézal et al. 2018). However, its Mrt phenotype is still temperature-dependent as the sterility occurs much earlier at 25°C.

      1. If intracellular bacteria are simply somatic, then how is it that they are transmitted to progeny. If they are released into the environment and then consumed by hatched larvae, this is soma-to-soma transmission.

      __Response: __These microsporidia (which are eukaryotes related to fungi) are indeed transmitted horizontally. To make this clear, we added: "colonizing its intestinal cells and being transmitted horizontally via defecation and ingestion of spores". The soma-to-germline interaction concerns the effect of microsporidia on germline maintenance.

      Minor: 1. 'We measured the mortal germline (Mrt) phenotype'. Mortal Germline (Mrt)

      __Response: __It is unclear as to whether phenotypes start with a capital letter when they are in full words. We did write phenotypes in previous works with a capital letter but have changed because C. elegans nomenclature rules (https://cgc.umn.edu/nomenclature) suggest that they should not: "Phenotypic characteristics can be described in words, e.g., dumpy animals or uncoordinated animals." For the mortal germline phenotype in particular, we find several ways to write it in articles (with 0, 1 or 2 capital letters, including the three reviewers). We are happy to change it if required.

      Reviewer #2

      Major comments: The authors claimed that the variants causing Mrt exist at intermediate frequency in the natural population but the evidence supporting this claim is rather limited.

      __Response: __Thank you for this comment as it helped us clarify the manuscript.

      To better explain the notion of intermediate frequency in the GWAS, we added an explanation of the principle of the GWAS (see above) and again in the Discussion: "The intermediate frequency of the candidate alleles derives from the GWAS approach, which cannot detect rare alleles, such as set-24, that are present in a single strain of the dataset."

      We also illustrated the frequency by adding a plot (Fig. 1F) showing the association of the most associated candidate SNP, with a visual depiction of the frequency. We further added in Results: "For SNPs with a high significance (p-4) in the fine mapping, the frequency of the Mrt associated allele was comprised between 21 and 41% in our GWAS strain set (Table S3); as an example, the Mrt allele of the associated SNP shown in Figure 1F (III:4677491) displayed a frequency of 29% in the restricted strain set. Over the global wild strain set with genotypes at CeNDR in 2020, these numbers are 17-58% and 39%, respectively. "

      To strengthen the claim, the authors should examine the distribution and frequency (perhaps coupled with phylogenetic analysis) of the Ch III haplotype in the wild isolates. The authors should also examine the GWAS peak for the signature of balancing selection (e.g., dN/dS ratio).

      __Response: __Thank you for this comment. The different associated SNPs in Table S3 differ in their allele frequency (Table S3), hence they belong to different haplotypes. We added a supplementary Figure S2 with an analysis of the haplotype structure. Those at a low frequency (around 20%) belong to the same haplotype (e.g. JU775 and MY10) but some associated alleles are present in more haplotypes (40-50%), such as JU1793. Even if we neglect recombination, the history of mutations in the region is complex and there is not a single associated haplotype. We now show the genotypes of these different haplotypes at all SNPs in Table S3. We also added Table S4 that shows the co-occurrence of relevant haplotypes in local populations.

      Concerning tests of balancing selection, without knowing the causal polymorphism and linked haplotype, this is far reaching. We only feel confident to say that the causal polymorphism(s) is present at a significant frequency. We added however the fact that irrespective of which polymorphisms are causal, both alleles were found to coexist locally.

      Results: relevant text was added at the end of the GWAS section.

      Discussion: "The co-occurrence of relevant chromosome III haplotypes on multiple continents and in local populations (Table S4) is suggestive of balancing selection; however, a linked locus other than that causing the Mrt phenotype may be involved."

      Does JU775 carry polymorphisms in genes that are known to be involved in Mrt? These genes may genetically interact with the Ch III variant, as suggested by the partial penetrant phenotypes of the introgressed lines. It would be helpful to have a table summarize the variation in these genes.

      __Response: __It is difficult to deduce much from a genomic variant analysis, so we refrain from showing tables of polymorphisms beyond that used for the fine GWAS mapping in Table S3. For example, a non-synonymous SNP may or may not alter protein activity and cis-regulatory elements are difficult to assess. Moreover, an obviously null allele may be compensated by another polymorphism in the background. The JU775 alleles and bam files are publically available from CeNDR (Erik Andersen's lab): https://caendr.org/data/data-release/c-elegans/latest

      It is curious to me that for experiments with HT115, the expression of the RNAi vectors was induced with IPTG. Is this step necessary? It is known that even the backbone of L4440 could trigger a non-specific RNAi response (PMID: 30838421). I wonder if activating exogenous RNAi response is required for Mrt rescue.

      __Response: __Indeed: this experiment was initially aimed at testing RNAi sensitivity of JU775, thus IPTG was added on the plate (Figure 7, panel B). We therefore repeated the memory experiment with OP50 and without IPTG, with a similar result (Figure 7, panel A).

      In figure 7, it appears that the worms transferred from MG1655/HT115 to OP50 showed an even stronger rescue (higher Mrt value) than the ones constantly on MG1655/HT115. This suggests to me that fluctuations in food composition may strongly affect epigenetic inheritance. Please clarify as this is very interesting, if true.

      __Response: __Note: This answers the comment above (IPTG is not required).

      We indeed noticed this strong rescue but do not wish to make a point as we did no attempt to reproduce this result in the exact same conditions. The experiment in panel B does not show this effect.

      Optional - Numerous studies have shown that SKN-1 regulates metabolism in response to food composition and availability (PMID: 23040073). Additionally, some recent studies have indicated a role of SKN-1 in epigenetic inheritance triggered by exogenous RNAi. In particular, SKN-1 promotes stress-induced epigenetic resetting (PMID: 33729152). I wonder if SKN-1 modulates Mrt based on bacterial diet.

      __Response: __We tested skn-1b/c hypomorphic and gain-of-function mutants in the N2 background on E. coli OP50 and did not see an effect of the skn-1 allele.

      Minor comments Line 47: typo "...they defined..."

      __Response: __We did mean "thus defined".

      Line 100-101: weird sentence structure. Please consider rephrasing.

      __Response: __We simplified to "a wild C. elegans strain can keep the memory of its culture on a suppressing bacterial strain."

      Line 138-139: I don't quite understand what "intermediate-frequency chromosome III alleles" means here. Some SNPs were found in Ch III 4-6Mb? Please expand.

      __Response: __We rephrased to: "because this isolate carries the chromosome III alleles associated in the GWAS analysis with the Mrt phenotype (Table S3)."

      Line 213 - it was unclear to me why the assay was performed at 23C instead of 25C. I later learned in the method section that microsporidia cannot be cultured at 25C. I think it will be helpful to add that information when microsporidia is introduced to improve clarity.

      __Response: __We added: " We used a temperature of 23°C because these microsporidia kill C. elegans too rapidly at 25°C."

      Reviewer #3.

      Minor points 1. Could the authors please define "experimental blocks"

      __Response: __We added the following sentence in Results: "Each Mrt assay started at a certain date constitutes an experimental block."

      1. Legend to supplementary snp table should be completed: define AF, impact, modifier, moderate, AA1, AA2...

      __Response: __This is added in the first sheet of the table. We also simplified the table and removed some of these columns.

      1. Please define "intermediate-frequency allele"

      __Response: __We added in Results: "GWAS can only detect polymorphisms that are at intermediate frequencies in the panel, i.e. cases where both alleles occur at frequencies higher than 5%." We also added below: " "For SNPs with a high significance (p-4) in the fine mapping, the frequency of the Mrt associated allele was comprised between 21 and 41% in our GWAS strain set (Table S3); as an example, the Mrt allele of the associated SNP shown in Figure 1F (III:4677491) displayed a frequency of 29% in the restricted strain set."

      1. Figure 7 legend: Authors should be more specific in describing the figure: After 10 (A panel), 13 or 20 generations (B panel) on the K-12 strain... What is E. coli OP50 start 'G10'? the 15° stock?

      __Response: __We changed to: " After 10 (A panel), 13 or 20 generations (B panel) on the K-12 strain" and added some details in:

      "A control from a 15°C culture maintained without starvation ("15°C stock") was bleached in parallel (labeled "E. coli OP50 start "G10" " in the graph of panel A)."

      Optional: Did the authors attempt to rescue the Mrt phenotype with individual metabolites (eg Vit B12...)? These are not straight forward experiments and most likely part of a future study.

      __Response: __We indeed tested several metabolites that are known to differ in C. elegans raised on E. coli OP50 versus K-12 strains for their effect on the Mrt phenotype. None was able to rescue the mortal germline phenotype. However, especially in these long multigenerational experiments, it is difficult to know whether the metabolites are stable. We monitored vitamin B12 activity by using an acdh-1::GFP reporter that is known to be repressed by vitamin B12 - so we are confident of this negative result, which we now show in Figure S4. As cell wall lipopolysaccharide (LPS) differ between E. coli K-12 and B strains, we also tested the E. coli LPS mutants, which had no eff

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      • The nematode C. elegans is at the forefront of research on transgenerational epigenetic inheritance. In this work the authors studied the effects of natural genetic variations on multigenerational inheritance, using the temperature-sensitive Mortal germline phenotype (Mrt) as a paradigm in C. elegans. In ts Mrt mutants, animals become progressively sterile at 25{degree sign}C (stressful temperature) over subsequent generations and, importantly, this phenotype is reversible. The present study originated from the authors' previous observation that multiple C. elegans wild isolates display a ts-Mrt phenotype when cultured in the lab, raising the question of whether this intrinsically deleterious phenotype may be suppressed in the wild, and how natural genetic variation affects this phenotype.

      • By comparing 132 wild isolates of C. elegans, the authors found a wide distribution in ts-Mrt phenotypes ranging from 3 to 20 generations to reach sterility at 25{degree sign}C. The variance among a restrictive set of 115 replicates was low for strong Mrt values and high at intermediate trait values. Given this distribution, the authors analyzed the data using generalized linear mixed models. This reviewer is unable to evaluate the appropriateness of these models. They then performed GWAS mapping combined with analysis of introgression lines and identified a QTL on chromosome III between 4.66 and 6 .49Mb that includes a number of potentially interesting candidates that were not further analyzed in this work.

      • Because the authors noticed that the Mrt phenotype commonly appears after bleaching the culture, a treatment that kills associated microbes, they then tested the impact of naturally associated microbes on the Mrt phenotype. They found that freshly isolated strains such as JU3224 could be propagated for more than 20 generations at 25{degree sign}C with their associated microbes, while after bleaching on OP50 (bacteria commonly used in lab culture) they developed a Mrt phenotype at 25{degree sign}. They then fed the isolates with naturally associated bacteria isolated in the lab-either their own or from other isolates. Reassociation of single bacterial clones, or a mix of these, fully or partially rescued the Mrt phenotype. Importantly, bacteria isolated from one strain was able to rescue the Mrt of another strain, suggesting common mechanisms of action in rescuing the Mrt phenotype. Surprisingly Microsporidia, usually detrimental to C. elegans, also rescued the Mrt phenotype. These results show that infection of somatic tissues can influence the germline.

      • ts Mrt mutations so far identified affect nuclear small RNA pathways, small RNA amplification and histone modifications in the germline. The authors further show that the Mrt phenotype of laboratory mutants in small RNA inheritance or chromatin factors such as the set-2 histone methyltrasferase is also suppressed by culture on bacteria other than E. coli OP50.

      • Finally, the authors tested whether animals have a memory of their past bacterial environment by shifting animals of the C. elegans JU775 strain that had been cultured for several generations at 25{degree sign}C on an E. coli K-12 strain (on which their Mrt phenotype was suppressed) to the laboratory E. coli OP50, which usually reveals the Mrt phenotype. Lines that were propagated for 10-20 generations at 25{degree sign}C on an E. coli K-12 strain (MG1655 or HT115) showed a rescued phenotype when transferred back on OP50, consistent with a multigenerational memory of the bacterial environment.

      • All experiments are well executed, clearly presented and of the highest standard.

      Significance

      C. elegans is an excellent model system to study transgenerational inheritance. However, most studies on epigenetic inheritance in this system are carried out under standard laboratory conditions, and the phenotypes followed often not very robust (stress resistance, longevity..) raising questions as to their interpretation. This work is an important contribution to the field because it reveals how a widely studied phenotype (the Mrt phenotype) relates to natural isolates. The results reported demonstrate a clear link between the environment and the multigenerational transmission of non-genetic information. They also raise interesting questions on the ability of a species to transiently provide environmental cues to a variable number of generations. Finally, these results offer hints that that the Mrt phenotype may result from inherited metabolic changes, as observed using other experimental paradigms in C. elegans, including starvation. This work will therefore be of interest to a wide audience interested in epigenetic inheritance and the environment, soma-germline communication, and host pathogen interactions.

      Minor points

      1. Could the authors please define "experimental blocks"

      2. Legend to supplementary snp table should be completed: define AF, impact, modifier, moderate, AA1, AA2...

      3. Please define "intermediate-frequency allele"

      4. Figure 7 legend: Authors should be more specific in describing the figure: After 10 (A panel), 13 or 20 generations (B panel) on the K-12 strain... What is E. coli OP50 start 'G10'? the 15{degree sign} stock?

      Optional:

      Did the authors attempt to rescue the Mrt phenotype with individual metabolites (eg Vit B12...)? These are not straight forward experiments and most likely part of a future study.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      In this study, Frézal et al. reported novel interactions between microbes and C. elegans in the regulation of epigenetic inheritance. By screening for 132 isotypes, the authors found natural genetic variants that contribute to the mortal germline (Mrt) phenotype. The authors further found that naturally associated gut bacteria, microsporidia, and E. coli K12 could rescue Mrt phenotype in wild isolates as well as in epigenetic mutants. Finally, the authors showed that the epigenetic memory of bacterial environment could propagate transgenerationally. I find this paper highly intriguing as it provides valuable insights into the impact of the environment on epigenetic inheritance and its effects on evolution within ecologically relevant contexts.

      Major comments:

      • The authors claimed that the variants causing Mrt exist at intermediate frequency in the natural population but the evidence supporting this claim is rather limited. To strengthen the claim, the authors should examine the distribution and frequency (perhaps coupled with phylogenetic analysis) of the Ch III haplotype in the wild isolates. The authors should also examine the GWAS peak for the signature of balancing selection (e.g., dN/dS ratio).

      • Does JU775 carry polymorphisms in genes that are known to be involved in Mrt? These genes may genetically interact with the Ch III variant, as suggested by the partial penetrant phenotypes of the introgressed lines. It would be helpful to have a table summarize the variation in these genes. <br /> It is curious to me that for experiments with HT115, the expression of the RNAi vectors was induced with IPTG. Is this step necessary? It is known that even the backbone of L4440 could trigger a non-specific RNAi response (PMID: 30838421). I wonder if activating exogenous RNAi response is required for Mrt rescue.

      • In figure 7, it appears that the worms transferred from MG1655/HT115 to OP50 showed an even stronger rescue (higher Mrt value) than the ones constantly on MG1655/HT115. This suggests to me that fluctuations in food composition may strongly affect epigenetic inheritance. Please clarify as this is very interesting, if true.

      • Optional - Numerous studies have shown that SKN-1 regulates metabolism in response to food composition and availability (PMID: 23040073). Additionally, some recent studies have indicated a role of SKN-1 in epigenetic inheritance triggered by exogenous RNAi. In particular, SKN-1 promotes stress-induced epigenetic resetting (PMID: 33729152). I wonder if SKN-1 modulates Mrt based on bacterial diet.

      Minor comments:

      • Line 47: typo "...they defined..."

      • Line 100-101: weird sentence structure. Please consider rephrasing.

      • Line 138-139: I don't quite understand what "intermediate-frequency chromosome III alleles" means here. Some SNPs were found in Ch III 4-6Mb? Please expand.

      • Line 213 - it was unclear to me why the assay was performed at 23C instead of 25C. I later learned in the method section that microsporidia cannot be cultured at 25C. I think it will be helpful to add that information when microsporidia is introduced to improve clarity.

      Significance

      This study beautifully demonstrates how diet composition affects epigenetic inheritance. This study is rigorous (replicated by 3 different labs) and the data is solid. Using natural wild isolates and naturally associated microbes, the authors described how diet composition affects germline mortality and epigenetic inheritance. Interestingly, the authors showed that Mrt phenotype might be the result of standard lab cultivation conditions and it was masked when the worms were fed on naturally associated bacteria and microsporidia. Overall, the findings are very interesting and novel. While mechanistic insights are currently lacking, it is outside the scope of this paper. This paper provides an interesting paradigm to study how genetic and environmental variation influence epigenetic inheritance and evolution. I believe this paper will be of great interest to audiences across many fields of biology, including quantitative biology, evo-devo, ecology, and genetics and epigenetics.

      My field of expertise: C. elegans biology, epigenetic inheritance, genetics.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary:

      Frezal, Felix and colleagues study 132 wild isolates of the C. elegans species and demonstrate that the majority of these strains will become sterile within 20 generations if grown at 25oC. This is a very thorough analysis of a multigenerational trait that the authors show is commonly found in wild C. elegans strains. The authors use GWAS to identify a peak on chromosome III that is enriched in strains that become sterile at 25oC. Consistently genetic crosses place this segment of chromosome III from a Mrt wild strain into the N2 background resulted in a strong Mrt phenotype. The authors noticed that bleaching of wild C. elegans strains to remove associated bacteria promoted the Mrt phenotype. Remarkably, the authors show that growth of bleached wild strains on bacteria isolated from the wild strains prior to bleaching is sufficient to suppress the Mrt phenotype. These results were obtained with two wild isolates and with multiple species of wild bacteria, strongly supporting the authors' conclusions. The authors also show that independent types of intracellular bacteria that infect the intestine can partially suppress the Mrt phenotypes. The authors also show partial to strong rescue of temperature-sensitive epigenetic mutants set-2, set-24 and nrde-2 by wild bacteria. Remarkably, the authors demonstrate that growth of the introgressed JU775 strain on a N2 background can be grown on suppressor bacteria for 10 to 20 generations, then bleached and placed on OP50, then there is a multigenerational memory of the suppressor bacteria. This intriguing result is consistent with bacteria having an epigenetic effect on C. elegans Mrt phenotypes, which are themselves in some cases caused by epigenetic defects.

      Comments for the authors:

      1. 'For GWAS, the strains that were fertile after 20 generations were considered non-Mrt.'

      One aspect of Fig 1D that could be clarified are the dots at generation 21. If these represent strains that were always fertile at generation 21, then perhaps give these a different color to indicate that sterility was never observed?

      1. 'The mean Mrt values of strains ranged from sterile at 3 generations to fertile after 20 generations at 25oC, with a skewed distribution toward high values (Figure 1B).'

      Based on Table S2, part of the explanation for this skewed distribution in later generations is that some strains became sterile rapidly for some blocks, whereas the same strain did not become sterile in other blocks. For example, JU1200, JU360, PB303. I suggest providing a second color for Fig. 1D for strains that sometimes displayed sterility and sometimes did not.

      For those that were sometimes fertile and sometimes sterile, I suggest creating a graph in Figure 1 that shows generations at sterility or lack of sterility, color coded by block. This will allow the significance of strains with high generation Mrt values to be better appreciated for readers who do not look at the supplementary table.

      1. The GWAS section could benefit from a simple explanation of the premise of GWAS for non-specialist readers.

      2. One problem might be that the Mrt phenotype is widespread among wild strains. To the authors' credit, they consider results observed in different laboratories as valid, even when the results do not agree. If the Mrt phenotype is influenced by the environment, then some laboratory environments might result in 'false negative' Mrt results that could be ignored in favor of positive results from another lab that appear strong. Might focusing on strains with a set of strong positive results from one lab allow the authors to draw stronger GWAS conclusions?

      3. The authors' perform GWAS based on the variance of the Mrt phenotype data. Would the GWAS data be more illuminating if the authors only considered strains that become sterile fairly rapidly, within 10 generations. The authors might then have a second category that included strains that become sterile from generation 11-20. If the genetic basis for the Mrt phenotypes is the same, then GWAS of strains that become sterile in less than 10 generations might yield similar peaks as GWAS for strains that become sterile between generations 11-20.

      4. 'We did not investigate whether a second locus present in JU775 on the right arm of Chr III might have a lesser effect.'

      5. It might be interesting to test the memory of growth on beneficial bacteria on JU4134, which had a Mrt phenotype that was strongly suppressed by the beneficial bacteria.

      6. The Mrt phenotype of mutants in small RNA inheritance and histone modifying enzymes 'appears however distinct from that of the prg-1/piwi mutant (for which the cause of sterility is debated), especially the latter does not show temperature dependence and is suppressed by starvation.'

      While it is true that the cause of sterility is debated for the prg-1/piwi mutant, this mutant is defective for small RNA silencing and likely has parallels with some defects in histone modifying enzymes. Anecdotal reports suggest that starvation might affect the Mrt phenotype or longevity of histone modifying enzyme mutants. Moreover, the cause of sterility is not clear for small RNA inheritance and histone modifying enzyme mutants. It is fair to say that the distinction between temperature-sensitivity or lack of temperature sensitivity of small RNA mutants is not understood. Could the authors please comment here about whether any of the wild strains display sterility at 20oC.

      1. If intracellular bacteria are simply somatic, then how is it that they are transmitted to progeny. If they are released into the environment and then consumed by hatched larvae, this is soma-to-soma transmission.

      Minor comments:

      1. 'We measured the mortal germline (Mrt) phenotype'. Mortal Germline (Mrt)

      Significance

      All in all, this is an interesting and well-written manuscript that represents a considerable amount of work and demonstrates that a temperature-sensitive multigenerational sterility phenotype is widespread among wild C. elegans strains. This Mrt phenotype is modulated by the food they consume or by intracellular bacterial parasites that reside in somatic intestinal cells. This may mean that the intestine is a major modulator of the Mrt phenotype, which may be a consequence of lab culture conditions and may not occur for wild strains in the wild. Nevertheless, the phenotype or phenotypes are intriguing and likely relevant to natural variation.

      The limitations of this manuscript include a lack of understanding of the precise genes involved or if small RNAs or metabolites from bacteria are involved. But this manuscript represents an enormous effort and raises many interesting points that will be addressed in future efforts.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity):

      Major comments:

      1. A control group of mice fed chow diet is needed to distinguish the effects of the genotype from those caused by diet. What is the phenotype of regular chow-fed mice in terms of energy metabolism and thermogenesis?

      We are sincerely grateful to Reviewer 1 for raising an important question regarding the need for a control group of mice fed chow diet.

      To address this concern, we have conducted experiments on mice fed a regular chow diet and measured their phenotype in terms of energy metabolism and thermogenesis. In addition to be sure that the phenotype also is present in when we compared littermates we have included as control both to chow-fed CD4-Cre and littermates (MKK3/6f/f). Our findings reveal that MKK3/6CD4-KO mice fed a chow diet presented an increased brown adipose tissue (BAT) thermogenesis compared with CD4-Cre and littermates. This phenotype is similar to the observed in HFD-fed mice. Also, these results indicate that the same phenotype is observed when we compared with littermates including an extra control in the study.

      To further investigate the effect on energy metabolism, we utilized metabolic cages. The data from these experiments align with the increased thermogenesis observed in MKK3/6CD4-KO mice fed a chow diet, as they also demonstrated increased energy expenditure. We thank the reviewer for this suggestion as we believe that these new data strengthen our conclusion significantly.

      We have thoughtfully incorporated these essential findings into in Supplementary Figure 2C-D of the manuscript.

      1. While an increase in BAT temperature (as demonstrated here by infrared imaging) in line with increased thermogenesis, it will be critical to verify this hypothesis by indirect calorimetry. Energy expenditure, food intake, and activity measures should be added for regular and DIO mice. Please follow the guidelines for ANCOVA analysis and measurements explained in PMID: 22205519 and PMID: 21177944.

      We are grateful to Reviewer 1 for bringing up an essential point concerning the need to verify our hypothesis on increased BAT temperature and thermogenesis through indirect calorimetry. We acknowledge the importance of including energy expenditure, food intake, and activity measures for both regular and DIO mice to strengthen our study.

      To address this valuable suggestion, we have taken immediate action. We utilized metabolic cages in mice under chow diet. The data from these experiments align with the increased thermogenesis observed in MKK3/6CD4-KO mice fed a chow diet, as they also demonstrated increased energy expenditure, without differences in food intake or locomotor activity. We thank the reviewer for this suggestion as we believe that these new data strengthen our conclusion significantly. These new data are now in Supplementary Figure 2A-B.

      In addition, we have initiated a new experimental group of age-matched mice on HFD, which we will carefully feed for 8 weeks. Following this dietary period, we will subject the mice to metabolic cage analysis, allowing us to obtain accurate data on energy expenditure, food intake, and activity levels. These additional measurements will provide a comprehensive understanding of the metabolic changes induced by MKK3/6 deficiency in T cells under different dietary conditions.

      1. That the phenotype is still seen at isothermal housing is interesting but should be backed up by direct assessment of thermogenic capacity (see PMID: 21177944). In the end, it could also be increased heat loss, independently of heat production. If the browning is cause or consequence remains unclear, then.

      Thank you for raising this important point. Indeed, it is essential to corroborate the observed phenotype with direct assessments of thermogenic capacity to gain a comprehensive understanding of the underlying mechanisms. The study mentioned in PMID: 21177944 highlights the significance of evaluating thermogenesis directly to support the findings.

      According to your suggestion, we plan to house the animals at 30 ºC for four weeks and subsequently inject norepinephrine to evaluate thermogenesis capacity while measuring brown adipose tissue (BAT) activation. This approach should provide valuable insights into the thermogenic potential of the animals under isothermal conditions.

      However, we will not be able to conduct the experiment in metabolic cages at 30 ºC due to the constraint that our system does not allow 30 ºC temperature. For this reason, we will measure BAT temperature to analyze this experiment.

      1. Regarding the in vitro data, a thermogenic phenotype should be functionally verified by Seahorse analysis.

      We thank Reviewer 1 for raising an important point concerning the need for functional verification of the thermogenic phenotype observed in our in vitro data using Seahorse analysis.

      In response to this valuable suggestion, we performed Seahorse analysis in differentiated adipocytes treated with or without IL-35 for 48 hours. The results demonstrated a slight increase in basal metabolism and a heightened response to isoproterenol (ISO) stimulation of β3 adrenergic receptors in adipocytes after IL-35 treatment. These findings provide functional evidence supporting the thermogenic phenotype induced by IL-35 in adipocytes.

      We have thoughtfully included this essential data in Figure 2 of this revision plan, allowing reviewers and the scientific community to comprehensively evaluate and validate the functional implications of our findings.

      1. Mechanistically, there is epistasis type of experiment that IL-35 influences Ucp1 levels via ATF2 as the data remain associative in nature.

      Thank you for your valuable comment. We agree that to establish a mechanistic link between IL-35 and Ucp1 levels will improve the strength of the manuscript.

      To delve deeper into the mechanism through which IL-35 influences Ucp1 expression, we focused on the role of ATF2, a transcription factor known to be involved in regulating UCP1 levels (PMID: 11369767 and PMID: 15024092). In our investigation, we treated adipocytes with IL-35 both in the presence and absence of an inhibitor targeting the ATF2 pathway. The results were illuminating as we observed a significant reduction in the expression of Ucp1 when the ATF2 pathway was inhibited.

      These findings indicate that ATF2 is indeed a crucial mediator of the effects of IL-35 on Ucp1 levels. By inhibiting the ATF2 pathway, we demonstrate a direct functional link between IL-35 and the expression of Ucp1, providing mechanistic insights into the regulatory role of IL-35 in thermogenesis. We included new results in Figure 7F.

      1. What are other consequences of injecting IL-35? Is it good or bad? What is the therapeutic potential in DIO mice? Also, in these experiments (Fig. 7) indirect calorimetry as described would be supportive of the claims.

      Regarding the consequences of injecting IL-35, we have already performed experiments to analyze its effect. Our findings indicate that IL-35 increases thermogenesis in BAT (Figure 7), suggesting that it may play a role in promoting energy expenditure, which could be beneficial in combating diet-induced obesity (DIO) in mice. Importantly, we did not observe any negative effects of IL-35 in our experiments.

      Based on these promising results, we are expecting the therapeutic potential of IL-35 in DIO mice. By promoting thermogenesis in BAT, IL-35 may offer a novel approach to manage obesity and related metabolic disorders. However, we acknowledge that further comprehensive studies are needed to fully understand its therapeutic benefits and potential side effects.

      In our future works, we plan to evaluate a targeted delivery system for IL-35. We are currently generating IL-35 loaded metal-organic frameworks (MOFs) labeled with adipose tissue-specific peptides. This innovative strategy aims to enhance the delivery of IL-35 to adipose tissue, potentially maximizing its effects in the relevant areas. Our ongoing work with IL-35 loaded MOFs may offer a promising avenue for targeted delivery.

      Minor comments:

      1. The authors claim that their HFD-fed MKK3/6CD4-KO mice are protected against hyperglycemia, but only fasted/fed blood glucose tests are performed. Lower glucose levels could be explained due to a hyperinsulinemic state in response to growing insulin resistance in the presence of HFD. It would be sensible to perform both glucose and insulin tolerance tests to back up your statement.

      Thank you for your insightful comment. We agree that to support our claim of protection against hyperglycemia in HFD-fed MKK3/6CD4-KO mice, further tests are necessary beyond fasted/fed blood glucose measurements.

      In response to your suggestion, we conducted both glucose tolerance tests (GTT) and insulin tolerance tests (ITT) in HFD-fed MKK3/6CD4-KO mice. We did not observed differences in glucose tolerance and but ITT showed significantly enhanced insulin sensitivity compared to control mice. These findings provide evidence that the protection against hyperglycemia in HFD-fed MKK3/6CD4-KO mice is not solely due to a hyperinsulinemic state, but rather indicates genuine improvements in glucose handling and insulin response.

      We have thoughtfully included these crucial data in the revised version of the manuscript, both in the main text and Supplementary Figure 4. We extend our appreciation to the reviewer for this valuable suggestion, which has enhanced the scientific rigor and completeness of our study.

      1. Please provide the loading control for p38 and S6 blots (Figure 6G).

      Thank you for the comment. The loading control we used for P p38 and P S6 blots in Figure 6G is β-actin. Due to the limited amount of sample available, we can only use β-actin as the loading control. The sample amount obtained is very limited, and we can only provide enough lysate to run a couple of blots from the same sample. Running several western blots with the same sample is almost impossible given the constraint of the sample availability. We apologize for this limitation, but it is necessary to avoid using too many mice for ethical reasons, as the samples come from a large number of mice.

      1. Statistical test from Figure 7B should be a t-test, since it is only comparing 2 variables (PBS vs IL-35), and not a 2-way ANOVA as described in the legend.

      We sincerely thank the reviewer for the comment. It was indeed a mistake in the text. While we have performed a t-test, there was an error in the legend that we have now corrected. We apologize for any confusion this may have caused and appreciate the opportunity to rectify the oversight.

      1. Label correctly the panels in the figures -examples: Fig 3, panels C and D are interchanged; reference in the text to Fig S1G even though the figure only as panels A-F; Fig 7 legend referes to the statistical test of panel E when the figure only has A-D.

      We sincerely apologize for any mistakes in our manuscript that may have caused difficulties while reading the article and potentially led to misleading results. We are grateful to Reviewer #1 for bringing these errors to our attention. Thanks to their diligent review, we have been able to identify and rectify the issues in our manuscript. The necessary corrections have been made, ensuring the accuracy and reliability of our research. We greatly appreciate the reviewer's valuable feedback and contribution to improving the quality of our work.

      1. There are several typos along the text, please revise (example: page 4;line 4 -"tremorgenic")

      We apologize for the presence of any typos in the initial version of the article. We have thoroughly revised the manuscript to correct these errors. Thank you for bringing this to our attention and helping us improve the accuracy and clarity of our work.

      Reviewer #1 (Significance):

      The manuscript is well written, and the research conducted properly, even though a thorough analysis of energy metabolism in mice and cells is missing and the mechanistic claims are based on relatively thin data.

      The immune system and inflammation play important roles for obesity and insulin resistance, yet the roles they play in thermogenic adipocytes remains unclear. This work adds novel aspects to this relationship.

      Reviewer #2 (Evidence, reproducibility and clarity):

      This manuscript by Nikolic et al sought to investigate the role of p38 activation in adipose tissue Treg cells and obesity. They found that the expression of p38a, its upstream kinase MKK6, and downstream substrate ATF2 was upregulated specifically in adipose T cells associated with human obesity. They generated T cell-specific knockout MKK3/6 in mice and found these animals were protected from diet-induced obesity as a result of increased BAT thermogenesis. Mechanistically, loss of p38a activation promoted adipose tissue accumulation of Treg cells, leading to elevated IL-35 availability and UCP1 expression.

      Major comments:

      1. They attributed the obesity protection to energy expenditure; however, food intake and intestinal absorption were never tested. Immune cells particularly Treg cells are important modulates of nutrient uptake.

      We are sincerely grateful to Reviewer #2 for this crucial comment, highlighting the importance of assessing not only energy expenditure but also food intake and intestinal absorption in our study.

      In response to this valuable suggestion, we have initiated an HFD experiment to comprehensively examine food intake and intestinal absorption. For food intake analysis, we are employing metabolic cages, which will allow us to monitor and quantify the amount of food consumed by the mice accurately. Additionally, we plan to follow the methodology outlined in the study by Kraus et al. (PMID: 27110587) to measure lipid content in feces, enabling us to evaluate intestinal absorption.

      By conducting these additional experiments, we aim to gain a deeper understanding of the potential role of Treg cells, known immune modulators of nutrient uptake, in our observed obesity protection phenotype.

      1. At thermoneutrality, BAT is inactive even though UCP1 expression is still present (not activated). MKK3/6 deficiency in T cells still confer protection against obesity at thermoneutrality suggests it regulates other energy balance components in addition to BAT thermogenesis.

      Thanks for the comment. We believe that the effects of IL35 on thermogenesis are likely partly mediated by alternative mechanisms, as we did not observe an increase in UCP1 gene expression in BAT in vivo (Figure 3D of the manuscript), and the increase in thermogenesis is still present even at thermoneutrality where UCP1 is inactive (Figure 4E of the manuscript). This suggests that IL35 might regulate other alternative pathways that control BAT thermogenesis.

      While our current findings provide valuable insights, further experiments may be necessary to fully understand the underlying mechanisms. For instance, conducting experiments with transgenic mice expressing IL35 or using IL35 knockout (KO) mice could shed more light on the specific pathways through which IL35 exerts its effects on thermogenesis and energy balance.

      In conclusion, we hypothesize that IL35's effects on thermogenesis are mediated partly by alternative mechanisms beyond UCP1 activation, and its ability to enhance thermogenesis even at thermoneutrality highlights its potential as a regulator of energy balance. We plan to further investigate the specific mechanisms through which IL35 impacts thermogenesis and energy balance. To achieve this, we will consider conducting experiments with transgenic mice expressing IL35 or using IL35 knockout (KO) mice in follow up studies. This is now discussed in our manuscript.

      1. Loss of adipose Treg cells (such as Pparg KO, Foxp3-DTR) did not lead to obvious obesity phenotypes. Gain-of-function Treg cells (such as adoptive transfer, IL-2/IL-2 Ab) did not results in profound obesity protection as observed in MKK3/6 CD4-KO mice. It suggests that MKK3/6 KO in T cells causes other immune defects (besides Tregs).

      We agree with the referee's assessment that the lack of obvious obesity phenotypes in above mentioned animal models. The results we observed in our MKK3/6CD4-KO mice suggest that p38 signaling pathway in T cells may modulate their function, leading to an upregulation of IL35 expression, which could be a contributing factor to the significant obesity protection observed in MKK3/6CD4-KO mice. We believe that IL35's effects on energy balance and thermogenesis are critical components of the observed protection against obesity in this model.

      Regarding the studies with PPAR KO in Treg cells, it is important to note that they did not specifically focus on the effect of thermogenesis. While they observed a general tendency of increased fat deposition when treated with a PPAR agonist in the Treg deficient PPAR KO mice, these findings were not extensively studied in that particular paper. Thus, additional research is necessary to specifically evaluate thermogenesis in these mice and further understand the role of PPAR in Treg-mediated thermogenic processes.

      We also acknowledge the presence of contradictory results from loss-of-function experiments of Treg cells in mice. The observed metabolic changes may be context-dependent, and the impact of Treg cells on metabolism might vary under different physiological conditions. For instance, in lean conditions where adipose tissue inflammation is low, a decrease in VAT Treg cells might not lead to significant metabolic changes. However, under certain circumstances, such as obesity, VAT Treg cells may play a critical role in regulating metabolism. In this context increasing that population that is reduced during obesity could results in improve metabolic performance.

      In conclusion, our findings suggest that the lack of p38 activation in Treg cells may prevent the dramatic down-regulation and loss of function observed in Treg cells during obesity. This preservation of Treg function could be a significant factor driving the observed protection against obesity in MKK3/6CD4-KO mice.

      While further studies are required to elucidate the precise timing and spatial aspects of the specific functions of adipose-resident Treg cells, it is evident that these cells play a crucial role in maintaining immune and metabolic homeostasis. They achieve this, in part, by regulating adipose inflammation, insulin sensitivity, lipolysis, and thermogenesis. This is now discussed in our manuscript.

      1. The increase in IL-35 seemed to be very moderate, compared to the metabolic phenotypes. It raises the question if IL-35 is responsible for BAT activation and reduced weight gain. It is unclear what systemic and local levels of IL-35 were reached after recombinant IL-35 treatment (Fig. 7B). IL-35 antibody blockade experiment in KO mice is recommended.

      Physiological changes in cytokines can indeed have a significant impact on the metabolic profile due to their continuous and intricate interactions. Even minor alterations in the overall cytokine milieu can result in substantial changes in metabolism (doi.org/10.1073/pnas.1215840110). In fact, it is well-established that in humans, small changes in cytokine profiles between genders, in obesity, and during aging can play a critical role in the development of pathology. These cytokines often operate in a chronic manner, exerting long-term effects on various physiological processes (doi.org/10.1038/s41467-020-14396-9).

      In summary, the dynamic interplay of cytokines in metabolism can lead to significant metabolic changes even with subtle alterations in their levels. While the increase in IL-35 may appear moderate, our findings using recombinant IL35 indicate that IL-35 increases thermogenesis in BAT, suggesting that it may play a role in promoting energy expenditure, which could be beneficial in combating diet-induced obesity (DIO) in mice. Importantly, we did not observe any negative effects of IL-35 in our experiments.

      1. IL-35 induced p-ATF2 is acute and transient (Fig. 7D) and it was able to increase BAT temperature in just 4 h (Fig. 7B). However, Ucp1 transcription and translation generally take much longer time (e.g. 2d in Fig. 7C). IL-35 may increase energy expenditure through UCP1-independent mechanisms.

      Thanks for the comment. As previously mentioned, we believe that the effects of IL35 on thermogenesis are might be mediated by alternative mechanisms, as we did not observe an increase in UCP1 gene expression in BAT, and the increase in thermogenesis is still present even at thermoneutrality where UCP1 is inactive. This suggests that IL35 might regulate other alternative pathways that control BAT thermogenesis.

      While our current findings provide valuable insights, further experiments may be necessary to fully understand the underlying mechanisms. For instance, conducting experiments with transgenic mice expressing IL35 or using IL35 knockout (KO) mice could shed more light on the specific pathways through which IL35 exerts its effects on thermogenesis and energy balance. We plan to further investigate the specific mechanisms through which IL35 impacts thermogenesis and energy balance. To achieve this, we will consider conducting experiments with transgenic mice expressing IL35 or using IL35 knockout (KO) mice in follow up studies. This is now discussed in our manuscript.

      Minor comments:

      1. The gating of Treg cells should exclude CD25- cells. Single positive (CD25+ or Foxp3+) cells are progenitors of Tregs. In addition to number, phenotypic activation of Treg cells should also be determined.

      Thank you for the comment. We have reanalyzed our data by excluding CD25- cells and included now in the figure 5A of the manuscript and new supplementary figure 7 of revised manuscript. We also checked CD69+ and KLRG1+ Treg cells and observed no differences between genotypes. We also included figures in this revision plan (Figure 5 and 6).

      1. ATF is also important for adipogenesis, is the adipogenic differentiation of BAT SVF cells affected by MKK3/6 KO or IL-35 treatment?

      We appreciate the reviewer's observation regarding the importance of ATF in adipogenesis. To investigate this aspect further, we performed in vitro differentiation of adipocytes and treated them with IL-35 in the presence or absence of an inhibitor targeting the upstream activator of ATF.

      The results were compelling, as IL-35 treatment led to an increase in the expression of adipogenic markers, including Pparg, Adipoq, Leptin, and Perilipin. In contrast, inhibiting ATF activation resulted in a reduction of these adipogenic markers. These findings provide strong evidence that ATF plays a significant role in mediating the effects of IL-35 on adipogenesis.

      We have thoughtfully included these essential data in Figure 7G of the manuscript. We extend our gratitude to the reviewer for their keen observation, which has enhanced the scientific depth and completeness of our study.

      1. Metabolic cage experiments are desired to determine whole-body energy balance, including food intake, physical activity, and heat production.

      To address this valuable suggestion, we have taken immediate action. We utilized metabolic cages in mice under chow diet. The data from these experiments align with the increased thermogenesis observed in MKK3/6CD4-KO mice fed a chow diet, as they also demonstrated increased energy expenditure, without differences in food intake or locomotor activity. We thank the reviewer for this suggestion as we believe that these new data strengthen our conclusion significantly. The new data are included in Supplementary figure 2 A-B.

      In addition, we have initiated a new experimental group of age-matched mice on HFD, which we will carefully feed for 8 weeks. Following this dietary period, we will subject the mice to metabolic cage analysis, allowing us to obtain accurate data on energy expenditure, food intake, and activity levels. These additional measurements will provide a comprehensive understanding of the metabolic changes induced by MKK3/6 deficiency in T cells under different dietary conditions.

      1. Total UCP1 expression (both RNA and protein) in the whole BAT from an animal should determined (since BAT is smaller in KO mice).

      Thank you for this comment. Yes, we have measured UCP1 expression in the whole BAT from the animals. It is in the figure 3C and 3D and here. Although in vitro studies indicated that IL35 increase UCP1 in adipocytes we were not able to find an increase of this protein in BAT

      We believe that the effects of IL35 on thermogenesis are likely partly mediated by alternative mechanisms, as we did not observe an increase in UCP1 gene expression in BAT in vivo, and the increase in thermogenesis is still present even at thermoneutrality where UCP1 is inactive (Figure 4E of the manuscript). This suggests that IL35 might regulate other alternative pathways that control BAT thermogenesis.

      1. Fig. 6C, IL-35-expressing Treg cells should be quantified from adipose tissue.

      We appreciate the referee's suggestion to quantify IL-35-expressing Treg cells from adipose tissue in Fig. 6C. While we agree that this would be valuable information, we encountered technical challenges that made it impractical to measure IL-35 directly in Treg cells from the visceral adipose tissue (VAT).

      One of the main technical challenges we encountered is the low number of Treg cells present in the adipose tissue, making it difficult to obtain sufficient cell material for accurate quantification of IL-35. Treg cells are relatively rare compared to other immune cell populations in the adipose tissue, and their extraction and analysis can be technically demanding.

      Reviewer #2 (Significance):

      The manuscript is innovative in define the novel role of p38 activation in the T cell compartment and its metabolic regulation. The involvement of Treg cells in adipose tissue homeostasis has been well documented and Treg cell-derived IL-35 has been demonstrated in immune regulation. The authors provided a relatively thorough description of the altered metabolism in these Mkk3/6 CD4-KO mice; however, the reviewer has doubts if Treg cells and IL-35 are primary mechanisms of the observed protection from obesity. The manuscript would be much stronger if the model were Treg cell-specific KO and/or IL-35 deficiency in Treg cells reverses obesity resistance conferred by MKK3/6 deficiency. It also suspected that BAT thermogenesis is not the major reason, as BAT deficiency or UCP1 KO results in much milder phenotypes in mice, even at thermoneutrality.

      Reviewer #3 (Evidence, reproducibility and clarity):

      Specific comments:

      1. It's important to use proper controls for mouse metabolic studies. The authors stated that CD4-Cre and MKK3/6 CD4-KO mice are all in the C57B/6L background. However, it would appear that these two lines were bred separately. The difference in the genetic background, despite minor, can lead to the observed phenotype, notably weight gain. Since the metabolic phenotypes seem to be driven by the weight difference, it is even more critical to include additional controls to validate the findings. For instance, crossing MKK3/6 f/f with one copy of CD4-Cre with MKK3/6 f/f to generate age-matched MKK3/6 CD4-KO and MKK3/6 f/f controls should be used to repeat major in vivo studies similar to those in Fig. 2-4.

      We thank the reviewer for the comment. Although, every control is important using conditional mice, there are several papers indicating that all the cre expression lines have for their own effects that could be important in metabolism and there are several articles that strongly recommended to use cre+ lines as a control. For that reason, we have used the cre expressing line as a control because we really think is the best one (Jonkers and Berns, 2002). In fact, Jackson laboratory recommend to use cre expressing line as a control to avoid side effects that cre overexpression could have in the tissue of interest (https://biokamikazi.files.wordpress.com/2014/07/cre-lox-imp-notes.pdf).

      However, as this reviewer suggested, we checked that similar results were obtained using littermates as controls and we have now included these data in the manuscript (Supplementary Figure 2D).

      1. The assessment of adipose tissue immune cell population in Fig. 5 was conducted after HFD-induced obesity. As mentioned above, the change in Treg and M2 cell percentage could be due to the body weight difference. The experiment should be repeated (with proper controls) in normal chow and after a few weeks of HFD when Treg numbers start to decline.

      Thank you for the comment. We currently performing short HFD experiment to check Treg and M2 cell population in adipose tissue using the littermates as controls.

      In addition, we checked those cell populations in adipose tissue infiltrates in mice fed chow diet and observed no differences in M2 macrophage population between mice, while the percentage of Treg cells was actually lower in MKK3/6CD4-KO mice ND-fed mice (Fig 12 of revision plan). This result suggests that higher accumulation of Treg cells in mice lacking p38 activation in T cells are specific of obese state and strengthen our hypothesis that DIO protection in MKK3/6CD4-KO mice is due to Treg cell population.

      1. Data related to the mechanistic link in Fig. 6/7 are not robust and require a large amount of additional work to substantiate the claim. First of all, the role of IL-35 in BAT thermogenesis remains unclear. It's somewhat surprising to see a single dose of IL-35 i.v. injection is sufficient to increase BAT temperature in Fig. 7B. Minimally, the authors need to demonstrate that IL-35 treatment (perhaps after a few daily doses) is able to increase browning/beiging of fat cells and improve cold tolerance when placing the mice at 4 degree of several hours (and up to 3 days). Serum FGF21 level should also be measured after/during IL-13 treatment. Secondly, ATF2 knockout or knockdown in brown preadipocytes should be employed to demonstrate that IL-35 induced UCP1 and FGF21 expression is ATF2 dependent. Another key experiment is to use IL-35 deficient Treg model to definitively demonstrate the requirement of Treg IL-35 to maintain thermogenesis. However, this can be done in a follow up study.

      We are grateful for all the insightful comment provided by Reviewer #3. We understand the concern, but we have the limitations in performing several sequential i.v. injections in our animal facility due to ethical permissions. In light of this constraint, we have devised an alternative approach to evaluate the role of IL-35 in adaptive thermogenesis.

      To address this, we conducted a cold tolerance test in both control mice and MKK3/6CD4-KO mice, which express higher levels of IL-35. Our findings revealed that MKK3/6CD4-KO mice exposed to cold conditions were able to preserve their body and brown adipose tissue (BAT) temperature, while the temperature of control CD4-Cre mice gradually dropped during the cold challenge.

      The data from this cold tolerance test support our hypothesis and demonstrate the role of IL-35 in promoting adaptive thermogenesis, leading to enhanced temperature maintenance in MKK3/6CD4-KO mice. These observations have been included in Figure 7B of the manuscript, and detailed results are available in Figure 11 of this revision plan.

      We appreciate the reviewer's valuable input, which has encouraged us to explore alternative experimental approaches to address the research question effectively.

      We agree with the reviewer #3 that using IL-35 deficient Treg model would be great approach to confirm our results, but we think that now with the additional experiments we have performed, we strength our findings that IL-35 has a novel role in controlling adipose tissue thermogenesis.

      Reviewer #3 (Significance):

      Dissipating energy as heat through brown or beige adipocyte-mediated thermogenesis is believed to be an effective way to combat obesity. The current study aims to characterize the p38 signaling pathway in T cells as a potential target to modulate browning or beiging of adipose tissues. This would be of interest to the basic biomedical research community, particularly in the area of immunometabolism. A major limitation is the concern of improper controls for the mouse models, which makes data interpretation difficult. In addition, the mechanistic studies lack in depth analyses to support the conclusion.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Nikolic et al. examine the metabolic outcome of T cell specific deletion of MKK3/6 (MKK3/6 CD4-KO), which are the main activators of p38. Previous studies have demonstrated that MKK3/6 CD4-KO leads to Treg expansion and that Tregs in adipose tissues are associated with improved metabolic homeostasis. In line with these observations, the authors show that MKK3/6 CD4-KO mice gain less weight and have more active brown fat thermogenesis on a HFD both at the room temperature and 30C housing conditions. They also find more Tregs and M2 macrophages in eWAT of MKK3/6 CD4-KO. All of the metabolic parameters are compared to CD4-cre mice as the wild type controls. Mechanistically, the authors suggest that reduced p38 activation by MKK3/6 CD4-KO leads to increased IL-35 production by Tregs, which induces beiging/browning of adipose tissues to promote metabolic health.

      The authors have spent most of the efforts conducting metabolic phenotyping of MKK3/6 CD4-KO mice. One potential issue is whether the non-littermate CD4-cre mice are the proper controls for the comparison. In addition, the mechanistic link of the IL-35-ATF2-UCP1/FGF21 axis has only been superficially addressed.

      Specific comments:

      1. It's important to use proper controls for mouse metabolic studies. The authors stated that CD4-Cre and MKK3/6 CD4-KO mice are all in the C57B/6L background. However, it would appear that these two lines were bred separately. The difference in the genetic background, despite minor, can lead to the observed phenotype, notably weight gain. Since the metabolic phenotypes seem to be driven by the weight difference, it is even more critical to include additional controls to validate the findings. For instance, crossing MKK3/6 f/f with one copy of CD4-Cre with MKK3/6 f/f to generate age-matched MKK3/6 CD4-KO and MKK3/6 f/f controls should be used to repeat major in vivo studies similar to those in Fig. 2-4.
      2. The assessment of adipose tissue immune cell population in Fig. 5 was conducted after HFD-induced obesity. As mentioned above, the change in Treg and M2 cell percentage could be due to the body weight difference. The experiment should be repeated (with proper controls) in normal chow and after a few weeks of HFD when Treg numbers start to decline.
      3. Data related to the mechanistic link in Fig. 6/7 are not robust and require a large amount of additional work to substantiate the claim. First of all, the role of IL-35 in BAT thermogenesis remains unclear. It's somewhat surprising to see a single dose of IL-35 i.v. injection is sufficient to increase BAT temperature in Fig. 7B. Minimally, the authors need to demonstrate that IL-35 treatment (perhaps after a few daily doses) is able to increase browning/beiging of fat cells and improve cold tolerance when placing the mice at 4 degree of several hours (and up to 3 days). Serum FGF21 level should also be measured after/during IL-13 treatment. Secondly, ATF2 knockout or knockdown in brown preadipocytes should be employed to demonstrate that IL-35 induced UCP1 and FGF21 expression is ATF2 dependent. Another key experiment is to use IL-35 deficient Treg model to definitively demonstrate the requirement of Treg IL-35 to maintain thermogenesis. However, this can be done in a follow up study.

      Significance

      Dissipating energy as heat through brown or beige adipocyte-mediated thermogenesis is believed to be an effective way to combat obesity. The current study aims to characterize the p38 signaling pathway in T cells as a potential target to modulate browning or beiging of adipose tissues. This would be of interest to the basic biomedical research community, particularly in the area of immunometabolism. A major limitation is the concern of improper controls for the mouse models, which makes data interpretation difficult. In addition, the mechanistic studies lack in depth analyses to support the conclusion.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      This manuscript by Nikolic et al sought to investigate the role of p38 activation in adipose tissue Treg cells and obesity. They found that the expression of p38a, its upstream kinase MKK6, and downstream substrate ATF2 was upregulated specifically in adipose T cells associated with human obesity. They generated T cell-specific knockout MKK3/6 in mice and found these animals were protected from diet-induced obesity as a result of increased BAT thermogenesis. Mechanistically, loss of p38a activation promoted adipose tissue accumulation of Treg cells, leading to elevated IL-35 availability and UCP1 expression.

      Major comments:

      1. They attributed the obesity protection to energy expenditure; however, food intake and intestinal absorption were never tested. Immune cells particularly Treg cells are important modulates of nutrient uptake.
      2. At thermoneutrality, BAT is inactive even though UCP1 expression is still present (not activated). MKK3/6 deficiency in T cells still confer protection against obesity at thermoneutrality suggests it regulates other energy balance components in addition to BAT thermogenesis.
      3. Loss of adipose Treg cells (such as Pparg KO, Foxp3-DTR) did not lead to obvious obesity phenotypes. Gain-of-function Treg cells (such as adoptive transfer, IL-2/IL-2 Ab) did not results in profound obesity protection as observed in MKK3/6 CD4-KO mice. It suggests that MKK3/6 KO in T cells causes other immune defects (besides Tregs).
      4. The increase in IL-35 seemed to be very moderate, compared to the metabolic phenotypes. It raises the question if IL-35 is responsible for BAT activation and reduced weight gain. It is unclear what systemic and local levels of IL-35 were reached after recombinant IL-35 treatment (Fig. 7B). IL-35 antibody blockade experiment in KO mice is recommended.
      5. IL-35 induced p-ATF2 is acute and transient (Fig. 7D) and it was able to increase BAT temperature in just 4 h (Fig. 7B). However, Ucp1 transcription and translation generally take much longer time (e.g. 2d in Fig. 7C). IL-35 may increase energy expenditure through UCP1-independent mechanisms.

      Minor comments:

      1. The gating of Treg cells should exclude CD25- cells. Single positive (CD25+ or Foxp3+) cells are progenitors of Tregs. In addition to number, phenotypic activation of Treg cells should also be determined.
      2. ATF is also important for adipogenesis, is the adipogenic differentiation of BAT SVF cells affected by MKK3/6 KO or IL-35 treatment?
      3. Metabolic cage experiments are desired to determine whole-body energy balance, including food intake, physical activity, and heat production.
      4. Total UCP1 expression (both RNA and protein) in the whole BAT from an animal should determined (since BAT is smaller in KO mice).
      5. Fig. 6C, IL-35-expressing Treg cells should be quantified from adipose tissue.

      Referees cross-commenting

      I agree with Reviewer #1. In addition to energy metabolism and mechanistic action of IL-35, more rigor characterization of adipose Treg cells is needed.

      Significance

      The manuscript is innovative in define the novel role of p38 activation in the T cell compartment and its metabolic regulation. The involvement of Treg cells in adipose tissue homeostasis has been well documented and Treg cell-derived IL-35 has been demonstrated in immune regulation. The authors provided a relatively thorough description of the altered metabolism in these Mkk3/6 CD4-KO mice; however, the reviewer has doubts if Treg cells and IL-35 are primary mechanisms of the observed protection from obesity. The manuscript would be much stronger if the model were Treg cell-specific KO and/or IL-35 deficiency in Treg cells reverses obesity resistance conferred by MKK3/6 deficiency. It also suspected that BAT thermogenesis is not the major reason, as BAT deficiency or UCP1 KO results in much milder phenotypes in mice, even at thermoneutrality.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      In this study, Nikolic et al. show a novel role for p38 signaling in Treg cells, which impacts adipocytes through IL-35. This mechanism seems to be important for adipose tissue browning and metabolic health and could be potentially therapeutically exploited.

      Major comments:

      1. A control group of mice fed chow diet is needed to distinguish the effects of the genotype from those caused by diet. What is the phenotype of regular chow-fed mice in terms of energy metabolism and thermogenesis?
      2. While an increase in BAT temperature (as demonstrated here by infrared imaging) in line with increased thermogenesis, it will be critical to verify this hypothesis by indirect calorimetry. Energy expenditure, food intake, and activity measures should be added for regular and DIO mice. Please follow the guidelines for ANCOVA analysis and measurements explained in PMID: 22205519 and PMID: 21177944.
      3. That the phenotype is still seen at isothermal housing is interesting but should be backed up by direct assessment of thermogenic capacity (see PMID: 21177944). In the end, it could also be increased heat loss, independently of heat production. If the browning is cause or consequence remains unclear, then.
      4. Regarding the in vitro data, a thermogenic phenotype should be functionally verified by Seahorse analysis.
      5. Mechanistically, there is epistasis type of experiment that IL-35 influences Ucp1 levels via ATF2 as the data remain associative in nature.
      6. What are other consequences of injecting IL-35? Is it good or bad? What is the therapeutic potential in DIO mice? Also, in these experiments (Fig. 7) indirect calorimetry as described would be supportive of the claims.

      Minor comments:

      1. The authors claim that their HFD-fed MKK3/6CD4-KO mice are protected against hyperglycemia, but only fasted/fed blood glucose tests are performed. Lower glucose levels could be explained due to a hyperinsulinemic state in response to growing insulin resistance in the presence of HFD. It would be sensible to perform both glucose and insulin tolerance tests to back up your statement.
      2. Please provide the loading control for p38 and S6 blots (Figure 6G).
      3. Statistical test from Figure 7B should be a t-test, since it is only comparing 2 variables (PBS vs IL-35), and not a 2-way ANOVA as described in the legend.
      4. Label correctly the panels in the figures -examples: Fig 3, panels C and D are interchanged; reference in the text to Fig S1G even though the figure only as panels A-F; Fig 7 legend referes to the statistical test of panel E when the figure only has A-D.
      5. There are several typos along the text, please revise (example: page 4;line 4 -"tremorgenic")

      Referees cross-commenting

      I think we three reviewers are pretty much on the same page - mouse energy metabolism explored too little and the mechanistic insight a bit thin considering the relatively strong claims.

      Significance

      The manuscript is well written, and the research conducted properly, even though a thorough analysis of energy metabolism in mice and cells is missing and the mechanistic claims are based on relatively thin data.

      The immune system and inflammation play important roles for obesity and insulin resistance, yet the roles they play in thermogenic adipocytes remains unclear. This work adds novel aspects to this relationship.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      We thank the reviewers for their thoughtful comments. We were delighted that the reviewers found our manuscript and results “solid”, “important”, “well-written”, “thoughtful”, “critical addition to the literature”, that the “design of (these) experiments is high in quality” and “conclusions are convincing and the experiments are well executed”.

      We were thrilled the reviewers appreciated that “this manuscript provides solutions to technical limitations to observe mRNA in vivo” by approaching such limitations “in a thoughtful study where many of the salient features of MS2 epitope tagging are systematically measured” and “it provides a greatly improved tool to track mRNA by live imaging” that “also alerts of experimental noise that can be found and can be specific for each gene/transcript

      We will address all the concerns raised by the reviewers. Most of the comments concern text edits. In addition, we will add the following to the Results section:

      1. Quantitation of observed phenotypes in Figures 1C-D and 2C-D;

      2. Quantitation of cytoplasmic transcripts in Figure 1G-L.

      Quantitation will be performed as previously done in Tocchini et al., 2021.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Review of: "An adapted MS2-MCP system to visualize endogenous cytoplasmic mRNA with live imaging in Caenorhabditis elegans"<br /> Authors: Cristina Tocchini and Susan Mango

      The MS2-MCP imaging platform is an essential imaging system that enables dynamic quantification of mRNA transcription, abundance, location, and turnover in living biological systems. In the last ten or so years, this approach has been used in extremely successful ways in Drosophila embryos to dissect both the regulatory logic underpinning early transcriptional organization and activation with unprecedented resolution and, furthermore, how active mRNA localization outside of the nucleus impacts pattern formation. The authors correctly point out that full implementation of this tool has been suspiciously lacking in the C. elegans community for some time (aside from a few noted implementations).

      In this manuscript, Tocchini and Mango directly approach this deficit in a thoughtful study where many of the salient features of MS2 epitope tagging are systematically measured. Specifically, the authors use CRISPR genomic engineering to tag two separate dosage-sensitive, developmental genes and study the expression and function of these genes within the context of the MS2/MCP-GFP system. The authors demonstrate that the location of the MS2 epitope insertion within the endogenous 3'UTR is an important design consideration for functional, downstream implementation of the imaging system. In both cases, insertion of the MS2 hairpins near the end of the open reading frame of either gene results in overt and specific developmental phenotypes that phenocopy previously characterized loss of function alleles of each gene. The design of these experiments is high in quality in that they measure both the levels of cytoplasmic abundance of the various epitope-tagged mRNAs as well as the protein expression levels for these transgenes (by monitoring the levels of GFP expression (each MS2-tagged gene encodes a functional GFP-tagged allele). In two clear transgene examples, they demonstrate that the loss of function phenotypes of the proximally-tagged (closest to the ORF) transgenes disrupt mRNA levels and expression and reduce the proper localization of these mRNAs. This may be why previous attempts at implementing this important imaging system have failed.

      The authors then characterize the cellular systems that cause the differential expression of MS2-tagged transgenes. The authors note that previous studies on simpler systems and in C. elegans have suggested that the nonsense-mediated mRNA decay (NMD) pathway limits the expression of mRNAs with exceptionally long 3'UTRs. Tocchini and Mango then use C. elegans NMD mutants to demonstrate that ablation of this natural RNA degradation system corrects the developmental and gene expression defects associated with the reduction of function MS2 insertion alleles. These experiments are complete and compelling as they are validated at all levels (GFP expression (via quantification of GFP expression) and mRNA expression, and mRNA localization levels (via in situ hybridization).

      The authors then make the case that the type and expression levels of the MCP-GFP fusion protein are also essential features that need to be optimized for an effective imaging system. The authors suggest that optimal visualization of endogenous genes requires the surprisingly low-level expression of the MCP-GFP fusion protein. The authors use a novel transgene that differs from the conventional system. Specifically, the Tocchini system employs a 2xMCP ORF fused to 2xmCherry ORFs fusion. This transgene lacks the NLS typically used to localize exported mRNAs in the cytoplasm and also encodes two MCPs that may or may not facilitate dimerization on the MS2 hairpins. They demonstrate that endogenous, epitope-tagged transgenes can be visualized in developing embryos and that tethering this 2xMCP fusion to the reporter transcript does not alter RNA expression levels. While the authors demonstrate that visualization is possible with this system, it is hard to tell if this fusion protein dramatically improves over other available systems without a direct comparison. For instance, measuring the signal-to-noise (S/N) ratio of localized 2xMCP-2xmCherry would be a good addition and support the author's claims. If it were an exceptional system, these calculations should exceed the well-characterized and quantified MCP-GFP system described in Lee et al. 2019 ((Lee et al., 2019). It is just too hard to know if this is a dramatic element that should now be included in future RNA localization experiments.

      Minor critiques:

      1. The authors should provide more details in the experimental description of the MS2-tagged alleles (or in the figure images). It needs to be clarified in the main text how many MS2 hairpins there are, though this can be found in the materials and methods. In addition, it would be nice to know if these were any of the variations of MS2 hairpins that have already been optimized in some other way to increase or decrease structure or RNA metabolism defects in other systems. Specifically, are these hairpins the newest versions, V6 or V7, described in manuscripts from the Singer laboratory (e.g., (Tutucci et al., 2018))? For aficionados of this imaging system, it would be important to qualify each of the potential new features that make the results in this manuscript so clear and important.
      2. For people that are colorblind (or have reduced ability to distinguish some colors from others (like me, a reviewer)), it would be nice to have the MS2 illustrations in Figures 1A and B not have that color within the black, normal UTR. It's picky, but I had to ask someone what color that was.

      References:

      Lee, C., Shin, H., and Kimble, J. (2019). Dynamics of Notch-Dependent Transcriptional Bursting in Its Native Context. Dev Cell 50, 426-435 e424.

      Tutucci, E., Vera, M., Biswas, J., Garcia, J., Parker, R., and Singer, R.H. (2018). An improved MS2 system for accurate reporting of the mRNA life cycle. Nat Methods 15, 81-89.

      Significance

      In summary, this is a well-written and critical addition to the literature that will hopefully increase the implementation of this system in C. elegans research. The systematic approach to getting a new experimental platform up and running certainly has a place in the canon. Aside from the missing elements regarding the putative improvements and/or direct comparisons between different MCP fusion proteins, the manuscript is solid, important, and nearly ready to go.

      It is an advance and will, as noted above, likely serve to help implement this system by other C. elegans reserachers.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      The transparency of C. elegans invites us to push the limit of live imaging. In this context, observation of endogenous mRNA using the high affinity between MS2 RNA hairpins (from a bacteriophage) and a protein (MCP) that can be fluorescently labeled. In the time of CRISPR-Cas, the editing of endogenous genes is feasible and authors use it to insert 24 MS sequences (about 650 bp in total) at the UTRs of a couple of genes. Once they found a way to insert the MS2 sequences at the UTRs, although with phenotypic consequences that are solved in mutants defective in Non-mediated decay of transcript, they tune the expression of the MCP using a heat-shock promoter with a leaking expression at 25C and its location in the cytoplasm avoiding nuclear location signal (NLS) of the protein.

      Major comments:

      They present solutions for live imaging endogenous mRNA that would be useful for colleagues interested in this technique but also show experimental noises that would be specific to each gene/transcript/UTR. In the end, the best value of this technique is to observe "real" or physiological levels but to reach this point they need to use a mutant background (NMD mutants), which may alter the "real" scenario. They found a smart way to title the article using "An adapted..." but it would be more realistic/honest to mention in the title that this is happening only in NMD mutant backgrounds. I also have doubt ion the use of the acronym MS2-MCP in the title. What about something like "Visualization of endogenous cytoplasmic mRNA with live imaging in C. elegans embryos requires an inactive Non-mediated decay"?

      In any case, the conclusions are convincing and the experiments are well executed. I do not find the need for any essential experiments if they are clearer through the manuscript (from title and abstract to discussion) that this technique may need to be optimized and (and maybe validated with FISH) for each specific transcript, and developmental stage cell type where NMD and polyadenylation may work differently. Another source of experimental noise may come from the use of mcherry, which is known for forming aggregates in some cells/stages.(would this artefactual aggregation occur in figure 1L?)<br /> The only experiment that I missed, not essential but easy to perform, is a better description of the slight developmental delay of dlg-1 MS2 v2 animals. Size measures? Time until they lay embryos?

      In this sense, although is not the main purpose of the article, they could highlight the fact that this is an additional option to produce hippomorphic alleles of essential genes.

      Regarding methods, I miss information about the CRISPR-Cas efficiency of inserting the MS2 sequences at the UTRs. Sizes are "small" and can facilitate the insertion of dsDNA repair template, but it would be useful to know what efficiency would be expected. It would be good to mention somewhere how frequent are GG PAM sequences at UTRs sequences (probably less common than in other regions). In this sense, the use of minimal PAM Cas9 variants (Vicencio et al, Nat Comm 2022) would be necessary.

      Minor comments:

      In the abstract, line 33, remove epithelial? I do not think this is relevant in this sentence.<br /> In figure 4, panels B and C, add the two different embryonic stages on the left side. Then, it wouldn't be necessary to read the legend to understand the figure.

      Referees cross-commenting

      I find useful and reasonable the comments of my colleagues

      Significance

      I work in C. elegans on diverse topics, with an interest in RNA, and I have used FISH in C. elegans in the past. I find this study useful to expand the C. elegans toolbox in C. elegans. This manuscript provides solutions to technical limitations to observe mRNA in vivo in the cytoplasm, but also alerts of experimental noise that can be found and can be specific for each gene/transcript.

      It is focused on the C. elegans embryo, which is the system of interest for the authors. I miss a bit of discussion at least on the use of this methodology in other stages. One of the interesting aspects of observing mRNA in vivo is the capacity to manipulate the environment. Such capacity is very limited in embryos but feasible in larvae or adults with the use of microfluidics.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary:

      The authors of this manuscript have adapted the MS2-MCP system to visualize endogenous cytoplasmic mRNA with live imaging in Caenorhabditis elegans. They have identified some of the issues that might have prevented the MS2-MCP system's adaptation to C. elegans. Specifically, they have identified that the length of the 3'UTR, which is significantly increased upon the insertion of the MS2 sequences, impacts the mRNAs' stability. They have also shown that removing the nonsense-mediated decay pathway can prevent the destabilization of the MS2 transcripts. Moreover, they have also optimized the MCP expression to avoid nuclear retention of the MS2 transcripts and mislocalization of the mRNAs.

      Major comments:

      • The authors show that the insertion of 24xMS2 in two endogenous genes, spc-1 and dlg-1, causes some phenotypes such as slow growth, lack of coordination (Unc), small body size (Sma), and reduced brood sizes. However, only an image example is provided in Fig. 1 C, D, and quantifying all these phenotypes would be nice. Same in Fig. 2C, D.
      • Similarly, the reduction in mRNA spots from smFISH in Fig. 1 G-L is difficult to visualize by eyes, and proper smFISH quantification will help interpret the results.
      • The authors also claim a reduction in cytoplasmic RNAs and increased signal in nuclear RNAs in Fig. 1J, L. A proper quantification of nuclear and cytoplasmic smFISH will help interpret the results.
      • In Fig. 3D-F, the authors quantified the signal of nuclear smFISH. However, it is unclear to me in what samples or conditions the statistical test is performed. For example, do the three stars in Fig. 3D refer to the significant decrease or increase of NMD strain compared to the WT? What about the stars in Fig. 3E? The authors should indicate what samples they compare in the statistical test.

      Minor comments:

      • on line 195, the authors reference Fig. 3A. However, it should be Fig. 4A.
      • In Fig. 4B, C the authors can add close to the image of the embryos the developmental stages. This will help the reader identify the embryo's developmental stage in the figure's upper and lower parts.
      • The authors can expand a bit the discussion on how their method differs (advantages and disadvantages) from the MASS system by Hu et al., 2023.

      Significance

      This manuscript will help the C. elegans community to adapt and use the MS2-MCP system to visualize endogenous mRNAs by live imaging. Their finding could also be adapted to other animal model systems. At the moment, only one published report has described the usage of the MS2-MCP system in C. elegans (by Hu et al., 2023), which combined the MS2 and Suntag systems. In this way, Hu et al., 2023 could shorten the length of MS2 insertion. I am unsure if this is why they do not observe any impact on endogenous mRNA tagged with MS2. However, they only track one gene, and it is possible that different 3'UTR will react differently to the insertion of MS2 repeats. Another manuscript (Kinney et al., BioRxiv 2023) showed the usage of the MS2-MCP-GFP system to track miRNA transcription. In this case, the insertion of the MS2 repeats in the transgenic lin-4 miRNA precursor rescued lin-4 mutation. In this manuscript, Tocchini and Mango identified possible issues in inserting MS2 repeats in endogenous 3'UTR. They have overcome this potential issue by changing the position of the insert in the 3'UTR and by removing the nonsense-mediated decay pathway to prevent destabilization of the mRNA-MS2 transcripts. One possible limitation is that possible system users need to work in a mutant background for the nonsense-mediated decay pathway, which is not ideal. However, it provides a greatly improved tool to track mRNA by live imaging. Therefore, their improved methodology will certainly contribute to expanding the use of MS2-MCP system in C. elegans.

      I have expertise in C. elegans biology and transcription, but I do not have expertise in the imaging system, and therefore, I cannot fully judge the methodology they have used and the quality of the imaging system.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1:

      In this paper, authors report that radiation, acidic pH, hypoxia, and drugs that interfere with lipid synthesis, all of which affect lipid droplets (LD), also affect the production of small extracellular vesicles (sEVs). In addition, they also report that LD content and sEV secretion are also modulated in CR-CSCs. Authors conclude that sEV formation and secretion is directly linked to LDs, and that their studies may open the way to new clinical perspectives. However, some important issues need to be addressed before the paper can be considered for publication.

      My main concern is that the notion that LDs and sEVs are linked remains vague. Do cells contain more LDs and secrete more sEVs because these two pathways are selectively up-regulated via some mechanisms that controls both pathways in a concerted manner? Or do cells with more LDs and more sEVs also contain more of everything, perhaps as a result of metabolic activity?

      We appreciate the Reviewer's observations. Indeed, this comment represents the main pillar of the entire manuscript. We have attempted to uncover the molecular mechanism behind this novel and intriguing organelle connection. First of all, we have adapted the manuscript emphasizing that the LD – sEV connection might be direct or indirect. Our omic data suggested that some proteins belonging to the RAB family, mainly Rab18, Rab7a and Rab5c, could play a pivotal role in the LDs-sEVs axis. To strengthen those results, we have performed additional experiments by silencing the expression of the three candidate Rabs. Rab5c seems to be a good candidate to modulate the LD-sEV connection. We believe that Rab5c is not the only contributor to the LD-sEV connection but is part of a whole set of different elements that regulate this axis. However, it is quite challenging to rule out other molecular candidates as co-contributors to this phenomenon, especially when considering cellular metabolic pathways.

      We recognize that external stimuli, such as radiation, pH, and lipid-interfering drugs, may exert their effects on other cellular organelles, even though we have strived to analyze each individual phenomenon rigorously. We are confident that our work lays the foundation for further research in the field.

      A direct corollary of this issue is whether increased sEV secretion reflects more endosomes and lysosomes (e.g. LysoTracker-positive compartments) or whether sEV secretion is selectively up-regulated.

      Thanks to the Reviewer’ suggestion, we have analyzed both the lysosome and endosome contents in our experimental cell systems. These data are now included in the manuscript in Figure S8. We have observed that it is unlikely that lysosomes are directly involved in the LD – sEV connection. However, the expression of Rab7a, a regulator of the late endosomal pathway, correlated with the LD content of the cells and their sEV release. Therefore, the endosomal pathway might be a good candidate to contribute to this LD – sEV connection.

      At one point, authors argue that cells that secrete more sEVs also contain more MVBs, but this issue remains elusive. To what extent is the increase in LDs and sEVS correlated in particular with an increase in endosome-lysosomes, and ER-Golgi (LDs originate from the ER)?

      We thank the Reviewer for this comment. We agree that the analyses of sEVs secreted in the media might not reflect the MVB content in the cells. However, two experiments, one on Panc01 cells and another one on MCF7 cells, showed that the number of MVBs, assessed by confocal microscopy using CD63 staining (MCF7) or CD63 and Alix plasmids (PANC-01), was directly correlated with the number of released sEVs in the media (Figure Fig S3C and 4J).

      In addition, we included additional experiments assessing the lysosome content in HT29 LDHigh and LDLowcells. Hereby, we confirmed that HT29 LDHigh cells showed a higher LD content than HT29 LDLow cells. Inversely, by studying the lysotracker area per cell, we showed that HT29 LDLow population has a higher lysosomal content as compared to their counterpart, HT29 LDHigh cells (test = Wilcoxon rank sum test with continuity correction_ W = 85127, p-value = 7.255e-07 for LDs and W = 49321, p-value = 1.14e-11 for Lysotracker). However, we could not demonstrate a clear correlation between the number of LDs in the cell and the lysotracker signal.

      Finally, we have also studied the expression of GM130, a Golgi-shaping protein (Ref. 1) and Rab7, a late-endocytic protein (Fig S8C). While the expression of Rab7 (endosome) seemed to correlate with the LD and sEV contents, the expression of GM130 (Golgi) gave back no coherent results. Indeed, it was inversely correlated to the LD and sEV amount, in accordance with what was already reported elsewhere (Ref 2 and 3)

      • Nakamura N. Emerging new roles of GM130, a cis-Golgi matrix protein, in higher order cell functions. J Pharmacol Sci. (2010) 112:255–64. Doi: 10.1254/jphs.09R03CR
      • Lydia-Ann L.S. Harris, James R. Skinner, Trevor M. Shew, Nada A. Abumrad, Nathan E. Wolins. _Monoacylglycerol disrupts Golgi structure and perilipin 2 association with lipid droplets.___Doi.org/10.1101/2021.07.09.451829
      • Alvin Kamili, Nuruliza Roslan, Sarah Frost, Laurence C. Cantrill, Dongwei Wang, Austin Della-Franca, Robert K. Bright, Guy E. Groblewski, Beate K. Straub, Andrew J. Hoy, Yuyan Chen, Jennifer A. Byrne; TPD52 expression increases neutral lipid storage within cultured cells. J Cell Sci 1 September 2015; 128 (17): 3223–3238. Doi: 10.1242/jcs.167692

      Authors conclude that the data with lipid inhibitors strengthen the connection between LDs and sEVs (Fig 2 and S2). However, is this regulation selective, or does it merely reflect the general effect of these inhibitors on membrane-related processes? The same comment applies to the role of iron metabolism after knockdown of ferritin heavy chain (Fig 3 and S3), acidic pH and X-ray radiation (Fig 4 and S4).

      We thank the Reviewer for the interesting observation. As previously mentioned, we cannot rule out other potential contributors to the LDs-sEVs connection upon lipid inhibitor treatments and/or the others external stimuli applied to our cell systems.

      The data presented in this manuscript merely represent a novel and unexplored (at least so far) organelle connection, direct or indirect, with a broad clinical implication. As the membrane-related processes (such as Endosomes, Golgi apparatus, Exosome (sEV) pathway, Lysosomes and Autophagosome) are all interconnected, in our opinion, it might be quite challenging to make such a definitive statement.

      Such assertion would require extensive further investigation to relate each organelle to the LDs and/or sEVs. However, with our research, we hope to open the door to a new era of investigations regarding the sEV – LDs connection.

      OTHER COMMENTS

      1) Which cell line is used for sEV analysis (markers vs contaminants (Fig S1B)? In any case, the data should be shown for both cell types.

      Our method to isolate sEVs is a standardized method that was already published by our group and collaborators in 2020 (M. Bordas, et al., Optimized Protocol for Isolation of Small Extracellular Vesicles from Human and Murine Lymphoid Tissues. Int J Mol Sci (2020) https:/doi.org/10.3390/ijms21155586.). This protocol was validated on human and mouse tissues, much more complex samples than cell culture supernatant.

      Figure S1C was modified, as requested by the Reviewer, including new data for HT29, Panc01 and MCF7 cell lines to broaden the panel. Those results confirmed the good purity of sEV samples isolated from cell culture supernatant.

      2) The Tsg101 blot is not impressive (Fig S1B): the difference between cells and sEVs is not easy to see. It would be nice if blots were quantified.

      Indeed, the signal obtained for TSG101 for sEVs derived from Panc01 cell line is quite weak. It is important to remember that not all sEV markers are highly expressed in all cell lines and their derived sEVs. Some cell line-derived sEVs show a low or high expression of the diverse sEV markers. To answer the Reviewer #1’s comment, we quantified the expression of TSG101 in Panc01-derived sEVs. The quantification showed that TSG101 is 6.8 times more expressed on Panc01-dervied sEVs as compared to the cell line. However, since the expression is quite low, this quantification should be taken with some caution.

      In light of the Reviewer ‘comment, we have performed the Western Blot analysis on other cell lines (HT29 and MCF7), and we have replaced TSG101 marker with CD9 marker (Figure S1C).

      3) From Fig 1B it cannot be concluded that the size of sEVs ranges from 30 to 200nm: the micrograph only shows a few structures.

      We appreciate the Reviewer's comment and have attempted to provide more clarity. Firstly, we want to highlight that TEM micrographs of sEVs typically show the donut shape, a unique feature of sEVs imaged with TEM, as well as a size range. In Figure 1B micrograph, the sEV size is approximately 100 nm. The size distribution of LoVo and HT29-derived sEVs can be observed from the NTA size measurements in Figure S1B. Indeed, the peak size is 148 nm for LoVo-derived sEVs and 135 nm for HT29, which aligns with the sEV sizes presented in Figure 1B. We have also included multiple micrographs here under. As the number of Supplementary Figures is already large, we have decided to not include those micrographs in the manuscript. The average size of LoVo-derived sEVs, based on TEM micrograph analysis, was 94 ± 41.10 nm, while the average size of HT29-derived sEVs was 76.41 ± 44.22 nm. The size discrepancy between the two methods (NTA versus TEM) can be ascribed to the dehydration step required for TEM, which results in a reduction of the actual sEV size.

      4) HT29 cells contain far more LDs than LoVo cells (Fig 1A). Similarly, sEV proteins (CD63, CD81, CD9, Hsc-70) are more abundant in HT29 sEVs than in LoVo sEVs (Fig 1D). However, the sEV preparation from HT29 cells contains only approx. 50% more total protein than LoVo sEVs (Fig S1D-E). Are sEVs prepared from LoVo cells far more contaminated with cell debris etc.. than sEV fractions from HT29 cells?

      We are confident that our EV isolation method allows us to achieve high yield and excellent purity. It is possible that a lower number of sEVs in samples may lead to increased protein contamination during ultracentrifugation. However, size exclusion chromatography should minimize this protein contamination. It is important to note that the NTA method is significantly more sensitive and accurate than Qubit protein quantification. Consequently, protein concentration and particle concentration should not be directly compared.

      5) LD staining should be shown for the corresponding populations of cells with high/low CD63 (Fig 1E). Cells in culture can be somewhat heterogeneous, but the difference between low and high CD63 is quite extreme (Fig 1E). Is such high heterogeneity also observed with other proteins of the endocytic and biosynthetic pathways? Authors conclude that cells containing high CD63 levels also contain more MVBs (Fig 1E): are all late endosomal proteins (e.g. LAMP1, RAB7) upregulated in cells with high CD63?

      We thank the Reviewer for this comment, and we totally agree with the Reviewer that it would be better to have the LD and CD63 staining on the same images. Unfortunately, the staining for CD63 on LD540-sorted HT29 cells requires a permeabilization step that interferes with the cellular lipid part and could therefore negatively affect the LD imaging by confocal microscopy. To prove that the HT29 LDHigh and HT29 LDLowcontain high and low LD amount respectively, we sorted HT29 cells based on the LD content and, soon after, we observed them at the confocal microscopy. We thus added new images in Figure S1F, corresponding to the LD fluorescence detection. The readers will also appreciate the explanation regarding the inability of observing both LDs and CD63 staining on the same confocal images under the line 165 – 166:

      As the staining for CD63 required a permeabilization step, and therefore lipid digestion, it was not possible to assess both LDs and CD+MVBs on the same micrographs “.

      In addition, we have added confocal images representing HT29 cells sorted based on their LD content and stained with Hoechst and Lysotracker. A quantification of the Lysotracker fluorescence per cell and the correlation with the number of LDs can also be appreciated in Figure S8A-B.

      Finally, we performed Western Blot analysis to examine Rab7a expression under various conditions described in our manuscript (Figure S8C). In general, Rab7 expression corresponded with LD content, indicating that cells with high LD content exhibited higher Rab7 expression, while cells with low LD amount showed lower Rab7 expression, except for Triacsin-C. The Reviewer can now appreciate the quantification in the graphs provided below (not included in the manuscript).

      Regarding the heterogeneity of LDs, CD63+MVBs, or lysotracker among the cell population, we have indeed noticed heterogeneity observable in these three types of staining in HT29, particularly in the HT29 LDHighpopulation.

      6) Inhibitors of lipid synthesis reduce LD formation (Fig 2B), sEV production and CD63 / CD81/ CD9 secretion (Fig2C-D, Fig S2B). Are the cellular levels of these (and other endosomal) proteins also reduced after inhibitor treatment? Does the stimulation of LD formation with oleic acid also stimulate CD63 synthesis and sEV production?

      We thank the Reviewer for this very interesting comment. To answer this question, we have added a supplementary figure (Figure S2A, S2B) showing the cellular expression of CD63 upon LD inhibition or stimulation.

      During the planning of our experiments, we discussed about the possibility of using oleic acid to induce the formation of Lipid Droplets, which was ultimately not done. This is because the use of oleic acid would have more strongly stimulated the triglyceride pathway, as extensively discussed elsewhere (Mejhert N. et al., The lipid droplet knowledge portal: a resource for systematic analyses of lipid droplet biology, Developmental Cell, 2022). Since Lipid Droplets are made by cholesterol esters and triglycerides, we preferred to use other stimuli (hypoxia, radiation), all of them already discussed in literature, to induce both pathways simultaneously, resulting in the Lipid Droplet formation/induction.

      7) It seems that pH and irradiation increase sEV markers far more significantly (Fig 4 B-C and Fig S4A-E) than FTH1 depletion decreases sEV markers (Fig 3 D-E). In fact, authors mention that they cannot exclude a contamination of sEVs with small apoptotic / autophagic vesicles after irradiation (Fig 4). To facilitate comparison, it would be nice to also show the number of secreted particles per cell (like after FTH1 depletion Fig 3D), as well as the distribution of possible contaminants (e.g. Fig S1). Also, authors state that the increase in the number CD63+ MVBs after irradiation is shown, but this is not the case.

      We apologize to the Reviewer because, in fact, one figure was missing (Figure 4). We have rectified this by increasing the quality of Figure 4 and have added representative images for each acquisition of the number of MVBs, either positive for CD63 or Alix, in transfected Panc01 cells X-ray irradiated (8 Gy) or not (0Gy). In addition, a similar experiment was performed in MCF7 cells transduced with shRNA or shFTH1. CD63+ MVBs were assessed in both cell line and the number of CD63+ puncta (MVBs) were quantified by ImageJ. The results, although not significative, illustrated a trend for MCF7 shFTH1 to contain less CD63+ MVBs than MCF7 shRNA. Furthermore, the quantification of sEVs released in the conditioned media was performed in three independent experiments and demonstrated that significantly less particles (sEVs) were released by MCF7 shFTH1 than MCF7 shRNA.

      8) Are the proteomic data (Fig 6) with LDlow and LDhigh cells obtained after cell sorting, as in Fig 1E? Did authors compare the proteome of LoVo and HT29 sEVs? How do the protein profiles (in particular proteins involved in lipid metabolism) obtained under different conditions compare with each other, in particular after irradiation (Fig4N) and knockdown of ferritin heavy chain (Fig 3, Fig S3)? It would also be interesting to compare these data with the data obtained in CR-CSCs culture under hypoxia (Fig S5). Are common proteins involved in sEV production and LD biosynthesis identified in the analysis of these biological processes? Is there a common set of proteins/genes revealed by this analysis, which may potentially control sEV production and LD biosynthesis?

      We thank the reviewer for this interesting comment.

      Proteomic analyses have been performed on the following conditions:

      • Panc01 (0 Gy – 6 Gy – 8 Gy) for sEV samples
      • MCF7 (shFTH1 and MCF7 shRNA)
      • MCF7 (0 Gy and 6 Gy)
      • MCF7 (Normoxia and Hypoxia)
      • H460 (0 Gy and 6 Gy)
      • H460 (Normoxia and Hypoxia)

      RNA sequencing was performed on the following conditions:

      • CR-CSCs (#4, #8, #21)

      Based on all those data, we have analyzed the sEV pathway and how this pathway was modulated in the conditions with high LD content and low LD content. We therefore came up with several proteins, presented in Figure S7. Based on this analysis, we have decided to further investigate the role of RAB18, RAB5c and RAB7a in the connection between LDs and sEVs. Those additional results can be found in Figure 6 and Figure S7A (originally Figure 6). We have found that RAB5c, but not RAB7a or RAB18, seems to be a good candidate to intervene in the LD – sEV connection.

      Minor comments

      1) Some parts of the text are still a bit rough, and should be read and corrected carefully. For example: i) isn't it obvious that a common source of lipids builds up the membrane of sEVS, much like any other membrane (line 90, p.2); ii) what does this sentence mean: "LD have been considered as mere fat storage organelles for a long time, although important evidence could be traced back to the early 1960's". Important evidence for what? iii) why is the acronym AdExo used? iv) (line 138) the text should probably be "sEVs released during 72h were studied" and not "released sEVs were studied ... 72 h after seeding".

      We apologize to the Reviewer if some parts of the paper were a bit rough. We have re-read the entire manuscript and corrected all the parts that needed revision work.

      2) The captions are far too small in most figures and diagrams (for example X and Y axis in Fig 1C-D, text in Fig 1E; Fig S1; Fig 3C proteins in the heatmap).

      We agree with the Reviewer. All images and their captions were properly revised.

      3) The color code for LoVO and HT29 cells is reversed in Fig S1D-E

      The mistake was corrected.

      4) In Fig 1D, I cannot see CD81 in the LoVo blot.

      In the image below, it is possible to see the LoVo blot.

      5) Wording is not adequate in following sentences: "62.7% of proteins related to the exosomal pathway are downregulated in MCF7 shFTH1 cells" (line 233) and a few lines below: ".. the expression of almost all exosomal markers was downregulated in MCF7 shFTH1 cells" (line 239). Does 62.7% represent all proteins?

      We apologize to the reviewer for the mistake. We rephrased this sentence.

      6) In Fig 3E authors compare sEV markers secreted by cells treated with shFTH1 or control shRNA. The Anx5 and CD63 blots are not very convincing (quantification would be helpful).

      We apologize to the Reviewer for this issue. These Western Blot analyses were performed only once, therefore a quantification in the manuscript would not be relevant. However, we report here the results of the quantification. The expression of Annexin V was 1.58 times higher in MCF7 shRNA than MCF7 shFTH1, while the expression of CD63 was 1.34 time higher in MCF shRNA as compared to MCF7 shFTH1.

      7) The micrographs in Fig 4L are too small: gold particles cannot be seen, even in the high magnification views.

      We thank the Reviewer for her/his comment. We have moved the micrograph and the quantification histogram to the Figure S6. Now, it is possible to discriminate easily gold nanoparticles.

      8) The micrographs showing ALIX and CD63 (Fig 4J) in irradiated and unirradiated Panc01 cells should be shown for comparison.

      We followed the Reviewer’ suggestion as it is possible to note in the Figure below.

      Reviewer #2:

      This manuscript describes a relationship between lipid droplet presence in cells and small EV secretion. First, correlations are done between number of lipid droplets and numbers of EVs secreted. Then chemical inhibitors of lipid droplet biosynthesis pathways were shown to reduce small EV secretion. Then various processes known to target lipid droplets, including iron metabolism, irradiation, hypoxia, low pH are used to show concordant effects on lipid droplets and small EV secretion. Proteomic analysis of EVs and cells subjected to some of the treatments are also performed. Overall, it is an interesting line of investigation and the data overall seem solid. Several flaws exist, which can probably be fixed. These include the use of different cell lines for different experiments. It makes it a bit difficult to connect everything together. It could be fixed by adding some extra cell lines to some experiments - for example taking the MCF7 and Panc-01 cells for which proteomics was performed and redoing some of the correlative and causative experiments from Figs 1 and 2.

      We appreciate the Reviewer's insightful observation. Following her/his suggestion, we have conducted additional experiments on MCF7, H460 and PANC-01 cell lines to enhance data consistency and facilitate a smoother transition between different sections of the paper.

      It also would be good to have some more direct evidence of the connection between lipid droplets and EV secretion - one could argue that this was already done in Fig 2 with the chemical inhibitors, I wonder if there is a genetic way to do it too?

      We totally agree with the Reviewer. Indeed, starting from our proteomic data we highlighted some genes belonging to the RAB family as potential candidates to interfere with the LD – sEV connection. The Reviewer can now appreciate in Figure 6 and Figure S7, the results from the additional experiments we carried out on RAB5c, RAB7a and RAB18 silencing in HT29 cells. The former Figure 6 has been moved in the Supplementary part (Figure S7).

      Some tightening up of the writing (especially the Discussion) and the resolution of the figures would also improve the manuscript.

      We apologize to the Reviewer for this issue. We have now re-prepared all Figures by increasing their resolution, as well as reviewing the entire manuscript with the aim of making the reading smoother and simpler.

      Overall, it is a nice piece of work but there are many minor things to be fixed.<br /> <br /> Specific Comments:

      The sentence in the Introduction: "The non-endosomal pathway generates sEVs devoid of<br /> CD63, CD81 and CD9 or sEVs enriched in ECM and serum-derived factors (7)." is not well-supported and should be removed. The idea that you can classify membrane of origin based on markers has not been proven, but rather assumed.

      We agree with the Reviewer. We have rephrased the sentence.

      We thank the Reviewer for this comment. In response to this, we have generated correlation graphs for several of our experiments:

      • HT29 (CTL – Triacsin-C - PF-06424439) in Figure 2E
      • PANC-01 (CTL – 2 – 4 – 6 – 8 Gy) in Figure 4K
      • CR-CSCs (#4, #8, #21) in Figure 5E

      The Method used for EV purification should be stated in the Results rather than referring to a reference and a Supplemental Figure (S1A) that is too low of a resolution to see.

      Our method to isolate sEVs is a standardized methods that was already published by our group and collaborators in 2020 (M. Bordas, et al., Optimized Protocol for Isolation of Small Extracellular Vesicles from Human and Murine Lymphoid Tissues. Int J Mol Sci (2020) https:/doi.org/10.3390/ijms21155586.). This protocol was validated on human and mouse tissues, much more complex samples than cell culture supernatant.

      In regard to the Reviewer’s comment, we have added a better description of the protocol in the Results part, referring to the Material and Method. For this reason, we decided to keep the sEV protocol in the SI section. We apologize for the low quality of the Figure S1. In agreement with the Reviewer suggestion, we have modified the image by increasing its quality.

      Fig 1B would be better to have an image in which the EVs are not aggregated.

      We thank the Reviewer for this comment and have modified the Figure accordingly.

      Fig 3 is interesting but jumps cell lines. For better continuity, some of the experiments from Figs 1 and 2 should be repeated in the MCF7 cells to connect with the proteomics.

      In agreement with the Reviewer’ comment, we decided to perform additional experiment on MCF7, using Triacsin-C. The Reviewer can now appreciate the results in Figure 2F, Figure 2G and Figure S2E.

      Fig 3C is too low resolution to read, please export at higher resolution.

      We are sorry for the low-quality Figure. We have modified the image accordingly.

      Please provide all the raw proteomics data as a supplementary spreadsheet.

      We have provided all the raw data regarding our proteomic analyses.

      Fig 4 panels are low resolution

      We apologize for the low-resolution Figure. We have modified the figure by increasing the quality.

      Fig 4 again adds new cell lines with H460 and Panc-01

      We thank the reviewer for this comment. In this regard, we have performed additional experiment:

      • Western Blot: comparison cellular and exosomal markers (Figure S1C)
      • MCF7 (CTL - Triacsin) (Figure 2F, Figure 2G and Figure S2E)
      • Western Blot: analysis of RAB7a, GM130

      The images corresponding to 4J should be shown in a Supp Figure somewhere

      We thank the reviewer for pointing out this oversight. We have added the confocal images corresponding to the Figure 4J below the quantification.

      The statements: "In addition, the exosomal nature of Panc01-derived vesicles was demonstrated by an analysis of CD63+ or Alix+ multivesicular bodies (MVBs) in unirradiated (0 Gy) or irradiated (8 Gy) pancreatic cancer cells (Fig 4J). Moreover, we confirmed a clear correlation between cellular LD content and sEV biogenesis, as represented in Fig 4K." are overly conclusive. For 4J, one can make a statement about the MVBs but not the EVs as that's not what was measured there. Likewise for 4K, what was measured was how many EVs were released not how many were formed. While the data are suggestive of alteration of exosome biogenesis, they are not conclusive.

      We agree with the reviewer and have performed the necessary changes in the manuscript. The reviewer can see the changes under the lines 282 – 284:

      “In addition, the analysis of CD63+ or Alix+ multivesicular bodies (MVBs) in unirradiated (0 Gy) or irradiated (8 Gy) pancreatic cancer cells revealed an increased number of MVBs after irradiation (Figure 4J).”

      Western blot is always capitalized by convention - Western not western.

      We have corrected it accordingly.

      Fig 5A is too small and low resolution - suggest eliminating and just put info in methods.

      We are sorry for the low-resolution image. We have followed the Reviewer suggestion. The graphical method has been now moved to the Supplementary Figure S6.

      Fig 5G, many of the genes shown are frequently EV cargoes but most not involved in exosome biogenesis - not sure where the label of Exosome pathway came from but it is not very compelling. Only ANXA2, Arf6, and Rab5C seem related and they are barely elevated.

      We completely agree with the Reviewer's comment. As a result, we have revised the heatmap title to "Exosomal Cargoes and Pathways" instead of "Exosomal Pathway".

      Most main figures and all supplementary figures are extremely low res - please fix.

      We are very sorry for the low-quality figures. We have revised all Figures (main text and SI) by increasing their quality.

      Fig 6 is first mentioned in the Discussion - it should be described in the Results before that (or alternatively removed).

      We agree with the Reviewer. Our initial idea was to mention perspectives of analyses that could be carried ulteriorly. Nevertheless, we have performed additional experiments in order to get insight on the mechanism involved in the LD – sEV connection. Indeed, based on our proteomic data, we have analyzed the sEV pathway and how this pathway was modulated in the conditions with high LD content and low LD content. We therefore came up with several proteins, presented in Figure S7A (originally Figure 6). Based on this analysis, we have decided to further investigate the role of RAB18, RAB5c and RAB7a in the connection between LDs and sEVs. Those additional results can be found in Figure 6 and Figure S7 in the Results section. We have found that RAB5c, but not RAB7a or RAB18, seems to be a good candidate to intervene in the LD – sEV connection.

      Table S1, also first mentioned in the Discussion, is missing. Either describe in the Results section or remove the callout to it.

      Our apologies for that. The Table S1 has been now mentioned in the Results section and has been properly uploaded.

      The discussion is too dense with too many trains of thought, often many different directions in the same paragraph. It needs to be streamlined, with a central thought for each paragraph and good transitions between the paragraphs.

      We apologize to the Reviewer if the Discussion part was a bit confusing. We rewrote the paragraph, streamlining it and making the transitions between its paragraphs smoother.

      Reviewer #2 (Significance (Required)):

      <br /> Strengths of this manuscript are the interesting connection between lipid droplets and exosomes and the number of experiments to address it.

      <br /> Limitations: use of different cell lines for different figures, overall descriptive nature with regard to direct demonstration of connection to lipid droplets -- it's kind of done in Fig 2, but could be possibly bolstered.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      This manuscript describes a relationship between lipid droplet presence in cells and small EV secretion. First, correlations are done between number of lipid droplets and numbers of EVs secreted. Then chemical inhibitors of lipid droplet biosynthesis pathways were shown to reduce small EV secretion. Then various processes known to target lipid droplets, including iron metabolism, irradiation, hypoxia, low pH are used to show concordant effects on lipid droplets and small EV secretion. Proteomic analysis of EVs and cells subjected to some of the treatments are also performed. Overall, it is an interesting line of investigation and the data overall seem solid. Several flaws exist, which can probably be fixed. These include the use of different cell lines for different experiments. It makes it a bit difficult to connect everything together. It could be fixed by adding some extra cell lines to some experiments - for example taking the MCF7 and Panc-01 cells for which proteomics was performed and redoing some of the correlative and causative experiments from Figs 1 and 2. It also would be good to have some more direct evidence of the connection between lipid droplets and EV secretion - one could argue that this was already done in Fig 2 with the chemical inhibitors, I wonder if there is a genetic way to do it too? Some tightening up of the writing (especially the Discussion) and the resolution of the figures would also improve the manuscript. Overall, it is a nice piece of work but there are many minor things to be fixed.

      Specific Comments:

      The sentence in the Introduction: "The non-endosomal pathway generates sEVs devoid of<br /> CD63, CD81 and CD9 or sEVs enriched in ECM and serum-derived factors (7)." is not well-supported and should be removed. The idea that you can classify membrane of origin based on markers has not been proven, but rather assumed.

      Fig 1A and B - to better support the idea of a correlation between LD formation and EV release, more than two cell lines should be used and a linear correlation plot with R2 value shown. Likewise, it would be very interesting to see whether there is really a correlation between LD content and CD63-endosome positivity in a similar manner, given the results in Fig 1E. Also, it would be good to see LD and CD63 in the same cells for Fig 1E from the sorted populations.

      The Method used for EV purification should be stated in the Results rather than referring to a reference and a Supplemental Figure (S1A) that is too low of a resolution to see.

      Fig 1B would be better to have an image in which the EVs are not aggregated.

      Fig 3 is interesting but jumps cell lines. For better continuity, some of the experiments from Figs 1 and 2 should be repeated in the MCF7 cells to connect with the proteomics.

      Fig 3C is too low resolution to read, please export at higher resolution.

      Please provide all the raw proteomics data as a supplementary spreadsheet

      Fig 4 panels are low resolution

      Fig 4 again adds new cell lines with H460 and Panc-01

      The images corresponding to 4J should be shown in a Supp Figure somewhere

      The statements: "In addition, the exosomal nature of Panc01-derived vesicles was demonstrated by an analysis of CD63+ or Alix+ multivesicular bodies (MVBs) in unirradiated (0 Gy) or irradiated (8 Gy) pancreatic cancer cells (Fig 4J). Moreover, we confirmed a clear correlation between cellular LD content and sEV biogenesis, as represented in Fig 4K." are overly conclusive. For 4J, one can make a statement about the MVBs but not the EVs as that's not what was measured there. Likewise for 4K, what was measured was how many EVs were released not how many were formed. While the data are suggestive of alteration of exosome biogenesis, they are not conclusive.

      Western blot is always capitalized by convention - Western not western.

      Fig 5A is too small and low resolution - suggest eliminating and just put info in methods.

      Fig 5G, many of the genes shown are frequently EV cargoes but most not involved in exosome biogenesis - not sure where the label of Exosome pathway came from but it is not very compelling. Only ANXA2, Arf6, and Rab5C seem related and they are barely elevated.

      Most main figures and all supplementary figures are extremely low res - please fix.

      Fig 6 is first mentioned in the Discussion - it should be described in the Results before that (or alternatively removed).

      Table S1, also first mentioned in the Discussion, is missing. Either describe in the Results section or remove the callout to it.

      The discussion is too dense with too many trains of thought, often many different directions in the same paragraph. It needs to be streamlined, with a central thought for each paragraph and good transitions between the paragraphs.

      Significance

      Strengths of this manuscript are the interesting connection between lipid droplets and exosomes and the number of experiments to address it.

      Limitations: use of different cell lines for different figures, overall descriptive nature with regard to direct demonstration of connection to lipid droplets -- it's kind of done in Fig 2, but could be possibly bolstered.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      In this paper, authors report that radiation, acidic pH, hypoxia, and drugs that interfere with lipid synthesis, all of which affect lipid droplets (LD), also affect the production of small extracellular vesicles (sEVs). In addition, they also report that LD content and sEV secretion are also modulated in CR-CSCs. Authors conclude that sEV formation and secretion is directly linked to LDs, and that their studies may open the way to new clinical perspectives. However, some important issues need to be addressed before the paper can be considered for publication.

      My main concern is that the notion that LDs and sEVs are linked remains vague. Do cells contain more LDs and secrete more sEVs because these two pathways are selectively up-regulated via some mechanism that controls both pathways in a concerted manner? Or do cells with more LDs and more sEVs also contain more of everything, perhaps as a result of metabolic activity? A direct corollary of this issue is whether increased sEV secretion reflects more endosomes and lysosomes (e.g. LysoTracker-positive compartments) or whether sEV secretion is selectively up-regulated. At one point, authors argue that cells that secrete more sEVs also contain more MVBs, but this issue remains elusive. To what extent is the increase in LDs and sEVS correlated in particular with an increase in endosome-lysosomes, and ER-Golgi (LDs originate from the ER)? Authors conclude that the data with lipid inhibitors strengthen the connection between LDs and sEVs (Fig 2 and S2). However, is this regulation selective, or does it merely reflect the general effect of these inhibitors on membrane-related processes? The same comment applies to the role of iron metabolism after knockdown of ferritin heavy chain (Fig 3 and S3), acidic pH and X-ray radiation (Fig 4 and S4)

      Other comments

      1. Which cell line is used for sEV analysis (markers vs contaminants (Fig S1B)? In any case, the data should be shown for both cell types.
      2. The Tsg101 blot is not impressive (Fig S1B): the difference between cells and sEVs is not easy to see. It would be nice if blots were quantified.
      3. From Fig 1B it cannot be concluded that the size of sEVs ranges from 30 to 200nm: the micrograph only shows a few structures.
      4. HT29 cells contain far more LDs than LoVo cells (Fig 1A). Similarly, sEV proteins (CD63, CD81, CD9, Hsc-70) are more abundant in HT29 sEVS than in LoVo sEVs (Fig 1D). However, the sEV preparation from HT29 cells contains only approx. 50% more total protein than LoVo sEVs (Fig S1D-E). Are sEVs prepared from LoVo cells far more contaminated with cell debris etc.. than sEV fractions from HT29 cells?
      5. LD staining should be shown for the corresponding populations of cells with high/low CD63 (Fig 1E). Cells in culture can be somewhat heterogeneous, but the difference between low and high CD63 is quite extreme (Fig 1E). Is such high heterogeneity also observed with other proteins of the endocytic and biosynthetic pathways? Authors conclude that cells containing high CD63 levels also contain more MVBs (Fig 1E): are all late endosomal proteins (e.g. LAMP1, RAB7) upregulated in cells with high CD63?
      6. Inhibitors of lipid synthesis reduce LD formation (Fig 2B), sEV production and CD63 / CD81/ CD9 secretion (Fig2C-D, Fig S2B). Are the cellular levels of these (and other endosomal) proteins also reduced after inhibitor treatment? Does the stimulation of LD formation with oleic acid also stimulate CD63 synthesis and sEV production?
      7. It seems that pH and irradiation increase sEV markers far more significantly (Fig 4 B-C and Fig S4A-E) than FTH1 depletion decreases sEV markers (Fig 3 D-E). In fact, authors mention that they cannot exclude a contamination of sEVs with small apoptotic / autophagic vesicles after irradiation (Fig 4). To facilitate comparison, it would be nice to also show the number of secreted particles per cell (like after FTH1 depletion Fig 3D), as well as the distribution of possible contaminants (e.g. Fig S1). Also, authors state that the increase in the number CD63+ MVBs after irradiation is shown, but this is not the case.
      8. Are the proteomic data (Fig 6) with LDlow and LDhigh cells obtained after cell sorting, as in Fig 1E? Did authors compare the proteome of LoVo and HT29 sEVs? How do the protein profiles (in particular proteins involved in lipid metabolism) obtained under different conditions compare with each other, in particular after irradiation (Fig4N) and knockdown of ferritin heavy chain (Fig 3, Fig S3)? It would also be interesting to compare these data with the data obtained in CR-CSCs culture under hypoxia (Fig S5). Are common proteins involved in sEV production and LD biosynthesis identified in the analysis of these biological processes? Is there a common set of proteins/genes revealed by this analysis, which may potentially control sEV production and LD biosynthesis?

      Minor comments

      1. Some parts of the text are still a bit rough, and should be read and corrected carefully. For example: i) isn't it obvious that a common source of lipids builds up the membrane of sEVS, much like any other membrane (line 90, p.2); ii) what does this sentence mean: "LD have been considered as mere fat storage organelles for a long time, although important evidence could be traced back to the early 1960's". Important evidence for what? iii) why is the acronym AdExo used? iv) (line 138) the text should probably be "sEVs released during 72h were studied" and not "released sEVs were studied ... 72 h after seeding".
      2. The captions are far too small in most figures and diagrams (for example x and Y axis in Fig 1C-D, text in Fig 1E; Fig S1; Fig 3C proteins in the heatmap).
      3. The color code for LoVO and HT29 cells is reversed in Fig S1D-E
      4. In Fig 1D, I cannot see CD81 in the LoVo blot.
      5. Wording is not adequate in following sentences: "62.7% of proteins related to the exosomal pathway are downregulated in MCF7 shFTH1 cells" (line 233) and a few lines below: ".. the expression of almost all exosomal markers was downregulated in MCF7 shFTH1 cells" (line 239). Does 62.7% represent all proteins?
      6. In Fig 3E authors compare sEV markers secreted by cells treated with shFTH1 or control shRNA. The Anx5 and CD63 blots are not very convincing (quantification would be helpful).
      7. The micrographs in Fig 4L are too small: gold particles cannot be seen, even in the high magnification views.
      8. The micrographs showing ALIX and CD63 (Fig 4J) in irradiated and unirradiated Panc01 cells should be shown for comparison.

      Significance

      The topic of the paper is clearly interesting, since the mechanisms that regulate sEV formation and secretion are not fully understood and since the notion that their fate is linked to LDs is potentially exciting.

      My expertise: subcellular organization, endocytosis, membrane traffic, organelle biogenesis

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      We thank the reviewers for their thoughtful comments. Here we provide a point-by-point response to their reviews. All additional experiments that are present in the revised manuscript, or that we plan to include in the final manuscript, are numbered.

      __Reviewer #1 (Evidence, reproducibility and clarity (Required)):

      The concept introduced by this paper is exciting and novel. However, the current paucity of presented data can lead to incorrect interpretations of the findings and speculations that might not hold true after a more rigorous assessment of the observed phenomenon.

      The premise of this study builds upon an interaction between the PAXT complex and nuclear YTH domain containing proteins. However, figures 1B and C should be improved. The interacting band for the ZFC3H1 presented in panel B does not seem to match the size of the construct used in panel C. Is the Flag version of ZFC3H1 expressing a smaller isoform for this protein? __

      The reviewer is correct in that endogenous ZFC3H1 (which migrates at 250kD with a minor band at 150kD, see Figure 1B in the initial manuscript) appears to differ from the FLAG-tagged construct as expressed from a plasmid transfected in HEK293 cells (which migrates as two bands at 180kD and 200kD, see Figure 1C in the initial manuscript). For the endogenous protein, the predicted molecular weight is 226kD and the 250kD band disappears when cells are transduced with lentivirus containing shRNAs against ZFC3H1 (see Figure 4A in the initial manuscript), indicating that it is the correct protein. Both the 250kD endogenous protein (*) and the 200kD overexpressed protein (**) in transfected HEK293 and U2OS cells are detected in immunoblots using anti-ZFC3H1 antibodies (see Figure 1 in this document) indicating that the over-expressed protein is indeed ZFC3H1.

      [ Figure 1]

      _Figure 1. Molecular Weight Size Comparison of Endogenous ZFC3H1 and FLAG-ZFC3H1 (1-1233). _Lysates from HEK293 and U2OS that were either untransfected or transfected with FLAG-ZFC3H1 (1-1233) plasmid. We labelled the bands corresponding to the endogenous ZFC3H1 “*” and FLAG-ZFC3H1 (1-1233) “**”.

      We have sequenced the plasmid, and discovered that it contains an additional sequence inserted within the middle of the ZFC3H1 cDNA with a premature stop codon. As such, the version of the protein that is expressed from the plasmid only contains amino acids 1-1233 of the endogenous protein and is missing amino acids 1234-1989. The deleted region only contains TPR repeats, and is not known to interact with any of the well characterized interactors of ZFC3H1 (Wang, Nuc Acid Res 2021, Figure 3). We have renamed this construct FLAG-ZFC3H1 (1-1233). Given these new considerations, our results are consistent with the idea that the N-terminal portion of ZFC3H1 interacts with U1-70K, YTHDC1 and YTHDC2. We will change the text to reflect this.

      We are currently in the process of deleting the small insertion to obtain a plasmid that encodes a full length version of human ZFC3H1. For the final manuscript:

      Experiment #1) We will repeat the co-immunoprecipitation experiment with the full length FLAG-ZFC3H1 to determine whether it interacts with YTHDC1 and YTHDC2. This will take a few weeks.

      __Also, the YTHDC1-2 interaction in panel C is not as convincing considering the negative controls lane show some degree of binding. __

      Although the reviewer is correct that there is substantial background binding in the YTHDC1 immunoblot, we disagree with their characterization of the results with the YTHDC2 immunoblot (see Figure 1B-C in the initial manuscript). In the new manuscript we have included:

      Experiment #2) A new co-immunoprecipitate of the FLAG-tagged ZFC3H1 (1-1233) from HEK293 cells under more stringent conditions where the background level of YTHDC1 binding to beads is negligible. We have already completed this experiment (see Figure 1D in the revised manuscript).

      __Additionally, can the authors test if their RNaseA treatment worked? __

      In the new manuscript we have included:

      Experiment #3) A new co-immunoprecipitate of FLAG-YTHDC1 immunoblotted for eIF4AIII from HEK293 cell lysates. We find that without RNAse, there is some eIF4AIII in the precipitates but that the levels diminish substantially after RNAse A/T1 treatment. We have already completed this experiment (see Figure 1B in the preliminary revised manuscript).

      __Why do you need 18 hours to observe the nuclear export of your modifiable construct when inhibiting METTL3 in figure 3? Is it possible that your observation is secondary to phenotypes these cells develop as a result of blocking METTL3? __

      We treated cells for this period of time so that during the expression of the reporter, all of the newly synthesized mRNA is expressed in the absence of m6A methyl transferase activity. For shorter treatment times, it is unclear whether the bulk of the reporter mRNA, which would be synthesized before the treatment, would lose any pre-existing m6A marks, making a negative result hard to interpret. Previously we found that although 50% of intronic polyadenylated (IPA) transcripts from our reporters are rapidly degraded, about 50% are stable and are nuclear retained over extended periods of time (see Lee at al., PLOS ONE 2015; https://journals.plos.org/plosone/article?id=10.1371/journal.pone.0122743 Figure 3B-G). We believe that the bulk of the reporter mRNA that we are visualizing is stable and accumulates in the nucleus. Given that METTL3-depletion inhibits nuclear retention and that versions of the IPA reporter that lack m6A modification motifs are exported, we think that the most likely interpretation of the 18 hour STM2457 treatment experiments is that the lack of methyltransferase activity had a direct effect, rather than an indirect effect, on nuclear retention. We would be open to performing more experiments if the editors insist, however we ordered STM2457 four weeks ago and it has yet to arrive from Sigma-Millipore. Performing this experiment may substantially delay our ability to resubmit the manuscript in a timely manner.

      __Is ALKBH5 nuclear and/or cytoplasmic in the cell system used? __

      According to The Human Protein Atlas, ALKBH5 is predominantly nuclear in U2OS cells, with some present in the cytoplasm (https://www.proteinatlas.org/ENSG00000091542-ALKBH5/subcellular#human).

      In the revised manuscript we have included:

      Experiment #4) Data from subcellular fractionation demonstrating that ALKBH5 is present in both the nucleus and cytoplasm that we have already performed (see Figure 4J in the preliminary revised manuscript).

      __Reviewer #1 (Significance (Required)):

      The study is highly significant __

      ------

      __Reviewer #2 (Evidence, reproducibility and clarity (Required)):

      Summary: In the manuscript by Lee et al. entitled "N-6-methyladenosine (m6A) Promotes the Nuclear Retention of mRNAs with Intact 5'Splice Site Motifs", the authors provide evidence that m6A modifications within specific regions of transcripts can confer nuclear retention. These results are important because they add to our understanding of how m6A modifications can contribute to post-transcriptional regulation. Although the authors do not quite come out and say this, data seem to be accumulating to suggest that the location of the m6A modifications within a given transcript can dictate the functional consequences of those modifications.__

      We thank the reviewer for pointing this out. We have included a few sentences in the new preliminary revised manuscript pointing out that the location of the m6A modification in IPA transcripts, with respect to intact 5’SS and poly(A) signals, may play a role in promoting nuclear retention.

      __The current work builds on previous findings from these authors identifying factors critical for retention of intronic polyadenylated (IPA) transcripts. The present study identified m6A modification as one of the signals for the retention of such transcripts. The authors use reporters for their analysis and also examine validated endogenous IPA transcripts. The data presented supports the conclusions albeit they show a surprising finding for one of the m6A erasers, ALKBH5. However, there is some controversy over the mechanism by which ALKBH5 functions and whether the m6A mark is truly reversible, so these results may continue to add to this point of view.

      Major Comments: One experiment that might add to the argument would be overexpression of Mettl3 as compared to catalytically inactive Mettl3. The prediction would be that the reporter transcript with intact DRACH sequences would be even more retained in the nucleus in a manner that depends on Mettl3 catalytic activity. For some of the data presented, the reporter is already wholly nuclear so no difference could be detected, but in the U2OS cells shown in Figure 2B, it appears that an increase in nuclear localization might be evident. Such an experiment would add an orthogonal approach to demonstrate that the methylation by Mettl3 is required for retention. If such an experiment would work with the endogenous IPA transcripts shown in Figure 4, but these transcripts may already be too nuclear to detect any increase in nuclear retention.

      __

      We have performed two experiments that try to address this but they gave negative results:

      Experiment #5) We have over-expressed wildtype and a methyl transferase mutant FLAG-METTL3 and assessed the nuclear export/retention of ftz-Δi-5’SS mRNA. There was no effect (see Figure 2 in this document).

      [Figure 2]

      __Figure 2. Over-expression of METTL3 does not increase the nuclear retention of ftz-Δi-5’SS. __U2OS cells were co-transfected with ftz-Δi-5’SS reporter and either FLAG-METTL3 or FLAG-METTL3-D395A, which lacks methyl-transferase activity (Wang, Mol Cell 2016). Cells were fixed, stained for ftz mRNA by fluorescent in situ hybridization and METTL3 using anti-FLAG antibodies. The nuclear and cytoplasmic distribution of ftz mRNA was quantified as described in the manuscript. Note that this is the average of one independent experiment (each bar consisting of the average of at least 50 cells). We plan to repeat this two more times, but we anticipate that these will show the same result.

      We could include this negative data as a supplemental figure. We believe that there are two possible reasons for this experimental result. First, as the reviewer points out, the reporter transcripts are already too nuclear to detect any significant change. Second, METTL3 is part of a larger complex that includes several proteins including METTL14, WTAP and potentially other proteins (for example see Covelo-Molares, Nuc Acid Res 2021). We may need to co-express all of these proteins to see an effect.

      Experiment #6) We have also expressed versions of ftz-Δi and ftz-Δi-5’SS mRNA with optimized m6A modification (i.e. DRACH) motifs (AGACT) to enhance methylation (“e-m6A-ftz”). We only observed a slight increase in nuclear retention but it is not significant (see Figure 2A,C in the revised manuscript).

      Again, this result could be explained by the fact that the reporter is too nuclear to detect any significant increase in retention. We had originally performed this in parallel with the no-m6A-ftz-Δi-5’SS reporters but did not report this negative data in the original manuscript.

      __Some rather minor changes to the presentation of the data could enhance the impact of this study.

      Specific Comments:

      The primary question in this manuscript is comparing reporters with m6A site (intact DRACH sequences) to those without. For this reason, organizing the data to the +/- DRACH sites are adjacent to one another might make the most sense. This point is evident in Figure 1C where perhaps simply changing the order of the bars presented to put the ones directly compared adjacent would be preferable. Then the p-value would compare sets of data directly adjacent to one another. __

      We thank the reviewer for this suggestion and we have made these changes to the figures in the preliminary revised manuscript.

      __While the authors show representative fields/cells for most assays, they do an excellent job of providing quantitation as well. One exception is Figure 3D, which shows a single cell image for the most key panel (the 5'SS-containing reporter upon Mettl3 depletion). If there is not a field with more cells, the authors could create a montage. __

      In the revised manuscript, we have replaced this image with one containing multiple cells expressing the reporter.

      __Minor Comments:

      Figure presentation:

      The text in a number of the figures is VERY small (Figures 1B,1C, and 4A) for example. __

      We have fixed this in the new manuscript.

      __Figure 3A includes the label "shRNA:" at the top, but these cells are treated with Mettl3 inhibitor and there does not appear to be any shRNA employed, so this seems like a labeling error. __

      We have fixed this in the new manuscript.

      __In Figure 3C, the immunoblot of Mettl3, there are three bands that all disappear completely upon knockdown of Mettl3- are these all Mettl3? This should at least be mentioned in the legend and perhaps indicated in the figure. The authors do mention in the text employing shRNAs to target multiple Mettl3 isoforms, so likely this is the case. __

      We have clarified these issues in the new manuscript.

      __Minor points (some really minor to just polish the presentation for clarity):

      The word "since" should only be used if there is a time element- otherwise the word "as" is preferable.

      For example on p. 4, the sentence: "Since inhibition of mRNA export typically enhances the nuclear retention of RNAs with intact 5'SS motifs (Lee et al. 2020),.." would more precisely read "As inhibition of mRNA export typically enhances the nuclear retention of RNAs with intact 5'SS motifs (Lee et al. 2020),..". __

      We thank the reviewer for pointing this out. We have fixed this issue in the revised manuscript.

      __Reviewer #2 (Significance (Required)):

      Summary: In the manuscript by Lee et al. entitled "N-6-methyladenosine (m6A) Promotes the Nuclear Retention of mRNAs with Intact 5'Splice Site Motifs", the authors provide evidence that m6A modifications within specific regions of transcripts can confer nuclear retention. These results are important because they add to our understanding of how m6A modifications can contribute to post-transcriptional regulation. Although the authors do not quite come out and say this, data seem to be accumulating to suggest that the location of the m6A modifications within a given transcript can dictate the functional consequences of those modifications.

      This study would be of significant interest to those that study gene expression in any context as well as cell biologists as the data add to our understanding of export of mRNA from the nucleus. This work also adds to our understanding of the biological consequences of m6A modification, which is an area of significant interest. In my opinion, the authors could make a broader conclusion that we do, which is that the location of the modification significantly dictates function- an extension of previous findings mostly focused on processed mRNA transcripts. __

      -------

      __Reviewer #3 (Evidence, reproducibility and clarity (Required)):

      Quality control of mRNA is vital for all types of cells. In eukaryotic cells, nuclear export of misprocessed mRNAs containing the 5' splice site is prevented. In this manuscript, Lee and colleagues demonstrate that the nuclear retention of intronic polyadenylated transcripts is dependent on m6A modification. Based on the results shown in yeast, they perform immunoprecipitation experiments and demonstrate the interaction between ZFC3H1, a component of the PAXT complex, and YTHDC1 and YTHDC 2, nuclear YTH RNA-binding proteins that recognize m6A-modified transcripts. The study also shows the interaction of U1-70K with YTHDC1 and with ZFC3H1. Depletion of YTHDC1/2 prevents the nuclear retention of IPA transcripts. Additionally, CLIP-seq analysis is performed, demonstrating that m6A modification is enriched around the 5' splice site motif and the 3' polyadenylation site in IPAs. From these observations, they conclude that m6A modification contributes to the quality control of mRNA by promoting nuclear retention of misprocessed transcripts.

      Major Points 1. The interaction between ZFC3H1 and YTHDC1 is clearly shown by immunoprecipitation of FLAG-tagged YTHDC1 in Figure 1B. However, the co-purification of YTHDC1 with FLAG-tagged ZFC3H1 in Figure 1C is rather ambiguous. Additionally, the immunoprecipitated samples do not appear to show signals corresponding to FLAG-tagged ZFC3H1, making it unclear if the immunoprecipitation is working. It is essential to provide a better quality result to clarify these observations. __

      Please see our responses to reviewer #1. We have repeated the co-immunoprecipitation of FLAG-ZFC3H1 (1-1233) with YTHDC1 under more stringent conditions and have reduced the background binding (see Figure 1B and D in the new manuscript). We have also determined why the FLAG-ZFC3H1 is smaller than expected as the construct contains a premature stop codon. As explained above, we are in the midst of generating a full-length FLAG-ZFC3H1 and we plan to repeat the co-immunoprecipitation with this new construct.

      2. While the authors demonstrate that the m6A modification is dispensable for the targeting of IPA reporter transcripts to the nuclear speckles, it would be valuable to investigate whether m6A is required for their exit from the nuclear speckles. Do reporter transcripts with m6A motifs remain in the nuclear speckles at later time points?

      We have now analyzed the colocalization of nuclear speckles (SC35) with ftz-Δi-5’SS, which contains both a 5’SS and DRACH motifs, and no-m6A-ftz-Δi-5’SS, which contains a 5’SS but lacks DRACH motifs, at steady state – i.e. after 18-24 hours of transfection (as opposed to at early time points as shown in Figure 2D-E of the initial manuscript). Unexpectedly, we see that both mRNAs continue to colocalize with nuclear speckles, although the no-m6A-ftz-Δi-5’SS mRNA is well exported from the nucleus and its signal in nuclear speckles is faint (see Figure 2F-H in the new manuscript).

      Previously, we observed that ftz-Δi-5’SS required the 5’SS motif to remain in nuclear speckles at these later time points (Lee PLOS ONE 2015 and Lee RNA 2022). Upon closer inspection, ftz-Δi-5’SS mRNA also accumulates in additional nuclear foci that are not SC35-positive. Our new results may indicate that m6A marks promote the transfer of mRNAs from nuclear speckles to other foci, but more data is required to make a firm statement. Given this, we plan to conduct further experiments which may take a month to complete:

      Experiment #7) We are now assessing whether these additional ftz-Δi-5’SS foci correspond to either YTHDC-positive foci which were previously shown to partially overlap nuclear speckles and sequester m6A-rich mRNAs (Cheng Cancer Cell 2022), or “pA+ RNA foci” which accumulate MTR4/ZFC3H1-targetted RNAs when the nuclear exosome is inhibited (Silla Cell Reports 2018). These foci are enriched in ZFC3H1. We plan on co-staining ftz-Δi, ftz-Δi-5’SS, no-m6A-ftz-Δi and no-m6A-ftz-Δi-5’SS with SC35, YTHDC1 and ZFC3H1 to determine whether m6A may help to transfer mRNAs from nuclear speckles to YTHDC1 or ZFC3H1-enriched foci.

      __3. Figures 5B and 5C suggest that ZFC3H1 is required for the degradation of IPA transcripts. However, the range of the vertical axis is inappropriate and it is difficult to assess the extent of the increase in expression levels. Please adjust the vertical axis range for improved clarity. __

      We thank the reviewer for the feedback we have added additional graphs with an expanded vertical axis to demonstrate that ZFC3H1 is required for the degradation IPA transcripts.

      Minor Points 1. page 4, line 2 "RNAse" should be corrected to "RNase".

      We thank the reviewer for catching this error. We have fixed this.

      __ 2. page 7, line 5: Is the statement "prevents the nuclear export and decay of non-functional and misprocessed RNAs" correct? m6A modification promotes the decay of such RNAs. __

      We thank the reviewer for pointing this out. We have altered the text to clarify that m6A promotes decay.

      __3. Figure 2E: ftz-∆i should be ftz-∆i-5'SS. __

      We thank the reviewer for catching this error. We have fixed this.

      __4. Figure 5A: It would be helpful to indicate the number of IPA transcripts analyzed. __

      We have included this information.

      __Reviewer #3 (Significance (Required)):

      Overall, the work is sound and generally well-controlled. This study advances our understanding of the quality control of misprocessed transcripts in higher eukaryotes. This reviewer suggests a few points for clarification or improvement. __

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Quality control of mRNA is vital for all types of cells. In eukaryotic cells, nuclear export of misprocessed mRNAs containing the 5' splice site is prevented. In this manuscript, Lee and colleagues demonstrate that the nuclear retention of intronic polyadenylated transcripts is dependent on m6A modification. Based on the results shown in yeast, they perform immunoprecipitation experiments and demonstrate the interaction between ZFC3H1, a component of the PAXT complex, and YTHDC1 and YTHDC 2, nuclear YTH RNA-binding proteins that recognize m6A-modified transcripts. The study also shows the interaction of U1-70K with YTHDC1 and with ZFC3H1. Depletion of YTHDC1/2 prevents the nuclear retention of IPA transcripts. Additionally, CLIP-seq analysis is performed, demonstrating that m6A modification is enriched around the 5' splice site motif and the 3' polyadenylation site in IPAs. From these observations, they conclude that m6A modification contributes to the quality control of mRNA by promoting nuclear retention of misprocessed transcripts.

      Major Points

      1. The interaction between ZFC3H1 and YTHDC1 is clearly shown by immunoprecipitation of FLAG-tagged YTHDC1 in Figure 1B. However, the co-purification of YTHDC1 with FLAG-tagged ZFC3H1 in Figure 1C is rather ambiguous. Additionally, the immunoprecipitated samples do not appear to show signals corresponding to FLAG-tagged ZFC3H1, making it unclear if the immunoprecipitation is working. It is essential to provide a better quality result to clarify these observations.
      2. While the authors demonstrate that the m6A modification is dispensable for the targeting of IPA reporter transcripts to the nuclear speckles, it would be valuable to investigate whether m6A is required for their exit from the nuclear speckles. Do reporter transcripts with m6A motifs remain in the nuclear speckles at later time points?
      3. Figures 5B and 5C suggest that ZFC3H1 is required for the degradation of IPA transcripts. However, the range of the vertical axis is inappropriate and it is difficult to assess the extent of the increase in expression levels. Please adjust the vertical axis range for improved clarity.

      Minor Points

      1. page 4, line 2 "RNAse" should be corrected to "RNase".
      2. page 7, line 5: Is the statement "prevents the nuclear export and decay of non-functional and misprocessed RNAs" correct? m6A modification promotes the decay of such RNAs.
      3. Figure 2E: ftz-∆i should be ftz-∆i-5'SS.
      4. Figure 5A: It would be helpful to indicate the number of IPA transcripts analyzed.

      Significance

      Overall, the work is sound and generally well-controlled. This study advances our understanding of the quality control of misprocessed transcripts in higher eukaryotes. This reviewer suggests a few points for clarification or improvement.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary: In the manuscript by Lee et al. entitled "N-6-methyladenosine (m6A) Promotes the Nuclear Retention of mRNAs with Intact 5'Splice Site Motifs", the authors provide evidence that m6A modifications within specific regions of transcripts can confer nuclear retention. These results are important because they add to our understanding of how m6A modifications can contribute to post-transcriptional regulation. Although the authors do not quite come out and say this, data seem to be accumulating to suggest that the location of the m6A modifications within a given transcript can dictate the functional consequences of those modifications.

      The current work builds on previous findings from these authors identifying factors critical for retention of intronic polyadenylated (IPA) transcripts. The present study identified m6A modification as one of the signals for the retention of such transcripts. The authors use reporters for their analysis and also examine validated endogenous IPA transcripts. The data presented supports the conclusions albeit they show a surprising finding for one of the m6A erasers, ALKBH5. However, there is some controversy over the mechanism by which ALKBH5 functions and whether the m6A mark is truly reversible, so these results may continue to add to this point of view.

      Major Comments: One experiment that might add to the argument would be overexpression of Mettl3 as compared to catalytically inactive Mettl3. The prediction would be that the reporter transcript with intact DRACH sequences would be even more retained in the nucleus in a manner that depends on Mettl3 catalytic activity. For some of the data presented, the reporter is already wholly nuclear so no difference could be detected, but in the U2OS cells shown in Figure 2B, it appears that an increase in nuclear localization might be evident. Such an experiment would add an orthogonal approach to demonstrate that the methylation by Mettl3 is required for retention. If such an experiment would work with the endogenous IPA transcripts shown in Figure 4, but these transcripts may already be too nuclear to detect any increase in nuclear retention.

      Some rather minor changes to the presentation of the data could enhance the impact of this study.

      Specific Comments:

      The primary question in this manuscript is comparing reporters with m6A site (intact DRACH sequences) to those without. For this reason, organizing the data to the +/- DRACH sites are adjacent to one another might make the most sense. This point is evident in Figure 1C where perhaps simply changing the order of the bars presented to put the ones directly compared adjacent would be preferable. Then the p-value would compare sets of data directly adjacent to one another.

      While the authors show representative fields/cells for most assays, they do an excellent job of providing quantitation as well. One exception is Figure 3D, which shows a single cell image for the most key panel (the 5'SS-containing reporter upon Mettl3 depletion). If there is not a field with more cells, the authors could create a montage.

      Minor Comments:

      Figure presentation:

      The text in a number of the figures is VERY small (Figures 1B,1C, and 4A) for example.

      Figure 3A includes the label "shRNA:" at the top, but these cells are treated with Mettl3 inhibitor and there does not appear to be any shRNA employed, so this seems like a labeling error.

      In Figure 3C, the immunoblot of Mettl3, there are three bands that all disappear completely upon knockdown of Mettl3- are these all Mettl3? This should at least be mentioned in the legend and perhaps indicated in the figure. The authors do mention in the text employing shRNAs to target multiple Mettl3 isoforms, so likely this is the case.

      Minor points (some really minor to just polish the presentation for clarity):

      The word "since" should only be used if there is a time element- otherwise the word "as" is preferable.

      For example on p. 4, the sentence: "Since inhibition of mRNA export typically enhances the nuclear retention of RNAs with intact 5'SS motifs (Lee et al. 2020),.." would more precisely read "As inhibition of mRNA export typically enhances the nuclear retention of RNAs with intact 5'SS motifs (Lee et al. 2020),..".

      Significance

      Summary: In the manuscript by Lee et al. entitled "N-6-methyladenosine (m6A) Promotes the Nuclear Retention of mRNAs with Intact 5'Splice Site Motifs", the authors provide evidence that m6A modifications within specific regions of transcripts can confer nuclear retention. These results are important because they add to our understanding of how m6A modifications can contribute to post-transcriptional regulation. Although the authors do not quite come out and say this, data seem to be accumulating to suggest that the location of the m6A modifications within a given transcript can dictate the functional consequences of those modifications.

      This study would be of significant interest to those that study gene expression in any context as well as cell biologists as the data add to our understanding of export of mRNA from the nucleus. This work also adds to our understanding of the biological consequences of m6A modification, which is an area of significant interest. In my opinion, the authors could make a broader conclusion that we do, which is that the location of the modification significantly dictates function- an extension of previous findings mostly focused on processed mRNA transcripts.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      The concept introduced by this paper is exciting and novel. However, the current paucity of presented data can lead to incorrect interpretations of the findings and speculations that might not hold true after a more rigorous assessment of the observed phenomenon.

      The premise of this study builds upon an interaction between the PAXT complex and nuclear YTH domain containing proteins. However, figures 1B and C should be improved. The interacting band for the ZFC3H1 presented in panel B does not seem to match the size of the construct used in panel C. Is the Flag version of ZFC3H1 expressing a smaller isoform for this protein? Also, the YTHDC1-2 interaction in panel C is not as convincing considering the negative controls lane show some degree of binding. Additionally, can the authors test if their RNaseA treatment worked?

      Why do you need 18 hours to observe the nuclear export of your modifiable construct when inhibiting METTL3 in figure 3? Is it possible that your observation is secondary to phenotypes these cells develop as a result of blocking METTL3?

      Is ALKBH5 nuclear and/or cytoplasmic in the cell system used?

      Significance

      The study is highly significant

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      1. General Statements

      We would like to thank the editorial staff and the reviewers for their handling of our manuscript. We were very pleased with the timely communications from Review Commons, and we are grateful to have been assigned this insightful and constructive group of reviewers.

      The reviewers were well-suited to evaluate our work based on their stated areas of expertise (cancer biology, image analysis, machine learning, cell-based screening, etc.). As such, we received thoughtful and constructive feedback, which we have already incorporated into our attached revision. We are confident that these reviews have improved our manuscript.

      Our goal with this manuscript is to present a proof-of-concept study where high-content imaging and morphological profiling are used to characterize drug resistance in clonal cell lines. The main criticism from reviewers was that our original manuscript may have overstated our method’s ability to discriminate the signal of bortezomib resistance and that any extension beyond cultured cells (to patient samples for example) would require significant follow-up studies. The reviewers suggested that such work would be beyond the scope of our study, and recommended toning down our language to better reflect the limitations of this proof-of-concept work. We have embraced this suggestion, extensively revising our text, and we now believe our language and tone more accurately reflects our results. The reviewers also suggested follow-up computational analyses to more robustly characterize the bortezomib resistance signature. We have performed these analyses and added their description to our revised manuscript. We feel that these analyses have improved understanding of the signature, and will help a reader to gain a deeper understanding of our results and methodology.

      The reviewers also suggested several minor changes; many of which we embraced fully, but others that we chose not to incorporate. We felt that a lack of clarity in our text contributed to these reviewer suggestions. In these cases, we improved clarity in the text and responded to each comment point-by-point in the “prefer not to carry out” section. Further, we address all reviewer comments in the following document point-by-point, grouped by common themes across reviewers (e.g., tone, clarity, analyses, etc.).

      Lastly, a common theme among reviewer comments was their appreciation for our strong methodology and data transparency (examples pasted below). We are extremely gratified by this observation as we feel this is a particular strength of our manuscript. In addition, we were pleased to see reviewers engaged by our work, acknowledging the interest this manuscript is likely to generate among a broad range of scientific disciplines.

      Examples of reviewer appreciation of our strong methodology and data transparency:

      Reviewer 1: “However, this does not imply that the same approach can not achieve the goal, perhaps by using other cell painting markers for bortezomib-sensitivity, or with the same markers to assess sensitivity of different drugs. The cell painting + analysis approaches are not new and the clinical impact is questionable, but the technical aspects (data, analysis) are exceptional and the concept may hold as I described above.”

      Reviewer 2: “The paper is well written, and the text is clear, as is the presentation of data and transparency of methods being utilized. The methods were applied appropriately and followed established standards in the field. The paper's premise is timely and interesting, addressing a pressing issue in cancer therapy: making informed treatment decisions fast, based on markers found in tumors early in tumor development, and using image-based screening for characterizing drug resistance before treatment could be an option. A fascinating bit of the manuscript is the description of the feature selection from the screen is done systematically, considering the technical and biological variability and technical artifacts and modeling covariates using linear models seems a very appropriate way of doing so and could serve as another proof of concept that this is indeed the most robust way of modeling and removing signal of technical covariates from the data.”

      Reviewer 3: “The strengths of this study are the machine learning best practice and detailed methodology. The experiments could be reproduced and statistical analysis is more than adequate. The analysis takes into account batch effects, well position, differences in cell numbers, and other sources of technical variation that complicate high-content image analysis. It is a good exemplar of how unsupervised morphological profiling can be applied to imaging data. The major limitation is the generalizability of this particular method for patient samples. This could be addressed in the Discussion.”

      1. Description of the planned revisions

      We have incorporated all planned revisions.

      1. Description of the revisions that have already been incorporated in the transferred manuscript

      Text revisions already carried out

      1. [Text revision] We have materially toned down our claims in the manuscript in two distinct areas: A) model performance and B) potential clinical application. A) Model performance. We specifically balanced our discussion of the discriminative signal of the Bortezomib Signature. While the signature adequately separated never-before-seen wildtype and resistant clones with metrics well above randomly permuted baselines (accuracy near 80%, average precision about 70%, area under the ROC curve (AUROC) about 84%), there were many limitations that we should have more explicitly highlighted. For example, many individual profiles were incorrectly classified, some clones were predicted entirely incorrectly, and many profiles did not receive Bortezomib Signature scores above the randomly permuted baseline. We have more clearly discussed these limitations and used more balanced language (see key examples of text-based changes below). Additionally, we modified a figure (now Figure 3) to include boxplots of clones that explicitly show the Bortezomib Signature scores of each well profile and permit examination of the strength of the signature for each clone (previously found in Figure 2-Supplement 9). Lastly, we add a new supplementary figure (now Figure 5-Supplement 1) that describes a feature space analysis of misclassified samples. Please note that this figure rearrangement and new analysis helped to balance our claims, but were also performed in response to other tangential reviewer comments. B) Clinical application. In the abstract, introduction, and discussion, we further emphasized that this work is a proof of concept, and that more advances must be made prior to clinical application.

      We made these changes in direct response to the following reviewer comments:

      Reviewer 1 - Major Comment 1 (relevant excerpts)

      While I am convinced that the signature captures morphological phenotypes associated with drug resistance, at the cumulative scale, the discriminative signal of a single cell type seems weak… With Fig. 4, the data fully supports the argument that the bortezomib-signature encodes bortezomib-resistance, but the signal is weak. Thus statements such as "We found the Bortezomib Signature could predict whether a cell line was bortezomib-resistant or bortezomib-sensitive" (line #172) and the specificity statements in the abstract" (line #28) are not supported by the data in my opinion. I would recommend the authors to tune down these and other related statements throughout the manuscript.

      Reviewer cross-commenting - Reviewer 1

      My main critic is regarding "over selling" a weak discriminative signal. Specifically, I am not convinced that the major claims regarding predicting sensitivity and specificity at the single cell types scales are supported by the data. Since reviewer #2 and #3 did not raise this concern I think it is worth discussion here.

      Once these statements are tuned down - I think no significant additional work is needed to make the point that they can measure a discriminative signal. If they want to make these claims, perhaps they'd like to collect more data to gain statistical power (but I am not optimistic this will work at the single cell level).

      Personally, I was happy with the authors' choice of cell lines not included in the training dataset. I am not convinced that additional cell lines + validations are necessary for making the point of a proof of principle.

      Reviewer cross-commenting - Reviewer 2

      I agree that, perhaps, my major criticism of the paper was the manuscript's 'overselling' of claims that were only weakly supported by the data. Yes, if the authors tune down their claims and clearly state that this is an interesting starting point and proof of concept study, it might be ok to publish with only minor revisions. If the claims should be more generalized, then this study needs more data supporting the conclusions and the method's predictive power.

      Reviewer 2 - Major Comment 8

      Lastly, I find some misfits between the question, the model used, and the conclusions drawn. The authors start by exploring the problem of bortezomib resistance in cancer treatment, which they say is a devastating issue for patients with, e.g., multiple myeloma. Yet, the authors use HCT116 as their model cell line, a microsatellite instable, colorectal cell line with several intrinsic mutations that make it a difficult model to address physiologically relevant medical problems after all. The authors then go on to suppose that their method might be suitable to diagnose resistance in patient samples, but I am not convinced this conclusion can be speculated based on data from HCT cells. I suggest the authors test their approach on at least two other cell lines (maybe from different tissues) and benchmark their results against a dataset of digital pathology where such predictions are made from stained and analyzed tissue slices. This way, after a thorough benchmark against related third-party data sets, the method would significantly gain relevance, the paper would appeal to a broader audience, and the advance gains more merit.

      Reviewer 3 - Major Comment 5

      It is not clear from the Discussion whether this type of analysis is more broadly applicable to cell lines derived from patients, rather than selected from a parental cell line, or if this approach would be more efficient than genotyping or next-gen sequencing. How many replicates and ground truth cell lines would be necessary for predictive confidence?

      We edited the last two sentences of the abstract to tone down specificity claims (“provide evidence”) and clarify that we are establishing a “proof-of-concept framework”.

      • This signature predicted bortezomib resistance better than resistance to other drugs targeting the ubiquitin-proteasome system. Our results establish a proof-of-concept framework for the unbiased analysis of drug resistance using high-content microscopy of cancer cells, in the absence of drug treatment.

      We revised the last paragraph of the introduction to contrast bortezomib predictions with ixazomib/CB-5083 predictions, and to remove claims about “using microscopy to guide therapy”.

      • This morphological signature correctly predicted the bortezomib resistance of seven out of ten clones not included in the signature training dataset. Overall, our results establish a proof-of-concept framework for identifying unbiased signatures of drug resistance using high-content microscopy. The ability to identify drug-resistant cells based on morphological features provides a valuable orthogonal method for characterizing resistance in the absence of drug treatment.

      To tone down claims in the figures, we added boxplots to Figure 3 (previous Figure 2) showing specific distribution of signature scores per well profile and updated Figure 4 legend (previous Figure 3).

      • Figure 4. Bortezomib Signature has limited ability to characterize clones resistant to other ubiquitin-proteasome system inhibitors.

      We modify the following text in the discussion to tone down claims of specificity and clinical utility:

      • This Bortezomib Signature correctly predicted the bortezomib resistance of seven out of ten clones not included in the training dataset and was more specific to bortezomib-resistance given its limited ability to identify clones that were resistant to other UPS-targeting drugs.

      Though it is unclear whether this method can be extended to patient samples, where identifying intrinsic drug resistance in cells prior to treatment has the potential to improve targeted cancer therapy, our results are an encouraging proof of concept. We expect that further refinement may develop Cell Painting as a tool for identifying drug-resistant cells, perhaps even guiding strategies to overcome intrinsic resistance.

      1. [Text revision] We defined LD50 in text (originally line #97), changed description of resistant clone selection to remove main text references to LD90 (originally line #87), and stated drug concentrations used for selection in Methods. We also defined LD90 in the Methods and described its role in determining the drug concentrations to use for clone selection. This change was in response to the following comments:

      Reviewer 1 - Minor Comment 2

      What is LD90 (line #87)? LD50 (line #97)?

      Reviewer 2 - Minor Comment 5

      What was the LD 90 per drug on HCT cells? Rather than LD90 foldchanges, absolute concentrations should be used in the results and discussion to allow the reader to vet the conclusions.

      • To determine the appropriate drug concentrations to use in order to isolate drug-resistant clones, we performed proliferation assays on HCT116 parental cells with our drugs of interest: bortezomib (proteasome inhibitor), ixazomib (proteasome inhibitor), or CB-5083 (p97 inhibitor) (Fig. 1-Supplement 1 A-D).
      • We characterized the bortezomib-resistant clones and found that the median lethal doses (LD50s) were ~2.8- to ~9-fold that of HCT116 parental cells (Fig. 1-Supplement 2 B).
      • Briefly, HCT116 cells were plated in 150 mm dishes and grown in the presence of the desired drug at a concentration that resulted in the death of the majority of cells (selection concentrations: bortezomib, 12 nM; ixazomib, 150 nM; CB-5083, 600 and 700 nM).
      • Using the data from our proliferation assays, we calculated the median lethal dose (LD50) for each of our drugs of interest by fitting data of normalized growth vs. log[drug concentration] to a sigmoidal dose-response curve using GraphPad Prism (v.9.2.0) (Fig. 1-Supplement 1 D).

      • [Text revision] We thank the reviewer for allowing us an opportunity to improve clarity on the clones we used. We now describe the total number of clones generated and removed unnecessary references to specific clones for ease of reading (originally lines #96-98) (We maintain all references to specific clones in the figures, legends, supplement, and methods)

      Reviewer 1 - Minor Comment 3

      It was not clear to me in the text which and how many cell lines were evaluated and the reader is forced to go to the SI. For example, "(BZ01-10 and BZ clones A and E)" (line #96-97) and "wild-type clones (WT01-05, 10, and 12-15)" (line #98) appeared when presenting the results without a clear explanation and made it harder for me to follow. Summary of the data (for example, based on Figure 2-Supplement 8) can be briefly mentioned in the text to make it more clear for the reader.

      We added the following to the second paragraph of the results:

      • Together these methods provided a total of twelve bortezomib-resistant, five ixazomib-resistant, five CB-5083-resistant, and twelve bortezomib-sensitive clones as well as HCT116 parental cells for our experiments.

      [Text revision] We removed duplicate text (originally lines #115-125).

      Reviewer 1 - Minor Comment 5

      1. Lines #104-111 were duplicated in lines #114-122.

      Reviewer 3 - Minor Comment 4

      Ten lines of text are duplicated on page 5.

      Reviewer 2 - Minor Comment 4

      on page 5, paragraph 4, there is a sizeable copy-and-paste error of text being identically replicated.

      1. [Text revision] We provided more intuition of the Bortezomib Signature in the results section (originally lines #150-151).

      Reviewer 1 - Minor Comment 6

      The "Bortezomib Signature" is a critical measurement but is only briefly mentioned in lines 150-151 ("..based on the direction-sensitive ranking method for phenotype analysis, singscore (Foroutan et al., 2018)"). Please provide more information/intuition.

      • We used these 45 features to compute a rank-based resistance score or “Bortezomib Signature” for each well profile based on the direction-sensitive method called singscore (Foroutan et al. 2018). Singscore ranks these 45 resistance-related features on a per sample basis and calculates a normalized score between -1 and 1, with higher values expected for bortezomib-resistant clones and lower values expected for bortezomib-sensitive clones.

      • [Text revision] We clarified that DNA sequencing had been performed solely on clones A and E in a previous study (originally lines #88-90). Furthermore, one of the strengths of our approach is that it can identify resistant clones in an unbiased fashion prior to molecular characterization. It is beyond scope to perform these sequencing studies in the present paper.

      Reviewer 2 - Minor Comment 3

      The authors talk about validating the mutation - PSMB5 by RNA-seq. However, the data for the genotyping/sequencing/characterization of these newly generated BZ-resistant lines are missing.<br />

      In the results, we clarify DNA sequencing that was previously performed on clones A and E

      • We also isolated bortezomib-sensitive (wild-type; WT) clones by dilution of the HCT116 parental cell line and acquired two bortezomib-resistant clones (BZ clones A and E) both with mutations in PSMB5 identified by RNA sequencing performed in previous work (Fig. 1-Supplement 1 E) (Wacker et al. 2012).

      In the last paragraph of the discussion, we highlight the strength of our unbiased approach

      • Together, our work has demonstrated the potential for morphological profiling with Cell Painting to be used as an unbiased method to characterize resistance in the absence of drug treatment. Our results indicate that different mechanisms of bortezomib resistance may generate distinct morphological profiles; with larger and broader training datasets, it may be possible to identify signatures for distinct mechanisms of bortezomib resistance as well as signatures of resistance to other drugs. Though it is unclear whether this method can be extended to patient samples, where identifying intrinsic drug resistance in cells prior to treatment has the potential to improve targeted cancer therapy, our results are an encouraging proof of concept. We expect that further refinement may develop Cell Painting as a tool for identifying drug-resistant cells, perhaps even guiding strategies to overcome intrinsic resistance.

      • [Text revision] We thank the reviewers for their suggestions. We agree that the description of the experimental design was somewhat unclear and have provided greater detail and clarity, particularly regarding the generation of clones. We used the HCT116 parental cell line to generate drug-resistant clones by identifying single surviving cells after drug treatment and allowing these cells to expand prior to isolating colonies for experimentation. We did not perform experiments to confirm whether these “clones” were isogenic and can not exclude cell migration during expansion or genetic drift as convoluting factors. However, we have provided greater detail in the descriptions of our method for clone isolation in order to address this concern.

      Reviewer 1 - Minor Comment 1

      More information in Fig. 1's legend would be helpful to follow the experimental design. I found it hard to follow in its current form and had to go back to carefully reading the main text to fully understand.

      Reviewer 2 - Minor Comment 6

      The description of the resistant clonal populations is confusing. As I understand, no single-cell clones were isolated during the selection procedure. Thus, the training lines are not yet isogenic clones but oligoclonal sub-populations of the parental cell line. The authors could provide more details here and discuss the different characteristics of their sub-populations, e.g., their growth kinetics or molecular alterations.

      We bolstered the description in the results.

      • We first isolated and characterized drug-resistant cells (Fig. 1 A). To isolate drug-resistant clones, we used an approach we have described previously (Wacker et al. 2012; Kasap, Elemento, and Kapoor 2014) and the HCT116 cell line. These cancer cells express multidrug resistance pumps at low levels and are mismatch repair deficient, providing a genetically heterogeneous polyclonal population of cells (Umar et al. 1994; Papadopoulos et al. 1994; Teraishi et al. 2005) allowing for isolation of drug-resistant clones in 2-3 weeks. We hypothesize that a rapid selection of resistance could favor the isolation of clones with intrinsic resistance. To determine the appropriate drug concentrations to use in order to isolate drug-resistant clones, we performed proliferation assays on HCT116 parental cells with our drugs of interest: bortezomib, ixazomib, or CB-5083 (Fig. 1-Supplement 1 A-D). We also isolated bortezomib-sensitive (wild-type; WT) clones by dilution of the HCT116 parental cell line and acquired two published bortezomib-resistant clones (BZ clones A and E) both with mutations in PSMB5 identified by RNA sequencing performed in previous work (Fig. 1-Supplement 1 E) (Wacker et al. 2012). We characterized the bortezomib-resistant clones and found that the median lethal doses (LD50s) for bortezomib were ~2.8- to ~9-fold that of HCT116 parental cells (Fig. 1-Supplement 2 B). In contrast, bortezomib-sensitive clones had LD50s for bortezomib that ranged from ~0.7- to ~1.2-fold that of HCT116 parental cells (Fig. 1-Supplement 2 A). Together these methods provided a total of twelve bortezomib-resistant, five ixazomib-resistant, five CB-5083-resistant, and twelve bortezomib-sensitive clones as well as HCT116 parental cells for our experiments.

      We also updated the legend for Figure 1A.

      • Figure 1. Experimental design for using Cell Painting to examine morphological profiles of drug-resistant cells. (A) Graphic of the experimental workflow: we isolated drug-resistant clones by treating parental HCT116 cells with a high dose of the desired drug and then expanded them for experiments. We isolated drug-sensitive clones by diluting HCT116 cells and then expanded them for experiments. We then performed proliferation assays on select clones to screen for multidrug resistance. Next, we performed Cell Painting on both drug-resistant and -sensitive clones, using multiplexed high-throughput fluorescence microscopy of fixed cells followed by feature extraction and morphological profiling to search for features that contribute to a signature of drug resistance.

      • [Text revision] We clarified that the Bortezomib Signature did not correspond to well position (originally lines #155-157).

      Reviewer 1 - Minor Comment 9

      Line #155-156: "We found that the pattern of Bortezomib Signatures corresponded to the cell identity plate layout", the word "not" is missing before "corresponded".

      We found that the pattern of Bortezomib Signatures did not correspond to well position relative to the plate (Fig. 2-Supplement 7 B), indicating that the well position for each clone was not strongly contributing to its Bortezomib Signature.

      1. [Text revision] We explicitly described the result that some misclassified clones (WT10, WT15, and BZ06) did not have unexpected bortezomib sensitivity as determined by proliferation assays. We also moved the supplementary figure to an updated Figure 3 to better highlight this result (described below in “Figure revisions already carried out”). Lastly, we add a new figure (Figure 5-Supplement 1) to more explicitly analyze the misclassified lines (described below in “New analyses already carried out”).

      Reviewer 3 - Minor Comment 3

      The bortezomib sensitivity of the WT lines used in the last experiments was determined and did not seem to be greater than parental. This could be mentioned in the text; the figure raises the question and the answer is provided, but it's in the supplemental material.

      While the Bortezomib Signature correctly characterized the bortezomib sensitivity of most clones, it consistently misclassified others (WT10, WT15, and BZ06) (Fig 5-Supplement 1 A). Proliferation assays conducted in earlier experiments showed that WT10 and WT15 were sensitive to bortezomib while BZ06 was resistant (Fig. 1-Supplement 2 A and B). By comparing these incorrect predictions with high-confidence correct predictions, we observed differences that varied by clone type, suggesting unique morphology may be driving each of these misclassifications (Fig. 5-Supplement 1 B and C). These results are consistent with the Bortezomib Signature being generalizable to clones not included in the training dataset and suggest that morphological profiling has the potential to identify bortezomib-resistant clones based on the morphological features of cells in the absence of drug treatment.

      1. [Text revision] We clarified that the metrics (accuracy and average precision) were based on median Bortezomib Signature scores of all replicate well-level profiles per clone. We can compare samples based on rank, and difference from 95% confidence interval of permuted data. There is no current way for our method to assign a likelihood. Also note that we have updated the discussion to discuss alternative metrics (see Reviewer 1 - Minor Comment 7) These are very important distinctions, and we are grateful to the reviewer for bringing them up.

      Reviewer 3 - Major Comment 3

      The study classifies cells as binary sensitive or resistant, but would results be improved by scoring based on likelihood of being resistant/sensitive?

      Reviewer 3 - Minor Comment 2

      It is not clear whether the accuracy was based on a percentage of replicates per cell line that were classified correctly or whether that was referring to classification of the cell line overall as sensitive/resistant.

      • We next examined whether the Bortezomib Signature was able to predict the bortezomib resistance of a clone based on morphological profiling data (Fig. 3 A-E and Fig. 3-Supplement 2 A and B). We called the clone bortezomib-resistant if the median Bortezomib Signature of all replicate well profiles was greater than zero and bortezomib-sensitive if the median Bortezomib Signature less than zero. In the training dataset, the Bortezomib Signature correctly predicted the bortezomib resistance of all ten clones, with median Bortezomib Signatures for eight out of ten clones beyond the 95% confidence interval for the randomly permuted data (Fig. 3 A). The accuracy of the Bortezomib Signature was 88% while the average precision was 81% for the training dataset (Fig. 3-Supplement 2 A and B) (see Methods). The signature performed similarly well in the validation dataset (Fig. 3 B), with an accuracy of 92% and an average precision of 89% (Fig. 3-Supplement 2 A and B). In the test dataset the Bortezomib Signature correctly predicted the bortezomib resistance of all clones, though only HCT116 parental cells had a median Bortezomib Signature outside the 95% confidence interval for the randomly permuted data (Fig. 3 C). The test dataset had an accuracy of 80% and an average precision of 68% (Fig. 3-Supplement 2 A and B). Similarly, in the holdout dataset the Bortezomib Signature had an accuracy of 78% and an average precision of 69% (Fig.3 -Supplement 2 A and B), and correctly predicted the bortezomib resistance of twelve out of thirteen clones, with WT01 misclassified as bortezomib-resistant (Fig. 3 D). In the holdout dataset, four of the twelve correctly characterized clones had median Bortezomib Signatures outside the 95% confidence interval for the randomly permuted data.

      We also mirrored language when discussing the ixazomib and CB-5083 results.

      • However, only two of the four correctly identified ixazomib-resistant clones and one of the three CB-5083-resistant clones had median Bortezomib Signatures outside the 95% confidence interval of the randomly permuted data. The area under the ROC (AUROC) curve for ixazomib-resistant and CB-5083-resistant clones (0.63 and 0.60, respectively) was lower than those calculated for the training, validation, test, and holdout datasets. In addition, many of the Bortezomib Signatures for well profiles of ixazomib- and CB-5083-resistant clones, particularly those for CB-5083-resistant clones, landed within the 95% confidence interval of the randomly permuted data. These results suggest that the Bortezomib Signature is not a general signature of UPS-targeting drug resistance and instead has some specificity for bortezomib.

      • [Text revision] We added an explicit note that our image analysis pipelines are also publicly available. Our reporting of our data processing pipelines are documented fully and well above standards in our field. Linking the publicly-available resources with these methods maximizes reproducibility.

      Reviewer 1 - Minor Comment 10

      Additional details on the processing steps in the analysis pipeline in the Methods will be highly appreciated.

      We include all image analysis pipelines at https://github.com/broadinstitute/profiling-resistance-mechanisms (G. Way et al. 2023).

      1. [Text revision] We have compared our approach to the on-disease/off-disease scores as introduced in (Heiser et al. 2020). We agree with the reviewer that a discussion of these two methods would help clarify our phenotypic signature concept. The on/off score is about the degree to which a perturbation pushes disease towards a healthy state. In this case we have 3 sets of data: healthy samples (used for training), disease samples (used for training), and the sample we want to score, which should be of the form "disease + perturbation". With our approach, based on singscore, we also have 3 sets of data: sensitive samples (used for training), resistance samples (used for training), and the sample we want to score. Here, our sample we want to score could be anything, not necessarily of the form "resistance + perturbation". Furthermore, singscore does not have the concept of orthogonality to resistance/sensitivity. This would become relevant if we were exploring perturbations or conditions that would induce a resistant cell line to become sensitive, but we are not doing that here. There are other statistical differences (projection vs. rank based etc.) but the key difference is the applicability of the method to the specific problem at hand.

      Reviewer 1 - Minor Comment 7

      How is the Bortezomib Signature related to the "on-disease"/"off-disease" scores described in https://www.biorxiv.org/content/10.1101/2020.04.21.054387v1.full? Are there other alternatives used for similar binary phenotypic signatures? What is the justification for using these measurements? I would love to see this generalized concept explicitly discussed in the Discussion.

      We added the following to the discussion.

      • The Bortezomib Signature is conceptually similar to the on-disease/off-disease score (Heiser et al. 2020). Both require three phenotypic measurements: a target phenotype representing ideal, a disease phenotype, and a new phenotype to classify. However, our approach is technically different (non-parametric compared to linear projection) and our goals are different (phenotypic classification compared to perturbation alignment). Other methods also enable phenotype labeling, but they focus on single-sample annotation without regard to a target phenotype (Wawer et al. 2014; Rohban et al. 2017; Simm et al. 2018; Nyffeler et al. 2020).

      Figure revisions already carried out

      1. [Figure revision] We moved all boxplots from the original Fig. 2-Supplement 9 to the main text (also splitting Fig. 2 into Fig. 2 and 3). From the original Figure 2, we moved the accuracy and average precision bar graphs to the supplement. We also note that this change increases transparency of the discriminative signal of our signature.

      Reviewer 1 - Minor Comment 8

      I would highly recommend showing the Bortezomib Signatures from Figure 2-Supplement 9. in Fig. 2. This was the main measurement used throughout the manuscript and in my opinion, it is very important to consistently visualize the data along the manuscript, for clarity and easier reader interpretation.

      1. [Figure revision] We adjusted the position of the legend in the accuracy and average precision bar graphs (originally Fig. 2 C and D, now Fig. 3-Supplement 2) for clarity. We also note that keeping the bar chart here is standard best practice (compared to a dot plot).

      Reviewer 1 - Minor Comment 4

      I found the visualization in Fig. 2C-D not intuitive (it is properly explained in the legend). I suggest replacing the accuracy colorbar with a color marker to make it more distinct from the random permutation (|--*--|) The location of the text "mean +- SD of 100 random permutation" made me first think that it is linked to the holdout.

      1. [Figure revision] We changed the point distribution in the boxplots (from expanded to standard) to minimize overlap with the boxplot lines. We also updated the legend text to indicate that individual points in boxplots represent the Bortezomib Signature for well profiles. Note, we paste a representative example of this change above (new Figure 3).

      Reviewer 3 - Minor Comment 1

      I found the box plots somewhat difficult to interpret (especially where the WT lines had a lot of overlap with the red shaded area). Do the points in these charts correspond to replicate wells?

      We also update the figure legend.

      • Plots show values for individual well profiles (points), range (error bars), 25th and 75th percentiles (box boundaries), and median.

      • [Figure revision] [Response to Reviewer 2 - Major Comment 7] We thank the reviewer for allowing us an opportunity to clarify the mechanism. We feel that it is beyond scope of this manuscript to disentangle the molecular alterations that cause bortezomib resistance based on our Cell Painting insights. This wet lab experimental process is arduous and cost prohibitive, and we argue that one of the benefits of taking a morphology approach to resistance status is that we can detect resistant cells (and therefore cells that won’t die when presented with a treatment) without knowing the molecular mechanism.

      Nevertheless, the reviewer has encouraged us to enhance the ability for a reader to view and interpret the signature to perhaps more easily facilitate future work. Previously, we presented our signature in text form in Figure 2-Supplement 4 and in heatmap form in Figure 2-Supplement 5. Here, we add a new figure (Figure 2-Supplement 6; pasted below) which will improve interpretability.

      Reviewer 2 - Major Comment 7:

      Next to feature importance, the authors do not discuss (or I missed) what biology the features represent. Such the reader is left wondering what the actual mechanism of bortezomib resistance could be and if cell painting could shed light on the molecular alterations that cause the treatment resistance. While reviewing, I thus wondered which audience the authors targeted with their manuscript. A more focused analysis of their data that highlights aspects of the study either for the machine learning community, the cell biology community, or the precision oncology community would greatly benefit the manuscript's impact. In its current form, the study's findings seem diluted and spread across a wide range of research questions.<br />

      • Figure 2-Supplement 6. Bortezomib Signature visualized by CellProfiler features. Visualization of CellProfiler features contributing to the Bortezomib Signature. Features with high values (mean signature estimates) in resistant cells are purple while features with low values in resistant cells are green. The mean signature estimates were based on Tukey's Honestly Significant Difference test score and the number in each box represents the number of features used to calculate the mean signature estimate.

      Additionally, we add the following to the results section:

      • We then examined the grouping of features across compartments and channels and found radial distribution features were higher in resistant cells (Fig 2-Supplement 6).

      The code change to generate the signature visualization summary is available at: https://github.com/broadinstitute/profiling-resistance-mechanisms/pull/131

      New analyses already carried out

      1. [New analysis] [Response to Reviewer 2 - Major Comment 5] We agree that a systematic analysis of feature selection methods will provide additional insights not already in the manuscript. Therefore, we have performed two new computational experiments to compare our linear modeling feature selection approach against other standard approaches. We demonstrate that our linear modeling approach is effective at isolating the core differences between resistant and sensitive classes.

      Specifically, we performed two analyses: A) UMAP and B) k-means cluster analysis. We analyzed profiles defined by four different feature selection approaches: 1) Using all traditional CellProfiler features; 2) Using the traditional CellProfiler feature selection approach (removing low variance features, high correlating features, etc.); 3) Using 45 random features (same size as Bortezomib Signature); and 4) Using only the bortezomib signature features. We performed Fisher’s exact tests to derive odds ratios of cluster membership by resistance status and calculated Silhouette widths to quantify relative proximity of clusters.

      This analysis generates a new supplementary figure (see below), and demonstrates that the linear-modeling-based feature selection isolated the features driving the differences between the clone types (resistance vs. wildtype) while the standard approaches do not as effectively separate.

      Reviewer 2 - Major Comment 5:

      A fascinating bit of the manuscript is the description of the feature selection from the screen is done systematically, considering the technical and biological variability and technical artifacts and modeling covariates using linear models seems a very appropriate way of doing so and could serve as another proof of concept that this is indeed the most robust way of modeling and removing signal of technical covariates from the data. Yet, I wondered why the authors do not discuss other means of feature selection or dimensionality reduction; further, they need to show how the features cluster the cell lines or why impact (information content) different features deliver. For an audience interested in the technical aspects of cell painting analysis and machine learning based on the data, that would, IMHO, be the most exciting questions.

      • Figure 3-Supplement 3. Benchmarking linear-modeling feature selection to separate clones by bortezomib resistance. Uniform Manifold Approximation and Projection (UMAP) analysis of the qualitative separability of (A) resistance status and (B) Bortezomib Signature scores across four different feature spaces. (C) k-means clustering from k=2 to k=14 of average odds ratio, maximum odds ratio (Fisher’s exact test), and Silhouette width using Bortezomib Signature features.

      Additionally, we add the following to the results section:

      • We then compared our linear-modeling approach to feature selection against other feature spaces and found that the Bortezomib Signature clusters same-type clones (bortezomib-resistant vs. bortezomib-sensitive) with higher enrichment compared to the full feature space, standard feature selection (see Methods), or a random selection of 45 features (Fig 3-Supplement 3).

      And methods section, describing this analysis:

      • We were also interested in comparing the ability of different feature spaces to cluster clones of the same type (resistant vs. sensitive). This analysis would determine if the Bortezomib Signature features, which we derived using linear modeling to isolate biological from technical variables, had a greater ability to cluster. We compared the Bortezomib Signature against three other feature spaces: 1) the full feature space, 2) standard feature selection (see Image data processing methods), and 3) 45 randomly selected features. We performed two analyses using these four feature spaces including Uniform Manifold Approximation and Projection (UMAP) (McInnes et al. 2018) and k-means clustering. For UMAP, we used default umap-learn parameters to identify two UMAP coordinates per feature space. We then visualized the clusters by their resistance status and Bortezomib Signature score. The UMAP analysis represents a qualitative analysis. Next, we applied k-means clustering with 25 initializations across a range of 2-14 clusters (k). Prior to clustering and for each feature space, we applied principal component analysis (PCA) and transformed each feature space into 30 principal components. This step was necessary to compare k-means clustering metrics, which are sensitive to the feature space dimensionality. We applied a Fisher’s exact test to each cluster using a two-by-two contingency matrix that specified cluster membership for each clone classification (resistant vs. sensitive). We visualized the mean odds ratio and max cluster odds ratio for each feature space across k. A high odds ratio tells us that the feature space effectively clusters clones of the same resistance status. Lastly, we calculated Silhouette width (the average proximity between samples in one cluster to the second nearest cluster) for each feature space across k.

      The code change to derive the UMAP coordinates, perform clustering, and generate the figure is available at https://github.com/broadinstitute/profiling-resistance-mechanisms/pull/132

      1. [New analysis] [Response to Reviewer 3 - Major Comment 1] We thank the reviewer for this suggestion, which allowed us to explore the misclassified samples in more depth. We added a new supplementary figure in which we summarized all bortezomib clones (wildtype and resistant) in their accuracy based on the bortezomib signature (panel A). We did not include training set samples in this analysis. Using samples that were consistently incorrectly classified with high confidence (three samples: WT15, BZ06, WT10) we performed two separate two-sample Kolmogorov–Smirnov (KS) tests. Specifically, we compared high incorrect wildtype to high correct wildtype and high incorrect resistant to high correct resistant. Our results indicate that most bortezomib signatures were significantly different between correct and incorrect assignments (panel B), and that the signature features varied between resistant and wildtype misclassification tests (panel C).

      Reviewer 3 - Major Comment 1:

      While the claims are largely substantiated, there are a few points where further consideration would improve the manuscript. Several cell lines were mis-classified with what appears to be a high degree of certainty. Can the authors tell what was driving those predictions? Was there something in the morphological signature that weighed more heavily in those cases?

      • Figure 5-Supplement 1. Examining the accuracy of clone classification and misclassification of clones. (A) Proportion of high-confidence correct, low-confidence correct, low-confidence incorrect, and high-confidence incorrect predictions of well profiles across clones in the test, holdout, and validation sets. High-confidence predictions (high) had a Bortezomib Signatures greater (resistant clones) or less than (sensitive) the 95% confidence interval of randomly permuted data while low-confidence predictions (low) had Bortezomib Signatures within the 95% confidence interval of randomly permuted data. (B) Visualization of Kolmogorov-Smirnov (KS) test statistic means of feature groups across channels and cellular compartments. (C) Plot of the KS test statistic means for feature groups in bortezomib-resistant vs. -sensitive cells. Each feature group is color coded by the imaging channel.

      Additionally, we add the following to the results section:

      • While the Bortezomib Signature correctly characterized the bortezomib sensitivity of most clones, it consistently misclassified others (WT10, WT15, and BZ06) (Fig 5-Supplement 1 A). Proliferation assays conducted in earlier experiments showed that WT10 and WT15 were sensitive to bortezomib while BZ06 was resistant (Fig. 1-Supplement 2 A and B). By comparing these incorrect predictions with high-confidence correct predictions, we observed differences that varied by clone type, suggesting unique morphology may be driving each of these misclassifications (Fig. 5-Supplement 1 B and C). These results are consistent with the Bortezomib Signature being generalizable to clones not included in the training dataset and suggest that morphological profiling has the potential to identify bortezomib-resistant clones based on the morphological features of cells in the absence of drug treatment.

      And methods section, describing this analysis:

      Some profiles were consistently predicted incorrectly with high confidence but in the opposite direction (see Figure 5-Supplement 1). For a well-level profile to be categorized as high-confidence (in either the correct or incorrect directions), it needed to score beyond the 95% confidence interval of the randomly permuted data range. For example, a high-confidence incorrect resistant profile would have a Bortezomib Signature below 95% confidence interval of the randomly permuted data. To evaluate the features driving the differences in these samples, we applied two-sample Kolmogorov–Smirnov (KS) tests per Bortezomib Signature feature. We applied these tests to two separate groups: 1) misclassified bortezomib-sensitive vs. high-confidence accurate bortezomib-sensitive and 2) misclassified bortezomib-resistant vs. high-confidence accurate bortezomib-resistant.

      The code change to generate the UMAP coordinates and figure is available at https://github.com/broadinstitute/profiling-resistance-mechanisms/pull/130

      Description of analyses that authors prefer not to carry out

      1. [Response to Reviewer 2 - Minor Comments 1 and 2]: These are interesting suggestions! Still, we prefer not to speculate on the biological mechanism of the Bortezomib signature. Connecting morphological features identified as contributing to the Bortezomib Signature by Cell Painting to specific biological pathways would demand considerable cell-based assays to validate. In addition, our analyses suggest that the features contributing to the Bortezomib Signature are spread across a range of cellular compartments and channels, making it difficult to pin down specific mechanisms or pathways as likely contributors to bortezomib resistance. However, we are adding a figure to increase interpretability of the signature, which will aid in developing future hypotheses. Note that the signature was not possible to detect by eye (Fig. 2 A).

      Reviewer 2 - Minor Comment 1:

      There could be some speculation on the mechanism of Bortezomib resistance concerning the literature with the existing image data. For example, Bortezomib resistance is connected to serine synthesis and how a particular feature could contribute to the known mechanism.<br />

      Reviewer 2 - Minor Comment 2:

      Along the same lines, the authors could show that larger cells lead to resistance with microscopic images.

      2. [Response to Reviewer 2 - Major Comment 8]: We appreciate the reviewer’s concern that our work using HCT116 clonal cells lines may not directly reflect results from patient samples. Our choice was based on previously published work demonstrating the efficiency with which HCT116 cells generate resistant clones due to diminished DNA mismatch repair and decreased expression of drug efflux pumps. Since our work is a proof of concept rather than a comprehensive demonstration of translating morphological profiling into clinical practice, we believe that experiments using multiple patient cell lines from different tissues as well as digital pathology records to be beyond the scope of this work. We instead chose to tone down the language of our manuscript to more clearly acknowledge the limitations of our work and clarify this as a proof of concept.

      Reviewer 2 - Major Comment 8 (relevant excerpt):

      I suggest the authors test their approach on at least two other cell lines (maybe from different tissues) and benchmark their results against a dataset of digital pathology where such predictions are made from stained and analyzed tissue slices. This way, after a thorough benchmark against related third-party data sets, the method would significantly gain relevance, the paper would appeal to a broader audience, and the advance gains more merit.<br />

      3. [Response to Reviewer 3 - Major Comment 2]: The bortezomib sensitivity of ixazomib- and CB-5083-resistant clones was not determined, and hence can not be ruled out as a possible explanation for their high Bortezomib Signature scores. However, we prefer not to conduct additional proliferation assays for the misclassified clones (IX02, WT06, CB14, CB16) in the presence of bortezomib to determine whether coincidental bortezomib resistance might explain the signature performance. Our rationale is that three other misclassified clones (WT10, WT15, and BZ06) had the expected bortezomib sensitivity in proliferation assays (Fig. 1-Supplement 2), meaning that additional proliferation assays may not reveal any insights regarding the signature performance.

      Reviewer 3 - Major Comment 2:

      Was the bortezomib sensitivity of the IX (or CB) resistant cell lines determined? If there were differences, this could explain some of the variation in the morphological signatures. This could be easily done in one or two growth experiments.

      4. [Response to Reviewer 2 - Major Comment 7]: Thank you for pointing this out. Our goal is to keep the study multi-disciplinary. We are adding a figure to increase interpretability of the signature, and adding text-based clarifications.

      Reviewer 2 - Major Comment 7 (relevant excerpt):

      While reviewing, I thus wondered which audience the authors targeted with their manuscript. A more focused analysis of their data that highlights aspects of the study either for the machine learning community, the cell biology community, or the precision oncology community would greatly benefit the manuscript's impact. In its current form, the study's findings seem diluted and spread across a wide range of research questions.<br />

      5. [Response to Reviewer 2 and 3 - Major Comments 6 and 4]: We prefer not to expand the scope of the model to predict other drug signatures. This would require a substantial amount of work to generate the appropriate drug-resistant clones, collect the imaging data, and analyze it, and we think it important to convey the purpose of our paper is proof of concept. We do not feel that the time invested in performing this analysis would result in adequate returns beyond what we already demonstrate.

      Reviewer 2 - Major Comment 6.

      Interestingly, the Bortezomib signature is specific to the drug and not a broad range of proteasomal inhibitors. However, seeing the common features between all the proteasomal inhibitors would be interesting.

      Reviewer 3 - Major Comment 4

      There was some predictive ability of the Bortezomib Signature for ixazomib resistance. Were there some features that were correlated with IX-resistance, i.e. UPS pathway, versus specific to bortezomib? Do the features suggest anything about resistance mechanisms or is the feature set too abstruse to interpret?

      References

      Foroutan, Momeneh, Dharmesh D. Bhuva, Ruqian Lyu, Kristy Horan, Joseph Cursons, and Melissa J. Davis. 2018. “Single Sample Scoring of Molecular Phenotypes.” BMC Bioinformatics 19 (1): 404.

      Heiser, Katie, Peter F. McLean, Chadwick T. Davis, Ben Fogelson, Hannah B. Gordon, Pamela Jacobson, Brett Hurst, et al. 2020. “Identification of Potential Treatments for COVID-19 through Artificial Intelligence-Enabled Phenomic Analysis of Human Cells Infected with SARS-CoV-2.” bioRxiv. https://doi.org/10.1101/2020.04.21.054387.

      McInnes, Leland, John Healy, Nathaniel Saul, and Lukas Großberger. 2018. “UMAP: Uniform Manifold Approximation and Projection.” Journal of Open Source Software 3 (29): 861.

      Nyffeler, Johanna, Clinton Willis, Ryan Lougee, Ann Richard, Katie Paul-Friedman, and Joshua A. Harrill. 2020. “Bioactivity Screening of Environmental Chemicals Using Imaging-Based High-Throughput Phenotypic Profiling.” Toxicology and Applied Pharmacology 389 (January): 114876.

      Rohban, Mohammad Hossein, Shantanu Singh, Xiaoyun Wu, Julia B. Berthet, Mark-Anthony Bray, Yashaswi Shrestha, Xaralabos Varelas, Jesse S. Boehm, and Anne E. Carpenter. 2017. “Systematic Morphological Profiling of Human Gene and Allele Function via Cell Painting.” eLife 6 (March). https://doi.org/10.7554/eLife.24060.

      Simm, Jaak, Günter Klambauer, Adam Arany, Marvin Steijaert, Jörg Kurt Wegner, Emmanuel Gustin, Vladimir Chupakhin, et al. 2018. “Repurposing High-Throughput Image Assays Enables Biological Activity Prediction for Drug Discovery.” Cell Chemical Biology 25 (5): 611–18.e3.

      Wacker, Sarah A., Benjamin R. Houghtaling, Olivier Elemento, and Tarun M. Kapoor. 2012. “Using Transcriptome Sequencing to Identify Mechanisms of Drug Action and Resistance.” Nature Chemical Biology 8 (3): 235–37.

      Wawer, Mathias J., Kejie Li, Sigrun M. Gustafsdottir, Vebjorn Ljosa, Nicole E. Bodycombe, Melissa A. Marton, Katherine L. Sokolnicki, et al. 2014. “Toward Performance-Diverse Small-Molecule Libraries for Cell-Based Phenotypic Screening Using Multiplexed High-Dimensional Profiling.” Proceedings of the National Academy of Sciences of the United States of America 111 (30): 10911–16.

      Way, Gregory, Yu Han, David Stirling, and Shantanu Singh. 2023. Broadinstitute/profiling-Resistance-Mechanisms: Analysis for Preprint. Zenodo. https://doi.org/10.5281/ZENODO.7803787.

      Way, Gregory P., Maria Kost-Alimova, Tsukasa Shibue, William F. Harrington, Stanley Gill, Federica Piccioni, Tim Becker, et al. 2021. “Predicting Cell Health Phenotypes Using Image-Based Morphology Profiling.” Molecular Biology of the Cell 32 (9): 995–1005.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary

      This study aimed to determine a morphological signature based on CellPainting that could predict resistance of multiple myeloma cells to bortezomib. Bortezombi-resistant clones were derived from parental wild type cells and tested for multidrug resistance. Best practice machine learning methods were applied to identify features that were not correlated with technical variation or nonspecific heterogeneity between cell lines. A Bortezomib Signature comprised of 45 selected features complied using a robust ranked method was described. This signature score was then used to classify cell lines, including wild-type and bortezomib-resistant clones not included in the training set and clonal lines selected for resistance to other drugs. The Bortezomib Signature performed better than chance for predicting resistance to this drug in validation and holdout datasets, and for an independent dataset of clones not used in the initial training. Some predictive power was observed for a drug targeting the same pathway (UPS), but with lower accuracy.

      Major comments

      While the claims are largely substantiated, there are a few points where further consideration would improve the manuscript.<br /> Several cell lines were mis-classified with what appears to be a high degree of certainty. Can the authors tell what was driving those predictions? Was there something in the morphological signature that weighed more heavily in those cases?<br /> Was the bortezomib sensitivity of the IX (or CB) resistant cell lines determined? If there were differences, this could explain some of the variation in the morphological signatures. This could be easily done in one or two growth experiments.<br /> The study classifies cells as binary sensitive or resistant, but would results be improved by scoring based on likelihood of being resistant/sensitive?<br /> There was some predictive ability of the Bortezomib Signature for ixazomib resistance. Were there some features that were correlated with IX-resistance, i.e. UPS pathway, versus specific to bortezomib? Do the features suggest anything about resistance mechanisms or is the feature set too abstruse to interpret?<br /> It is not clear from the Discussion whether this type of analysis is more broadly applicable to cell lines derived from patients, rather than selected from a parental cell line, or if this approach would be more efficient than genotyping or next-gen sequencing. How many replicates and ground truth cell lines would be necessary for predictive confidence?

      Minor comments

      I found the box plots somewhat difficult to interpret (especially where the WT lines had a lot of overlap with the red shaded area). Do the points in these charts correspond to replicate wells?<br /> It is not clear whether the accuracy was based on a percentage of replicates per cell line that were classified correctly or whether that was referring to classification of the cell line overall as sensitive/resistant.<br /> The bortezomib sensitivity of the WT lines used in the last experiments was determined and did not seem to be greater than parental. This could be mentioned in the text; the figure raises the question and the answer is provided, but it's in the supplemental material.<br /> Ten lines of text are duplicated on page 5.

      Significance

      The strengths of this study are the machine learning best practice and detailed methodology. The experiments could be reproduced and statistical analysis is more than adequate. The analysis takes into account batch effects, well position, differences in cell numbers, and other sources of technical variation that complicate high-content image analysis. It is a good exemplar of how unsupervised morphological profiling can be applied to imaging data. The major limitation is the generalizability of this particular method for patient samples. This could be addressed in the Discussion.

      The advances presented here are largely technical, as machine learning best practices are implemented to address a specific type of drug resistance.

      This study would be especially valuable to computationally-oriented biologists and has some translational appeal, which could be broadened by including some more information in the Discussion regarding the logistics of this approach in the clinic and whether the signatures identified have interpretability as well as predictive power.

      Areas of expertise: image analysis, systems biology, cancer biology

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary:

      In the study reviewed herewith, Kelley et al. present an interesting approach to evaluate whether image-based profiling via the cell painting assay can enable a model to predict if a cell line is resistant to the anti-cancer drug bortezomib (a specific inhibition of the proteasome). Therefore, they employed the human colorectal cancer-derived derived cell line HCT116 and used a previously published selection protocol to derive bortezomib resistant clonal subpopulation. These sub-populations were then subjected to cell painting, and it was tested (by dividing the data set into training, test, validation, and holdout) if a machine learning model could be trained on selected image-derived features that would predict if the cell line is indeed resistant to bortezomib treatment. This worked in 7 out of ten cases and two commercially available resistant HCT116 clones. Theoretically, this implies that image-based cellular profiling of tissue materials could work as a fast and reliable method to gauge the treatment resistance of cancer patients before treatment and open new avenues for personalized medicine.

      Major comments:

      • the paper is well written, and the text is clear, as is the presentation of data and transparency of methods being utilized.
      • the methods were applied appropriately and followed established standards in the field.
      • The paper's premise is timely and interesting, addressing a pressing issue in cancer therapy: making informed treatment decisions fast, based on markers found in tumors early in tumor development, and using image-based screening for characterizing drug resistance before treatment could be an option.
      • A fascinating bit of the manuscript is the description of the feature selection from the screen is done systematically, considering the technical and biological variability and technical artifacts and modeling covariates using linear models seems a very appropriate way of doing so and could serve as another proof of concept that this is indeed the most robust way of modeling and removing signal of technical covariates from the data.
      • Yet, I wondered why the authors do not discuss other means of feature selection or dimensionality reduction; further, they need to show how the features cluster the cell lines or why impact (information content) different features deliver. For an audience interested in the technical aspects of cell painting analysis and machine learning based on the data, that would, IMHO, be the most exciting questions.
      • Interestingly, the Bortezomib signature is specific to the drug and not a broad range of proteasomal inhibitors. However, seeing the common features between all the proteasomal inhibitors would be interesting.
      • Next to feature importance, the authors do not discuss (or I missed) what biology the features represent. Such the reader is left wondering what the actual mechanism of bortezomib resistance could be and if cell painting could shed light on the molecular alterations that cause the treatment resistance. While reviewing, I thus wondered which audience the authors targeted with their manuscript. A more focused analysis of their data that highlights aspects of the study either for the machine learning community, the cell biology community, or the precision oncology community would greatly benefit the manuscript's impact. In its current form, the study's findings seem diluted and spread across a wide range of research questions.
      • Lastly, I find some misfits between the question, the model used, and the conclusions drawn. The authors start by exploring the problem of bortezomib resistance in cancer treatment, which they say is a devastating issue for patients with, e.g., multiple myeloma. Yet, the authors use HCT116 as their model cell line, a microsatellite instable, colorectal cell line with several intrinsic mutations that make it a difficult model to address physiologically relevant medical problems after all. The authors then go on to suppose that their method might be suitable to diagnose resistance in patient samples, but I am not convinced this conclusion can be speculated based on data from HCT cells. I suggest the authors test their approach on at least two other cell lines (maybe from different tissues) and benchmark their results against a dataset of digital pathology where such predictions are made from stained and analyzed tissue slices. This way, after a thorough benchmark against related third-party data sets, the method would significantly gain relevance, the paper would appeal to a broader audience, and the advance gains more merit.

      Minor comments

      • There could be some speculation on the mechanism of Bortezomib resistance concerning the literature with the existing image data. For example, Bortezomib resistance is connected to serine synthesis and how a particular feature could contribute to the known mechanism.
      • Along the same lines, the authors could show that larger cells lead to resistance with microscopic images.
      • The authors talk about validating the mutation - PSMB5 by RNA-seq. However, the data for the genotyping/sequencing/characterization of these newly generated BZ-resistant lines are missing.
      • on page 5, paragraph 4, there is a sizeable copy-and-paste error of text being identically replicated.
      • What was the LD 90 per drug on HCT cells? Rather than LD90 foldchanges, absolute concentrations should be used in the results and discussion to allow the reader to vet the conclusions.
      • The description of the resistant clonal populations is confusing. As I understand, no single-cell clones were isolated during the selection procedure. Thus, the training lines are not yet isogenic clones but oligoclonal sub-populations of the parental cell line. The authors could provide more details here and discuss the different characteristics of their sub-populations, e.g., their growth kinetics or molecular alterations.

      Referees cross-commenting

      I agree that, perhaps, my major criticism of the paper was the manuscript's 'overselling' of claims that were only weakly supported by the data. Yes, if the authors tune down their claims and clearly state that this is an interesting starting point and proof of concept study, it might be ok to publish with only minor revisions. If the claims should be more generalized, then this study needs more data supporting the conclusions and the method's predictive power.

      Significance

      In its current form, the manuscript describes an incremental technical advance that addresses an interesting premise for precision oncology but explores it on a single cell line. The manuscript is timely in that it adds to a growing body of research around the utilization of image-based profiling in various areas of biomedical research using cell painting to create reference datasets for training advanced machine learning models for diagnosis. This research might be interesting and relevant for cell biology, precision medicine, oncology, image-based profiling, and machine-learning communities.

      My expertise includes image-based profiling, cell-based screening, assay development, functional genomics, cancer research, drug discovery, machine learning, and data science. There is no aspect of the study I cannot review.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      The authors use Cell Painting, a high-content image-based phenotypic assay, to distinguish between clonal cancer cell lines that are resistant versus sensitive to a proteasome inhibitor anti-myeloma drug called bortezomib. The authors characterized a high-dimensional cell morphology signature for bortezomib-resistance, evaluated it on an independent subset of cell lines, and evaluated specificity in respect to other drugs targeting the ubiquitin-proteasome system. The authors thus propose image-based morphology characterization as an alternative method for characterizing drug resistance.

      Strengths: solid methodology - cell lines validation of drug resistance, extensive data collection, thorough validation of the analysis pipeline, avoiding potential confounders, biases and proper data partitioning to test and hold-out (what the authors refer to as "machine learning best practices").

      Weakness: weak discriminative signal. Some aspects of the writing could be improved to make the manuscript easier to follow (see Minor comments).

      Major comments:

      While I am convinced that the signature captures morphological phenotypes associated with drug resistance, at the cumulative scale, the discriminative signal of a single cell type seems weak. Specifically, it is not clear whether the signature can effectively capture the drug resistance of a single cell line. In Figure 2-Supplement 9, considering the test (C) and the holdout (D), only 1/9 BZ clones' median signatures were beyond the 95% confidence interval, with 4/6 and 2/6 WT cell types with median signatures beyond the positive and negative 95% confidence interval correspondingly. When defining bortezomib-sensitivity according to the median signatures' sign (>0 or <0) of a cell line, Figure 2-Supplement 9 shows that in the test+holdout there are 9/9 correct bortezomib-resistance (BZ) and 6/7 correct bortezomib-sensitive (WT) predictions. However, similar discrimination levels also appeared in the other drugs (ixazomib, CB-5083), making the statements about specificity less grounded. When the authors evaluate the AUROC they report ~0.6 (line #194) for the non-specific (ixazomib, CB-5083) drugs versus ~0.75 for bortezomib-resistance (line #202). With Fig. 4, the data fully supports the argument that the bortezomib-signature encodes bortezomib-resistance, but the signal is weak. Thus statements such as "We found the Bortezomib Signature could predict whether a cell line was bortezomib-resistant or bortezomib-sensitive" (line #172) and the specificity statements in the abstract" (line #28) are not supported by the data in my opinion. I would recommend the authors to tune down these and other related statements throughout the manuscript. An alternative would be to increase the number of wells and see whether this weak signal can indeed be statistically amplified with many replicates to make a robust and specific characterization of a cell line's bortezomib-sensitivity (but I assume this is a lot of work and probably out of scope of this manuscript). I think it is also important to discuss in more detail the interpretation of these results (including Figure 2-Supplement 9), in this context, in the Discussion.

      Minor comments:

      Suggested clarifications (some might be less relevant if the manuscript is designed for experts in the more clinical domain who are familiar with these terms / style):

      1. More information in Fig. 1's legend would be helpful to follow the experimental design. I found it hard to follow in its current form and had to go back to carefully reading the main text to fully understand.
      2. What is LD90 (line #87)? LD50 (line #97)?
      3. It was not clear to me in the text which and how many cell lines were evaluated and the reader is forced to go to the SI. For example, "(BZ01-10 and BZ clones A and E)" (line #96-97) and "wild-type clones (WT01-05, 10, and 12-15)" (line #98) appeared when presenting the results without a clear explanation and made it harder for me to follow. Summary of the data (for example, based on Figure 2-Supplement 8) can be briefly mentioned in the text to make it more clear for the reader.
      4. I found the visualization in Fig. 2C-D not intuitive (it is properly explained in the legend). I suggest replacing the accuracy colorbar with a color marker to make it more distinct from the random permutation (|--*--|) The location of the text "mean +- SD of 100 random permutation" made me first think that it is linked to the holdout.
      5. Lines #104-111 were duplicated in lines #114-122.
      6. The "Bortezomib Signature" is a critical measurement but is only briefly mentioned in lines 150-151 ("..based on the direction-sensitive ranking method for phenotype analysis, singscore (Foroutan et al., 2018)"). Please provide more information/intuition.
      7. How is the Bortezomib Signature related to the "on-disease"/"off-disease" scores described in https://www.biorxiv.org/content/10.1101/2020.04.21.054387v1.full? Are there other alternatives used for similar binary phenotypic signatures? What is the justification for using these measurements? I would love to see this generalized concept explicitly discussed in the Discussion.
      8. I would highly recommend showing the Bortezomib Signatures from Figure 2-Supplement 9. in Fig. 2. This was the main measurement used throughout the manuscript and in my opinion, it is very important to consistently visualize the data along the manuscript, for clarity and easier reader interpretation.
      9. Line #155-156: "We found that the pattern of Bortezomib Signatures corresponded to the cell identity plate layout", the word "not" is missing before "corresponded".
      10. Additional details on the processing steps in the analysis pipeline in the Methods will be highly appreciated.

      Referees cross-commenting

      My main critic is regarding "over selling" a weak discriminative signal. Specifically, I am not convinced that the major claims regarding predicting sensitivity and specificity at the single cell types scales are supported by the data. Since reviewer #2 and #3 did not raise this concern I think it is worth discussion here.

      Once these statements are tuned down - I think no significant additional work is needed to make the point that they can measure a discriminative signal. If they want to make these claims, perhaps they'd like to collect more data to gain statistical power (but I am not optimistic this will work at the single cell level).

      Personally, I was happy with the authors' choice of cell lines not included in the training dataset. I am not convinced that additional cell lines + validations are necessary for making the point of a proof of principle.

      Significance

      Cell Painting was applied to many applications, but as far as I am aware this is the first attempt for an image-based phenotypic characterization of drug resistance. While the authors established that this approach can measure, to some extent, bortezomib-sensitivity, at the current state of the results, I am not convinced that cell painting can be practically used to assess bortezomib-sensitivity of a single cell line. However, this does not imply that the same approach can not achieve the goal, perhaps by using other cell painting markers for bortezomib-sensitivity, or with the same markers to assess sensitivity of different drugs. The cell painting + analysis approaches are not new and the clinical impact is questionable, but the technical aspects (data, analysis) are exceptional and the concept may hold as I described above.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity):

      Summary: Forer and Otsuka provide first-rate evidence for tethers fixed in place between separating anaphase chromosomes using electron tomography. The authors traced the anaphase movement of a number of living cells before fixation for examination using electron tomography. The manuscript is clearly written and provides an excellent introduction and discussion of the known literature. The reader will have an excellent background to see the importance of this work.

      Major comments:<br /> - Are the claims and the conclusions supported by the data or do they require additional experiments or analyses to support them?

      No further experiments are needed. The data are very supportive, and extremely clear.<br /> - Are the data and the methods presented in such a way that they can be reproduced? Yes.<br /> - Are the experiments adequately replicated and statistical analysis adequate? Yes.

      Minor comments:<br /> - Are prior studies referenced appropriately? Yes.<br /> - Are the text and figures clear and accurate? Absulotely.<br /> - Do you have suggestions that would help the authors improve the presentation of their data and conclusions?

      The authors are to congratulated on their major contribution to this study on tethers between separated daughter chromosomes. It is a tpur deforce to go from the living cells to fixing and identifying the same separated chromosomes using electron tomography to see the ultrastructure of the fibers seen fir.

      Referees cross-commenting<br /> Thank you reviewer #2. The manuscript should be published. It is an excellent contribution.

      We thank the reviewer for the appreciation of the clarity and quality of our work.

      Reviewer #1 (Significance):

      Provide contextual information to readers (editors and researchers) about the novelty of the study, its value for the field and the communities that might be interested.

      This manuscript is the first to use electron tomography to identify the tethers between separated anaphase chromosomes. Forer and the laetMichael Berns and their co-authors have published a number of papers using phase microscopy and lasers to report on the physical nature and elastic properties of these fibres in the past. Forer and Otsuka have presented first-rate evidence for the reality of these structures using electron tomography. This manuscript should highlighted in the published journal.<br /> The chemical identity of these fibers as the authors state is unclear.

      The following aspects are important:

      • Audience: describe the type of audience ("specialized", "broad", "basic research", "translational/clinical", etc...) that will be interested or influenced by this research; how will this research be used by others; will it be of interest beyond the specific field?

      This exciting contribution will be read by anyone interested in mitosis. It will be of interest to all Cell Biologists because of the careful manner in which the living cells were studied before they were fixed for examination using electron tomography. The readers will be dreaming how they can use this process on their Cell Biology problems._

      • Please define your field of expertise with a few keywords to help the authors contextualize your point of view. Indicate if there are any parts of the paper that you do not have sufficient expertise to evaluate.

      I am a cell Biologist who has made contributions, both in light microscopy and in transmission microscopy on diving cells, both in tissue culture and in situ in aviav and zebrafish embryos.

      We thank the reviewer for appreciating the significance of our work.

      Reviewer #2 (Evidence, reproducibility and clarity):

      In this paper, the authors use light microscopy and electron tomography to study anaphase chromosomes in crane fly spermatocytes. They find that there are two "tether" structures that connect telomeres of sister chromatids. One tether is thicker (denser) and extends between sister chromatids during early but not late anaphase, whereas a second, less-dense tether maintains contact with both sister chromatids in all examined stages of anaphase. The paper makes arguments as to what the tethers could or could not be. Specifically, they are too numerous to be ultrafine DNA bridges seen in various normal or abnormal segregation events and they also do not affect anaphase chromosome motion the same way ultrafine DNA bridges do.

      Major comments:<br /> The major claim that there are tethers that connect sister chromatids in anaphase is supported by the data. Moreover, the data resolves two types of tethers on the basis of their density. While it is unclear what the composition of the tethers are, the paper makes a convincing case that they cannot be the DNA ultrafine bridges seen in other studies. The discussion has sufficient caveats that most readers will see that more work is needed to identify the composition of the two tethers. In my opinion, no further experiments are needed to support the modest claims of this paper. Therefore, I only have minor comments that may hopefully improve the paper's clarity.

      We thank the reviewer for the positive evaluation of our work.

      Minor comments:<br /> It was argued that the tethers reported here were also seen in other species and cellular contexts, where the imaging work was done with projection EM imaging. Presumably, what is new here is the usage of electron tomography. It would help readers if the authors explained why the electron tomography done here was essential to arrive at key conclusions.

      Thank you for the useful comment. We have added the explanation of why electron tomography was critical to visualise small tether structures to the last paragraph of the Discussion on page 7.

      p.3 mitochondria appeared to be fixed properly ... (e.g., Figs. 1C, 2B) - I don't see any mitochondria in any figures. Perhaps this observation should be noted as "not shown"?

      We thank the reviewer for pointing this out. We have added an electron micrograph of mitochondria to the Supplementary Figure 1.

      p.3 The images shown in Figs. 1, 2, 4 - The figures should be called out in the order; in this case, Fig 3 has not been called out yet.

      We have corrected the order of the figures.

      p.4 we did not find any other connecting structures - Because the sample was processed by traditional EM methods, it's safer to add a caveat that other connecting structures could be missed if they were disrupted by sample prep or if they did not pick up stain as well as the two structures presented in this paper.

      We have clarified that our sample was chemically fixed in the first paragraph of the Discussion on page 4. Because the details of how our samples were prepared are described in the Method section, we did not add further details to this paragraph.

      p.7 we expect such structures to be commonly seen in other cell types as well if they are examined carefully - Instead of saying that examinations should be done "carefully", it would be more helpful to specify how other cell types should be examined. This work shows that the bridges can be found if the cells are either sectioned parallel to the spindle axis or if a sufficiently large volume is sampled.

      We have now clarified that 3D electron microscopy techniques such as electron tomography are critical to visualise small tether structures in the last paragraph of the Discussion on page 7.

      Please use consistent spelling/hyphenation of ultrafine/ultra-fine and word choice (strands vs. bridges).

      Referees cross-commenting<br /> I agree with my co-reviewers's comments and have no further suggestions._

      Reviewer #2 (Significance):

      This may be the first use of electron tomography to study the structural details of tethers that connect chromosomes in anaphase cells. The data is of sufficient quality to reveal differences in density. Namely, one class of tether appears to be an extension of the chromosome while the other class is composed of thin filaments. This study is novel in that it characterizes a mitosis-associated complex that is poorly studied compared to the microtubule-based spindle apparatus and the kinetochore. Hopefully, the tethers will draw more attention and further characterization by methods like super-resolution microscopy and cryo-electron microscopy. My expertise is in chromatin, mitotic machines, and cryo-electron tomography.

      We thank the reviewer for appreciating the novelty and the impact of our work.

      Reviewer #3 (Evidence, reproducibility and clarity):

      Summary:

      Tethers between telomeres of chromosomes in anaphase were inferred from earlier studies of laser microbeam cutting experiments. The current paper presents images from electron tomography of crane fly spermatocytes that substantiates the earlier inference. The authors deduce that the darker filaments and the lighter filaments that they visualize may be the structural tethers at telomeres.

      Major comments:

      The experiments are carefully done, and the conclusions are appropriately worded to qualify any caveats. This short communication is well-presented, and I have only a few comments._

      We thank the reviewer for appreciating the clarity and quality of our work.

      The authors should expand their list of references on bridges to include those listed by Warecki et al (Curr Biol 33:1-17, 2023; refs 15-26, etc).

      We do not think it is necessary to expand the list of references for ultra-fine DNA bridges. In the article we submitted, we discussed the Warecki at al article in the penultimate paragraph of the Discussion; we concluded that the bridges that Warecki at al described are different from ours in having so few per cell that they couldn’t be tethers, and further that there was no evidence that those bridges were elastic. For those reasons, we do not find discussion of those proteins relevant to tethers, any more than would listing all the proteins associated with ultra-fine DNA bridges be relevant to the elastic tethers.

      In the Discussion, we discussed data suggesting that a known elastic protein titin was present; that is as far as we wanted to go on speculation of what the elastic component of tethers might be.

      The authors present arguments that the tethers are not the DNA bridges observed by others. However, they should try to address this experimentally by treatment of their preparations with DNase to see if the thick and/or thin filaments disappear.

      While we agree that it would be important to identify the components of the tethers, we are concerned that those experiments are beyond the scope of this manuscript. Nevertheless, we appreciate the constructive suggestion for the future research direction.

      Moreover, they should discuss in more detail the possible functions of (DNA) bridges, including the recent model from Bill Sullivan's lab (Warecki et al, Curr Biol, 2023) that they help to retain fragments of broken chromosomes. In addition, the authors should summarize the various proteins that may be associated with the bridges (as enumerated in the Warecki et al 2023 paper).

      As we describe above, we concluded that the bridges Warecki at al described are different from the tethers that we report in our manuscript. Therefore, we do not think it is necessary to expand the discussion on the proteins and functions associated with ultra-fine DNA.

      The authors could add a sentence to the Results or Discussion of whether the thicker tethers might become stretched as anaphase progresses to become the thinner tethers (Fig. 4G).

      We thank the reviewer for this suggestion. We actually mentioned this possibility in the third paragraph of our Discussion on page 7.

      The authors may want to add a few sentences to the Discussion about the "chromosomal bouquet" stage of leptotene of meiosis prophase I where the telomeres of chromosomes seem pulled together and associate with the nuclear envelope --- they could speculate if this might also be due to the tethers that they describe in spermatocytes.

      This is a very interesting possibility. While we would refrain from adding this speculation to our manuscript as it is beyond the scope of the main points, it is certainly an interesting avenue of future research.

      Minor comments:

      A few additional comments are as follows:

      p. 2 last sentence of first paragraph -modify the wording about "no structural evidence that identifies physical connections between separating telomeres", since there is some information from genetic and cell biology light microscopy experiments. Perhaps simply change "structural" to "ultrastructural".

      We have changed the wording as the reviewer recommended

      p. 6, 5th line of second paragraph - change "ribosome DNA" to "ribosomal DNA"

      We have corrected it.

      Figure 1D - add the chromosome to the right of the schematic model (as suggested by Fig. 1B).

      We are sorry for the confusion. In Figure 1D, the left half of the tethers are 3D modelled and shown. We have clarified this point by modifying the legend of Figure 1D

      p. 17 (Methods), line 10 of first paragraph - state if this is light or heavy Halocarbon oil (give details).

      It is a mixture of heavy and light Halocarbon oil. We have clarified it on page 17.

      p. 17 (Methods), line 12 of first paragraph- state the concentration for fibrinogen and for thrombin.

      As we wrote in the original manuscript, the procedures are described in detail in our previous publication (Forer A. & Pickett-Heaps J. (2005) Fibrin clots keep non-adhering living cells in place on glass for perfusion or fixation. Cell Biology International 29: 721–730). Nonetheless, to clarify this point, we have modified the text on page 17.

      p. 17 (Methods), line 4 of second paragraph - is there any data to show that the filaments (tethers) occur if there is no cold shock?

      Yes, we do see similar filamentous structures in the sample without cold shock. For your information, we show one of the electron micrographs below. In our manuscript, we show the data from the samples prepared with cold shock, because it better visualizes the filamentous structures. We now show these electron micrographs in the Supplementary Figure 2.

      Referees cross-commenting<br /> I concur with Reviewers #1 and #2 that this is a fine paper that should be published. My detailed comments submitted with my review are simply meant as revisions to further strengthen this paper.

      We thank the reviewer for supporting the publication of our manuscript.

      Reviewer #3 (Significance):

      Strengths: This is an important conceptual advance and the carefully done ultrastructural imaging provides the foundation for future studies that could delve into the molecular composition and functional significance of the tethers at telomeres of anaphase chromosomes seen here by 3D electron microscopy.

      Limitations: the molecular composition and functional roles are not yet known for the tethers seen here by 3D electron microscopy, but to do so would involve an entire new program of experimentation.

      Advances: there have only been two earlier ultrastructural papers on tethers at telomeres, and the tethers were peripheral to the main focus of those papers. Thus, the current paper extends our ultrastructural information about tethers.

      Audience: this work is of importance for scientists who study the mechanics of chromosome movement on spindles, including regulation to combat aneuploidy. This work will also be important for a broader audience to inform them about transmission of the hereditary information to daughter cells._

      We thank the reviewer for appreciating the significance and the impact of our work.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary:

      Tethers between telomeres of chromosomes in anaphase were inferred from earlier studies of laser microbeam cutting experiments. The current paper presents images from electron tomography of crane fly spermatocytes that substantiates the earlier inference. The authors deduce that the darker filaments and the lighter filaments that they visualize may be the structural tethers at telomeres.

      Major comments:

      The experiments are carefully done, and the conclusions are appropriately worded to qualify any caveats. This short communication is well-presented, and I have only a few comments. The authors should expand their list of references on bridges to include those listed by Warecki et al (Curr Biol 33:1-17, 2023; refs 15-26, etc). The authors present arguments that the tethers are not the DNA bridges observed by others. However, they should try to address this experimentally by treatment of their preparations with DNase to see if the thick and/or thin filaments disappear. Moreover, they should discuss in more detail the possible functions of (DNA) bridges, including the recent model from Bill Sullivan's lab (Warecki et al, Curr Biol, 2023) that they help to retain fragments of broken chromosomes. In addition, the authors should summarize the various proteins that may be associated with the bridges (as enumerated in the Warecki et al 2023 paper).

      The authors could add a sentence to the Results or Discussion of whether the thicker tethers might become stretched as anaphase progresses to become the thinner tethers (Fig. 4G).

      The authors may want to add a few sentences to the Discussion about the "chromosomal bouquet" stage of leptotene of meiosis prophase I where the telomeres of chromosomes seem pulled together and associate with the nuclear envelope --- they could speculate if this might also be due to the tethers that they describe in spermatocytes.

      Minor comments:

      A few additional comments are as follows:

      p. 2 last sentence of first paragraph -modify the wording about "no structural evidence that identifies physical connections between separating telomeres", since there is some information from genetic and cell biology light microscopy experiments. Perhaps simply change "structural" to "ultrastructural".

      p. 6, 5th line of second paragraph - change "ribosome DNA" to "ribosomal DNA"

      Figure 1D - add the chromosome to the right of the schematic model (as suggested by Fig. 1B).

      p. 17 (Methods), line 10 of first paragraph - state if this is light or heavy Halocarbon oil (give details).

      p. 17 (Methods), line 12 of first paragraph- state the concentration for fibrinogen and for thrombin.

      p. 17 (Methods), line 4 of second paragraph - is there any data to show that the filaments (tethers) occur if there is no cold shock?

      Referees cross-commenting

      I concur with Reviewers #1 and #2 that this is a fine paper that should be published. My detailed comments submitted with my review are simply meant as revisions to further strengthen this paper.

      Significance

      Strengths: This is an important conceptual advance and the carefully done ultrastructural imaging provides the foundation for future studies that could delve into the molecular composition and functional significance of the tethers at telomeres of anaphase chromosomes seen here by 3D electron microscopy.

      Limitations: the molecular composition and functional roles are not yet known for the tethers seen here by 3D electron microscopy, but to do so would involve an entire new program of experimentation.

      Advances: there have only been two earlier ultrastructural papers on tethers at telomeres, and the tethers were peripheral to the main focus of those papers. Thus, the current paper extends our ultrastructural information about tethers.

      Audience: this work is of importance for scientists who study the mechanics of chromosome movement on spindles, including regulation to combat aneuploidy. This work will also be important for a broader audience to inform them about transmission of the hereditary information to daughter cells.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      In this paper, the authors use light microscopy and electron tomography to study anaphase chromosomes in crane fly spermatocytes. They find that there are two "tether" structures that connect telomeres of sister chromatids. One tether is thicker (denser) and extends between sister chromatids during early but not late anaphase, whereas a second, less-dense tether maintains contact with both sister chromatids in all examined stages of anaphase. The paper makes arguments as to what the tethers could or could not be. Specifically, they are too numerous to be ultrafine DNA bridges seen in various normal or abnormal segregation events and they also do not affect anaphase chromosome motion the same way ultrafine DNA bridges do.

      Major comments:

      The major claim that there are tethers that connect sister chromatids in anaphase is supported by the data. Moreover, the data resolves two types of tethers on the basis of their density. While it is unclear what the composition of the tethers are, the paper makes a convincing case that they cannot be the DNA ultrafine bridges seen in other studies. The discussion has sufficient caveats that most readers will see that more work is needed to identify the composition of the two tethers. In my opinion, no further experiments are needed to support the modest claims of this paper. Therefore, I only have minor comments that may hopefully improve the paper's clarity.

      Minor comments:

      It was argued that the tethers reported here were also seen in other species and cellular contexts, where the imaging work was done with projection EM imaging. Presumably, what is new here is the usage of electron tomography. It would help readers if the authors explained why the electron tomography done here was essential to arrive at key conclusions.

      p.3 mitochondria appeared to be fixed properly ... (e.g., Figs. 1C, 2B) - I don't see any mitochondria in any figures. Perhaps this observation should be noted as "not shown"?

      p.3 The images shown in Figs. 1, 2, 4 - The figures should be called out in the order; in this case, Fig 3 has not been called out yet.

      p.4 we did not find any other connecting structures - Because the sample was processed by traditional EM methods, it's safer to add a caveat that other connecting structures could be missed if they were disrupted by sample prep or if they did not pick up stain as well as the two structures presented in this paper.

      p.7 we expect such structures to be commonly seen in other cell types as well if they are examined carefully - Instead of saying that examinations should be done "carefully", it would be more helpful to specify how other cell types should be examined. This work shows that the bridges can be found if the cells are either sectioned parallel to the spindle axis or if a sufficiently large volume is sampled.

      Please use consistent spelling/hyphenation of ultrafine/ultra-fine and word choice (strands vs. bridges).

      Referees cross-commenting

      I agree with my co-reviewers's comments and have no further suggestions.

      Significance

      This may be the first use of electron tomography to study the structural details of tethers that connect chromosomes in anaphase cells. The data is of sufficient quality to reveal differences in density. Namely, one class of tether appears to be an extension of the chromosome while the other class is composed of thin filaments. This study is novel in that it characterizes a mitosis-associated complex that is poorly studied compared to the microtubule-based spindle apparatus and the kinetochore. Hopefully, the tethers will draw more attention and further characterization by methods like super-resolution microscopy and cryo-electron microscopy. My expertise is in chromatin, mitotic machines, and cryo-electron tomography.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary: Forer and Otsuka provide first-rate evidence for tethers fixed in place between separating anaphase chromosomes using electron tomography. The authors traced the anaphase movement of a number of living cells before fixation for examination using electron tomography. The manuscript is clearly written and provides an excellent introduction and discussion of the known literature. The reader will have an excellent background to see the importance of this work.

      Major comments:

      • Are the claims and the conclusions supported by the data or do they require additional experiments or analyses to support them?

      No further experiments are needed. The data are very supportive, and extremely clear.<br /> - Are the data and the methods presented in such a way that they can be reproduced? Yes.<br /> - Are the experiments adequately replicated and statistical analysis adequate? Yes.

      Minor comments:

      • Are prior studies referenced appropriately? Yes.
      • Are the text and figures clear and accurate? Absulotely.
      • Do you have suggestions that would help the authors improve the presentation of their data and conclusions?

      The authors are to congratulated on their major contribution to this study on tethers between separated daughter chromosomes. It is a tpur deforce to go from the living cells to fixing and identifying the same separated chromosomes using electron tomography to see the ultrastructure of the fibers seen fir.

      Referees cross-commenting<br /> Thank you reviewer #2. The manuscript should be published. It is an excellent contribution.

      Significance

      Provide contextual information to readers (editors and researchers) about the novelty of the study, its value for the field and the communities that might be interested.

      This manuscript is the first to use electron tomography toidentify the tethers between separated anaphase chromosomes. Forer and the laetMichael Berns and their co-authors have published a number of papers using phase microscopy and lasers to report on the physical nature and elastic properties of these fibres in the past. Forer and Otsuka have presented first-rate evidence for the reality of these structures using electron tomography. This manuscript should highlighted in the published journal.<br /> The chemical identity of these fibers as the authors state is unclear.

      The following aspects are important:

      • Audience: describe the type of audience ("specialized", "broad", "basic research", "translational/clinical", etc...) that will be interested or influenced by this research; how will this research be used by others; will it be of interest beyond the specific field?

      This exciting contribution will be read by anyone interested in mitosis. It will be of interest to all Cell Biologists because of the careful manner in which the living cells were studied before they were fixed for examination using electron tomography. The readers will be dreaming how they can use this process on their Cell Biology problems.<br /> - Please define your field of expertise with a few keywords to help the authors contextualize your point of view. Indicate if there are any parts of the paper that you do not have sufficient expertise to evaluate.

      I am a cell Biologist who has made contributions, both in light microscopy and in transmission microscopy on diving cells, both in tissue culture and in situ in aviav and zebrafish embryos.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity):

      The manuscript by Rigger and Brenner details the role of vimentin network, in advancing OA pathogenesis by exacerbating premature senescence. The data is well presented and the study of interest, in that there is little known about vimentin in cartilage biology.<br /> The authors used OA derived cartilage explants and chondrocytes cultures, were graded for severity and compared accordingly. Figure 1 shows that markers of senescence are increased with structural damage, which is well established and consistant with the literature. Using a DOX model the authors induce premature senescence and exhibit a disrupted vimentin network. However, upon KD of CDKN2A, a marker of senescence, but did not observe complete reversal of CSV presentation.<br /> Next the authors show in figure 4 and 5, that the reduction or dismemberment of vimentin structures are linked to senescence and may act as contributing factors.<br /> Figures 6 and 7 then go on to show that upon advanced passage chondrocytes lose their vimentin network, and tend to senesce and mineralize.

      Reviewer #1 (Significance):

      Strength:<br /> This is a very novel study showing a link between vimentin and senescence in chondrocytes. The data are in line with other data. The work is clearly written structured and well displayed.

      Author´s response:<br /> We thank reviewer #1 for their interest in our work and their overall positive report.

      Suggestions for improvement:

      While the study is very thorough ought in describing the markers of senescence and vimentin network, it lacks insight regarding mechanism which isn't completely deciphered. Are there links to key transcription factors?

      Author´s response:<br /> The transcriptional regulation of vimentin in human cells is very complex. The VIM promoter region comprises multiple elements, such as a NF-kB- binding site, a PEA3-binding site and two AP1-binding sites (Zhang et al., 2003). Moreover, it was recently demonstrated that redox signaling is involved in vimentin expression at the wound margin after tissue injury in zebra fish (LeBert et al., 2018). However, it has also been reported that IL-1ß stimulation results in reduced gene expression of vimentin via p38-signalling in cartilage degeneration and OA progression (see manuscript REF. 36,37).

      In our study, we observed that enhanced CSV levels are associated with a decreased vimentin gene expression, indicating a lower stability of the mRNA or decreased transcription of VIM in senescent chondrocytes (maybe due to enhanced p38-signalling as mentioned above). Since the transcriptome in senescent cells is radically changed, this question cannot be answered easily.

      In future studies, we will rather try to clarify the underlying mechanism of vimentin externalization. There are still many questions to be answered: is the CSV anchored in the cell membrane (which anchor protein?) and is there still a connection to the intracellular vimentin network? Which proteins are involved in the externalization process: maybe comparable to phosphatidylserine exposure, mediated by flippases, scramblases, and lipid transfer proteins or rather by vesicles?

      Literature mentioned above (not included in manuscript):

      LeBert et al., 2018: Damage-induced reactive oxygen species regulate vimentin and dynamic collagen-based projections to mediate wound repair. DOI: 10.7554/eLife.30703

      Zhang et al., 2003: ZBP-89 represses vimentin gene transcription by interacting with the transcriptional activator, Sp1. DOI: 10.1093/nar/gkg380

      It is also unclear if disruption of the network is more detrimental than KD in promoting senescence.

      Author´s response:<br /> KD of Vimentin led to a gradually decrease of intracellular Vimentin content and consequent stress. The cells were analyzed 7 days after induction of the KD and exhibited a stable senescent phenotype, comparable to Doxorubicin-treated chondrocytes (treated with very low concentrations over several days to produce only mild but ongoing stress). These models might reflect the pathophysiologic situation: We think that cellular stress due to mechanical impact and subsequent oxidative stress/ low-grade inflammation might lead to a gradual disruption or re-organization of the vimentin network, which is accompanied by decreased vimentin gene expression.

      In case of the disruption of the vimentin network by Simvastatin, the stress response was very intense and rapid (24 h), and was only conducted as a proof-of-principle experiment. Despite the upregulation of some senescence-associated markers, we don`t think that permanent Simvastatin treatment would be suitable to obtain a stable senescent phenotype, but rather expect the cells to die due to excessive stress.

      It would have been good to include models OA murine models to understand these processes better, and make a stronger physiological connection with OA of the joint.

      Author´s response:<br /> The CSV antibody is only suitable for human cells and cannot be used for immunohistochemistry. Therefore, all previous reports of CSV are based on human (isolated) cells. At the current time point, it would not be possible to stain CSV in joints of mice after induction of PTOA due to the methodological limitations. We actually tested the CSV-antibody in isolated lapine chondrocytes and found a high percentage of CSV-positive cells, even at low passages. Although stress increased the amount of CSV-positive lapine cells, we did not consider the results as reliable due to the high percentage in un-stressed cells, which might result from unspecific antibody binding.

      Overall, we think that the usage of clinical OA samples is convincing and reflect the pathophysiologic situation in the human OA joint.

      Reviewer #2 (Evidence, reproducibility and clarity):

      The manuscript provides solid evidence for an association between cell surface vimentin (CSV) and chondrocyte senescence. Human cartilage and cultured chondrocytes are used with a wide range of approaches to provoke senescence: natural osteoarthritis, traumatic loading ex vivo, doxorubicin to cells in monolayer, vimentin siRNA, and simvastatin. In contrast, relatively little was done to try and interrupt or reverse the role of CSV in senescence, with CDKN2A siRNA representing one attempted intervention. The manuscript is well written and the data are presented in a logical and clear manner, with a high likelihood of being reproduced in subsequent studies.

      Author´s response:<br /> We thank reviewer #2 for their interest in our work and their mainly positive report.<br /> Regarding their comment on our attempts to reverse CSV on senescent chondrocytes, we would like to add the following: Reversal of cellular senescence is a very ambitious challenge. But in fact, we are currently preparing a manuscript in which we characterize an appropriate senolytic strategy to “rejuvenate” human chondrocytes and plan to use this approach to reduce the amount of senescent and thus CSV-positive cells in future experiments.

      _Major comments:

      In the doxorubicin experiments, the senescent cells show a spread morphology as expected. Given the importance of vimentin in cell spreading (as the authors own data show), the possibility that spread morphology itself (and not senescence) leads to CSV should probably be examined. This could perhaps be achieved by plating with different concentrations of fibronectin or other matrix proteins that produce a spread morphology to a degree that matches the doxo. If the cells remain spread for ~10 days but don't become senescent and don't have CSV, this would provide further support for a direct relationship.

      Author´s response:<br /> We agree that cell spreading is associated with various cellular processes (for example by the YAP signaling pathway). Moreover, we would like to thank the reviewer for the proposed experiment.

      Seeding of cartilage cells on fibronectin coated plates is a commonly used procedure to isolate chondrogenic stem progenitor cells, due to their higher affinity to fibronectin. The cells are usually cultured for several days on the coated plates and do not exhibit a flattened, senescent-like phenotype (as we observe for Doxorubicin-treated cells), but an elongated, fibroblast-/ stem cell-like shape. Our results (Figure 6E) demonstrate that CSPC have no increased CSV levels, despite their elongated (not flat) morphology.

      There are some findings supporting the assumption that CSV leads to enhanced cell adhesion, but not that adhesion or cell spreading promotes CSV: we included experiments with HeLa (low CSV levels) and SaOS-2 (high CSV levels), which demonstrated that high CSV levels are associated with increased plastic adhesion (Figure S5). In line with this, we demonstrated that higher CSV levels on chondrocytes were associated with enhanced fibronectin and vitronectin binding, which might explain increased plastic adhesion. Moreover, Simvastatin stimulation and subsequent cellular stress by Vimentin disruption resulted in enhanced CSV but did not lead to cell spreading (Actin not affected, cells rather elongated, not flattened).

      Minor comments:

      The CSV antibody and staining method appeared to have generated some signal from debris, which makes it challenging to assess the localization of true staining. Presumably the true staining would be present only on the cell surface. While the widefiled view is appreciated, perhaps insets with a higher magnification would clarify.

      Author´s response:<br /> In Figure 2h and Figure 2i, we provide insets of the IF-staining and an exemplary image made by scanning electron microscopy (SEM). CSV is not localized on debris – Figure 2h, actually represents the cell surface. The magnified, Doxo-treated cell is highly senescent and thus flattened. The uneven (rather spotted) staining pattern of CSV and the unusual shape of the cell might suggest that this is debris, not the cell membrane.

      For figure 1k, it is a bit surprising that CDKN2A would peak so early after injury and then drop off. Most studies in other systems show a gradual increase in CDKN2A levels with persistent stress as opposed to a rapid increase in response to acute stress. Could the drop-off be due to preferential death of these cells? The CSV % in 1m was taken from 7d after trauma (plus 7 days in monolayer it appears). Further discussion on the timing of traditional senescence markers as compared to the emergence of CSV would be useful.

      Author´s response:

      We would like to thank the reviewer for this comment. That CDKN1A was induced by mechanical trauma without significant decrease at the later time points was in line with the P53 expression, which we detected via immunohistochemistry (IHC; positive staining of chondrocyte nuclei in cartilage). P53 and P21 are regarded as interconnected senescence markers. Interestingly, P53 is not regulated on gene expression level upon cartilage trauma or Doxorubicine stimulation – but there is a significant increase in P53 nuclear translocation.

      Although such a discrepancy between gene expression and protein activity has not been reported in case of P16 or P21, we plan to investigate the dynamics of these cell cycle regulators and its connection to CSV after cartilage trauma in more detail in future studies.

      We included the following statement in the discussion part:

      “In the current study, we observed that CSV on chondrocytes was reduced by siRNA-mediated silencing of CDKN2A and increased after Doxo treatment or cartilage trauma. While we confirmed that mRNA levels of both CDKN1A and CDKN2A were significantly enhanced upon injury but exhibited different expression levels over time, we determined CSV-positive cells only at one time point after ex vivo cartilage trauma. Future studies might also consider earlier and later time points after cartilage injury to identify a potential time-dependent peak or decline in CSV-positive chondrocytes. In this way a potential association between CSV and the expression levels of CDKN1A and CDKN2A, which are thought to play differential roles in initiating and maintenance of senescence, respectively [50], might be clarified.”

      [50] Stein G, Drullinger L, Soulard A, and Dulić V. Differential Roles for Cyclin-Dependent Kinase Inhibitors p21 and p16 in the Mechanisms of Senescence and Differentiation in Human Fibroblasts. Mol Cell Biol. 1999;19(3): 2109–2117. https://doi.org/10.1128/mcb.19.3.2109.

      There is no CSV staining shown for figures 4 and 5. While the quantification of CSV was done by flow cytometry, it would nice confirmation to see the increase in CSV on the surface of cells with either siRNA for vimentin or the simvastatin.

      Author´s response:

      CSV-IF of simvastatin-treated chondrocytes is provided in Figure 5 (b). We did not perform exemplary staining of CSV after VIM-KD, because the quantification was performed via flow cytometry.

      Reviewer #2 (Significance):

      The strengths of the study include a rigorous design and the establishment of a potential new cell surface marker of chondrocyte senescence. The main limitation is that the conclusions are largely descriptive in nature.

      If CSV is confirmed as a robust marker of senescence, this would be of value to the field. While this marker has been explored previously in other systems, there is value in this manuscript given the wide range of contexts investigated for a cell type in which senescence likely has an important role.

      Reviewer #3 (Evidence, reproducibility and clarity):

      This study presents a sound piece of science in the puzzle about extracellular vimentin in the differentiation/dedifferentiation of human chondrocytes and senescence and osteoarthritis. Eventhough, no mechanism is elucidated, the results clearly point towards a correlation of the amount of extra cellular vimentin and the level of chondrocyte senescence, and therefore signs of osteoarthritic changes in the cultivated chondrocytes. The methods applied are state-of-the art and provide the means to generate meaningful results in this experimental setting. The paper is concise and clearly written, there are only minor remarks.

      Author´s response:

      We thank reviewer #3 for their interest in our work and their overall positive report.

      Minor comments:

      1. The main clue of the paper is extra cellular vinemtin around chondrites in culture, please provide better pictures (1g) to support this. Why is the extra cellular staining seen so broad and not concentrated on the cells surface? The picture chosen imply a huge amount of vimentin to be externilized in disease states. It also indicates that in diseased chondrocytes no intact or semi-intact vimentin network is found intracellular. Please comment.

      Author´s response:

      In Figure 1g, CSV is located on the cell membrane. The pattern of the staining was surprising to us, as well. CSV was not equally distributed on the membrane, but rather represented an inconsistent pattern. Sometimes the staining was located at the filopodia of the cells, sometimes the whole cell was covered by spots. We also observed this on cancer cells, which was in line with other studies using this antibody. It remains unclear whether the distribution of the CSV has any effect. But we assume that the high abundance in filopodia might be connected with cell adhesion and mobility, which was positively associated with CSV.

      Yes, chondrocytes isolated from highly degenerated tissue exhibited higher CSV levels as compared to cells derived from macroscopically intact regions. Although we did not investigate the vimentin network of these cells, our observations in Doxo-treated cells imply, indeed, that intracellular vimentin might be altered in diseased chondrocytes. According to this, Blain et al (Ref. 13) reported that there is a disassembly of the intracellular vimentin network in OA chondrocytes, which can disturb the chondrocyte phenotype and contributes to the development of OA (see discussion).

      1. In the doxo experiment no extracellular vimentin is found? Please explain.

      Author´s response:

      Doxo-treated cells are highly positive for CSV (= extracellular vimentin on membrane). However, the intracellular vimentin is strongly decreased and some cells seem to be negative. We have not clarified the underlying mechanism by now, but it seems that senescence/ disease progression negatively affects the transcription of vimentin and, at the same time, promotes the externalization of the existing intracellular vimentin. Altogether, this might result in a decline in intracellular vimentin.

      1. The SEM picture is showing what. IGH? The red dots are colloidal gold particles? In any case the quantity of stain gathered EM level would not correlate to the huge amount seen in LM staining. Please comment.

      Author´s response:

      For the SEM analysis, a gold particle-coated secondary antibody was used. The positive signal usually appears in white and was subsequently colored via a software. In IF and ICC staining, we had a signal amplification due to the biotin-streptavidin system and the magnification makes, of course, a huge difference.

      1. Why the ICC in Fig. 3c? The siRNA is not detected in the KD? A reduction of Vimentin could be shown via WB.

      Author´s response:

      In Figure 3c, the KD of P16 was confirmed on protein level. In addition to the gene expression analysis, we chose the ICC (IF) to confirm that there is a decline in active (nuclear) CDKN2A. In case of P53, we made the experience that gene expression and the amount of cytoplasmic/ nuclear protein might not be consistent.

      In Figure 4, we confirmed the successful KD of vimentin on mRNA and protein level (flow cytometry plus IF). Of course, WB would also be possible, but we decided to use the methods in which the antibody was well established and we wanted to visualize the disturbance of the intracellular vimentin network upon KD.

      1. Fig. 4c, why are there no remnants of the vimentin networks seen in the chondrocytes? A Knock-down, not a KO is shown.

      Author´s response:

      In fact, most of the intracellular vimentin seems to be gone. However, there are some remnants (condensed fibers/ bundles) of the former vimentin network. We applied the VIM-KD over seven days. Usually, a KD experiment is only conducted for 2-3 days. But since we were not sure how stable the vimentin protein would be, we chose seven days. This long-lasting KD might have resulted in a strong decline of the protein. Moreover, the CSV levels on these cells were very high, indicating that existing vimentin was externalized and additionally decreased the amount of intracellular vimentin.

      1. Please comment of the concentration of simvastatin, why not nmolar?

      Author´s response:

      The concentration of Simvastatin was chosen in accordance with Trogden et al. (Ref. 26), who first described the effects of simvastatin on the vimentin network. A lower concentration might have had the advantage, that the effects were less severe, allowing a longer observation time than 24h. However, as a proof-of-principle model to demonstrate the connection between vimentin network collapse ant CSV expression, the concentration worked quite well.

      1. CSV+ is misleading in Fig. 6g, it's not an over expression.

      Author´s response:

      We would like to thank the reviewer for this comment and removed the “+” to make it less misleading.

      1. The concept of EMT is debatable, at least in kidney fibrosis, and chondrocytes are not epithelial cells. Please add a more critical discussion point.

      Author´s response: The authors agree with the reviewer’s argument that chondrocytes are no epithelial cells ant that the term EMT doesn’t seem to be appropriate. However, this is one leading hypothesis proposed by the working group of Prof. Mayán, who described CX43 and other EMT-markers on/ in senescent chondrocytes (see reference 31; more recently: Cell Death Dis. 2022;13(8):681. doi: 10.1038/s41419-022-05089-w).

      We added the following passage in the discussion part to indicate that this hypothesis is a controversial concept:

      “Nevertheless, the hypothesis that chondrocytes might undergo an EMT-like process remains controversially discussed, because chondrocytes are mesenchymal and not epithelial cells. In a recent review, Gems and Kern propose to consider senescent chondrocytes as activated and hyperfunctional remodeling cells occurring during OA progression [49]. Accordingly, chondrosenescence might represent an unsuccessful attempt of tissue repair. They further suppose that the senescent or activated chondrocytes are associated with a hypertrophic, bone-forming phenotype, following the process of bone development rather than hyaline cartilage formation. In line with this, we observed that CSV was associated with enhanced osteogenic capacities and a decline in chondrogenic properties.”

      [49] Gems and Kern, 2022): Geroscience. 2022;44(5):2461-2469. doi: 10.1007/s11357-022-00652-x.

      Reviewer #3 (Significance):

      The manuscript provides novel insight in the role of intermediary filaments, i.e. vimentin, on chondrocyte senescence and osteoarthritic changes in vitro. It's strength is a thorough elucidation of the connection with a wealth of experimental data, a weakness is the missing elucidation, or first experiments in the direction, of the cell biological mechanism.<br /> It is well suited for a broad audience, because it deals with fundamental cell biological phenomena, definitely it's important for the OA /chondrocyte biology community.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      This study presents a sound piece of science in the puzzle about extracellular vimentin in the differentiation/dedifferentiation of human chondrocytes and senescence and osteoarthritis. Eventhough, no mechanism is elucidated, the results clearly point towards a correlation of the amount of extra cellular vimentin and the level of chondrocyte senescence, and therefore signs of osteoarthritic changes in the cultivated chondrocytes. The methods applied are state-of-the art and provide the means to generate meaningful results in this experimental setting. The paper is concise and clearly written, there are only minor remarks.

      Minor comments:

      1. The main clue of the paper is extra cellular vine tin around chondrites in culture, please provide better pictures (1g) to support this.<br /> Why is the extra cellular staining seen so broad and not concentrated on the cells surface? The picture chosen imply a huge amount of vimentin to be externilized in disease states. It also indicates that in diseased chondrocytes no intact or semi-intact vimentin network is found intracellular. Please comment.
      2. In the doxo experiment no extracellular vimentin is found? Please explain.
      3. The SEM picture is showing what. IGH? The red dots are colloidal gold particles? In any case the quantity of stain gathered EM level would not correlate to the huge amount seen in LM staining. Please comment.
      4. Why the ICC in Fig. 3c? The siRNA is not detected in the KD? A redyuction of Vimeo tin could be shown via WB.
      5. Fig. 4c, why are there so remenants of the vimentin networks seen in the chondrocytes? A Knock-down, not a KO is shown.
      6. Please comment of the concentration of simvastatin, why not nmolar?
      7. CSV+ is misleading in Fig. 6g, it's not an over expression.
      8. The concept of EMT is debatable, at least in kidney fibrosis, and chondrocytes are not epithelial cells. Please add a more critical discussion point.

      Significance

      The manuscript provides novel insight in the role of intermediary filaments, i.e. vimentin, on chondrocyte senescence and osteoarthritic changes in vitro. It's strength is a thorough elucidation of the connection with a wealth of experimental data, a weakness is the missing elucidation, or first experiments in the direction, of the cell biological mechanism.

      It is well suited for a broad audience, because it deals with fundamental cell biological phenomena, definitely it's important for the OA /chondrocyte biology community.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      The manuscript provides solid evidence for an association between cell surface vimentin (CSV) and chondrocyte senescence. Human cartilage and cultured chondrocytes are used with a wide range of approaches to provoke senescence: natural osteoarthritis, traumatic loading ex vivo, doxorubicin to cells in monolayer, vimentin siRNA, and simvastatin. In contrast, relatively little was done to try and interrupt or reverse the role of CSV in senescence, with CDKN2A siRNA representing one attempted intervention. The manuscript is well written and the data are presented in a logical and clear manner, with a high likelihood of being reproduced in subsequent studies.

      Major comments:

      In the doxorubicin experiments, the senescent cells show a spread morphology as expected. Given the importance of vimentin in cell spreading (as the authors own data show), the possibility that spread morphology itself (and not senescence) leads to CSV should probably be examined. This could perhaps be achieved by plating with different concentrations of fibronectin or other matrix proteins that produce a spread morphology to a degree that matches the doxo. If the cells remain spread for ~10 days but don't become senescent and don't have CSV, this would provide further support for a direct relationship.

      Minor comments:

      The CSV antibody and staining method appeared to have generated some signal from debris, which makes it challenging to assess the localization of true staining. Presumably the true staining would be present only on the cell surface. While the widefiled view is appreciated, perhaps insets with a higher magnification would clarify.

      For figure 1k, it is a bit surprising that CDKN2A would peak so early after injury and then drop off. Most studies in other systems show a gradual increase in CDKN2A levels with persistent stress as opposed to a rapid increase in response to acute stress. Could the drop-off be due to preferential death of these cells? The CSV % in 1m was taken from 7d after trauma (plus 7 days in monolayer it appears). Further discussion on the timing of traditional senescence markers as compared to the emergence of CSV would be useful.

      There is no CSV staining shown for figures 4 and 5. While the quantification of CSV was done by flow cytometry, it would nice confirmation to see the increase in CSV on the surface of cells with either siRNA for vimentin or the simvastatin.

      Significance

      The strengths of the study include a rigorous design and the establishment of a potential new cell surface marker of chondrocyte senescence. The main limitation is that the conclusions are largely descriptive in nature.

      If CSV is confirmed as a robust marker of senescence, this would be of value to the field. While this marker has been explored previously in other systems, there is value in this manuscript given the wide range of contexts investigated for a cell type in which senescence likely has an important role.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      The manuscript by Rigger and Brenner details the role of vimentin network, in advancing OA pathogenesis by exacerbating premature senescence. The data is well presented and the study of interest, in that there is little known about vimentin in cartilage biology.

      The authors used OA derived cartilage explants and chondrocytes cultures, were graded for severity and compared accordingly. Figure 1 shows that markers of senescence are increased with structural damage, which is well established and consistant with the literature. Using a DOX model the authors induce premature senescence and exhibit a disrupted vimentin network. However, upon KD of CDKN2A, a marker of senescence, but did not observe complete reversal of CSV presentation.

      Next the authors show in figure 4 and 5, that the reduction or dismemberment of vimentin structures are linked to senescence and may act as contributing factors.

      Figures 6 and 7 then go on to show that upon advanced passage chondrocytes lose their vimentin network, and tend to senesce and mineralize.

      Significance

      Strength:

      This us a very novel study showing a kink between vimentin and senescence e in chondrocytes. The data are in line with other data. The work is clearly written structured and well displayed.

      Suggestions for improvement:

      While the study is very thorough ought in describing the markers of senescence and vimentin network, it lacks insight regarding mechanism which isn't completely deciphered. Are there links to key transcription factors? It is also unclear if disruption of the network is more detrimental than KD in promoting senescence. It would have been good to include models OA murine models to understand these processes better, and make a stronger physiological connection with OA of the joint.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      The authors do not wish to provide a response at this time as only a revision plan is provided at this time.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Using an elegant set of experiments, the authors probe the function of PfPK2 in asexual blood stage development. Using a rapamycin inducible dimerizable Cre recombinase-based strategy for the conditional knockdown of PK2, coupled with immunofluorescence assay, live cell imaging and phosphoproteomics the authors demonstrate the role of PK2 in invasion of red blood cells and signalling pathways in the parasite. Results show that PK2 is critical for the formation of tight junctions between merozoites and host red blood cells and that PfPK2 regulates cGMP levels and calcium release.

      Major comments:

      1. Is PK2 expressed in other life cycle stages? The authors should discuss what is known about the role of PK2 in other stages of parasite development. Would PK2 inhibitors be expected to have activity against liver or sexual stages of the parasite?
      2. Have any inhibitors of PK2 been reported to date? If so, were these tested against the PK2 cKD parasite line? Do they show similar phenotypic effects to knockdown of PK2? Do these studies give any insight on how vulnerable PK2 is as a target? To what extent does PK2 need to be inhibited/knocked down to lead to parasite death/growth inhibition?
      3. What were the expression yields and purity for the recombinantly expressed wild-type PK2 and mutant/ΔRD and ΔCD deletions? Has the ATP Km for PK2 been reported/determined? How did the deletions affect protein expression and stability? Could these deletions be influencing protein folding/amount of active protein in the preparation rather than regulating catalytic activity? How were protein concentrations normalised for the assay?
      4. Does rapamycin have any effect on wild-type parasites? What controls were included to ensure that observed effects were a result of rapamycin-induced PK2 knockdown, rather than rapamycin acting on other pathways/interfering directly with parasite biology. This should be discussed for readers less familiar with this system. Presumably the cHA-PfPK2-loxP line was used as the control to account for this for at least some of the experiments.

      Minor comments:

      1. "Malaria contributes to almost 400,000 deaths globally (Hay et al, 2004)."It should be made clear that this is the number of deaths reported annually. The references and statistic should be updated based on the most recent WHO Malaria report.
      2. "Several of these kinases are regulated by second messengers like calcium, PIPs, cAMP, cGMP Plasmodium kinases like CDPKs, PKA, PKG etc and are crucial for diverse parasitic functions." This sentence could be reworded for clarity.
      3. "However, the precise function of this protein kinase in parasite life cycle has remained unknown." Add word "the" should be added before "parasite".
      4. "Present studies demonstrate that PfPK2 regulates the invasion of host erythrocyte by P. falciparum." Should be erythrocytes (plural).
      5. "but the sequence similarity is mainly between kinase domains of human CamKIand PfPK2 (~80%)". Change "mainly" to "highest".
      6. "PfPK2 has a long N-terminal extension compared to CamKIand a CaM binding domain was predicted within a postulated regulatory domain present downstream of the kinase domain, which is the case with CamKI (Kato et al, 2008)." Please check this sentence and reword for clarity.
      7. References to "www.plasmodb.org" should be replaced with specific references to the source of the information listed in the database.
      8. Figure 2b. Please explain the relevance of adding/excluding EDTA and the effect of this on PK2 activity.
      9. "Correct integration of the targeting plasmid at the desired locus was confirmed by PCR as amplicons of expected size were obtained parasites from a clone that lacked wild type PfPK2 (Fig. 2B)." Remove the word "parasites".
      10. "Successful expression of GFP-PfPK2 was indicated by Western blotting (Fig. 2D) as well as by IFA (Supp. Fig. S1A), which wasundetectable after rapamycin treatment suggesting efficient depletion of this kinase from the parasite (Fig. 2D and Supp. Fig. S1B)." Write IFA in full the first time it is used in the main text. "was undetectable" should be two words.
      11. "The release of another microneme protein EBA-175, which interacts with glycophorin A on the 11 surface of the erythrocytes (Sim et al, 1994), was also reduced (Fig. 5A, Supp. Fig. S3B). but its localization to the microneme was unaltered (Supp. Fig. S4)." Remove full stop mid-sentence.
      12. "Therefore, we focussed on a possible role of PfPK2 in the release of AMA1, which is involved in late inavsion events like Tight Junction formation and PfPK2 plays a role at this stage of invasion (Fig. 4)." Correct "inavsion" to "invasion".
      13. Figure 5b. What is the difference between upper and lower panels. Please label or include details in figure legend. Figure legends should also include details on the DAPI, DIC and merged images for readers less familiar with microscopy images.
      14. "AMA1 plays a role in the formation of Tight-Junction (TJ) is via its association with rhoptry neck proteins (RONs) that are released by the rhoptries "Just-in-Time" on to erythrocyte surface (Riglar et al, 2011)." Remove "is".
      15. "Interestingly, phosphorylation site S1214 which is regulated by PfPK2 (Fig. 6B, 7B) resides in this insert This phosphositeis conserved only in GC" Add full stop after "insert". Replace "phoshositeis" with "phoshosite is"
      16. Order of Materials and Methods should be consistent with order of the Results section. For example, "Expression of recombinant proteins" and "Kinase assays" sections should come first.
      17. Please ensure all abbreviations, panels and labels in Figures are clearly explained in the relevant figure legends.

      Significance

      This study provides new insight into the role of PK2 in Plasmodium falciparum asexual blood stage development. While this study gives provides information on parasite biology, it does not provide significant advances in terms of understanding the value of PK2 as a drug target. However, this study does lay the groundwork for future studies using PK2 inhibitors as tool compounds for phenotypic validation.

      Audience: Specialised audience interested in parasite biology/signalling pathways.

      Reviewer's expertise: Target-based malaria drug discovery, Plasmodium kinase inhibitors, Target deconvolution for phenotypic hits in Plasmodium, kinase assay development.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary:

      Provide a short summary of the findings and key conclusions (including methodology and model system(s) where appropriate). Please place your comments about significance in section 2.

      The authors present a biochemical and functional characterisation of the Protein Kinase 2 in Plasmodium falciparum (PfPK2). They state that PfPK2 is an active protein kinase, as shown by phosphorylation of a histone protein (Figure 1). The next 6 figures of the paper describe the functional characterisation of PfPK2. It is a protein essential for the propagation of asexual malaria parasites in the clinical blood stage (Figure 2). The protein is expressed in the schizont stage and deletion of the PfPK2 gene by the conditional DiCre-LoxP system prevents parasite invasion, while not affecting egress (Figure 3 and 4), reduces the protein levels of two invasion molecules (Figure 5), modifies phosphorylation status on a variety of schizont stage proteins that play a role in invasion (Figure 6). PfPK2 deletion also leads to the decrease in the molecule cGMP which plays a role in parasite invasion (Figure 7).

      Major comments:

      • Are the key conclusions convincing?

      The conclusions made in Figure 2, 3, 4 and 7 are adequate but I have added questions in my detailed analysis on these figures that should assist in increasing the confidence on the conclusions.<br /> In my view, conclusions made in Figure 1, 5 and 6 are not convincing as they are missing key quality control data including loading control. Detailed comments on each figure is attached.<br /> - Should the authors qualify some of their claims as preliminary or speculative, or remove them altogether?<br /> Conclusions drawn from Figures 1,5,6 should be supplemented with the experiments suggested, otherwise the claims should be marked as speculative.<br /> - Would additional experiments be essential to support the claims of the paper? Request additional experiments only where necessary for the paper as it is, and do not ask authors to open new lines of experimentation.<br /> The experiments suggested in the detailed figure comments are necessary for supporting the claims of the paper.<br /> - Are the suggested experiments realistic in terms of time and resources? It would help if you could add an estimated cost and time investment for substantial experiments.<br /> To improve the phosphoproteomics data, I would strongly suggest the creation or acquisition of a parasite line in which a related kinase has also been conditionally regulated and analysis of the phosphoproteome. I would suggest PfPKA made in the Wilde et al study. These experiments are required in order to qualify the statements made in the discussion section such as<br /> "A novel pathway was deciphered in which it is an upstream regulator of cGMP-calcium signalling axis, which is critical for invasion"<br /> - Are the data and the methods presented in such a way that they can be reproduced?

      How many times were the experiments mentioned in the following figures repeated?<br /> Figure 1C: PfPK2 phosphorylation assay<br /> Figure 1D: Kinase assay<br /> Figure 2B,C, D<br /> How many cells were counted in Figure 2F?<br /> How many biological replicates were used in Figure 6B?<br /> - Are the experiments adequately replicated and statistical analysis adequate?<br /> For figures not stated above, experiments were adequately replicated and analysis is adequate.

      Minor comments:

      • Specific experimental issues that are easily addressable.
      • Are prior studies referenced appropriately?<br /> Yes.
      • Are the text and figures clear and accurate?<br /> Axes on Figure 5A require editing as the values are unable to be seen clearly. Other figures are clear and well presented. Specific figure related comments are attached to this document.
      • Do you have suggestions that would help the authors improve the presentation of their data and conclusions?

      This data, in its current state, requires supplementation with necessary controls in order to be convincing and also additional experiments as suggested in the detailed figure related information.<br /> - Describe the nature and significance of the advance (e.g. conceptual, technical, clinical) for the field.<br /> The data included advances the field in characterising the role of PfPK2 and may influence the thinking of researchers studying Apicomplexan invasion biology. The PfPK2 phosphoproteomics and invasion imaging help knowledge gain in this area.<br /> - Place the work in the context of the existing literature (provide references, where appropriate).<br /> This work advances the data identified by Kato et al (2008) in the characterisation of PfPK2. Kato et al identified the activity of PfPK2, it's dependence on calmodulin and calcium for activity and the localisation of PfPK2 at the merozoite. Kato et al attempted to delete PK2 in the rodent malaria parasite P. berghei, but were unsuccessful, suggesting that the gene is essential for malaria parasites.<br /> Rawat et al recapitulate the kinase activity data and advance Kato et al's findings in creating a conditionally deletable PfPK2 parasite strain. Using this strain, they confirm the essentiality of the gene and characterise it's effect on invasion related processes in the parasite.<br /> - State what audience might be interested in and influenced by the reported findings.<br /> Researchers in Apicomplexan parasite biology may find this data interesting.<br /> - Define your field of expertise with a few keywords to help the authors contextualize your point of view. Indicate if there are any parts of the paper that you do not have sufficient expertise to evaluate.<br /> My expertise is in malaria parasite biology including parasite invasion, genetic manipulation and protein export.

      Specific Figure related comments:

      Figure 1:

      C: Where is the data on the purification of PfPK2? Looking at Image 1C, I don't know what has been added to the reaction to phosphorylate Histone IIa.<br /> I would like to see Coomassie or western blots of the His-PfPK2 purification and the presence of Histone IIa so that I can be convinced that only PfPK2 and not a contaminant E-coli protein is responsible for the histone phosphorylation.

      Without knowing this data, I cannot be convinced that PfPK2 is phosphorylating Histone IIa

      Da: Again, without a Coomassie of western blot to show that K140M/WT PfPK2 as well as Histone IIa have been included in the reaction, I can't know that it is the single point mutation that has led to the lack of autophosphorylation of PfPK2 and the lack of Histone IIa phorphorylation.<br /> Please include this data.

      Db:<br /> Once again, there is no Coomassie or western to show what proteins are included in each treatment. Please include there.<br /> Why is there a band of the size of 32P-HIIa in the WT lane lacking HIIa? Is there a breakdown product of PfPK2 that gets autophosphorylated?

      Figure 2:

      A, B: The PCR results are conclusive in showing integration of the plasmid.<br /> C: The PCR and western blots of excision are conclusive and show efficient excision.<br /> D: Ideally, I would like to see the growth of the 1G5DC parasite strain on DMSO and rapamycin as well alongside the PK2 parasite line. 200nM of rapamycin is double the concentration used by Collins et al 2013 when creating the 1G5DC strain.

      Supplementary Figure 2A: Figure is missing a loading control here. Rest of the figure is conclusive and the complementation has worked nicely. I expect that the reason the complemented parasite isn't behaving exactly like the DiCre strain is due to the fact the complementation is done via an episomal plasmid which may segregate differentially in daughter merozoites.<br /> Supp Figure 2C: The Y-axis is hidden behind the legend for the same axis.

      Figure 3

      A + B: In my view the data from Figure 3A and 3B are effectively the one figure and therefore does not require an A/B division. The data shows no change in schizont maturation while invasion rate of rapa treatment is ~50% of DMSO control.<br /> C: Data shows that parasite invasion is impaired when PK2 is not present and that the cytochalasin D treatment was successful.

      Figure 4:

      The data shows reduction in the invasion rate upon Rapa treatment and an increased rate of prolonged echinocytosis. The video stills demonstrate the defects nicely.

      The data suggest that upon Rapa treatment, 40% of all invasion events are successful (Figure 4B). This result does not seem to match with Figure 2E, where rapa treatment leads to a reduction in parasitemia much less that 40% of control.<br /> Can the authors comment on this discrepancy?

      Figure 5 and Supp Figure S3

      A loading control is required for culture supernatant and lysate in all conditions (Supp Figure S3A,B,C). Otherwise we cannot interpret the loss of AMA1/EBA175 shedding as well as the non-effect on RhopH3. Without this data, the interpretations made in Figure 5A cannot be assessed.<br /> B: The figure needs to be better labelled to represent the E64 treatment as well.<br /> The representative image for PK2-LoxP + Rapa + E64 does not show any level of AMA1 on the surface, but the graph to the right shows 30% of parasites with surface AMA1 levels. Can the authors attach a series of images of this condition, showing the variation of AMA1 on the parasite surface?<br /> C: This figure shows nicely that AMA1 levels are reduced during invasion in rapa treated parasites.

      Figure 6:

      A. Do the authors identify PfPK2 peptides also being depleted upon PK2 knockdown?<br /> If so, please update Figure 6B. If not, then it makes the autophosphorylation data from Figure 1D contradictory.

      The data from Figure 6B and 6C identifies many invasion related proteins that are hyper/hypo phosphorylated upon rapamycin addition. It would have been interesting to include another invasion related kinase as a control line in all conditions (For example PfPKAc) that would allow finely dissecting the involvement of PfPK2 vs PfPKA.

      Figure 7A:

      1. What happens to the total protein levels of PDE and GCa upon PfPK2 deletion? Does the level of cGMP reflect the increased stability or loss of these two antagonistic proteins?<br /> I.E Does PK2 depletion lead to lower GCa activity (=less cGMP production) or increased stability and action of PDE (= more degradation of cGMP)?

      Significance

      This work advances the data identified by Kato et al (2008) in the characterisation of PfPK2. Researchers in Apicomplexan parasite biology may find this data interesting.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary: The manuscript by Rawat et al presents study to define the mechanism of action of an essential Plasmodium protein kinase PfPK2 using a conditional knockout line. The data show that PfPK2 has a role in the invasion of RBC as well as in the intraerythrocytic maturation of the parasite. In addition, quantitative phosphoproteomics was done to identify potential cellular targets and pathway analysis. The manuscript reports several important findings. I have following comments and suggestions:

      Major:

      1. Page 3. The data of malaria mortality is from 2004! Current WHO data needs to be included.
      2. Page 4. As per recent analysis (Adderley and Doerig,2022), number of P. falciparum kinases is higher than 85.
      3. Page 5. Homology analysis references are quite old. Is there any update from recent comparative analysis or alpha fold-based structural similarity.
      4. Page 7, line 4. It should be "Conditional Knockout". Needs to be corrected elsewhere as well.
      5. Page 8, line 18. Based on the methods section, parasites were treated at the ring-stage- needs to stated here as well.
      6. Page 12, line 16. It would be helpful to define "significant" in terms of phosphorylation changes. p<0.5?
      7. Page 12, line 23. It is mentioned that PfPK2-depleted samples have an increase in PPM2 phosphorylation. Is there a reference for these specific phosphosites?
      8. It would be bit of a stretch to say that the differentially phosphorylated proteins are substrate/targets (as in Table 1). PfPK2 depletion occurred in ring stage, and the sample was collected at ~44 HPI, leaving ample time for compensatory changes.
      9. Table 1. Criteria used for selecting proteins for the table needs to be defined in the legend. Example: p<0.5 FC>|1.2|. It would be also useful to see the fold change difference in a column for each protein.
      10. Fig 6B. Looks like there are 8 phosphorites with a greater negative fold change than those annotated. Are these all "unknown proteins"? They may not fall into the invasion or signaling groups in Table 1, but they still need to be included.
      11. Fig 6C. In the STRING analysis, PfPK2 is included as a yellow diamond but not referenced in the legend. Also, it is not clear if PfPK2 was differentially phosphorylated or was placed in the STRING network manually. "CPPUF is a nice abbreviation for "conserved Plasmodium protein, unknown function" but it would be useful to include the PlasmoDb number.
      12. Fig 6C. What is the significance of the line thickness or the circle size? Needs to be specified in the legend.
      13. It is not clear how many replicates of small collection was done for phosphoproteomic analysis.

      Minor:

      1. Page 5, line 3. It would useful to include PlasmoDb ID of PfPK2.
      2. Page 5, line 17: should read "....it has long N and a C-terminal...."
      3. Page 7, line 21: should be " was undetectable"...
      4. Page 9, line 17. It would be better to rephrase as.. "To better understand defects in the invasion following PfPK2 depletion live cell imaging was performed".
      5. Page 11, line 5. should be " late invasion".
      6. Page 30. Expression of recombinant protein and kinases assay should be moved to the beginning of the Methods as it is mentioned first in the results.
      7. Page 31, line 10. Concentration of imidazole needs to be mentioned.

      Significance

      The manuscript defines the role of PFPK2 in parasite invasion as well as in asexual maturation. Although unambiguous PfPK2 targets have not been defined, but the phosphoroteomic analysis alludes to the role of PfPK2 in cellular processes, particularly in invasion.

      The manuscript will be of general interest to readership of the journals listed.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      We thank all Reviewers for their detailed and helpful comments and suggestions for this manuscript. The overall goal of this study is to interrogate which transcriptional and metabolic pathways lose oscillation when MYC is amplified or activated. We have now added additional replicates to our RNA-sequencing and nutrient transporter expression analyses, and have demonstrated that MYC disrupts oscillation of metabolic and biosynthetic gene expression, nutrient transporter oscillation, and metabolite pathways. On the suggestion of Reviewer #3, we have also strengthened this work by directly contrasting the transcriptional oscillations we observe in cancer cells with well-established primary cell models of transcriptional oscillation, MEFs and macrophages. We have carefully responded to each comment and have noted which Figures or lines in the manuscript address each comment. We hope that our revised manuscript is now suitable for publication.

      Reviewer #1:

      This is an interesting paper from the Altman, Weljie and Dang labs that furthers their previous publications looking at the effect of "oncogenic" Myc levels have on circadian gene expression. They provide compelling data that in neuroblastoma and osteosarcoma models of Myc amplification circadian gene expression and metabolite fluctuation are lost. The data is convincing and comprehensive and should be of broad general interest. There are a few major issues that need to be addressed.

      (Significance (Required)):

      The Myc oncogene is dysregulated in many cancer types, so there is considerable interest in the mechanisms that underpin its function as a transforming oncogene. This group of authors has previously described that Myc can disrupt circadian gene expression, which is linked to several types of cancer. This paper extends the authors previously findings by performing careful timed RNA-seq analysis and metabolomic analysis. The work is well done, and the findings justify the conclusions. This paper should be of interest to those who study Myc, circadian gene expression and cancer. Two key limitations are noted: 1) the cells that are analyzed are grown entirely in vitro in serum and nutrient replete media, 2) there is no direct evidence that the blockage of circadian gene expression by Myc is important for Myc-dependent transformation, although it seems likely. These two limitations do not detract from the significance of the manuscript.

      My expertise is in cancer-centric gene regulatory mechanisms

      We thank Reviewer #1 for finding the data compelling and comprehensive, and for suggesting key experiments and revisions. We also thank the Reviewer for noting that this will be of interest to those who study MYC and cancer (noted in the Significance section below). We have addressed all suggestions below.

      1) Is there a control (done here or in prior literature) showing that simply adding tam to cells doesn't change circadian gene expression?

      We previously performed this experiment in our 2015 Cell Metabolism paper (Altman and Hsieh, et al). In particular, by using either cells that did not express MYC-ER, or a control mutant MYC-ER that lacks transcriptional activity, we found the 4OHT did not blunt clock function in U2OS cells. This is now discussed in Lines 120-121 of the revised manuscript.

      2) All of the experiments rely on Myc:ER fusions. The authors should mine other datasets to determine of systems that rely on conditional expression of wt Myc drive a similar loss of circadian gene expression.

      This is an important point: the work in this manuscript focuses on cancers with endogenous levels of MYC where MYC-ER fusions drive changes in gene expression. In our previous works, we tested systems such as the Reviewer suggested, where endogenous MYC was conditionally overexpressed (Altman and Hsieh et al. Cell Metabolism 2015, Altman et al Nature Communications 2017). In these systems (liver cancer and Burkitt’s Lymphoma), elevated MYC also leads to dysfunction of circadian gene expression and oscillation. We reference this work on lines 123-126. In addition, two papers from the laboratory of Michael Brunner (Shostak et al Nature Communications 2016, Shostak et al Nature Communications 2017) used a tetracycline-inducible model of wild-type MYC in U2OS instead of MYC-ER. Their results largely mirrored our own, where BMAL1 was suppressed and molecular circadian oscillation was disrupted. These two papers are cited extensively throughout manuscript.

      3) The authors argue Myc:ER fusions mimic Myc amplification which is common in cancer and they discuss several previous papers that at least correlate the presence of amplified Myc with loss of circadian gene expression. It should be possible to test whether reduction of Myc by knockdown or using a Myc inhibitor restores circadian gene expression in a cell line known to be Myc amplified. This need not be an exhaustive RNA-seq experiment, but looking at a handful of circadian genes using a qPCR approach would be informative.

      Thank you for the suggestion. Our previous work in Burkitt’s Lymphoma and liver cancer cell lines, as well as analysis of data from primary liver cancer, drew on MYC Tet-OFF systems, where amplified MYC is suppressed by addition of tetracycline or doxycycline. In these experiments, the ‘control’ is amplified MYC, and the ‘experimental’ condition is tet- or dox-treated cells where MYC has been suppressed. In all cases, suppressing MYC led to predictable changes in circadian gene expression (suppression of REV-ERBα, upregulation of BMAL1, etc).

      To support these findings, upon the suggestion of the Reviewer, we identified a recent study where the PC3 prostate cancer cell line, known to harbor highly elevated endogenous MYC levels, was treated with the new generation MYC inhibitor MYCi361. We downloaded their RNA-sequencing data and performed differential expression analysis, and showed that several circadian genes were significantly altered upon treatment with the inhibitor, including BMAL1, PER2, and the REV-ERB genes. This is now included as Supplemental Figure S1C, and the text for this is on lines 126-136.

      4) The data pretty clearly shows that metabolites lose their periodicity MycON cells. Can these be linked back to loss of circadian expression of specific genes in those metabolic pathways? If so, are genes direct Myc transcriptional targets in other studies?

      Thank you for bringing this up. We now have computational evidence from multiple replicate circadian time-series experiments that MYC disrupts oscillation of the LAT1 amino acid transporter across multiple cell lines (Figure 6), and that MYC upregulates LAT1 and 4F2hc protein and mRNA (Supplemental Figure 6 and not shown). Indeed, LAT1 is known to be a direct MYC target, which is now mentioned on lines 493-499 of the discussion.

      In U2OS, where we performed our metabolomics studies, while LAT1 oscillates at the protein level, we saw less evidence of metabolic program oscillation at the transcriptional level, which we acknowledge on lines 274-277. This may be due to the fact that not all metabolic and protein oscillations arise solely from transcriptional oscillation, which we mention on lines 48-49 of the introduction, and revisit in lines 430-432 of the discussion. Nonetheless, our findings that MYC disrupts oscillation of nutrient transporters and metabolites fits in with the overall theme of this manuscript that MYC disrupts circadian control of metabolism.

      5) The finding Myc activation "releases" metabolic and biosynthetic pathways from circadian control implies that this must have something to do with Myc-dependent transformation. A priori, it is not obvious why this should be the case. Do metabolic precursors and biosynthetic molecules become periodically limiting when their levels oscillate in MycOFF cells? In MycOn cells do the non-oscillating metabolites, provide a growth advantage? This is a difficult question to address and one that is certainly beyond the scope of this manuscript. The authors should address this issue in their discussion.

      Thank you for the suggestion to discuss this idea in more detail. While we propose the hypothesis that circadian metabolic oscillations are limiting for tumor cells, testing this directly is indeed outside the scope of this current study. We address this issue on lines 506-518 of the discussion, where we contrast our hypothesis with the idea that alternate metabolic oscillations (those tied to cell cycle or faster-than-circadian) may arise in the absence of circadian control.

      Minor points<br /> 1) phrase "for the first time" is used multiple times in the discussion. Gets a bit redundant ( and loses impact). Consider revising.

      Thank you for this suggestion. We have revised our Discussion section accordingly.

      2) In figure 4, the periodicity in expression of the proteins in figure 4 is fairly clear, but it might be beneficial to bracket (or denote in some other way) the circadian fluctuation in expression.

      In response to other Reviewer comments, we performed multiple replicates of the nutrient transporter protein expression time-series, quantified protein, and calculated circadian oscillations. This is now presented in the new Figure 6.

      Referees cross-commenting

      I had not considered the important points raised by reviewer 3. The authors definitely need to address the concern over replicates and whether the gene expression of truly rhythmic. If not, this seems like a fatal flaw in the MS.

      We have carefully addressed both of these concerns by adding replicates to our RNA-sequencing and protein expression assays, and contrasting our gene expression oscillation findings with those in established primary cell models (Supplemental Figure 2A). Please see the Response to Reviewer 3 for more detail.

      Reviewer #2:

      (Significance (Required)):<br /> Using the circadian synchronized cancer cell lines, DeRollo and colleagues characterized the MYC oncoprotein role in metabolic role through the circadian clock disruption. Authors found that forced activation of MYC disrupts up to 85% of genes oscillation particularly nutrient transporter glycosylation and amino acid metabolism. This work addresses important questions in the circadian clock and cancer field through the oncogene activation, and the manuscript is well-written. However, there are a few concerns that should be addressed to improve the manuscript quality.

      We thank Reviewer #2 for their detailed and astute suggestions on demonstrating the degree of MYC overexpression and the synchronization / entrainment of our cells, as well as the suggestion to add and quantify multiple replicates. We also appreciate the Reviewer’s comments in the Significance section that the manuscript addresses important questions in the field and is well-written. We have individually addressed each comment below and made several revisions and additions in response to them.

      Major comments:<br /> 1) The 3 cell lines used in this paper, what is the expression levels of MYC protein under -OFF and -ON conditions? It is important to demonstrate this information through the western blot data. Since the 4-hydroxy tamoxifen was used to activate MYC, what is the vehicle/control for MYC OFF cells? Otherwise, it will be difficult to assess with everything observed on this manuscript under MYC-ON could be due to 4-hydroxy tamoxifen treatment.

      Thank you for this important consideration. We have clarified in the manuscript that the MYC-ER system is constitutively expressed, and when cells are treated with 4OHT, MYC-ER is activated and translocates to the nucleus, while MYC-OFF control cells are treated with ethanol as a vehicle (Lines 100-105). In response to the suggestion to quantify the degree of overexpression, we have also added new experiments to quantify the degree of MYC-ER overexpression, in Supplemental Figure 1A and 1B. Finally, we previously showed in our 2015 Cell Metabolism paper (Altman and Hsieh, et al) that 4OHT does not affect the molecular clock on its own. In particular, by using either cells that did not express MYC-ER, or a control mutant MYC-ER that lacks transcriptional activity, we found the 4OHT did not blunt clock function in U2OS cells. This is discussed in Lines 120-121 of the revised manuscript.

      2) In Fig. 1A, it is crucial to demonstrate that the circadian synchronization protocol is working by performing statistical analysis with at least 3 biological replicates. This should be performed by either cosinor analysis and/or JTK cycle analysis of all the canonical clock genes including BMAL1 (ARNTL), CLOCK, CRY1, CRY2, PER1, PER2, DBP and NR1D1. This reviewer would like to see both transcripts (qPCR) and protein levels (western blot data) of those clock genes expression pattern. Without these results, rest of the data will be hard to conclude the connection with the circadian/molecular clock.

      Thank you for bringing of the need for quantitation of circadian transcripts. In the new Figure 1, we have quantified and performed ECHO analysis (which is a parametric method of oscillation analysis used through the manuscript) on several circadian transcripts. We chose to specifically show the same n=2 input RNA that we used for RNA-seq for each cell lines. Additionally, our updated RNA-sequencing and analysis of oscillating genes in MYC-OFF shows that CRY2, PER2, PER3 oscillate in all three cell lines (Figure 2B). These findings agree with extensive literature by us and others that the molecular circadian clock is functional after dexamethasone entrainment in U2OS, SHEP, and SKNAS: Baggs et al Plos Biology 2009, Zhang et al Cell 2009, Hughes et al Plos Genetics 2009, Altman and Hsieh et al. Cell Metabolism 2015, Altman et al Nature Communications 2017, Shostak et al Nature Communications 2016, Shostak et al Nature Communications 2017.

      3) In Fig. 5A-C: It is important to repeat this western blot experiment at least 3 times and have the quantitation to demonstrate the circadian rhythmicity significance by probing to majority of the canonical clock proteins as discussed above.

      Thank you for this suggestion. The western blot experiments have now been repeated and have n=3-4 replicates, have been quantified, and oscillation assessed. We also quantified and plotted the oscillation of BMAL1 and REV-ERBα as comparisons. This is in the new Figure 6.

      Minor Comments:

      1) For all of the western blot images, authors need to show the molecular weight of the corresponding protein bands detected on the blots.

      All raw western blot images, including molecular weights, will either be published as Supplemental Material or on FigShare (with a persistent doi), depending on the preference of the Journal. A private link is available for Reviewers with all the relevant background data, including westerns with molecular weights.

      2) As a proof of concept, it would be interesting if knockout/knockout the MYC gene in these cell lines and look for the expression pattern of canonical clock gene expression levels to see whether it will help enhancing circadian rhythmicity. If authors cannot perform this experiment, it is important to address under discussion.

      Thank you for this idea. Reviewer 1 had a similar idea / comment. We addressed it through previous studies, and a new analysis of MYC-high PC3 prostate cancer cells treated with the MYC inhibitor MYCi361 (Supplemental Figure 1C). See response to Review 1 Major Point 3 for more details.

      3) It is not discussed, how many biological replicates were used for RNA-seq analysis?

      In the revised manuscript, each RNA-sequencing experiment is performed from n=2 biological replicates and n=13-14 time points per replicate.

      Referees cross-commenting

      I completely agree with reviewer #3. In fact, I have raised the similar points in my major comments #2 and # 3.

      As discussed below, we have responded to Reviewer comments with more replicates and new analyses of oscillating genes and proteins.

      Reviewer #3:

      Review of "MYC Disrupts Transcriptional and Metabolic Circadian Oscillations in Cancer and Promotes Enhanced Biosynthesis"<br /> DeRollo et al. attempt to find commonalities in how MYC affects transcriptional and metabolic programming by examining MYC-switchable U2OS, SHEP, and SKNAS cell lines. They claim that oncogenic MYC both represses transcriptional oscillation of many genes and supports rhythmic expression of other genes. They use RNA-Seq and UPLC-MS/MS with the appropriate bioinformatics analyses. In some cases, they employ qPCR and immunoblotting. In the three different cell lines, they observed that MYC either statically upregulates or downregulates oscillatory genes.

      (Significance (Required)):

      Myc is known to interfere with the rhythmic expression of core circadian clock genes. Myc seems to do this in order to rewire control-clock expression programs in favor of cell growth and proliferation.<br /> DeRollo et al intended to investigate which clock-controlled expression programs are deregulated by MYC. For this purpose, they investigated three cell lines. Unfortunately, it seems that clock-controlled genes are not really express with a (sufficiently) substantial amplitude in these cultured cells. It is therefore not possible to distinguish by RNA-seq the truly rhythmic genes from false positives. Therefore, it is not possible to reliably determine which metabolic rhythmic programs are deregulated by MYC.

      We thank Reviewer #3 for their important observations and suggestions with regards to the number of replicates employed, and the confidence in the oscillations we observe. We have responded to these comments in a detailed fashion by adding replicates to our RNA-sequencing and immunoblot, and by comparing the oscillations in our cell lines to established primary models of circadian oscillation to determine the amplitudes of oscillations we observed.

      As it stands, the work has technical and conceptual weaknesses.<br /> First, it is not clear how many replications the authors performed for RNA seq. For U2OS, this is explicitly stated. Replicate 1: four-hours sampling, ribosomal RNA depleted; replicate 2: two-hours sampling, polyA+ RNA. There do not appear to be replicates for the other two cell lines?

      Thank you for this observation. In the initial manuscript submission, we used n=2 replicates for U2OS and SHEP, and n=1 replicate for SKNAS (apologies for this not being clear). In response to your comment and others, we added a new biological replicate for RNA-sequencing for SKNAS, so each cell line now has n=2 replicates. This allowed us to more confidently identify oscillating genes across biological replicates for each cell line.

      U2OS cells are widely used in circadian research. These cells rhythmically express clock genes with a decent amplitude, which the authors confirmed by qPCR. However, the clock-controlled genes are generally expressed at a very low amplitude (in the range of standard deviation of RNA-Seq). It is therefore extremely difficult to identify and distinguish them from nonrhythmic genes by RNA seq. The fact that the authors find approximately the same number of rhythmic genes in MYC-OFF (Fig. 1B) as in MYC-ON (Supplemental Fig. 1B) and no overlap between the three cell lines tells that most of the genes shown in the heat maps are not truly rhythmic. Rather, they appear to represent those genes that are called rhythmic because a cosine wave happens to fit the data (better than a line). I suspect that true replicates (which are missing) would also show little or no overlap because most genes in these cells are probably not really rhythmic with any significant amplitude (and why would they be under constant conditions in a Petri dish?).

      Thus, if there are no rhythmic clock-controlled genes that can be clearly distinguished from non-rhythmic genes, there is no way to tell which rhythms are attenuated by MYC (apart from the core clock genes shown in Fig. 1A), or to identify potentially rhythmic pathways.

      Thank you for bringing up this important point about the confidence in the rhythmicity of genes examined. As is correctly noted and as brought up in response to Reviewer #2, cell lines such as U2OS, SKNAS, and SHEP have extensively been used for molecular clock studies (see Reviewer 2 Major Point 2). We also note that each cell line is now n=2 biological replicates, so all oscillating transcripts represent genes that were oscillating in both biological replicates. We take seriously the concern that the oscillating transcripts are not truly rhythmic, or of insufficient amplitude to be biologically significant. We employed a published algorithm, ECHO (De Los Santos et al, Bioinformatics 2019), which uses a conservative parametric approach to determine oscillations from sequencing data, and filters out genes that are too lowly expressed for oscillations to be determined, and those where a sharp increase or decrease in gene expression would preclude determining oscillations.

      To directly test the strength of our observed oscillations in MYC-OFF conditions, we downloaded and analyzed time-series RNA-sequencing data from two entrained primary cell models that are known to have robust transcriptional circadian oscillations: MEFs and macrophages. These two datasets were analyzed in the same fashion as our cell lines, using the ECHO parametric algorithm, and we plotted the median amplitude of oscillation for all transcripts that had significant circadian oscillation (Supplemental Figure 2A). We found that the median amplitude of oscillation was within the same range as those from primary cells: SHEP cells showed nearly identical median amplitude to MEFs, while U2OS and SKNAS had slightly higher median amplitudes than macrophages. These suggest that the oscillating transcripts we observe and measure represent true oscillations above background noise that are similar to those observed in primary cell models where transcriptomic oscillation has been extensively studied.

      With regards to oscillations observed in MYC-ON: we would first like to note that in some cases, there are fewer oscillating genes, especially in SKNAS, where there are less than half the number of oscillating genes in MYC-ON as compared to MYC-OFF. Nonetheless, the observation of emergent oscillations in MYC-ON cells is interesting, and we devote a paragraph to this in the Discussion (lines 457-482). We note that major perturbations to the molecular clock, such as DKO of REV-ERBα and β, result in emergent oscillations in the liver (Guan D et al, Science, 2020), and speculate that oscillations observed in MYC-ON may be from residual activity of CLOCK and BMAL1, which may occupy new sites when MYC is overexpressed.

      If there are no strong rhythmic clock-controlled genes, there are probably no strong rhythms in clock-controlled metabolism. Indeed, the authors found no overlap in rhythmic metabolites.

      Because metabolomics is far less sensitive than RNA-sequencing, we performed KEGG enrichment analysis on oscillating metabolites from both our replicates. We found that in MYC-OFF, the enriched metabolic pathways were identical in timing and identity of enriched pathways (Figure 8, Supplementary Figure 9), while these were quite divergent in MYC-ON. Thus, we concluded that common oscillating metabolic pathways in the absence of MYC are altered or disrupted by MYC activation.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Review of "MYC Disrupts Transcriptional and Metabolic Circadian Oscillations in Cancer and Promotes Enhanced Biosynthesis"<br /> DeRollo et al. attempt to find commonalities in how MYC affects transcriptional and metabolic programming by examining MYC-switchable U2OS, SHEP, and SKNAS cell lines. They claim that oncogenic MYC both represses transcriptional oscillation of many genes and supports rhythmic expression of other genes. They use RNA-Seq and UPLC-MS/MS with the appropriate bioinformatics analyses. In some cases, they employ qPCR and immunoblotting. In the three different cell lines, they observed that MYC either statically upregulates or downregulates oscillatory genes.<br /> As it stands, the work has technical and conceptual weaknesses.<br /> First, it is not clear how many replications the authors performed for RNA seq. For U2OS, this is explicitly stated. Replicate 1: four-hours sampling, ribosomal RNA depleted; replicate 2: two-hours sampling, polyA+ RNA. There do not appear to be replicates for the other two cell lines?<br /> U2OS cells are widely used in circadian research. These cells rhythmically express clock genes with a decent amplitude, which the authors confirmed by qPCR. However, the clock-controlled genes are generally expressed at a very low amplitude (in the range of standard deviation of RNA-Seq). It is therefore extremely difficult to identify and distinguish them from nonrhythmic genes by RNA seq. The fact that the authors find approximately the same number of rhythmic genes in MYC-OFF (Fig. 1B) as in MYC-ON (Supplemental Fig. 1B) and no overlap between the three cell lines tells that most of the genes shown in the heat maps are not truly rhythmic. Rather, they appear to represent those genes that are called rhythmic because a cosine wave happens to fit the data (better than a line). I suspect that true replicates (which are missing) would also show little or no overlap because most genes in these cells are probably not really rhythmic with any significant amplitude (and why would they be under constant conditions in a Petri dish?).<br /> Thus, if there are no rhythmic clock-controlled genes that can be clearly distinguished from non-rhythmic genes, there is no way to tell which rhythms are attenuated by MYC (apart from the core clock genes shown in Fig. 1A), or to identify potentially rhythmic pathways.<br /> If there are no strong rhythmic clock-controlled genes, there are probably no strong rhythms in clock-controlled metabolism. Indeed, the authors found no overlap in rhythmic metabolites.

      Significance

      Myc is known to interfere with the rhythmic expression of core circadian clock genes. Myc seems to do this in order to rewire control-clock expression programs in favor of cell growth and proliferation.<br /> DeRollo et al intended to investigate which clock-controlled expression programs are deregulated by MYC. For this purpose, they investigated three cell lines. Unfortunately, it seems that clock-controlled genes are not really express with a (sufficiently) substantial amplitude in these cultured cells. It is therefore not possible to distinguish by RNA-seq the truly rhythmic genes from false positives. Therefore, it is not possible to reliably determine which metabolic rhythmic programs are deregulated by MYC.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Major comments:

      1. The 3 cell lines used in this paper, what is the expression levels of MYC protein under -OFF and -ON conditions? It is important to demonstrate this information through the western blot data. Since the 4-hydroxy tamoxifen was used to activate MYC, what is the vehicle/control for MYC OFF cells? Otherwise, it will be difficult to assess with everything observed on this manuscript under MYC-ON could be due to 4-hydroxy tamoxifen treatment.
      2. In Fig. 1A, it is crucial to demonstrate that the circadian synchronization protocol is working by performing statistical analysis with at least 3 biological replicates. This should be performed by either cosinor analysis and/or JTK cycle analysis of all the canonical clock genes including BMAL1 (ARNTL), CLOCK, CRY1, CRY2, PER1, PER2, DBP and NR1D1. This reviewer would like to see both transcripts (qPCR) and protein levels (western blot data) of those clock genes expression pattern. Without these results, rest of the data will be hard to conclude the connection with the circadian/molecular clock.
      3. In Fig. 5A-C: It is important to repeat this western blot experiment at least 3 times and have the quantitation to demonstrate the circadian rhythmicity significance by probing to majority of the canonical clock proteins as discussed above.

      Minor Comments:

      1. For all of the western blot images, authors need to show the molecular weight of the corresponding protein bands detected on the blots.
      2. As a proof of concept, it would be interesting if knockout/knockout the MYC gene in these cell lines and look for the expression pattern of canonical clock gene expression levels to see whether it will help enhancing circadian rhythmicity. If authors cannot perform this experiment, it is important to address under discussion.
      3. It is not discussed, how many biological replicates were used for RNA-seq analysis?

      Referees cross-commenting

      I completely agree with reviewer #3. In fact, I have raised the similar points in my major comments #2 and # 3.

      Significance

      Using the circadian synchronized cancer cell lines, DeRollo and colleagues characterized the MYC oncoprotein role in metabolic role through the circadian clock disruption. Authors found that forced activation of MYC disrupts up to 85% of genes oscillation particularly nutrient transporter glycosylation and amino acid metabolism. This work addresses important questions in the circadian clock and cancer field through the oncogene activation, and the manuscript is well-written. However, there are a few concerns that should be addressed to improve the manuscript quality.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      This is an interesting paper from the Altman, Weljie and Dang labs that furthers their previous publications looking at the effect of "oncogenic" Myc levels have on circadian gene expression. They provide compelling data that in neuroblastoma and osteosarcoma models of Myc amplification circadian gene expression and metabolite fluctuation are lost. The data is convincing and comprehensive and should be of broad general interest. There are a few major issues that need to be addressed.

      1. Is there a control (done here or in prior literature) showing that simply adding tam to cells doesn't change circadian gene expression?
      2. All of the experiments rely on Myc:ER fusions. The authors should mine other datasets to determine of systems that rely on conditional expression of wt Myc drive a similar loss of circadian gene expression.
      3. The authors argue Myc:ER fusions mimic Myc amplification which is common in cancer and they discuss several previous papers that at least correlate the presence of amplified Myc with loss of circadian gene expression. It should be possible to test whether reduction of Myc by knockdown or using a Myc inhibitor restores circadian gene expression in a cell line known to be Myc amplified. This need not be an exhaustive RNA-seq experiment, but looking at a handful of circadian genes using a qPCR approach would be informative.
      4. The data pretty clearly shows that metabolites lose their periodicity MycON cells. Can these be linked back to loss of circadian expression of specific genes in those metabolic pathways? If so, are genes direct Myc transcriptional targets in other studies?
      5. The finding Myc activation "releases" metabolic and biosynthetic pathways from circadian control implies that this must have something to do with Myc-dependent transformation. A priori, it is not obvious why this should be the case. Do metabolic precursors and biosynthetic molecules become periodically limiting when their levels oscillate in MycOFF cells? In MycOn cells do the non-oscillating metabolites, provide a growth advantage? This is a difficult question to address and one that is certainly beyond the scope of this manuscript. The authors should address this issue in their discussion

      Minor points

      1. phrase "for the first time" is used multiple times in the discussion. Gets a bit redundant ( and loses impact). Consider revising.
      2. In figure 4, the periodicity in expression of the proteins in figure 4 is fairly clear, but it might be beneficial to bracket (or denote in some other way) the circadian fluctuation in expression.

      Referees cross-commenting

      I had not considered the important points raised by reviewer 3. The authors definitely need to address the concern over replicates and whether the gene expression of truly rhythmic. If not, this seems like a fatal flaw in the MS.

      Significance

      The Myc oncogene is dysregulated in many cancer types, so there is considerable interest in the mechanisms that underpin its function as a transforming oncogene. This group of authors has previously described that Myc can disrupt circadian gene expression, which is linked to several types of cancer. This paper extends the authors previously findings by performing careful timed RNA-seq analysis and metabolomic analysis. The work is well done, and the findings justify the conclusions. This paper should be of interest to those who study Myc, circadian gene expression and cancer. Two key limitations are noted: 1) the cells that are analyzed are grown entirely in vitro in serum and nutrient replete media, 2) there is no direct evidence that the blockage of circadian gene expression by Myc is important for Myc-dependent transformation, although it seems likely. These two limitations do not detract from the significance of the manuscript.

      My expertise is in cancer-centric gene regulatory mechanisms

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      We would like to thank the Review Commons editor and three reviewers for their enthusiastic response, including their constructive suggestions and appreciation of the high impact and originality of our study. We have completed the revisions and new analyses suggested by the reviewers, and we thank the reviewers for their suggestions to increase the impact and interest in this work and for guiding us towards this much improved manuscript.

      In this response letter, we present the response to each reviewer comment and associated revisions made to the text and figures as bullet points below the reviewers' text (black text).

      Reviewer #1 (Evidence, reproducibility and clarity (Required)):

      Summary:

      Yang et al. took advantage of recently published long-read-based genomic sequences of nearly homozygous genomes from complete hydatidiform moles to retrieve allelic sequences of LINE-1, the currently only active and autonomous retrotransposon of the human genome, and produced the repertoire of intact LINE-1 in a genome. The authors performed cell-culture-based retrotransposition assays measurements and in vivo fitness estimations of all identified intact LINE-1 to infer evolutionary dynamics. In this article, the authors further validate the major contribution of polymorphic LINE-1 to the de novo retrotransposition events in the human genome. They also described, at unprecedented resolution, allelic variations among LINE-1 loci and the potential impact of these variations to the interpretation of mutagenic potential of each LINE-1 locus.

      Major comments:

      1 - The key conclusions of the article are mostly convincing. However, it would be a substantial improvement to consolidate the data of the article with information about known active LINE-1s in germ cells or in cancer by using data from recent publications of the Devine and Tubio labs (for example PMID: 34772701, 32024998, 25082706). Across the article, no mention is made of the transductions generated during LINE-1 de novo retrotransposition, which is instrumental to monitor in vivo activity of a group of LINE-1 active copies. It would be of particular interest to make a link between in vitro activity from this study with LINE-1 classification based on their observed activity in cancer (PMID: 32024998, Figure 3b).

      • We thank this and the other reviewers for this suggestion. We agree that a more explicit comparison to the often-reported counts of 3’ transductions would be a valuable addition to our analyses. We have added the 3’ transduction counts from PMID:34772701, PMID:32024998 and PMID:25082706 to Table S2 (column Y, Z and AA), and made a comparison between these data and our Hamming-distance-based in vivo activity, as the new Figure S5. We found correlations between the two measurements in a significant proportion of LINE-1s, but some interesting exceptions exist which likely reflects the fact that most catalogued 3’ transductions come from cancer genomes, and cancer and germline cells represent distinct cellular environments in which distinct sets of LINE-1s are able to replicate (and leave 3’ transductions). In addition to the new figure (Figure S5), we have added a discussion paragraph focused on this interesting comparison.

      2 - The use of CHM1 BAC library Sanger sequencing validation and comparison with CHM13 and hg38 sequences is instrumental to support the building of LINE-1 repertoire in CHM1 genome, which is a valuable contribution of the article. The use of a distance-based metric to infer fitness of a LINE-1 is an interesting approach and allow to group LINE-1 copies based on their in vivo activity potential. Again, it would be beneficial to correlate the inferred fitness and retrotransposition activity of copies/alleles, when known, from the above-mentioned literature.

      • The sequence validation of LINE-1 sequences in CHM1 is an important point which we have addressed in the edited manuscript. Specifically, we used three forms of sequence validation including end-sequencing of one clone of each LINE-1 after it was cloned into the retrotransposition vector and whole-plasmid sequencing of select LINE-1s with discrepant activity amongst the three clones we assayed. In addition, we sequenced the entire LINE-1 sequence for four LINE-1s which had the largest number of mutations relative to their allelic counterpart in CHM13. Please see the above response to ‘Major comment 1’ for details of our new analysis comparing the previous literature to our data.

      3 - Some aspects of the writing of the article should be improved to better support the conclusions.

      • We thank the reviewer for providing these examples of parts of the text that were particularly difficult to read and comprehend. We have deeply streamlined and improved the text throughout the manuscript based upon detailed editing for readability and clarity by two experienced scientific writers. Below, we detail how we revised the particular sections presented by the reviewer, but we think the entire manuscript is now more succinct and clearer.

      • In general, the descriptions are dense, and details could be provided in a more direct way to lighten the results section. Several redundancies in the discussion can be combined to increase clarity.

      • We have spent considerable time tightening up the text, including removing several overlapping sections from the discussion which can be seen in the included version with changes tracked.

      • There is a lack of clarity in the description of how was handled each pair of alleles for which retrotransposition measurements vary between the study and the literature (last paragraph of the "Comprehensive measurement of LINE-1 in vitro activity in a human genome" section). It is not completely clear how the analysis was done and the way the data is presented in File S3 is not helping to support the conclusion. It could be useful to include some illustrative examples in a panel of Figure 2.

      • We agree that this description was hard to parse, and we have rewritten this and accompanying methods to simplify our explanation of these results. In addition, we have revised Figure 2 to show the data in much more detail. To further aid the logic flow related to this section, we moved the previous Figure 5B to Figure 2B, updated it with more suitable examples and edited the associated descriptions.

      • Regarding inferred in vivo activity, the text contains alternative description with the use of "fit" / "unfit", in vivo "active" / "inactive" or "no closely related LINE-1s" terms. The authors should find a way to clearly define and systematically use one set of terms to enhance clarity along the article. To parallel with in vitro active/inactive, it would be useful to use in vivo fit/unfit.

      • We thank the reviewer for this suggestion and agree with their suggested unified use of ‘in vivo fit/unfit’. To clarify and simplify these terms as much as possible, we added detailed explanations of in vivo / in vitro activity and systematically defined in vitro "active/inactive" (page 5, right column, line 50) and in vivo "fit/unfit" (page 8, left column, line 26) at their first appearance in the article, and we changed most instances of "in vivo activity" to "in vivo fitness" when context permits.

      4 - The authors suggest that in vitro activity can be predicted by integration of population frequency and in vivo activity (/fitness) (second paragraph of the "An analysis of LINE-1 evolutionary history [...] and in vivo activity" section). It would be beneficial to strengthen the writing of this section and ultimately validate/test the model by including data from some of the previous studies (e.g. Brouha 2003, Lutz 2003, Seleme 2006, Beck 2010, Rodriguez-Martin 2020, Chuang 2021).

      • We have thoroughly revised this section of the results (see response to ‘Major comment 3’ above), per the reviewers suggestion, to increase reader comprehension of this important data. In addition, we greatly appreciate the reviewer’s suggestion of a very interesting experimental direction – moving beyond a single long-read-based genome to many diverse genomes, and ultimately calculating the in vivo fitness of the LINE-1s from these diverse genomes. For a long time this has not been possible, but the recent publication of the Human Pangenome presents an opportunity to study this interesting question. Though beyond the scope of this paper, our lab is actively working on this fascinating question, and we appreciate the reviewer’s shared interest in this question.

      5 - The identification of adaptive mutations is only partially described and not strongly supported by experimental or analytical data. It would be interesting to explore the role of phylogenetically informative sites described in Figure 5B/C by testing non CHM1 alleles in retrotransposition assay (by introducing amino acid changes into the cloned CHM1 LINE-1 alleles) or by positioning the sites in ORF1p or ORF2p structure and/or domains to infer impact on functionality.

      • The reviewer rightly points out that this is one of the most interesting and novel findings of our manuscript. However, the testing of potentially adaptive mutations is potentially complicated and nuanced. Specifically, we don’t know the mechanism by which these mutations might be adaptive. It is possible that they simply increase in vivo germline retrotransposition activity and this increase would be reflected by an increase of in vitro retrotransposition activity. However, another possibility is that these adaptive phenotypes only show themselves in vivo or in the context of the host restriction factors expressed in the germline. We strongly agree with the reviewer that experimental and analytical data on the phylogenetic informative sites associated with the Figure 5 phylogeny is the key to finding out the mechanisms for these changes to affect LINE-1 activity/fitness, and we are, indeed, exploring this very question in the lab now with related projects. We respectfully suggest that these (extremely cool) experiments are beyond scope of this paper, but we have also added some more detailed description and analyses of the potentially adaptive LINE-1 variations from Figure 5 (from page 9, right column, line 50 to page 10, left column, line 5).

      Minor comments:

      1 - Regarding the in vitro retrotransposition assay, it would be beneficial to provide more data. The current Figure 2 could be enriched by the addition of data related to the variation in the replicates of the experiment (technical but mostly biological with the three clones per LINE-1 tested). Figure 2 could include a dashed line for 100% L1RP and 5% (since it is used as a threshold). It would be useful to provide an additional panel in Figure 2 to illustrate alleles of LINE-1 that are active in this study and compare the values obtained previously in other studies. Similarly, a supplemental table or alignment could be provided to document amino acid changes in the two alleles of each pair (see comment above in the Major Comment 5). The L1Hs subfamilies could also be included in the graph of Figure 2 to support the conclusions of remaining active old L1Hs at allelic forms in the human genome.

      • Upon consideration of this helpful comment, we now augment the presentation of our in vitro activity data with a remade Figure 2 with boxplots to show the variation of the data, as well as a horizontal dashed line showing the active-cutoffs and star signs showing which LINE-1s belong to L1Hs or L1PA2.

      2 - Also, the validation of cloning is not well described. The choice of PCR validation must be supported by more technical details on the design of the primers used to validate each copy. The authors should clearly state that the strategy chosen for retrotransposition assay does not rely on the transcription from LINE-1 5UTR but from an upstream strong promoter, ruling out the role of potential mutations in LINE-1 promoter.

      • As detailed above in the response to ‘Major Comment 1’, we used a combination of end sequencing, whole plasmid sequencing, and multi-read Sanger sequencing to validate the sequences of each LINE-1 cloned from a CHM1 clone. When cloning each LINE-1, we used a specific set of primers designed for the ends of the UTRs for each LINE-1. We have updated the methods and text to clarify this cloning step, and the sequences of these oligos are included in Table S2.
      • To clarify the fact that our retrotransposition assays use a common, strong promoter, we added text in several places stating this setup and discussing (paragraph that starts at page 11, right column, line 18) how 5'UTRs and other non-ORF factors can affect the rate of LINE-1 in vitro activity.

      3 - There are discrepancies with the reported numbers of LINE-1s between Figure 1A and Table S1: 154 vs. 151 in CHM1, 144 vs. 143 in CHM13, respectively.

      • We thank the reviewer for spotting this error on our part. The numbers in Figure 1 and the main text were correct, and we have revised Table S1 to reflect this data.

      4 - The choice of colors in Figure 3 is not perfectly clear and sometimes not as reported in the text (green highlight and orange highlight). Part of the Figure 3 legend is missing. It should include a description of the color code chosen for the right histogram.

      • We thank the reviewer for bringing this inconsistency to our attention. Based upon feedback from all reviewers, we have simplified the color scheme in Figure 3 and Figure 5 to focus on the core conclusions of these two figures. Specifically, in Figure 3, we have removed the quadrant shading and more clearly presented the cutoffs of ‘polymorphic/high frequency’ and ‘in vitro active/inactive’ as dashed lines in the scatter plot. In Figure 5, we have simplified to two colors – black for in vivo unfit and orange to show the in vivo fit LINE-1s which is also used in Figure 4 to show the definition of in vivo activity. These updated colors are now defined in the figure legends and main text, and we have made references to these colors consistent throughout.

      5 - For Figure 4, it would be useful to define in the legends the color code for the top histogram. To better read the scatter plot, the words "fit" and "unfit" could be added on each side of the vertical dashed line.

      • We thank the reviewer again for suggestions to improve the clarity of our figures. As mentioned above in ‘Minor comment 1’, we have removed unnecessary colors including the gradient of the histograms in Figure 3 and Figure 4, since the boundaries of each bin are already defined by the axis labels and tics. As suggested, we have also added ‘fit’ and ‘unfit’ labels to the dashed cutoff line in Figure 4 to clarify the meaning of this line.

      6 - In panel B of Figure 5, it seems that the color code and hot/cold description is not fully formatted.

      • This formatting error has been corrected.

      Reviewer #1 (Significance (Required)):

      In this article, Yang and colleagues present an unprecedented view of the allelic diversity of young LINE-1 copies related to variable retrotransposition activity in an individual genome. One key aspect of their work is the description of the presence of young active LINE-1 alleles that are absent or non-intact in other genome assemblies, while described at a lower scale in initial work from the Kazazian and Moran labs, cited in the manuscript. The work of Yang et al. demonstrates the requirement of multiple approaches and long-read-based sequencing of individual genomes to fully infer the mutagenesis risk of LINE-1 activity.

      The data and methods provided by the authors open the door to a more systematic analysis of mutations and rare allelic forms to understand both mechanistic aspects and evolution of LINE-1 retrotransposition in the human genome. The identification of rare allelic forms of old LINE-1 that retain activity despite previously being considered as inactive is particularly interesting in the light of LINE-1 evolution in the human genome. The authors also describe allelic diversity inside of the Ta1d subfamily, suggesting further diversification and emergence of LINE-1 subgroups. Together with the identification of nucleotide polymorphism among LINE-1 copies, these findings strengthen the notion of individual genomes with individual set of potentially mutagenic LINE-1 alleles.

      The findings and methods described in this article are of great interest to a wide audience including the fields of research focusing on human genome evolution, transposable elements, genomic instability, human genetic variation, and personalized medical diagnostic.

      Aurélien J. Doucet CNRS - Université Côte d'Azur

      Reviewer #2 (Evidence, reproducibility and clarity (Required)):

      This manuscript is an interesting and well-crafted study of LINE-1 activity at the single genome human genome level using long read-based haploid assemblies. The manuscript has some real gems and address critical aspects of LINE- biology that are typically not rigorously examined. The authors are to be commended for undertaking this exercise and for providing interesting perspectives that challenge the dogma that dominates the field in several areas. Despite the noted strengths of the contributions, the manuscript ignores the clear limitations inherent to the approaches taken and at times appears as dogmatic as the dogma that they themselves are trying to challenge. These deficiencies should be addressed before this manuscript is published.

      • We thank Reviewer 2 for their enthusiastic appreciation of the value and innovation of our manuscript. We also thank the reviewer for encouraging us to make careful consideration of the missing references relevant to our findings. We have had two researchers with experience in relevant fields edit our text for both readability, clarity, and proper inclusion of relevant references. We have added these throughout and taken careful effort to replace ‘dogmatic’ statements with clear presentations of the data and thorough referencing of the relevant literature.

      Several major and minor points to consider during revision include:

      Major:

      1. Several strategies have been published in the past that have confidently assign LINE-1s to specific loci despite use of shorter reads. These works should be acknowledged, even if as stated in the manuscript, use of longer reads will only continue to add confidence and validity to future assignments.

      2. We thank the reviewer for this suggestion, and we apologize for the omission of these important publications. As noted above, we have added numerous relevant references (reference 17-27 in the revised text) throughout the text including previous work that used short reads to confidently assign polymorphic/non-reference LINE-1s to specific loci. For example, we now cite the MELT pipeline to detect de novo L1 insertions with short reads (PMID: 28855259), and Iskow et al. 2010, which detects LINE-1s with junction fragment sequencing (PMID: 20603005). We have also added additional text to clarify that short reads are, indeed, often sufficient to place new LINE-1 insertions, while long reads are especially useful for resolving the sequence and location of these insertions. The new text (page 2, left column, line 22-30) presents the advantages/disadvantages of both short reads and long reads.

      3. One of the important requirements for precise quantification of LINE-1 activity and predicted risk scores cited in the manuscript was the need to predict activity based on sequence and location. This requirement, as posited in the manuscript, ignores the critical role of epigenetic control in the regulation of LINE-1 activity. As such, a discussion that acknowledges the critical roles of histone and DNA covalent modifications, and that integrates epigenomic insight into predictions of LINE-1 activity must be included in the manuscript.

      4. We thank the reviewer for suggesting this important discussion point. In response, we have expanded our discussion of this topic to place our data in the context of other literature on the effects of epigenomic regulation on in vivo LINE-1 activity, including histone and DNA modifications, as well as the effects of post transcriptional restriction factors (paragraph starting at page 11, right column, line 42).

      5. The limitations associated with the use of the CHMI were not addressed in the manuscript. While CHMI contain a paternal only genome, with no maternal contribution, the moles may arise from fertilization of an anuclear empty ovum by a haploid 23,X sperm or fertilization by two sperm giving rise to 46,XX or 46,XY karyotype. As such, generalizable conclusions about CHMI genetics should be carefully made given that the loss of maternal epigenetic imprinting and gain of paternally imprinted expression may result in abnormal gene expression, including that of LINE-1s. These variances will in turn impact LINE-1 activity profiles.

      6. We thank the reviewer for pointing out this confusingly written section of our manuscript, and we agree with the reviewer that LINE-1 activity measurements could be complicated in the CHM cell lines; however, all of our retrotransposition assays were carried out in the common background of 293T cells (chosen because of their low expression of know LINE-1 restriction factors (PMID: 25182477). We have modified the text (page 11, right column, line 52) to clarify these points.

      Minor

      1. Important citations of previously published work are not properly referenced throughout the manuscript. These are too numerous to identify individually, but the authors should carefully read the manuscript to ensure that proper documentation and reference to previous work is duly acknowledged.

      2. Please see our above response to ‘Major point 1’.

      There are several typos and missing prepositions that should be corrected. For instance, on page 7, the word "great" should be "greater".

      • Please see our above response to ‘Major point 1’ and Reviewer 1’s ‘Major comment 3’ for details on our in depth editing of the manuscript.

      Reviewer #2 (Significance (Required)):

      The contribution is highly significant as it challenges previously held concepts and advances our understanding of critical structure and function relationships of Line-1s.

      Reviewer #3 (Evidence, reproducibility and clarity (Required)):

      Yang et al. perform an in-depth analysis of potentially mobile source L1 alleles in a single human genome (CHM1) previously subjected to Pacbio whole genome sequencing. The retrotransposition efficiencies of source L1 alleles with intact ORFs were tested in vitro, and these efficiencies compared to a model of in vivo activity based on Hamming distance to other ORF-intact L1 alleles. Comparisons of CHM1 L1 alleles are made to CHM13 (used for the recent T2T reference assembly), and also to population-scale sequencing efforts to establish how widespread each source L1 allele is. These data showcase the advantages of being able to resolve L1 alleles with long-read sequencing, allowing the field to make much more accurate predictions of retrotransposition potential in a given genome. The core analyses appear robust and for the most part enough detail is provided to follow what was done.

      • We thank Reviewer 3 for their in depth reading and analysis of our manuscript and data, and for their enthusiasm about the importance of this work in the context of foundational research from their lab and many others in the field. We have carefully considered each comment and completed several new analyses of our data and related data from other publications. We feel that our manuscript is much improved with this new data, as detailed below. Comments:

      1) The text overlooks the potential importance of L1 5'UTR mutations in L1 activity and evolution, as per PMID:25274305, PMID:1701022, and other studies, as well as the impact of genomic context on source L1 activity, as per PMID:27016617, PMID: 33186547 etc. L1 promoter evolution is arguably a major driver of L1 lineage emergence.

      • We thank the reviewer for suggesting these important additions. To present the relevance of 5'UTR mutations on LINE-1 activity and evolution, we added a discussion paragraph (paragraph starting at page 11, right column, line 16) to address how 5'UTRs and other non-ORF factors can affect the rate of LINE-1 in vitro activity. Several key references have been added and discussed in the paragraph: PMID:25274305 reported the regulation of human LINE-1 by the evolution of its 5'UTR; PMID:1701022 was one of the earliest papers that found the effect the 5'UTR promoters on human LINE-1 retrotransposition; PMID: 27016617 and PMID: 33186547 reported specific L1 loci regulated by different promoters and was included in the discussion; PMID:9430649 was one of the examples of non-human LINE-1 lineages emerging because of different promoters and was cited in the added discussion paragraph. We have also added discussion points to make clear that genomic content has a clear role in the activity of source LINE-1s (paragraph starting at page 11, right column, line 42).

      2) The way the retrotransposition assay is done here (I think) removes parts of the UTRs as part of introducing L1s into retrotransposition vectors, meaning that the assay tests the biochemical activity of the ORFs. It would be helpful to readers to have a more detailed method for this assay, including the origins of the reporter plasmids, whether there is a CMVp boosting the L1 promoter etc, and some clarity about how much of each L1 was cloned into the assay.

      • We have added relevant details to the results (page 6, left column, line 5), discussion (page 11, right column, line 52), and methods (page 13, right column, line 16 and 30) sections to clarify the reviewer’s important points. The LINE-1s tested for in vitro activity were cloned in their entirety (UTRs and ORFs) but driven by both their native promoters in the 5'UTR as well as an upstream CMV promoter. Also, please see our response to Reviewer 1 ‘Minor comment 2’ above.

      3) Pacbio long-read sequencing has been used previously to locate and characterise L1 alleles in human DNA. The Introduction states: "These represent the first scalable methods to catalog LINE-1 locations and sequences in individual human genomes". The "first" here is questionable. Citations to PMID:31853540 and PMID:34772701 should be included. The latter is particularly relevant at it not only resolves source L1 sequences with PacBio sequencing but also summarises their retrotransposition efficiencies in vitro and population frequencies.

      • We apologize for leaving out these and other important references, and we agree that the “first” claim is unnecessary. We have added the references suggested for the reviewer as well as several other important references as detailed in the above response to Reviewer 2 ‘Major point 1’. In addition, we have revised the adjacent text and deleted any references to our work as the “first” in these approaches.

      4) I am very interested in the two source L1s (on chr7 and chr9) that were found here to be more active in vitro than L1RP (to my knowledge the most active such element isolated to date, or close to it). Is there anything unusual about these two L1s? A quick look at the supplemental suggested the chr9 element was 5' truncated, was it tested as such in vitro? Also I think it would be worth contrasting the assay (all in HEKs) used here to test efficiency with the assay used by Brouha ... I feel readers may be surprised to find two L1s more mobile than L1RP in one genome.

      • To provide more details about the two active L1s (chr7 and chr9), we investigated key changes that could be related to the in vitro activity of these elements and now show them in Figure 2B and File S3. In the process of this updated analysis and suggested modifications to Figure 2 by this reviewer and Reviewer 1, we saw that the chr7 L1, mentioned here, had one very high activity measurement pulling its activity above L1RP. As such, we decided to more rigorously normalize our data by using the positive and negative controls across all plates of each day instead of normalizing to the controls of individual plates, as we had previously done. In addition, for any L1 with discrepant activity among the three clones we assayed, we used whole plasmid sequencing to confirm the identity and consistency of all three clones. In three cases, we found that one or two of the three clones was the wrong L1, and hence excluded them for the in vitro activity calculation. After this validation and testing of additional clones, all clones from the same L1 have consistent in vitro activity (see updated Figure 2). The updated in vitro activity of the chr7 L1 is at 86.7% L1RP, and the chr9 L1 is at 261.4% L1RP in addition to the chr17 LINE-1 with 117% L1RP and two additional LINE-1s that have near-L1RP activity levels (Table S2, column S). These changes in L1 activity were updated in the text, figures, and supplemental materials. Also, we note that the chr9 element is 6019bp in length and was tested as such in vitro. Current work in the lab is attempting to understand the mechanisms of increased LINE-1 in vitro and in vivo activity, as described in detail in response to Reviewer 1’s ‘Major comment 5’.

      5) In several places it is mentioned how L1 alleles may differ from sequences provided in reference assemblies, and may therefore explain discrepancies between assay results here and in other studies (e.g. Brouha). The Seleme and Lutz papers are correctly mentioned here, but arguably the most complete demonstration of this concept, from PMID:31230816, is overlooked. This study reports a chr13 source L1 that was previously found to be inactive by Brouha, and with broken ORFs in the reference genome, has both mobile and immobile alleles in the human population. This L1 is actually in CHM13, but not CHM1, and is "hot" in some individuals and not others. There are several places in the manuscript where this earlier study is very relevant and it would be fair to ask it to be mentioned, especially as the results are concordant. The same concept is reinforced by an even more recent paper (PMID:35728967), except in macaque, showing that this is a general consideration for primate L1 lineages, and actually that source L1 is relatively old and yet jumps extremely well in vitro, which fits an observation made in the present study. Mutually supporting observations like these really add confidence that what is reported in the present study is robust.

      • We thank the reviewer for their suggestion to include these highly relevant and important papers; we apologize for this initial omission. We have now added several sentences to the introduction and discussion (top left paragraph page 11) in addition to citations of these papers.

      6) Hamming distance between ORF-intact source L1 alleles is used to assess in vivo activity. This seems reasonable. However, in other works, transductions have been used to identify families of very closely related L1s. I realise that many highly mobile source L1s will rarely generate insertions carrying transductions, and yet I wonder if any of the youngest L1s in the present study form transduction families, and whether estimates of in vivo activity based on transductions found in population-scale data would reconcile better with in vitro retrotransposition assay data.

      • We thank the reviewer for pointing out our exclusion of data on 3' transductions, the most commonly used surrogates of in vivo activity, while also acknowledging that only a small percent of new L1 retrotranspositions carry 3' transduction. Please see our above response to Reviewer 1’s ‘Major comment 1’ for details on our newly added comparison of our in vivo activity data to the 3' transduction-based somatic LINE-1 retrotransposition landscape of those reported in PMID:34772701, PMID:32024998 and PMID:25082706.

      7) In the Introduction, it is stated that L1 only transmits vertically. It may be prudent to mildly qualify this position, based on PMID:29983116.

      • The referenced text in the introduction has been changed from "LINE-1s only transmit vertically" to "LINE-1s generally transmit vertically with few exceptions", with the addition of the suggested citation.

      8) A column in Table S2 looks mislabelled: Column R should be CHM1 not CHM13?

      • We thank the reviewer for seeing this error. Column P (Column R in the previous version) of Table S2 is now correctly labeled as "CHM1 L1 intactness".

      Geoff Faulkner (University of Queensland)

      Reviewer #3 (Significance (Required)):

      This is a well-executed study of considerable interest to the mobile DNA field, and anyone working with long-read DNA sequencing. Its strengths are the genomic and bioinformatic analysis, leveraging the PacBio long-read data and BAC library available for CHM1 to full effect. One limitation (in current form) is its near-exclusive focus on ORFs to encapsulate how mobile a given L1 allele is, when genomic context and L1 promoter mutations could also contribute heavily. Although I liked the manuscript very much and enjoyed reviewing it, some of the conceptual advances are encroached upon by other work (including some very relevant and yet uncited literature). These issues can very likely be addressed via a revision, additional analyses may be required but not new experiments.

      Geoff Faulkner (University of Queensland)

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Yang et al. perform an in-depth analysis of potentially mobile source L1 alleles in a single human genome (CHM1) previously subjected to Pacbio whole genome sequencing. The retrotransposition efficiencies of source L1 alleles with intact ORFs were tested in vitro, and these efficiencies compared to a model of in vivo activity based on Hamming distance to other ORF-intact L1 alleles. Comparisons of CHM1 L1 alleles are made to CHM13 (used for the recent T2T reference assembly), and also to population-scale sequencing efforts to establish how widespread each source L1 allele is. These data showcase the advantages of being able to resolve L1 alleles with long-read sequencing, allowing the field to make much more accurate predictions of retrotransposition potential in a given genome. The core analyses appear robust and for the most part enough detail is provided to follow what was done.

      Comments:

      1. The text overlooks the potential importance of L1 5'UTR mutations in L1 activity and evolution, as per PMID:25274305, PMID:1701022, and other studies, as well as the impact of genomic context on source L1 activity, as per PMID:27016617, PMID: 33186547 etc. L1 promoter evolution is arguably a major driver of L1 lineage emergence.
      2. The way the retrotransposition assay is done here (I think) removes parts of the UTRs as part of introducing L1s into retrotransposition vectors, meaning that the assay tests the biochemical activity of the ORFs. It would be helpful to readers to have a more detailed method for this assay, including the origins of the reporter plasmids, whether there is a CMVp boosting the L1 promoter etc, and some clarity about how much of each L1 was cloned into the assay.
      3. Pacbio long-read sequencing has been used previously to locate and characterise L1 alleles in human DNA. The Introduction states: "These represent the first scalable methods to catalog LINE-1 locations and sequences in individual human genomes". The "first" here is questionable. Citations to PMID:31853540 and PMID:34772701 should be included. The latter is particularly relevant at it not only resolves source L1 sequences with PacBio sequencing but also summarises their retrotransposition efficiencies in vitro and population frequencies.
      4. I am very interested in the two source L1s (on chr7 and chr9) that were found here to be more active in vitro than L1RP (to my knowledge the most active such element isolated to date, or close to it). Is there anything unusual about these two L1s? A quick look at the supplemental suggested the chr9 element was 5' truncated, was it tested as such in vitro? Also I think it would be worth contrasting the assay (all in HEKs) used here to test efficiency with the assay used by Brouha ... I feel readers may be surprised to find two L1s more mobile than L1RP in one genome.
      5. In several places it is mentioned how L1 alleles may differ from sequences provided in reference assemblies, and may therefore explain discrepancies between assay results here and in other studies (e.g. Brouha). The Seleme and Lutz papers are correctly mentioned here, but arguably the most complete demonstration of this concept, from PMID:31230816, is overlooked. This study reports a chr13 source L1 that was previously found to be inactive by Brouha, and with broken ORFs in the reference genome, has both mobile and immobile alleles in the human population. This L1 is actually in CHM13, but not CHM1, and is "hot" in some individuals and not others. There are several places in the manuscript where this earlier study is very relevant and it would be fair to ask it to be mentioned, especially as the results are concordant. The same concept is reinforced by an even more recent paper (PMID:35728967), except in macaque, showing that this is a general consideration for primate L1 lineages, and actually that source L1 is relatively old and yet jumps extremely well in vitro, which fits an observation made in the present study. Mutually supporting observations like these really add confidence that what is reported in the present study is robust.
      6. Hamming distance between ORF-intact source L1 alleles is used to assess in vivo activity. This seems reasonable. However, in other works, transductions have been used to identify families of very closely related L1s. I realise that many highly mobile source L1s will rarely generate insertions carrying transductions, and yet I wonder if any of the youngest L1s in the present study form transduction families, and whether estimates of in vivo activity based on transductions found in population-scale data would reconcile better with in vitro retrotransposition assay data.
      7. In the Introduction, it is stated that L1 only transmits vertically. It may be prudent to mildly qualify this position, based on PMID:29983116.
      8. A column in Table S2 looks mislabelled: Column R should be CHM1 not CHM13?

      Geoff Faulkner (University of Queensland)

      Significance

      This is a well-executed study of considerable interest to the mobile DNA field, and anyone working with long-read DNA sequencing. Its strengths are the genomic and bioinformatic analysis, leveraging the PacBio long-read data and BAC library available for CHM1 to full effect. One limitation (in current form) is its near-exclusive focus on ORFs to encapsulate how mobile a given L1 allele is, when genomic context and L1 promoter mutations could also contribute heavily. Although I liked the manuscript very much and enjoyed reviewing it, some of the conceptual advances are encroached upon by other work (including some very relevant and yet uncited literature). These issues can very likely be addressed via a revision, additional analyses may be required but not new experiments.

      Geoff Faulkner (University of Queensland)

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      This manuscript is an interesting and well-crafted study of LINE-1 activity at the single genome human genome level using long read-based haploid assemblies. The manuscript has some real gems and address critical aspects of LINE- biology that are typically not rigorously examined. The authors are to be commended for undertaking this exercise and for providing interesting perspectives that challenge the dogma that dominates the field in several areas. Despite the noted strengths of the contributions, the manuscript ignores the clear limitations inherent to the approaches taken and at times appears as dogmatic as the dogma that they themselves are trying to challenge. These deficiencies should be addressed before this manuscript is published.

      Several major and minor points to consider during revision include:

      Major:

      1. Several strategies have been published in the past that have confidently assign LINE-1s to specific loci despite use of shorter reads. These works should be acknowledged, even if as stated in the manuscript, use of longer reads will only continue to add confidence and validity to future assignments.
      2. One of the important requirements for precise quantification of LINE-1 activity and predicted risk scores cited in the manuscript was the need to predict activity based on sequence and location. This requirement, as posited in the manuscript, ignores the critical role of epigenetic control in the regulation of LINE-1 activity. As such, a discussion that acknowledges the critical roles of histone and DNA covalent modifications, and that integrates epigenomic insight into predictions of LINE-1 activity must be included in the manuscript.
      3. The limitations associated with the use of the CHMI were not addressed in the manuscript. While CHMI contain a paternal only genome, with no maternal contribution, the moles may arise from fertilization of an anuclear empty ovum by a haploid 23,X sperm or fertilization by two sperm giving rise to 46,XX or 46,XY karyotype. As such, generalizable conclusions about CHMI genetics should be carefully made given that the loss of maternal epigenetic imprinting and gain of paternally imprinted expression may result in abnormal gene expression, including that of LINE-1s. These variances will in turn impact LINE-1 activity profiles.

      Minor

      1. Important citations of previously published work are not properly referenced throughout the manuscript. These are too numerous to identify individually, but the authors should carefully read the manuscript to ensure that proper documentation and reference to previous work is duly acknowledged.
      2. There are several typos and missing prepositions that should be corrected. For instance, on page 7, the word "great" should be "greater".

      Significance

      The contribution is highly significant as it challenges previously held concepts and advances our understanding of critical structure and function relationships of Line-1s.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary:

      Yang et al. took advantage of recently published long-read-based genomic sequences of nearly homozygous genomes from complete hydatidiform moles to retrieve allelic sequences of LINE-1, the currently only active and autonomous retrotransposon of the human genome, and produced the repertoire of intact LINE-1 in a genome. The authors performed cell-culture-based retrotransposition assays measurements and in vivo fitness estimations of all identified intact LINE-1 to infer evolutionary dynamics. In this article, the authors further validate the major contribution of polymorphic LINE-1 to the de novo retrotransposition events in the human genome. They also described, at unprecedented resolution, allelic variations among LINE-1 loci and the potential impact of these variations to the interpretation of mutagenic potential of each LINE-1 locus.

      Major comments:

      1. The key conclusions of the article are mostly convincing. However, it would be a substantial improvement to consolidate the data of the article with information about known active LINE-1s in germ cells or in cancer by using data from recent publications of the Devine and Tubio labs (for example PMID: 34772701, 32024998, 25082706). Across the article, no mention is made of the transductions generated during LINE-1 de novo retrotransposition, which is instrumental to monitor in vivo activity of a group of LINE-1 active copies. It would be of particular interest to make a link between in vitro activity from this study with LINE-1 classification based on their observed activity in cancer (PMID: 32024998, Figure 3b).
      2. The use of CHM1 BAC library Sanger sequencing validation and comparison with CHM13 and hg38 sequences is instrumental to support the building of LINE-1 repertoire in CHM1 genome, which is a valuable contribution of the article. The use of a distance-based metric to infer fitness of a LINE-1 is an interesting approach and allow to group LINE-1 copies based on their in vivo activity potential. Again, it would be beneficial to correlate the inferred fitness and retrotransposition activity of copies/alleles, when known, from the above-mentioned literature.
      3. Some aspects of the writing of the article should be improved to better support the conclusions.
        • In general, the descriptions are dense, and details could be provided in a more direct way to lighten the results section. Several redundancies in the discussion can be combined to increase clarity.
        • There is a lack of clarity in the description of how was handled each pair of alleles for which retrotransposition measurements vary between the study and the literature (last paragraph of the "Comprehensive measurement of LINE-1 in vitro activity in a human genome" section). It is not completely clear how the analysis was done and the way the data is presented in File S3 is not helping to support the conclusion. It could be useful to include some illustrative examples in a panel of Figure 2.
        • Regarding inferred in vivo activity, the text contains alternative description with the use of "fit" / "unfit", in vivo "active" / "inactive" or "no closely related LINE-1s" terms. The authors should find a way to clearly define and systematically use one set of terms to enhance clarity along the article. To parallel with in vitro active/inactive, it would be useful to use in vivo fit/unfit.
      4. The authors suggest that in vitro activity can be predicted by integration of population frequency and in vivo activity (/fitness) (second paragraph of the "An analysis of LINE-1 evolutionary history [...] and in vivo activity" section). It would be beneficial to strengthen the writing of this section and ultimately validate/test the model by including data from some of the previous studies (e.g. Brouha 2003, Lutz 2003, Seleme 2006, Beck 2010, Rodriguez-Martin 2020, Chuang 2021).
      5. The identification of adaptive mutations is only partially described and not strongly supported by experimental or analytical data. It would be interesting to explore the role of phylogenetically informative sites described in Figure 5B/C by testing non CHM1 alleles in retrotransposition assay (by introducing amino acid changes into the cloned CHM1 LINE-1 alleles) or by positioning the sites in ORF1p or ORF2p structure and/or domains to infer impact on functionality.

      Minor comments:

      1. Regarding the in vitro retrotransposition assay, it would be beneficial to provide more data. The current Figure 2 could be enriched by the addition of data related to the variation in the replicates of the experiment (technical but mostly biological with the three clones per LINE-1 tested). Figure 2 could include a dashed line for 100% L1RP and 5% (since it is used as a threshold). It would be useful to provide an additional panel in Figure 2 to illustrate alleles of LINE-1 that are active in this study and compare the values obtained previously in other studies. Similarly, a supplemental table or alignment could be provided to document amino acid changes in the two alleles of each pair (see comment above in the Major Comment 5). The L1Hs subfamilies could also be included in the graph of Figure 2 to support the conclusions of remaining active old L1Hs at allelic forms in the human genome.
      2. Also, the validation of cloning is not well described. The choice of PCR validation must be supported by more technical details on the design of the primers used to validate each copy. The authors should clearly state that the strategy chosen for retrotransposition assay does not rely on the transcription from LINE-1 5UTR but from an upstream strong promoter, ruling out the role of potential mutations in LINE-1 promoter.
      3. There are discrepancies with the reported numbers of LINE-1s between Figure 1A and Table S1: 154 vs. 151 in CHM1, 144 vs. 143 in CHM13, respectively.
      4. The choice of colors in Figure 3 is not perfectly clear and sometimes not as reported in the text (green highlight and orange highlight). Part of the Figure 3 legend is missing. It should include a description of the color code chosen for the right histogram.
      5. For Figure 4, it would be useful to define in the legends the color code for the top histogram. To better read the scatter plot, the words "fit" and "unfit" could be added on each side of the vertical dashed line.
      6. In panel B of Figure 5, it seems that the color code and hot/cold description is not fully formatted.

      Significance

      In this article, Yang and colleagues present an unprecedented view of the allelic diversity of young LINE-1 copies related to variable retrotransposition activity in an individual genome. One key aspect of their work is the description of the presence of young active LINE-1 alleles that are absent or non-intact in other genome assemblies, while described at a lower scale in initial work from the Kazazian and Moran labs, cited in the manuscript. The work of Yang et al. demonstrates the requirement of multiple approaches and long-read-based sequencing of individual genomes to fully infer the mutagenesis risk of LINE-1 activity. The data and methods provided by the authors open the door to a more systematic analysis of mutations and rare allelic forms to understand both mechanistic aspects and evolution of LINE-1 retrotransposition in the human genome. The identification of rare allelic forms of old LINE-1 that retain activity despite previously being considered as inactive is particularly interesting in the light of LINE-1 evolution in the human genome. The authors also describe allelic diversity inside of the Ta1d subfamily, suggesting further diversification and emergence of LINE-1 subgroups. Together with the identification of nucleotide polymorphism among LINE-1 copies, these findings strengthen the notion of individual genomes with individual set of potentially mutagenic LINE-1 alleles. The findings and methods described in this article are of great interest to a wide audience including the fields of research focusing on human genome evolution, transposable elements, genomic instability, human genetic variation, and personalized medical diagnostic.

      Aurélien J. Doucet CNRS - Université Côte d'Azur

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Responses to Reviewers’ Comments

      __Reviewer_ #1 (Evidence, reproducibility and clarity (Required)):_ __ *Summary In this study, using genetic labeling and deletion modeling, the authors discovered that Tbx1 myoprogenitors, rather than Pax3+/Myf5+ cells, give rise to supraclavicular brown adipose tissue (scBAT). This finding is both intriguing and significant. Furthermore, the genetic ablation of PPARγ or PRDM16, driven by Tbx1-Cre, reduced the size/weight of scBAT and its thermogenesis function capacity, supporting the importance of Tbx-1 cells. Interestingly, the authors found that human scBAT, located in the deep neck region, exhibits higher Tbx1 expression than subcutaneous neck WAT, potentially indicating a similar origin of scBAT in rodents and humans. Overall, this novel finding is exciting and could push the BAT field into a new phase. The manuscript is also well-written and organized.

      Comments 1. The authors assert that Myf5+ progenitors do not contribute to scBAT adipocytes. However, Figure 2C shows that 7% of medial scBAT are mG positive, suggesting a minor contribution of Myf5+ progenitors to scBAT. The conclusion that scBAT does not originate from Myf5+ precursor cells may be overly strong.*

      __Response____: __We thank the reviewer for pointing out the overstatement. We have rephased the conclusion from Figure 2 as “Myf5+ progenitors seldom give rise to scBAT adipocytes”.

      The authors claim that "prdm16 is dispensable for the development of scBAT." However, supporting data seems insufficient. The absence of a difference in scBAT weight does not guarantee that development was unaffected. Additional experiments, like H&E staining and immunofluorescence (IF) staining of RFP/GFP, could help demonstrate a similar number of GFP+ brown adipocytes in scBAT, thereby supporting this statement.

      Response____: __We do not have direct evidence that Prdm16 is dispensable for scBAT development; therefore, we have removed such statement in the revised manuscript. In addition, we provided H&E staining of scBAT, demonstrating evident adipose whitening and fibrosis in Prdm16ΔTbx1 mice (__Figure 7C). These results are consistent with previous findings that PRDM16 is required for the maintenance of brown adipocyte identity (PMID: 24703692).

      • In the "scBAT contributes to temperature maintenance in mice" section, this phenomenon only seems to apply to female mice (Figure 5). This conclusion may need adjustment to account for this sex difference. Moreover, for females, are there H&E staining results available for scBAT? For males, is there a change in Ucp1 expression? It could be beneficial to examine mRNA expression levels for additional thermogenic genes, such as Dio2, Prdm16, Cidea, and PGC-1a.*

      Response____: __Following the reviewer’s suggestion, we rephased our conclusion from Figure 5 to "scBAT contributes to temperature maintenance in __female mice". Also as requested, we provided the H&E staining of female scBAT (Figure 5E), showing a partial loss of brown adipose identity in PpargΔTbx1 mice. For males, we performed RT-qPCR and calculated total gene expression based on tissue weight. While Pparg, Prdm16 and Dio2 total expression was downregulated in Pparg-deficient scBAT, total Ucp1 expression was unchanged in scBAT and iBAT of male mice (Figure 5N, O). This might explain the similar cold tolerance between wildtype and PpargΔTbx1 males (Figure 5P).

      • Considering that mutant mice have smaller scBAT, how does this difference influence glucose homeostasis between WT and mutants?*

      Response____: __We performed glucose tolerance test and found no difference in glucose homeostasis of HFD PpargΔTbx1 mice (__Figure 6C). This could be due to intact iBAT in these animals, which is a much larger glucose consumer than scBAT.

      • In Figures 5G & M, it seems more accurate for the y-axis label to read "Body temperature change."*

      __Response____: __We changed the y-axis label. Thank you for the suggestions.

      • There is inconsistency in scale bar labeling: o Figure 3A vs 3D o Figure 4B vs 4C o Figure 5L, where scale bars are missing in the H&E staining images o Figure S1, where the scale bar is in the middle of the image*

      __Response____: __In the revised manuscript, we have provided consistent scale bars in all figures, including all these pointed by the reviewer. Thank you.

      • In Figure 3, the western blot data should be quantified, and the molecular weight (kDa) should be included to clarify the band's position.*

      Response____: __We have now indicated molecular weights in __Figure 3C, F and provided quantified UCP1 and PPARγ expression levels to the right of Western blots.

      • The statement "RT-qPCR revealed much higher levels of TBX expression in total lysates from deep neck BAT (Figure 8A)" could be clarified by adding "than neck WAT" at the end. *

      __Response____: __Done. Thank you for your suggestion.

      Reviewer____ #1 (Significance (Required)): Significance The major advancement of this study lies in the authors' novel finding of the embryonic lineage of brown adipocytes in the supraclavicular brown adipose tissue (scBAT) depot. They demonstrated that this lineage differs from those in the dorsal BAT depots by utilizing Myf5-Cre, Pax3-Cre, and TBX1-Cre reporter mouse models.

      __Reviewer____ #2 (Evidence, reproducibility and clarity (Required)): __ *This study aimed to investigate the developmental origin of brown adipose tissue (BAT) in the supraclavicular region and its implications for metabolic health. The authors have used genetic fate mapping in mice to trace the lineage of brown adipocytes in the supraclavicular region. The findings revealed that supraclavicular brown adipocytes do not originate from the Pax3+/Myf5+ epaxial dermomyotome, which is responsible for the development of interscapular BAT. Instead, most supraclavicular brown adipocytes were marked by the Tbx1+ lineage, indicating that the pharyngeal mesoderm is involved in their development. This work provides the first evidence that scBAT adipocytes do not share the same embryonic origins as iBAT fat cells. By identifying the location-specific myogenic progenitors for supraclavicular BAT versus interscapular BAT, the researchers shed light on the distinct developmental origins of different BAT depots. Overall, these findings provide new insights into the developmental origin of supraclavicular BAT and highlight the need to consider the anatomical locations and developmental origins in studying BAT development and function.

      Major comments: The manuscript addresses an important yet understudied area and provides convincing results that support the key conclusions. The authors also engage in extensive discussion, speculating on the broader implications of the findings and outlining the limitations of their study. The observation that the loss of Pparg in Tbx1-expressing cells leads to a reduction in supraclavicular BAT is intriguing. However, as the authors acknowledge, Tbx1 is expressed in inguinal white adipocytes as well. Therefore, it remains unclear how changes in Tbx1 in ingWAT in this model might affect BAT depots. The authors should provide further clarification on this matter and include additional discussion regarding the potential indirect effects and limitations of these models. *

      Response____: __We thank the reviewer for their overall enthusiasm about our manuscript, in which we primarily focus on Tbx1 lineage cells in BAT, the supraclavicular depot to be more specific. Tbx1-Cre mediated genetic deletion of the Pparg gene will inevitably perturb PPARγ function in other Tbx1-expressing tissues, such as scapular muscles and certain WAT depots. In a recent publication (PMID: 32240964), the Tbx1 gene was shown to be expressed in mouse inguinal WAT (iWAT) at a much higher level than interscapular BAT. TBX1 overexpression was sufficient to induce UCP1 protein in iWAT, while TBX1 deficiency rendered mice more sensitive to cold-induced body temperature drop. However, in our PpargΔTbx1 mice, we did not observe any significant loss of Pparg gene expression in iWAT (__Figure S3F), suggesting the possibility that Tbx1-positive cells are only a small proportion of iWAT. Nevertheless, we did not see any reduced expression of thermogenic genes in iWAT of PpargΔTbx1 mice (Figure S3F), indicating either that Tbx1-expressing cells are dispensable for thermogenesis or that other still-undefined mechanisms exist to compensate for the loss of PPARγ. We acknowledge the limitations of our animal models and have discussed these points in the second paragraph of Discussion.

      In Figure 5, it would be beneficial to include the expression of a broader range of thermogenic genes and proteins in both males and females. This would strengthen the argument that the loss of Pparg in Tbx1+ progenitors impairs scBAT function.

      Response____: __Thank you for the suggestion and we have now provided the expression of additional thermogenic genes including Ucp1, Prdm16, Ppargc1a, Cidea, and Dio2 in both males and females (__Figure 5F, G, N, and O). These results demonstrate that loss of PPARγ in Tbx1+ myoprogenitors impairs scBAT thermogenic function, with more profound impact on males than females.

      *Minor Comments: It would be more appropriate to present the results of the cold challenge studies as absolute body temperature rather than just the difference between groups. This is important because there could be baseline differences in basal body temperature among the experimental groups. *

      __Response____: __Thank you for the suggestion. The results summarized in Figure 5H and 5P were from several batches of independent experiments that were carried out at different times. Plotting absolute body temperature created a bigger variance that precluded us from drawing any conclusion. Therefore, we decided to plot the data as changes in body temperature in Figure 5H and 5P. As shown in Figure 8, we did not observe baseline differences in body temperature between wildtype and Prdm16ΔTbx1 mice, suggesting that loss of scBAT thermogenic function would not affect basal body temperature.

      *It is recommended to include representative images of cre-negative animals to validate the reagents and models used in the study. Including such images would enhance the reliability of the experimental approach and strengthen the overall validity of the findings. *

      Response____: __Representative images of Cre-negative mT/mG reporter animals have now been provided in __Figure S1. Thank you for the suggestion.

      The manuscript would benefit from placing the findings in the context of the broader field. For instance, discussing whether the loss of Prdm16 in Myf5 progenitors has a similar impact on interscapular BAT (iBAT) development or thermogenic function would provide valuable insights. Additionally, exploring the relationship between Tbx1 progenitors and adipocyte progenitors in adult BAT depots, such as Pdgfra+ and Trpv1+ progenitors, could further enhance our understanding of the developmental origins and functional characteristics of different adipocyte populations

      __Response____: __The reviewer raised an important point that we should place our finders in the context of the broader field. Indeed, it has been reported earlier that loss of PRDM16 in Myf5 progenitors leads to iBAT dysfunction in adult mice (PMID: 24703692), a similar phenotype to what we found in scBAT of Prdm16ΔTbx1 mice. The early embryonic development was not evidently perturbed in iBAT (using Myf5-cre) or scBAT (using Tbx1-cre), which might be compensated by other PRDM family protein, such as PRDM3. However, using Adipoq-Cre to knock out PRDM16 only leads to defects in beige adipocyte, but not classic BAT (PMID: 24439384). Together, these results demonstrate that PRDM16 in myoprogenitors is required for brown adipocyte identity maintenance during aging and similar transcriptional regulatory circuits control the differentiation and/or activity of both iBAT and scBAT. Further investigations are warranted to identify depot-specific regulations and functions of BAT. In adult BAT, progenitors that are marked by genes like Pdgfra and Trpv1 have been reported to contribute to cold-induced BAT recruitment and tissue homeostasis. While not the scope of our current research, future endeavors are needed to test if embryonic Tbx1+ myoprogenitors give rise to all or only some populations of adult BAT progenitors. These points have now been added to the revised Discussion section.

      Reviewer____ #2 (Significance (Required)): * Field of expertise: Adipose biology, Developmental biology, Thermogenesis

      The manuscript addresses an important and previously understudied area in the field of brown adipose tissue (BAT) development. The study challenges the long-standing assumption that Pax3+/Myf5+ progenitors in the dermomyotome are the sole developmental source of brown adipocytes in mice, including both interscapular BAT (iBAT) and other brown adipocytes.

      This work provides experimental evidence that the supraclavicular BAT (scBAT) adipocytes have a distinct embryonic origin compared to iBAT fat cells. While the developmental source of iBAT has been known for over a decade, this study demonstrates for the first time that scBAT adipocytes do not arise from the Pax3+/Myf5+ progenitors in the dermomyotome.

      The findings presented in this manuscript have the potential to make a lasting impact on the brown adipose tissue research community, particularly those interested in the developmental aspects of brown fat. *

      __Reviewer____ #3 (Evidence, reproducibility and clarity (Required)): __ *This manuscript explores the cell origin of supraclavicular brown adipose tissue (scBAT). UCP1-expressing brown/beige adipocytes are found in several anatomical locations but most studies have been focused on intrascapular brown adipose tissue (iBAT) in rodents, a depot that only exist in infants (but not adults) in humans. Very little is known about the biology (including the origin) of other BAT deposes. Here using multiple lineage tracing tools and conditional KO mice the authors convincingly demonstrate that about 50% of the scBAT cells originate from Tbx1+ myoprogenitors and provide functional evidence that ablation of Pparg or Prdm16 in Tbx1-cells blocks development of scBAT or affects thermogenic function of mice. The study design is straightforward, well carried out and the conclusion is supported by the data. The discovery has significant implication as brown adipocytes are mainly found in the supraclavicular region in the human body. I have few relative minor comments.

      Tbx-Cre lineage tracing indicates that 50% scBAT cells are marked by Tbx1. The authors discussed this may be either due to low recombination efficiency of Tbx1-Cre or due to multiple lineage origin of scBAT cells. The possibility of Cre recombination efficiency can be addressed easily by genomic DNA recombination analysis. *

      Response____: __To test recombination efficiency, we performed RT-PCR to detect wildtype and mutant Pparg mRNA using primers provided in the original report of the Pparg-floxed mice (PMID: 14660788). The mutant Pparg mRNA could only be detected in scBAT, not iBAT, of PpargΔTbx1 mice (__Figure S3A). The similar abundance of wildtype and mutant Pparg transcripts in PpargΔTbx1 scBAT indicates a 50% recombination rate. However, we currently cannot distinguish the following possible reasons for the incomplete recombination: 1) only half of scBAT adipocytes are Tbx1-lineage cells; 2) a portion of Tbx1-progeny cells have low Cre expression and thus no recombination; 3) The Tbx1 promoter does not fully recapitulate the endogenous Tbx1 gene expression. The identity of non-Tbx1 progeny adipocytes in the scBAT depot requires future investigations.

      *In the introduction the authors conclude that Myf5, Pax3 are "location markers" instead of cell identity markers. While I agree with this statement it is quite confusing. It implies that the study aims to identify cell identify markers but based on the analogy Tbx1 is also a location marker. *

      __Response____: __We apologize for the confusion caused by using “identity” to describe both lineage origins and functional cell types. Based on previous and our current studies, we propose Myf5, Pax3, and Tbx1 as location markers that distinguish BAT depots at different locations. They are not identity markers as these progenitors also give rise to other cell types, such as muscles, white adipocytes, fibroblasts, and others. We removed any statements indicating Tbx1 as an identity marker.

      *Figure 8 human expression compares TBX1 mRNA levels in neck fat to subcutaneous white fat. This is different from the comparison of dorsal and ventral BAT in mice. Therefore, the data are not particularly relevant to this study and should be moved to a supplemental figure. *

      __Response____: __We agree with the reviewer that the comparison of human TBX1 was not made between dorsal and ventral BAT. This is due the loss of dorsal BAT in adult humans, which make it almost impossible to make such direct comparison in humans. Nonetheless, the higher expression of TBX1 in deep neck BAT compared to subcutaneous neck WAT suggests that TBX1 might be functionally important for BAT thermogenesis. This is consistent with a recent report using transgenic Tbx1 mouse models that demonstrates the role of adipose TBX1 in thermogenic capacity (PMID: 32240964). We believe that data in Figure 8 is relevant to our current study, but we are also receptive to move it to the supplementary if the reviewer insists so.

      *Can the authors speculate/discuss on why there is no Pax3 labelled cells but 7% Myf5 labeled cells in scBAT? *

      __Response____: __We would like to clarify that both Pax3+ (6.7%) and Myf5+ (7.2%) progenitors label a similar percentage of brown adipocytes in the medial scBAT. These percentages are now provided in the revised text and can be visualized in Figure 1H and 2D. On the other hand, nearly none of lateral scBAT adipocytes were labelled by either Pax3+ or Myf5+ progenitors.

      Reviewer____ #3 (Significance (Required)): The discovery has significant implication as brown adipocytes are mainly found in the supraclavicular region in the human body.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      This manuscript explores the cell origin of supraclavicular brown adipose tissue (scBAT). UCP1-expressing brown/beige adipocytes are found in several anatomical locations but most studies have been focused on intrascapular brown adipose tissue (iBAT) in rodents, a depot that only exist in infants (but not adults) in humans. Very little is known about the biology (including the origin) of other BAT deposes. Here using multiple lineage tracing tools and conditional KO mice the authors convincingly demonstrate that about 50% of the scBAT cells originate from Tbx1+ myoprogenitors and provide functional evidence that ablation of Pparg or Prdm16 in Tbx1-cells blocks development of scBAT or affects thermogenic function of mice. The study design is straightforward, well carried out and the conclusion is supported by the data. The discovery has significant implication as brown adipocytes are mainly found in the supraclavicular region in the human body. I have few relative minor comments.

      Tbx-Cre lineage tracing indicates that 50% scBAT cells are marked by Tbx1. The authors discussed this may be either due to low recombination efficiency of Tbx1-Cre or due to multiple lineage origin of scBAT cells. The possibility of Cre recombination efficiency can be addressed easily by genomic DNA recombination analysis.

      In the introduction the authors conclude that Myf5, Pax3 are "location markers" instead of cell identity markers. While I agree with this statement it is quite confusing. It implies that the study aims to identify cell identify markers but based on the analogy Tbx1 is also a location marker.

      Figure 8 human expression compares TBX1 mRNA levels in neck fat to subcutaneous white fat. This is different from the comparison of dorsal and ventral BAT in mice. Therefore, the data are not particularly relevant to this study and should be moved to a supplemental figure.

      Can the authors speculate/discuss on why there is no Pax3 labelled cells but 7% Myf5 labeled cells in scBAT?

      Significance

      The discovery has significant implication as brown adipocytes are mainly found in the supraclavicular region in the human body.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      This study aimed to investigate the developmental origin of brown adipose tissue (BAT) in the supraclavicular region and its implications for metabolic health. The authors have used genetic fate mapping in mice to trace the lineage of brown adipocytes in the supraclavicular region. The findings revealed that supraclavicular brown adipocytes do not originate from the Pax3+/Myf5+ epaxial dermomyotome, which is responsible for the development of interscapular BAT. Instead, most supraclavicular brown adipocytes were marked by the Tbx1+ lineage, indicating that the pharyngeal mesoderm is involved in their development. This work provides the first evidence that scBAT adipocytes do not share the same embryonic origins as iBAT fat cells. By identifying the location-specific myogenic progenitors for supraclavicular BAT versus interscapular BAT, the researchers shed light on the distinct developmental origins of different BAT depots. Overall, these findings provide new insights into the developmental origin of supraclavicular BAT and highlight the need to consider the anatomical locations and developmental origins in studying BAT development and function.

      Major comments:

      The manuscript addresses an important yet understudied area and provides convincing results that support the key conclusions. The authors also engage in extensive discussion, speculating on the broader implications of the findings and outlining the limitations of their study.

      The observation that the loss of Pparg in Tbx1-expressing cells leads to a reduction in supraclavicular BAT is intriguing. However, as the authors acknowledge, Tbx1 is expressed in inguinal white adipocytes as well. Therefore, it remains unclear how changes in Tbx1 in ingWAT in this model might affect BAT depots. The authors should provide further clarification on this matter and include additional discussion regarding the potential indirect effects and limitations of these models.

      In Figure 5, it would be beneficial to include the expression of a broader range of thermogenic genes and proteins in both males and females. This would strengthen the argument that the loss of Pparg in Tbx1+ progenitors impairs scBAT function.

      Minor Comments:

      It would be more appropriate to present the results of the cold challenge studies as absolute body temperature rather than just the difference between groups. This is important because there could be baseline differences in basal body temperature among the experimental groups.

      It is recommended to include representative images of cre-negative animals to validate the reagents and models used in the study. Including such images would enhance the reliability of the experimental approach and strengthen the overall validity of the findings.

      The manuscript would benefit from placing the findings in the context of the broader field. For instance, discussing whether the loss of Prdm16 in Myf5 progenitors has a similar impact on interscapular BAT (iBAT) development or thermogenic function would provide valuable insights. Additionally, exploring the relationship between Tbx1 progenitors and adipocyte progenitors in adult BAT depots, such as Pdgfra+ and Trpv1+ progenitors, could further enhance our understanding of the developmental origins and functional characteristics of different adipocyte populations

      Significance

      Field of expertise: Adipose biology, Developmental biology, Thermogenesis

      The manuscript addresses an important and previously understudied area in the field of brown adipose tissue (BAT) development. The study challenges the long-standing assumption that Pax3+/Myf5+ progenitors in the dermomyotome are the sole developmental source of brown adipocytes in mice, including both interscapular BAT (iBAT) and other brown adipocytes.

      This work provides experimental evidence that the supraclavicular BAT (scBAT) adipocytes have a distinct embryonic origin compared to iBAT fat cells. While the developmental source of iBAT has been known for over a decade, this study demonstrates for the first time that scBAT adipocytes do not arise from the Pax3+/Myf5+ progenitors in the dermomyotome.

      The findings presented in this manuscript have the potential to make a lasting impact on the brown adipose tissue research community, particularly those interested in the developmental aspects of brown fat.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary

      In this study, using genetic labeling and deletion modeling, the authors discovered that Tbx1 myoprogenitors, rather than Pax3+/Myf5+ cells, give rise to supraclavicular brown adipose tissue (scBAT). This finding is both intriguing and significant. Furthermore, the genetic ablation of PPARγ or PRDM16, driven by Tbx1-Cre, reduced the size/weight of scBAT and its thermogenesis function capacity, supporting the importance of Tbx-1 cells. Interestingly, the authors found that human scBAT, located in the deep neck region, exhibits higher Tbx1 expression than subcutaneous neck WAT, potentially indicating a similar origin of scBAT in rodents and humans. Overall, this novel finding is exciting and could push the BAT field into a new phase. The manuscript is also well-written and organized.

      Comments

      1. The authors assert that Myf5+ progenitors do not contribute to scBAT adipocytes. However, Figure 2C shows that 7% of medial scBAT are mG positive, suggesting a minor contribution of Myf5+ progenitors to scBAT. The conclusion that scBAT does not originate from Myf5+ precursor cells may be overly strong.
      2. The authors claim that "prdm16 is dispensable for the development of scBAT." However, supporting data seems insufficient. The absence of a difference in scBAT weight does not guarantee that development was unaffected. Additional experiments, like H&E staining and immunofluorescence (IF) staining of RFP/GFP, could help demonstrate a similar number of GFP+ brown adipocytes in scBAT, thereby supporting this statement.
      3. In the "scBAT contributes to temperature maintenance in mice" section, this phenomenon only seems to apply to female mice (Figure 5). This conclusion may need adjustment to account for this sex difference. Moreover, for females, are there H&E staining results available for scBAT? For males, is there a change in Ucp1 expression? It could be beneficial to examine mRNA expression levels for additional thermogenic genes, such as Dio2, Prdm16, Cidea, and PGC-1a.
      4. Considering that mutant mice have smaller scBAT, how does this difference influence glucose homeostasis between WT and mutants?
      5. In Figures 5G & M, it seems more accurate for the y-axis label to read "Body temperature change."
      6. There is inconsistency in scale bar labeling:
        • Figure 3A vs 3D
        • Figure 4B vs 4C
        • Figure 5L, where scale bars are missing in the H&E staining images
        • Figure S1, where the scale bar is in the middle of the image
      7. In Figure 3, the western blot data should be quantified, and the molecular weight (kDa) should be included to clarify the band's position.
      8. The statement "RT-qPCR revealed much higher levels of TBX expression in total lysates from deep neck BAT (Figure 8A)" could be clarified by adding "than neck WAT" at the end.

      Significance

      The major advancement of this study lies in the authors' novel finding of the embryonic lineage of brown adipocytes in the supraclavicular brown adipose tissue (scBAT) depot. They demonstrated that this lineage differs from those in the dorsal BAT depots by utilizing Myf5-Cre, Pax3-Cre, and TBX1-Cre reporter mouse models.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity (Required)):

      Summary: This manuscript by Lan et al. addresses the still incompletely resolved question as to how branching morphogenesis of the embryonic mammary epithelium is regulated at the molecular and cellular level. Using (combinatorial) primary explant cultures of wildtype and genetically engineered mouse embryos, in which the authors have developed a unique expertise over many years, together with imaging and RNAseq analyses, they (i) show that the timing of epithelial branching is dictated by the biological age of the epithelium, but that an epithelial-mesenchymal interaction is required to bestow branching ability on the mammary epithelium somewhere between E13.5 and E16.5, (ii) seek to determine if and how lineage and cell proliferation affect branching, (iii) show that while salivary mesenchyme can promote growth (i.e. branching density) of the E16.5 mammary epithelium, the mode of branching (i.e. lateral branching vs tip-clefting) is an intrinsic property of the mammary epithelium, (iv) use transcriptomics to identify genes that are likely to control either mammary- or salivary gland specific growth and/or branching patterns, (v) hypothesize that low levels of WNT signaling in the mammary gland mesenchyme (due to relatively high expression of WNT signaling inhibitors) are responsible for mammary specific branching, (vi) show that hyperactivation of WNT/CTNNB1 signaling in the mesenchyme indeed induces hyperbranching, (vii) identify Eda and Igf1 as putative mediators and paracrine signaling factors that regulate branching of the mammary epithelium upon secretion from the mesenchyme downstream of WNT/CTNNB1 signaling and (viii) show that mammary gland branching is impaired in Igfr1 null embryos.

      Major comments: 1. Overall, this is a solid study that is well controlled and technically of high quality. The materials and methods should allow follow up and replication by others and the transcriptomic data have been made available via NCBI GEO. I think the authors convincingly demonstrate points (i), (iii), (iv) and (vi) and (viii). I have some questions regarding (ii), (v) and (vii) and (viii) that I will pose below.

      Our response:

      We thank the reviewer for the careful assessment and recognition of our work. In the subsequent sections, we have tried to address all the concerns raised by the reviewer.

      Re: (ii): The authors try to study the link between basal cell fate and branching. They use position of the cells (which they describe clearly and which is a good choice), since they cannot use specific markers due to the fact that the basal and luminal linages have not yet segregated at this point. This part of the manuscript is not the most straightforward to follow. The most obvious experiment would have been to focus on the location of the cells and their associated cell cycle profile - but the authors themselves have just recently published a pre-print (their REF #54, now also out in JCB) that is an in-depth study of the link between cell proliferation + cell motility and branching, but this only becomes apparent in the discussion. In that sense, Fig2 of the current manuscript is less novel, although it is nice to see that it holds up in a slightly different analysis.

      Our response:

      We thank the reviewer for acknowledging our recently published work, which is focusing on the active branching phase during late embryogenesis/around birth. In the current proliferation analysis, however, our focus was on a different aspect of embryonic mammary gland development: understanding the mechanism underlying the ability to acquire competence to branch, i.e. how the epithelium changes between late bud and sprout stages. Our data obtained from tissue recombination and 3D culture experiments suggest that heterotypic mesenchymes or mesenchyme-free 3D organoid culture conditions do not provide sufficient signals to support branching of mammary epithelia before E16.5. We have rephrased the text to better emphasize this point.

      Instead of focusing on the cell cycle markers, the authors turn to a K14-Eda mouse model - which shows precocious branching and a temporary reduction in K8 expression. They also analyze Eda-KO embryos. Quite frankly, I find the authors' reasoning difficult to follow here and I cannot deduce how these experiments really address the question at hand (i.e. how lineage and cell proliferation affect branching), so I hope they can rewrite this section of the paper to make the arguments more clear and easy to follow for the reader who, at this point, knows little about Eda. For example, the authors present the argument that K14-Eda mice show a transient reduction in K8 expression - but we don't know if that also really means a (temporary?) change in (future?) luminal cell fate. In fact, since Eda later also makes an appearance as a candidate factor to be secreted by the mesenchyme together with Igf1, I wonder if their K14-Eda data would not be better suited to underscore that point instead and if the authors should perhaps eliminate this section altogether and just refer to their prior work in REF #45. If the authors think the current data add something more, than they need to be more explicit about this (and then also introduce the link to REF #45 in the results section).

      __Our response: __

      We agree with all the reviewers in that this part of the manuscript was not mature enough and provided only indirect evidence on the potential link between lineage segregation and branching ability. This is an important question in the field that merits a study of its own and should be addressed with better tools than those available to us at present. As suggested by reviewers #1 and #3, we have omitted this part in the revised manuscript.

      Re: (v): Do the authors have any WNT/CTNNB1 target genes that they can include in their transcriptomics analysis to show that the WNT/CTNNB1 signaling levels are indeed lower in the mammary mesenchyme? Axin2 comes to mind, but there are some other negative feedback targets that are often induced across tissues, e.g. Rnf43 and/or Znrf3 and/or Sp5?E.g. to include in FIg6E?

      __Our response: __

      In the original manuscript (lines 339-342), we have performed the GSVA analysis comparing the KEGG database, and the significantly altered pathways comparing different mammary mesenchymes with salivary gland mesenchyme have been pooled and displayed as heatmap in Supplementary Fig 4b. The WNT signaling pathway is lower in the mammary mesenchyme, especially at E16.5.

      As suggested by the reviewer, we have analyzed Axin2, the most commonly used readout of WNT/CTNNB1 signaling activity in our RNA-seq data that we include as a __new Supplemental Fig. 4c __in the preliminary revised manuscript. Axin2 data indicate that Wnt/β-catenin signaling activity is lower in the E16.5 fat pad, where branching takes place, compared to younger stages of mammary gland and the salivary gland.

      Plan for the final revision:

      Additionally, we will provide expression data of a transgenic Wnt reporter from the same developmental stages and tissues that were used to generate the RNA-seq data.

      Re: (vii) and (viii): The authors convincingly show the phenotype of the Igfr1 KO mice, but I hope the authors concur that an epithelial only Igfr1 KO (or alternatively a mesenchymal only Igf1 KO, or epithelial/mesenchymal recombination experiments with WT vs IGFR1 null or IGF1 null tissue, or experiments with small molecule inhibitors of IGF1/IGFR1 signaling) would have given more solid mechanistic evidence regarding the presumed paracrine effect of IGF1 signaling. I am not asking the authors to perform another mouse experiment or even generate or use these conditional strains, but if the authors agree, then I do think this would merit some attention in the discussion section. See also my comments regarding Eda in point 1.

      Our response:

      As shown in the current manuscript, Igf1 is expressed in the mammary and salivary gland mesenchyme. This finding is in line with E14 in situ expression data available in Genepaint (https://gp3.mpg.de/results/Igf1) showing that overall in embryonic tissues, Igf1 is mainly produced in mesenchymal tissues. Of note, in Genepaint, a clear signal can be detected in the salivary gland mesenchyme, not the epithelium. Published E16 and E18 datasets indicate low level of Igf1 expression in the mammary epithelium (https://wahl-lab-salk.shinyapps.io/Mammary_snATAC/). Hence, we conclude that Igf1 is mainly produced by mesenchymal cells. Instead, Igf1r appears to be rather ubiquitously expressed.

      A previous study assessed BrdU incorporation in Igf1r-/- mammary buds at E14.5, and reported a specific proliferation defect in the epithelium, while no difference was detected in the mesenchyme (Fig. 9, Heckman et al., 2007; PMID:17662267). However, we cannot exclude the possibility of autocrine, mesenchymal Igf1/Igf1r signaling, which in turn could lead to upregulation of a paracrine factor to regulate epithelial growth.

      We agree with the reviewer in that novel conditional mouse models are beyond the scope of the current study. However, we do not think that small molecule drugs could be used to block Igf1r activity in a tissue-specific manner neither.

      Plan for the final revision:

      To further delineate the paracrine and/or autocrine role of Igf1/Igf1r pathway during mammary epithelial growth and branching, we will perform tissue recombination experiments between Igf1r-/- and control mammary epithelium and mesenchyme, as suggested by the reviewer.

      Minor comments: - A few minor spelling/grammar errors, including a couple of "the"s missing (first line of the abstract, and also preceding "Majority" in line 148.

      Our response:

      We apologize for these slips. They have been corrected in the revised manuscript.

      • Line 517-518: please also include the details for the Eda mice.

      Our response:

      We apologize for missing this important information in materials and methods. We have included a short introduction of the K14-Eda mice, a new reference for the original publication producing them, as well as the Jackson Laboratories strain number for Eda-/- (a.k.a. Tabby) mice in the revised manuscript.

      • 1f spelling error: separation

      Our response:

      The spelling error has been corrected in the revised manuscript.

      **Referees cross-commenting**

      Having read all three review reports I think they are pretty much in agreement, with shared questions about the inclusion/meaning/discussion of the lineage specification data and also agreement about the overall technical solidity of the data and this approach.

      I gather that reviewer #2 asks for more controls than myself or reviewer #3 and while I think all of their points are valid, in principle, I don't think all of these are required. I should add that I am inclined to trust the authors on their ability to separate mesenchyme and epithelium as they have been developing and optimising this system over many years.

      Our response:

      We are grateful to the reviewer for the reliance on the technical aspect of our experiments. We do routinely monitor tissue purity in the recombinants (for more details, see our response to reviewer #2). To demonstrate this, we have included new data in new Supplementary Fig. 1a,b and new Supplementary Fig. 3. We believe these additions will further enhance the validity of our findings and effectively address the concerns raised by reviewer 2.

      Reviewer #1 (Significance (Required)):

      General assessment: This is a carefully executed study in which an impressive amount of (combinatorial) embryonic mammary tissue explant experiments are combined with quantitative imaging and transcriptomics analysis.

      The main limitations of the work lie in the fact that the investigation of a potential link between branching and the cell cycle is not entirely novel, as the authors themselves recently published an nice pre-print (now also out in JCB) describing similar analyses. In addition, the mechanistic link between WNT/CTNNB1 signaling in the mesenchyme and the paracrine signaling activities of the presumed downstream effectors EDA and IGF, while plausible, is not yet complete. The work also does not yet addresses what exactly the branching identity is that is bestowed upon the mammary epithelium between E13.5 and E16.5 and how this then becomes an intrinsic (epigenetic?) feature of the mammary gland.

      Advance: This work provides more insight into the embryonic branching of the mammary gland - a stage of mammary gland development that is still poorly understood and that is, in general, understudied. In part, the work confirms prior work in the literature (their REF #19) regarding mammary and salivary gland tissue recombination experiments. It supplements this with a more elaborate time series of heterochronic and heterologous epithelium/mesenchyme explant cultures, using genetically engineered (and fluorescently labeled) mouse tissues to allow better and quantitative imaging. The transcriptomic analysis of different mesenchyme populations is also informative and allows the researchers to propose a putative mechanism for why the mammary gland branches differently from the salivary gland. The advance is both technical and functional, as well as conceptual, with some advance in terms of mechanism.

      Audience: This works should appeal to mammary gland biologists interested in the molecular and cellular mechanisms of (early) mammary gland development, as well as to a broader community of developmental biologists studying branching morphogenesis in tissues such as lung, kidney and salivary gland.

      My expertise: WNT signaling and mammary gland biology, at the intersection of developmental, stem cell and cancer biology

      __Reviewer #2 (Evidence, reproducibility and clarity (Required)): __

      The mammary gland is a branched structure that consists of a bilayered epithelium embedded in a specialized mesenchyme. In mice, at 11,5 days of embryogenesis, the ectoderm thickens forming 5 pairs of peculiar structures called placodes. During the following days, the placodes will grow and invaginate into the surrounding mammary mesenchyme and they will finally start to branch by the end of embryogenesis (E16). It has been suggested that the bidirectional communication between the growing mammary gland and the surrounding mesenchyme plays a pivotal role in the determination of each step of mammary gland development (placode formation, mammary bud invagination, gland outgrowth, branching). The role of different signalling molecules has already been shown, particularly for the placode growth and mammary bud invagination. Nevertheless, the pathways regulating embryonic mammary gland branching are still incompletely understood. In this manuscript, Lan and colleagues aim to decipher the correlation between different stages of mammary gland development such as proliferation, lineage segregation and ductal branching. Furthermore, they want to define which stage of mammary development is intrinsically determined by the epithelium and which one requires the supportive guidance of the mesenchyme. Lastly, they aim to discover the key signal for the growth and branching of mammary epithelium. To these purposes, they used an ex vivo model of heterochronic epithelial-mesenchymal recombination. In particular, they micro-dissected the epithelium and/or the mesenchyme from murine mammary glands at different stages of embryonic development (i.e. at E13,5 for the quiescent phase or 16,5 for branching phase) and explanted them together in different combinations using fluorescent reporters. To assess the role of the mesenchyme they also cultured the epithelium in a mesenchyme free 3d structure. Through this model they demonstrated that the presence of the mesenchyme is necessary for the priming of mammary epithelium for branching, since only E16,5 epithelial cells were able to grow and branch in a mesenchyme free 3D experiment. Nevertheless, intrinsic properties of the epithelium are necessary for the timing of branching, since E16,5 mesenchyme was not able to accelerate the outgrowth of E13,5 epithelia. In order to determine which epithelial properties are important, the authors correlated the beginning of cell proliferation in the embryonic mammary gland to the beginning of the branching phase. They indeed used the Fucci2a mouse model to carefully characterise the timing of mammary cells proliferation at different stages of embryonic development, concluding that the great majority of proliferating cells reside in the inner part of the mammary bud until E14,5, while in the external part at later stages. Regarding the importance of cell proliferation, Lan and colleagues claim that the beginning of the branching phase is not its direct consequence, thanks to the use of the K14Cre- Eda mouse model, known to have anticipated mammary gland development. Using this and the Eda-/- models, the authors also sustain that the branching occurs independently of the lineage specification of the epithelium. The use of salivary mesenchyme instead the mammary one was able to increase the number of branching of E16,5 mammary epithelium. Nevertheless, this model demonstrated that the branching pattern (side branching vs tip bifurcation) is an intrinsic feature of the epithelium. Lan and colleagues also defined the transcriptomic profiles of the mammary and salivary mesenchymes at different stages. In particular, they observed an increased expression of negative regulators of Wnt pathway in the mammary mesenchyme compared to the salivary mesenchyme. Moreover, using a mouse model where B-catenin is stabilised, they observed increased tip production in the mammary gland epithelium. They also showed that IGF1 production is increased after Wnt pathway activation and they tested its function, both treating their ex vivo cultures with exogenous IGF1 and using Igf1r-/- mouse models.

      Major comments 1- The great majority of the results of the manuscript are based on an ex vivo model of heterochronic epithelial-mesenchymal recombination. Since the authors are studying the effect of the mesenchyme of different stages on the epithelium (and vice versa), the purity of the two compartments after the dissection is particularly important. Although they said that the purity is evaluated (line 112), it would be important to show a control staining in which they use known markers of the mesenchyme with no colocalization with the fluorescent reporter of the epithelium.

      Our response:

      We agree with the reviewer that the purity of the separated tissues is very important for our conclusions. This is why we have used genetically labeled tissues in all recombination experiments: the epithelium and the mesenchyme were always isolated from embryos ubiquitously expressing GFP or tdTomato. We find this the most reliable way to assess the origin and purity of the isolated tissues. If there was any carry-over mesenchyme isolated with the GFP+ epithelium, this would be revealed as GFP+ mesenchymal cells in the recombinants consisting of otherwise tdTomato+ mesenchyme. And vice versa: any carry-over tdTomato+ epithelium isolated with the mesenchyme would be revealed as tdTomato+ epithelial cells in the recombinants. We apologize for not making this clear enough in the original manuscript. In the revised manuscript, we now provide confocal high-resolution images of the recombinant explants (new Supplementary Fig. 1a,b). The explants have been co-stained with the epithelial marker EpCAM, revealing a robust colocalization between the ubiquitously expressed florescent labels in the designated epithelial tissues and the EpCAM.

      2- Another important point for understanding the quality and impact of these findings is to assess the similarities and differences, if there are, between the in vivo mesenchyme and the ex vivo one. Indeed, once explanted and put in culture, mesenchymal cells could change their transcriptomic profile and consequently change their signals to the epithelium. The authors should assess the expression of the genes and pathways studied during embryonic development in vivo.

      Our response:

      The reviewer is correct in that the transcriptomes will likely undergo some changes when organs are cultured ex vivo. This is why RNA-seq was done on freshly isolated tissues. Regarding the potential changes taking place ex vivo, however, we do not consider them relevant with respect to the questions we are addressing in this study. The reason is (as reported in the manuscript) that all control recombinations (homochronic recombinations such as E13 epithelium + E13 mesenchyme, E16 epithelium + E16 mesenchyme etc.) branched essentially as in vivo. Therefore, we find the results and conclusions made from the tissue recombination experiments solid.

      3- The authors clearly showed that E16,5 epithelium is able to branch in a mesenchyme free 3D culture model, while epithelia from earlier stages don't. This led to the conclusion that mesenchyme is necessary for acquiring the branching ability. Nevertheless, the authors also said that early stages epithelia scarcely grow in the mesenchyme free 3D culture. Therefore, the lack of branching may be due to the lack of growth, if not the increase of death, of epithelial cells. The authors should quantify the size and the cell death of the epithelia in the different culture conditions and discuss better this point.

      Our response:

      The reviewer is correct in that one of the key functions of the mammary mesenchyme up to E16.5 may be to provide survival signals for the epithelium, and this might explain why epithelia younger than E16.5 fail to grow/branch when recombined with salivary gland mesenchyme and in mesenchyme-free organoid culture.

      Plan for the final revision:

      To address this issue, we will assess apoptosis in mammary epithelia cultured in the mesenchyme-free 3D culture organoid set-up.

      4- The Fucci2a model allowed to assess the proliferation of embryonic mammary epithelium, showing that the great majority of proliferating cells are basal, at late stages of development (line 182). As it has already been shown, lineage specification is a late process during mammary gland development. The fact that the proliferating cells reside at the external part of the bud does not mean that they are basal cells yet. A p63/K8 staining could be important to understand if the increased proliferation occured in already specified basal cells or not.

      __Our response: __

      Indeed, mammary lineage specification is a later process. As pointed out in the manuscript and by reviewer #1, the widely used basal and luminal lineage markers have not yet segregated to separate compartments at the developmental stages analyzed in our study, and therefore cannot be used as tools for this purpose. We would like to emphasize that in the manuscript, we analyzed the cells based on their position, and have used the term basal to indicate the basal position, not the prospective lineage. Accordingly, we used the term inner instead of luminal cells to indicate their location, not lineage. We have further clarified this point in the preliminary revised manuscript.

      5- The use of Fucci2a model showed that 20% of epithelial cells are proliferative at E13,5. This phase is considered as "quiescent" by the authors (line 120), but the moderate proliferation rate shown in this experiment demonstrated that it is not. A change of the nomenclature is needed.

      __Our response: __

      We have removed the word “quiescent” from the text.

      6- Through the use of K14-Eda and Eda-/- models, the authors claimed that the lineage specification is not a prerequisite for ductal branching. To support this point, they showed that the K14-Eda mice have an anticipated branching although the expression of K8 in the inner part of the bud is transitorily decreased. The authors link the K8 downregulation to a transient suppression of the luminal lineage, but this is clearly overclaimed. Although K8 is a known marker of luminal lineage, the downregulation of one marker is not sufficient to support their thesis. They should first check more markers and in particular critical regulators of luminal lineage as Notch1, Foxa1 and Elf5. Lately, the use of different models that drive embryonic epithelial cells to a forced lineage commitment (Notch1 or Δnp63 overexpression) would support more their claim. As additional evidence, the authors showed that Eda is able to promote basal cell signature. Firstly, the authors should better explain why this point would support their thesis. Secondly, the supplementary figure 2b does not show which genes are taken into account to define the basal signature. A list of these genes would be helpful, as well as staining for some representative proteins.

      Our response:

      We thank the reviewer for these constructive suggestions. We agree with all reviewers in that this part of the manuscript was not mature enough and provided only indirect evidence on the potential link between lineage segregation and branching ability. This is an important question in the field that merits a study of its own to be addressed with better tools than those available to us at present. As suggested by reviewers #1 and #3, we have omitted this part in the revised manuscript.

      7- The authors used the same mouse models to assess the importance of proliferation in the determination of ductal branching and they claimed that proliferation is not a sufficient feature. This conclusion was supported by two observations. The first one is the fact that the K14-Eda model shows an increased cell proliferation at early stages compared to wt, coupled with anticipated branching. Secondly, although having smaller glands compared to wt and showing a delay in ductal branching, Eda-/- mice have an epithelial proliferation rate very similar to wt. Again, the conclusion that proliferation is not sufficient for branching is overclaimed. Firstly, the authors should explain how the buds in wt and Eda-/- mice have different sizes although the similar proliferation (increased cell death?, cellular volume?). Secondly, to support the thesis that proliferation is not sufficient for branching, functional experiments should be performed (see point 12). For instance, the short-time treatments with inhibitors or promotors of proliferation may help to understand the effective role of proliferation in the determination of branching.

      Our response:

      We show that there is no direct link between onset of proliferation and acquisition of branching ability. However, we are not claiming that proliferation is not important for branching, as obviously new cells are needed as building blocks of growing tissues. In a recently published paper, we have assessed the role of proliferation in branch point formation in embryonic mammary glands. Using mitomycin C to block proliferation, we showed that initiation of new branches occurs even when proliferation is blocked (Myllymäki et al., JCB2023, PMID: 37367826).

      The reviewer was also asking why Eda-/- mammary primordia are smaller at E15.5-E16.5 despite similar proliferation rates. In the revised manuscript, we have quantified the volume of E13.5 Eda-/- and control mammary buds and show that Eda-/- buds are ~25% smaller (3.5 ± 0.8 x 105 µm3 in Eda-/- vs. 4.6 ± 0.7 x 105 µm3 in control, mean ± SD) already at the bud stage (new Supplementary Fig. 2c,d).

      We have also quantified the cellular size in Eda-/- and control mammary glands at E13.5 and E15.5 and found that mammary epithelial cells in Eda-/- embryos are ~15% smaller (new Supplementary Fig. 2e,f). Together, these data indicate that the smaller size of E15.5-E16.5 Eda-/- mammary glands is a combinatorial effect the smaller mammary anlage at E13.5 and smaller cell size. These findings, while interesting on their own, do not challenge our conclusions regarding the link between onset of proliferation and acquisition of branching ability.

      8- The heterotypic epithelial-mesenchymal recombination using the salivary gland is interesting. Nevertheless, some stainings to assess the purity of their systems are again required (e.g., marker of salivary epithelium to verify the purity of the mesenchyme and vice versa).

      __Our response: __

      As mentioned above, all tissue recombination experiments were performed so that the epithelium and the mesenchyme originated from genetically labelled embryos expressing different fluorescent proteins. In the revised manuscript, we provide confocal images of the salivary-mammary tissue recombinants (new Supplementary Fig. 3), confirming the purity of the tissue compartments used in these experiments.

      This model clearly showed that the mammary epithelium can form more branching when combined with the salivary mesenchyme. Moreover, the salivary epithelium preferentially branches through tip bifurcation, while mammary epithelium combined with the salivary mesenchyme has a mixed pattern of tip bifurcation and side branching (typical of the mammary gland). The authors thus concluded that the branching pattern is an intrinsic feature of the epithelium. However, a comparison between the percentage of tip bifurcation and side branching in the heterotypic combination and the homotypic combination between mammary epithelium and mammary mesenchyme is crucial to understand this point. Indeed, these results are not sufficient to exclude that the branching pattern is partially determined by intrinsic features and partially by extrinsic signals. The authors should carefully quantify the branching pattern in the homotypic combination and compare that to the heterotypic one. If the percentage of tip bifurcation do not change, their conclusion is correct; if this percentage increases in the heterotypic combination, it would be a sign of a partial effect of the signals of the mesenchyme.

      Our response:

      We thank the reviewer for raising this question. We have independently generated data on the type of mammary gland branching events in two papers with somewhat different culture and imaging conditions (Lindström et al., BiorXiv 2022 and Myllymäki et al., JCB, 2023, PMID: 37367826). Both analyses showed that in embryonic mammary glands, the majority of branching events (~70%) occurs by side-branching. These data are in line with the current study that we have now complemented to include also the mammary-mammary recombination experiments (revised Supplementary Video 1, revised Fig. 4b). Quantification of branching events revealed no significant difference in the type of branching events of mammary epithelia grown with salivary or mammary gland mesenchyme (revised Fig. 4c), further supporting our initial conclusions.

      9- Through the analysis of their transcriptomic data, Lan and colleagues found that the mammary mesenchyme expresses higher levels of negative regulators of Wnt pathway compared to the salivary mesenchyme. To demonstrate the value of their findings, they should confirm this in vivo, through staining of known Wnt proteins on the salivary and mammary mesenchymes at the embryonic stage.

      Our response:

      In mammals, there are 19 Wnt ligands, over a dozen secreted Wnt inhibitors, 10 Frizzled receptors, two Lrp co-receptors, and numerous other pathway modifiers that contribute to the net Wnt signaling activity in a complex manner. Furthermore, it has been “notoriously difficult to generate useful antibodies to vertebrate Wnt proteins...In general, these sera do not detect endogenous Wnt proteins in cell extracts, nor do they detect Wnt proteins in tissues by staining techniques. Hence, there are few data on Wnt protein distribution in intact vertebrate animals.” This is a direct citation from the Wnt Homepage, maintained by the Nusse Lab; https://web.stanford.edu/group/nusselab/cgi-bin/wnt/reagents#antibod.

      For all these reasons, we do not find this approach feasible nor informative.

      Instead, in the revised manuscript, we report the expression levels of Axin2, the most commonly used transcriptional readout of canonical Wnt activity in our RNA-seq data (new Supplementary Fig. 4c). Axin2 levels are lowest in the E16 fat pad where mammary branching takes place, much lower than in any other tissues analyzed in the study.

      Plan for the final revision:

      To complement these findings, we will additionally provide expression data of a transgenic Wnt reporter from the same developmental stages and tissues that were used to generate the RNA-seq data.

      10- Since the ability of the salivary mesenchyme to promote a higher rate of branching in the mammary epithelium, the authors wanted to assess what could be the role of Wnt signalling. To do so, they used a mouse model where B-catenin is stabilised, allowing an increased Wnt signalling in the mammary mesenchyme. As a result, they observed increased branching in the mammary epithelium. They also found that IGF1 is a ligand regulated by Wnt pathway in the mesenchyme. Therefore, the use of exogenous IGF1 in their ex vivo model was able to increase the branching of the mammary epithelium. Moreover, Igf1r-/- embryos showed a significant decrease of mammary gland branching. The conclusion based on these experiments was that the Wnt-Igf1-Igf1r axis plays a pivotal role in the promotion of mammary gland branching during embryogenesis. This conclusion is overclaimed for different reasons. Firstly, the normalization of the ductal branching to the body weight is insufficient to exclude that the impact of the Igf1r knockout may have severe consequences on the mammary gland formation, upstream of the ductal branching. Another parameter for this normalization is required (e.g., size of the bud before branching, proliferation status, etc).

      Our response:

      We agree with the reviewer in that Igf1r knockout may affect mammary gland formation in multiple ways, and also prior to onset of branching, as already indicated in the original manuscript: “…apart from one study reporting the smaller size of the E14 mammary bud in IGF-1R deficient embryos …” (line 398-399 in the revised version) and ‘…mammary gland 3 that was consistently absent.’ (line 414-415 in the revised version).

      To assess whether the reduced size and branching of E16.5/E18.5 Igf1r-/- mammary glands is merely a consequence of the smaller anlage, the revised manuscript includes new data reporting quantification of the volume of mammary gland 2 of Igf1r-/- and wild type littermate embryos at E13.5, E16.5, and E18.5 from 3D confocal images of whole mount EpCAM stained mammary glands. As can be seen from the new Fig. 7g-h, at E13.5, the mutant mammary buds are about 60% of the size of the controls, at E16.5, 25% and at E18.5 only 20 % revealing a progressive defect, indicative of a specific defect at the outgrowth and branching stage. This conclusion was validated by normalization to the body weight: at E13.5 the size of Igf1r-/- mammary anlage did not differ from that of the wild type embryos (p = 0.11), at E16.5 the sprouts were smaller in the mutants, though the difference did not reach statistical significance (p = 0.08), while at E18.5, the Igf1r-/- mammary glands were significantly smaller (p = 0.000021) (new Fig. 7i). We find these data compelling evidence for a specific role for Igf1r in outgrowth and branching of the embryonic mammary gland.

      The use of alternative models to specifically knockout the receptor in the epithelium or the ligand in the mesenchyme (e.g. viruses) would be even more useful to specifically focus on the role of this pathway for ductal branching excluding side effects.

      Our response:

      We thank the reviewer for this suggestion. Unfortunately, based on our experience, viral shRNA delivery is not sufficiently efficient for effective gene silencing, unlike Cre delivery for a gain-of-function approach (used in the current study to flox out exon 3 of beta-catenin) in case where the endogenous pathway activity is very low and therefore, targeting even a subset of cells is sufficient for upregulation of paracrine factors.

      Plan for the final revision:

      To address the question on the autocrine or paracrine role of Igf1r, we will perform tissue recombination experiments between Igf1r-/- and control mammary epithelium and mesenchyme.

      Another limit of this model is the fact that Igfr1 can be bound by Igf2 as well and we cannot exclude that this has an impact too (except if Igf2 is not expressed at this stage). A quantification of Igf2 expression may be useful.

      Our response:

      Indeed, we cannot exclude the possibility that Igf2 could also play a role (Igf2 expression was similar to Igf1 in our RNA-seq dataset, see Supplementary Fig. 5), but the connection of mesenchymal Wnt signaling activity was to Igf1, not Igf2 – in fact Igf2 was somewhat downregulated in Wnt3A treated sample reported by Wang et al. (Wang et al., 2021) (highlighted by an arrow in the revised Fig. 6). We have also clarified this point in the Discussion of the preliminary revised manuscript.

      11- From the experiments presented in this section it is clear that Wnt-Igf1-Igf1r axis has to be finely regulated to have the correct amount of ductal branching in the embryonic mammary epithelium. Nevertheless, the author just showed the RNA levels of Igf1 in the different compartments they have analysed. Stainings to see the effective presence of the ligand on the tissue is mandatory to clarify the role of this axis in the ductal branching in vivo.

      Our response:

      Igf1-Igf1r signaling plays a critical growth promoting function during embryonic and postnatal development. The expression of Igf1 at RNA and protein level has been detected in almost all tissues in humans (Daughaday et al., Endocr. Rev., 1989; PMID: 2666112). Given that Igf1 is a secreted protein and multiple Igf binding proteins (Igfpbs) (that regulate the bioactivity of Igf1 by sequestering it) are expressed in the mammary and salivary gland mesenchyme (Supplementary Fig. 5), we find it unlikely that Igf1 staining would provide any additional information to the current study, as they cannot be used to assess the source of Igf1, nor the location of the signaling activity.

      Furthermore, as underlined by the authors, this axis is specifically important and upregulated in the salivary gland. Due the limit of the Igf1R-/- model, we cannot exclude that, although Wnt-Igf1-Igf1r axis is able to increase the branching ability of mammary epithelium, the normal branching rate observed in wt mice is due to other pathways.

      Our response:

      We agree with the reviewer in that other pathways are also important in regulating normal mammary gland branching, for example, Eda/NF-κB and FGF pathways as we described in the Introduction. Our results do not exclude the possibility that also pathways other than Wnt regulate Igf1 expression. The reviewer is correct that if a paracrine factor is expressed in the salivary gland but not in the mammary mesenchyme, its physiological effect may be limited to the salivary gland. Indeed, cluster 5 identified by the mFuzzy analysis (Fig. 5f) is likely to include some genes like that. This is why we decided to focus on cluster 6 genes like Igf1. In the revised manuscript, we have better highlighted the difference between cluster 5 and 6 genes.

      Unfortunately, with the currently available tools, we cannot test the importance of the endogenous mesenchymal Wnt signaling activity by inactivating Wnt signaling activity specifically in the mesenchyme at the time point when branching begins. This would require an inducible mesenchymal Cre line (mesenchymal β-catenin is essential for the early fate specification of the primary mammary mesenchyme; Hiremath et al., 2012, PMID: 23034629), and conditional β-catenin null mouse. We do not have such mice available and we find that these experiments are beyond the scope of the current study.

      12- Lastly, once claimed to have found the key factor necessary for ductal branching promotion, the authors should also test if the proliferation and lineage segregations are unaffected in this context, confirming their dispensable role claimed in the initial part of the manuscript.

      __Our response: __

      Igf1/Igf1r is well-known for its growth promoting function via cell proliferation. We have no reasons to think that this would not be the case also in the mammary gland, and it was not our intention to give the impression that proliferation was not affected. In fact, Hiremath et al. (2012) already reported a defect in epithelial cell proliferation in Igf1rmammary buds at E14. Our key finding is that compared to other organs, the mammary gland is particularly sensitive to loss of Igf1r during branching morphogenesis. Finally, as pointed out earlier, better tools will be needed to assess the potential link between lineage segregation and onset of branching, a topic that we hope to address in the future.

      Minor comments: 1- An important paper on mammary gland ductal branching was published on Nature in 2017 by Scheele and colleagues and should be presented in the introduction, even though it is at later stages (after birth).

      Our response:

      We thank the reviewer for the suggestion. In the revised manuscript, we have added the findings from Scheele et al. 2017 in the introduction.

      2- In line 136 and 139 the authors referred to Fig 2 but it should be Fig 1

      Our response:

      We apologize for these slips. They have been corrected in the revised manuscript.

      3- The sentence on line 142 should be rephrased, since "advanced developmental stages" may be referred to pubertal development. The authors should specify that they are talking about embryonic development.

      Our response: We apologize for the potential misunderstanding. In the revised manuscript, we have used the phrase “advanced embryonic developmental stage” to describe our conclusion more precisely.

      Reviewer #2 (Significance (Required)):

      Overall, the authors concluded that embryonic mammary gland development and branching are extremely sensitive to the loss of IGF1, normally produced by the mesenchyme. The topic of the paper is interesting, the experimental approaches are well conceived, the data are convincing and the findings are of interest to developmental biologists. Nevertheless, there are some significant points that need to be further investigated before considering the manuscript suitable for publication:

      Our response:

      We thank the reviewer for the careful assessment and positive feedback of our manuscript. We have already addressed most of the points raised and most remaining ones will be addressed in the final revised manuscript.

      __Reviewer #3 (Evidence, reproducibility and clarity (Required)): __

      Here the authors use classical embryonic tissue recombination and pharmacological manipulation of explants in conjunction with cutting edge 3D imaging of tissue derived from highly sophisticated reporter and knock-out mouse models and state of the art transcriptomic analysis to masterfully delineate and dissect regulatory pathways critical for embryonic mammary development. Specifically, they set out to parse regulation of proliferation from that of branch patterning.

      While it has long been established that epithelial-mesenchymal interaction is necessary for mammary branching this work shows by heterochronic recombination that initiation mammary branching is not advanced by mesenchymal stage. By examining Fucci2a embryos the authors demonstrate that branching is preceded by a significant increase in basal cell-biased proliferation but, through further analysis of Eda gain and loss of function mice, conclude that proliferation per se does not cause branching. They show by heterotypic recombination with salivary tissue that early mammary epithelia rudiments require their own mesenchyme for survival and that although later E16.5 rudiments expand more robustly when in contact with salivary mesenchyme they nevertheless retain their characteristic mammary branch pattern. Thus, they establish that initiation and patterning are intrinsic properties of the epithelium but that early survival and later expansion/proliferation is regulated by the mesenchymal context. By transcriptomic comparison of mammary and salivary mesenchyme they reveal that genes encoding canonical Wnt attenuators and antagonists are highly expressed in early mammary mesenchyme and drop as branching ensues. The low expression of these negative regulators of Wnt signaling in salivary mesenchyme is proposed as an explanation for its growth and branch stimulating capability. In keeping with these observations, the authors show that experimental activation of mammary mesenchymal Wnt signaling augments both growth and branching. Lastly, they identify transcriptomic changes in IGF1 coincident with the initiation of mammary branching and confirm its role by extending analyses of the effects of gain and loss of function of IGF1 on embryonic mammary development.

      This is a thorough, well-constructed paper that adds new knowledge and important conceptual nuance and mechanistic insight to classical findings on branch patterning. This work is a technical tour de force and backed by solid quantitative and statistical analysis throughout. Their experimental approach is superb and the conclusions are sound. Their findings will be of great interest to the community of mammary gland biologists and to the wider field of embryologists focused on early development of a broad range of ectodermal appendages.

      I have some minor criticisms that I believe can be quickly remedied in a minor rewrite and suggestions for the authors consideration to improve the manuscript discussion as follows:

      Minor issues Abstract, line 37: The authors misuse the word "decompose" - it should be "deconstruct"

      __Our response: __ We thank the reviewer for pointing out our mistake, which we have corrected in the revised manuscript.

      Results, p7 line 48: Add "The" to the sentence: "The majority...."

      __Our response: __ Corrected it in the revised manuscript.

      P8 line 173 This sentence refers to Figure 2G which is a quantitative plot. I would suggest replacing the word "cluster" which implies a spatial organization with the word "subset" or "significant fraction" The spatial data in Fig 2d support basal bias but do NOT to my eye show any clustering - in fact the proliferative basal cells appear to be evenly dispersed within the basal layer.

      Our response:

      We thank the reviewer for highlighting this aspect. We agree that “significant fraction” is a more suitable term than “cluster”.

      P9 line 188: The statement on basal cell lineage specification needs a reference.

      __Our response: __

      Following the suggestions from reviewers #1 and 3, we have removed the content about lineage segregation in Results, together with this sentence.

      P10 line 201-216 I found the section on lineage specification (fig S2) weaker than the rest and a distraction from the main thrust of the paper making it difficult for the reader to focus. I suggest omitting this section and supplemental figures associated with it altogether.

      __Our response: __

      We agree with all reviewers in that this part of the manuscript was not mature enough and provided only indirect evidence on the potential link between lineage segregation and branching ability. This is an important question in the field that merits a study of its own that should be addressed with better tools than those available to us at present. As suggested by reviewers #1 and #3, we have omitted this part in the revised manuscript.

      P9 line 190: "displays precocious onset of branching" it is sufficient to say: displays precocious branching - the use of both "precocious" and "onset' is redundant.

      P10 line 229 Similarly, delete "the onset of branching was delayed" it is sufficient to say: branching was delayed.

      __Our response: __ Both sentences have been corrected it in the revised manuscript.

      P11 line 243: Delete "on the regulation of the" and substitute the word "to" in the sentence: "Next, we shifted our focus on the regulation of the branching pattern, which is thought to be determined by mesenchymal cues."

      __Our response: __ Corrected it in the revised manuscript.

      P11 line 241 subtitle and Figure 4 title: The disparity in titles here is jarring for the reader: Results text subtitle: "Salivary gland mesenchyme is rich in growth-promoting cues, but does not alter the mode of branch point formation of the mammary epithelium". Figure 4 Title: "Mammary mesenchyme is indispensable for the branching ability of the mammary gland". I suggest to the authors divide the figure as well as the text to make the two points indicated by their disparate titles separately.

      __Our response: __ We thank reviewer for the suggestion to clarify the Results part of the manuscript. As suggested, we have split the data under two separate subtitles, but due to limitations in figure numbers, we prefer to report these data in one figure panel.

      P12 line 279 From here on out the manuscript has a tendency to use the term "growth" ambiguously - in many instances it is unclear do they mean expansion, proliferation, increased branch number/ morphology?? Please try to clarify.

      __Our response: __

      Our aim is to use the term growth to mean tissue growth (expansion). We hope that this is clearer in the revised manuscript.

      P16 line 341 use word "prompted" instead of word "promoted"

      __Our response: __ We thank reviewer for spotting out the slip, which we have corrected in the revised manuscript.

      P16 line 382: include word "embryonic" before "mammary development"

      __Our response: __ We have modified the text in the revised manuscript.

      Discussion P18 line 416: Add the words "later stage (E16.5)" to the sentence: "Importantly, we demonstrate that salivary gland mesenchyme could only promote the growth of later stage (E16.5) mammary epithelium"

      __Our response: __ We thank reviewer for the suggestion. We have modified the text in the revised manuscript.

      P19 line 437: Given the authors statement "Instead, cell motility is critical for branch point formation in the mammary gland" they should consider a brief sentence mentioning their transcriptomic findings on cadherin 11 and Tenascin.

      __Our response: __ We thank the reviewer for appreciation of our transcriptomic data. In the revised manuscript, we have added the following text in discussion: “Accordingly, we observed significantly increased expression of cell migration promoting genes such as Cdh11 (encoding Cadherin 11), and Tnc (encoding Tenascin C) 60,61 in the E16.5 mesenchyme compared to E13.5 (Supplementary Table 2).”

      P19 line 451: Similarly, given their statement "This observation suggests that mammary epithelium itself carries the instructions dictating the mode of branching" they could consider their transcriptomic data on Ltbp1 in "mammary specific" clusters 7,8,9 as a matrix molecule initially expressed by mammary mesenchyme but which becomes expressed by luminal epithelial cells at precisely the time they acquire lineage specification and intrinsic branching capability.

      __Our response: __ This is an excellent suggestion. We have added following text in discussion: “It is worth noting that certain mesenchymal factors, such as Ltbp1, began transitioning towards epithelium-specific expression around E16.5 69. Exploring the potential impact of these factors on the self-instructed branching capacity of the mammary epithelium could yield valuable insights.”

      P20 lines 462-470 The authors should address their theory of Wnt suppression in the mammary mesenchyme in the context, albeit conflictingly, of earlier studies showing expression of Wnt signaling reporters, in either epithelial or mesenchymal locations during early stages.

      Our response: __ We thank reviewer for the suggestion. In the preliminary revised manuscript, we report Axin2 expression data as __new Supplementary Fig. 4c. Axin2 expression data suggest that Wnt/β-catenin activity is lowest in the E16.5 fat pad (where branching takes place) compared to all other tissues analyzed in the study.

      Plan for the final revision:

      For the final revised manuscript, we will additionally generate transgenic Wnt reporter expression data (see also our response to point 3 of Reviewer #1). These results will be discussed in light of the published Wnt reported literature in the final revised manuscript.

      Reviewer #3 (Significance (Required)):

      Here the authors use classical embryonic tissue recombination and pharmacological manipulation of explants in conjunction with cutting edge 3D imaging of tissue derived from highly sophisticated reporter and knock-out mouse models and state of the art transcriptomic analysis to masterfully delineate and dissect regulatory pathways critical for embryonic mammary development. Specifically, they set out to parse regulation of proliferation from that of branch patterning.

      This is a thorough, well-constructed paper that adds new knowledge and important conceptual nuance and mechanistic insight to classical findings on branch patterning. This work is a technical tour de force and backed by solid quantitative and statistical analysis throughout. Their experimental approach is superb and the conclusions are sound. Their findings will be of great interest to the community of mammary gland biologists and to the wider field of embryologists focused on early development of a broad range of ectodermal appendages.

      Our response:

      We much appreciate the positive evaluation of our manuscript. We have addressed all the feedback provided by the reviewer 3 in the preliminary revised manuscript, except the last point, which will be included in the final revision along with the new data on the Wnt reporter expression.

      Field of expertise: Embryonic and adult mammary development, Wnt signaling, cell adhesion

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Here the authors use classical embryonic tissue recombination and pharmacological manipulation of explants in conjunction with cutting edge 3D imaging of tissue derived from highly sophisticated reporter and knock-out mouse models and state of the art transcriptomic analysis to masterfully delineate and dissect regulatory pathways critical for embryonic mammary development. Specifically, they set out to parse regulation of proliferation from that of branch patterning.

      While it has long been established that epithelial-mesenchymal interaction is necessary for mammary branching this work shows by heterochronic recombination that initiation mammary branching is not advanced by mesenchymal stage. By examining Fucci2a embryos the authors demonstrate that branching is preceded by a significant increase in basal cell-biased proliferation but, through further analysis of Eda gain and loss of function mice, conclude that proliferation per se does not cause branching. They show by heterotypic recombination with salivary tissue that early mammary epithelia rudiments require their own mesenchyme for survival and that although later E16.5 rudiments expand more robustly when in contact with salivary mesenchyme they nevertheless retain their characteristic mammary branch pattern. Thus, they establish that initiation and patterning are intrinsic properties of the epithelium but that early survival and later expansion/proliferation is regulated by the mesenchymal context. By transcriptomic comparison of mammary and salivary mesenchyme they reveal that genes encoding canonical Wnt attenuators and antagonists are highly expressed in early mammary mesenchyme and drop as branching ensues. The low expression of these negative regulators of Wnt signaling in salivary mesenchyme is proposed as an explanation for its growth and branch stimulating capability. In keeping with these observations, the authors show that experimental activation of mammary mesenchymal Wnt signaling augments both growth and branching. Lastly, they identify transcriptomic changes in IGF1 coincident with the initiation of mammary branching and confirm its role by extending analyses of the effects of gain and loss of function of IGF1 on embryonic mammary development.

      This is a thorough, well-constructed paper that adds new knowledge and important conceptual nuance and mechanistic insight to classical findings on branch patterning. This work is a technical tour de force and backed by solid quantitative and statistical analysis throughout. Their experimental approach is superb and the conclusions are sound. Their findings will be of great interest to the community of mammary gland biologists and to the wider field of embryologists focused on early development of a broad range of ectodermal appendages.

      I have some minor criticisms that I believe can be quickly remedied in a minor rewrite and suggestions for the authors consideration to improve the manuscript discussion as follows:

      Minor issues

      Abstract, line 37: The authors misuse the word "decompose" - it should be "deconstruct"

      Results, p7 line 48: Add "The" to the sentence: "The majority...."

      P8 line 173 This sentence refers to Figure 2G which is a quantitative plot. I would suggest replacing the word "cluster" which implies a spatial organization with the word "subset" or "significant fraction" The spatial data in Fig 2d support basal bias but do NOT to my eye show any clustering - in fact the proliferative basal cells appear to be evenly dispersed within the basal layer.

      P9 line 188: The statement on basal cell lineage specification needs a reference.

      P10 line 201-216 I found the section on lineage specification (fig S2) weaker than the rest and a distraction from the main thrust of the paper making it difficult for the reader to focus. I suggest omitting this section and supplemental figures associated with it altogether.

      P9 line 190: "displays precocious onset of branching" it is sufficient to say: displays precocious branching - the use of both "precocious" and "onset' is redundant.

      P10 line 229 Similarly, delete "the onset of branching was delayed" it is sufficient to say: branching was delayed.

      P11 line 243: Delete "on the regulation of the" and substitute the word "to" in the sentence: "Next, we shifted our focus on the regulation of the branching pattern, which is thought to be determined by mesenchymal cues."

      P11 line 241 subtitle and Figure 4 title: The disparity in titles here is jarring for the reader: Results text subtitle: "Salivary gland mesenchyme is rich in growth-promoting cues, but does not alter the mode of branch point formation of the mammary epithelium". Figure 4 Title: "Mammary mesenchyme is indispensable for the branching ability of the mammary gland". I suggest to the authors divide the figure as well as the text to make the two points indicated by their disparate titles separately.

      P12 line 279 From here on out the manuscript has a tendency to use the term "growth" ambiguously - in many instances it is unclear do they mean expansion, proliferation, increased branch number/ morphology?? Please try to clarify.

      P16 line 341 use word "prompted" instead of word "promoted"

      P16 line 382: include word "embryonic" before "mammary development"

      Discussion P18 line 416: Add the words "later stage (E16.5)" to the sentence: "Importantly, we demonstrate that salivary gland mesenchyme could only promote the growth of later stage (E16.5) mammary epithelium"

      P19 line 437: Given the authors statement "Instead, cell motility is critical for branch point formation in the mammary gland" they should consider a brief sentence mentioning their transcriptomic findings on cadherin 11 and Tenascin.

      P19 line 451: Similarly, given their statement "This observation suggests that mammary epithelium itself carries the instructions dictating the mode of branching" they could consider their transcriptomic data on Ltbp1 in "mammary specific" clusters 7,8,9 as a matrix molecule initially expressed by mammary mesenchyme but which becomes expressed by luminal epithelial cells at precisely the time they acquire lineage specification and intrinsic branching capability.

      P20 lines 462-470 The authors should address their theory of Wnt suppression in the mammary mesenchyme in the context, albeit conflictingly, of earlier studies showing expression of Wnt signaling reporters, in either epithelial or mesenchymal locations during early stages.

      Significance

      Here the authors use classical embryonic tissue recombination and pharmacological manipulation of explants in conjunction with cutting edge 3D imaging of tissue derived from highly sophisticated reporter and knock-out mouse models and state of the art transcriptomic analysis to masterfully delineate and dissect regulatory pathways critical for embryonic mammary development. Specifically, they set out to parse regulation of proliferation from that of branch patterning.

      This is a thorough, well-constructed paper that adds new knowledge and important conceptual nuance and mechanistic insight to classical findings on branch patterning. This work is a technical tour de force and backed by solid quantitative and statistical analysis throughout. Their experimental approach is superb and the conclusions are sound. Their findings will be of great interest to the community of mammary gland biologists and to the wider field of embryologists focused on early development of a broad range of ectodermal appendages.

      Field of expertise: Embryonic and adult mammary development, Wnt signaling, cell adhesion

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      The mammary gland is a branched structure that consists of a bilayered epithelium embedded in a specialized mesenchyme. In mice, at 11,5 days of embryogenesis, the ectoderm thickens forming 5 pairs of peculiar structures called placodes. During the following days, the placodes will grow and invaginate into the surrounding mammary mesenchyme and they will finally start to branch by the end of embryogenesis (E16). It has been suggested that the bidirectional communication between the growing mammary gland and the surrounding mesenchyme plays a pivotal role in the determination of each step of mammary gland development (placode formation, mammary bud invagination, gland outgrowth, branching). The role of different signalling molecules has already been shown, particularly for the placode growth and mammary bud invagination. Nevertheless, the pathways regulating embryonic mammary gland branching are still incompletely understood. In this manuscript, Lan and colleagues aim to decipher the correlation between different stages of mammary gland development such as proliferation, lineage segregation and ductal branching. Furthermore, they want to define which stage of mammary development is intrinsically determined by the epithelium and which one requires the supportive guidance of the mesenchyme. Lastly, they aim to discover the key signal for the growth and branching of mammary epithelium.

      To these purposes, they used an ex vivo model of heterochronic epithelial-mesenchymal recombination. In particular, they micro-dissected the epithelium and/or the mesenchyme from murine mammary glands at different stages of embryonic development (i.e. at E13,5 for the quiescent phase or 16,5 for branching phase) and explanted them together in different combinations using fluorescent reporters. To assess the role of the mesenchyme they also cultured the epithelium in a mesenchyme free 3d structure. Through this model they demonstrated that the presence of the mesenchyme is necessary for the priming of mammary epithelium for branching, since only E16,5 epithelial cells were able to grow and branch in a mesenchyme free 3D experiment. Nevertheless, intrinsic properties of the epithelium are necessary for the timing of branching, since E16,5 mesenchyme was not able to accelerate the outgrowth of E13,5 epithelia. In order to determine which epithelial properties are important, the authors correlated the beginning of cell proliferation in the embryonic mammary gland to the beginning of the branching phase. They indeed used the Fucci2a mouse model to carefully characterise the timing of mammary cells proliferation at different stages of embryonic development, concluding that the great majority of proliferating cells reside in the inner part of the mammary bud until E14,5, while in the external part at later stages.

      Regarding the importance of cell proliferation, Lan and colleagues claim that the beginning of the branching phase is not its direct consequence, thanks to the use of the K14Cre- Eda mouse model, known to have anticipated mammary gland development. Using this and the Eda-/- models, the authors also sustain that the branching occurs independently of the lineage specification of the epithelium. The use of salivary mesenchyme instead the mammary one was able to increase the number of branching of E16,5 mammary epithelium. Nevertheless, this model demonstrated that the branching pattern (side branching vs tip bifurcation) is an intrinsic feature of the epithelium. Lan and colleagues also defined the transcriptomic profiles of the mammary and salivary mesenchymes at different stages. In particular, they observed an increased expression of negative regulators of Wnt pathway in the mammary mesenchyme compared to the salivary mesenchyme. Moreover, using a mouse model where B-catenin is stabilised, they observed increased tip production in the mammary gland epithelium. They also showed that IGF1 production is increased after Wnt pathway activation and they tested its function, both treating their ex vivo cultures with exogenous IGF1 and using Igf1r-/- mouse models.

      Major comments

      1. The great majority of the results of the manuscript are based on an ex vivo model of heterochronic epithelial-mesenchymal recombination. Since the authors are studying the effect of the mesenchyme of different stages on the epithelium (and vice versa), the purity of the two compartments after the dissection is particularly important. Although they said that the purity is evaluated (line 112), it would be important to show a control staining in which they use known markers of the mesenchyme with no colocalization with the fluorescent reporter of the epithelium.
      2. Another important point for understanding the quality and impact of these findings is to assess the similarities and differences, if there are, between the in vivo mesenchyme and the ex vivo one. Indeed, once explanted and put in culture, mesenchymal cells could change their transcriptomic profile and consequently change their signals to the epithelium. The authors should assess the expression of the genes and pathways studied during embryonic development in vivo .
      3. The authors clearly showed that E16,5 epithelium is able to branch in a mesenchyme free 3D culture model, while epithelia from earlier stages don't. This led to the conclusion that mesenchyme is necessary for acquiring the branching ability. Nevertheless, the authors also said that early stages epithelia scarcely grow in the mesenchyme free 3D culture. Therefore, the lack of branching may be due to the lack of growth, if not the increase of death, of epithelial cells. The authors should quantify the size and the cell death of the epithelia in the different culture conditions and discuss better this point.
      4. The Fucci2a model allowed to assess the proliferation of embryonic mammary epithelium, showing that the great majority of proliferating cells are basal, at late stages of development (line 182). As it has already been shown, lineage specification is a late process during mammary gland development. The fact that the proliferating cells reside at the external part of the bud does not mean that they are basal cells yet. A p63/K8 staining could be important to understand if the increased proliferation occured in already specified basal cells or not.
      5. The use of Fucci2a model showed that 20% of epithelial cells are proliferative at E13,5. This phase is considered as "quiescent" by the authors (line 120), but the moderate proliferation rate shown in this experiment demonstrated that it is not. A change of the nomenclature is needed.
      6. Through the use of K14-Eda and Eda-/- models, the authors claimed that the lineage specification is not a prerequisite for ductal branching. To support this point, they showed that the K14-Eda mice have an anticipated branching although the expression of K8 in the inner part of the bud is transitorily decreased. The authors link the K8 downregulation to a transient suppression of the luminal lineage, but this is clearly overclaimed. Although K8 is a known marker of luminal lineage, the downregulation of one marker is not sufficient to support their thesis. They should first check more markers and in particular critical regulators of luminal lineage as Notch1, Foxa1 and Elf5. Lately, the use of different models that drive embryonic epithelial cells to a forced lineage commitment (Notch1 or Δnp63 overexpression) would support more their claim. As additional evidence, the authors showed that Eda is able to promote basal cell signature. Firstly, the authors should better explain why this point would support their thesis. Secondly, the supplementary figure 2b does not show which genes are taken into account to define the basal signature. A list of these genes would be helpful, as well as staining for some representative proteins.
      7. The authors used the same mouse models to assess the importance of proliferation in the determination of ductal branching and they claimed that proliferation is not a sufficient feature. This conclusion was supported by two observations. The first one is the fact that the K14-Eda model shows an increased cell proliferation at early stages compared to wt, coupled with anticipated branching. Secondly, although having smaller glands compared to wt and showing a delay in ductal branching, Eda-/- mice have an epithelial proliferation rate very similar to wt. Again, the conclusion that proliferation is not sufficient for branching is overclaimed. Firstly, the authors should explain how the buds in wt and Eda-/- mice have different sizes although the similar proliferation (increased cell death?, cellular volume?). Secondly, to support the thesis that proliferation is not sufficient for branching, functional experiments should be performed (see point 12). For instance, the short-time treatments with inhibitors or promotors of proliferation may help to understand the effective role of proliferation in the determination of branching.
      8. The heterotypic epithelial-mesenchymal recombination using the salivary gland is interesting. Nevertheless, some stainings to assess the purity of their systems are again required (e.g., marker of salivary epithelium to verify the purity of the mesenchyme and vice versa). This model clearly showed that the mammary epithelium can form more branching when combined with the salivary mesenchyme. Moreover, the salivary epithelium preferentially branches through tip bifurcation, while mammary epithelium combined with the salivary mesenchyme has a mixed pattern of tip bifurcation and side branching (typical of the mammary gland). The authors thus concluded that the branching pattern is an intrinsic feature of the epithelium. However, a comparison between the percentage of tip bifurcation and side branching in the heterotypic combination and the homotypic combination between mammary epithelium and mammary mesenchyme is crucial to understand this point. Indeed, these results are not sufficient to exclude that the branching pattern is partially determined by intrinsic features and partially by extrinsic signals. The authors should carefully quantify the branching pattern in the homotypic combination and compare that to the heterotypic one. If the percentage of tip bifurcation do not change, their conclusion is correct; if this percentage increases in the heterotypic combination, it would be a sign of a partial effect of the signals of the mesenchyme.
      9. Through the analysis of their transcriptomic data, Lan and colleagues found that the mammary mesenchyme expresses higher levels of negative regulators of Wnt pathway compared to the salivary mesenchyme. To demonstrate the value of their findings, they should confirm this in vivo, through staining of known Wnt proteins on the salivary and mammary mesenchymes at the embryonic stage.
      10. Since the ability of the salivary mesenchyme to promote a higher rate of branching in the mammary epithelium, the authors wanted to assess what could be the role of Wnt signalling. To do so, they used a mouse model where B-catenin is stabilised, allowing an increased Wnt signalling in the mammary mesenchyme. As a result, they observed increased branching in the mammary epithelium. They also found that IGF1 is a ligand regulated by Wnt pathway in the mesenchyme. Therefore, the use of exogenous IGF1 in their ex vivo model was able to increase the branching of the mammary epithelium. Moreover, Igf1r-/- embryos showed a significant decrease of mammary gland branching. The conclusion based on these experiments was that the Wnt-Igf1-Igf1r axis plays a pivotal role in the promotion of mammary gland branching during embryogenesis. This conclusion is overclaimed for different reasons. Firstly, the normalization of the ductal branching to the body weight is insufficient to exclude that the impact of the Igf1r knockout may have severe consequences on the mammary gland formation, upstream of the ductal branching. Another parameter for this normalization is required (e.g., size of the bud before branching, proliferation status, etc). The use of alternative models to specifically knockout the receptor in the epithelium or the ligand in the mesenchyme (e.g. viruses) would be even more useful to specifically focus on the role of this pathway for ductal branching excluding side effects. Another limit of this model is the fact that Igfr1 can be bound by Igf2 as well and we cannot exclude that this has an impact too (except if Igf2 is not expressed at this stage). A quantification of Igf2 expression may be useful.
      11. From the experiments presented in this section it is clear that Wnt-Igf1-Igf1r axis has to be finely regulated to have the correct amount of ductal branching in the embryonic mammary epithelium. Nevertheless, the author just showed the RNA levels of Igf1 in the different compartments they have analysed. Stainings to see the effective presence of the ligand on the tissue is mandatory to clarify the role of this axis in the ductal branching in vivo. Furthermore, as underlined by the authors, this axis is specifically important and upregulated in the salivary gland. Due the limit of the Igf1R-/- model, we cannot exclude that, although Wnt-Igf1-Igf1r axis is able to increase the branching ability of mammary epithelium, the normal branching rate observed in wt mice is due to other pathways.
      12. Lastly, once claimed to have found the key factor necessary for ductal branching promotion, the authors should also test if the proliferation and lineage segregations are unaffected in this context, confirming their dispensable role claimed in the initial part of the manuscript.

      Minor comments:

      1. An important paper on mammary gland ductal branching was published on Nature in 2017 by Scheele and colleagues and should be presented in the introduction, even though it is at later stages (after birth).
      2. In line 136 and 139 the authors referred to Fig 2 but it should be Fig 1
      3. The sentence on line 142 should be rephrased, since "advanced developmental stages" may be referred to pubertal development. The authors should specify that they are talking about embryonic development.

      Significance

      Overall, the authors concluded that embryonic mammary gland development and branching are extremely sensitive to the loss of IGF1, normally produced by the mesenchyme. The topic of the paper is interesting, the experimental approaches are well conceived, the data are convincing and the findings are of interest to developmental biologists. Nevertheless, there are some significant points that need to be further investigated before considering the manuscript suitable for publication:

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary:

      This manuscript by Lan et al. addresses the still incompletely resolved question as to how branching morphogenesis of the embryonic mammary epithelium is regulated at the molecular and cellular level. Using (combinatorial) primary explant cultures of wildtype and genetically engineered mouse embryos, in which the authors have developed a unique expertise over many years, together with imaging and RNAseq analyses, they (i) show that the timing of epithelial branching is dictated by the biological age of the epithelium, but that an epithelial-mesenchymal interaction is required to bestow branching ability on the mammary epithelium somewhere between E13.5 and E16.5, (ii) seek to determine if and how lineage and cell proliferation affect branching, (iii) show that while salivary mesenchyme can promote growth (i.e. branching density) of the E16.5 mammary epithelium, the mode of branching (i.e. lateral branching vs tip-clefting) is an intrinsic property of the mammary epithelium, (iv) use transcriptomics to identify genes that are likely to control either mammary- or salivary gland specific growth and/or branching patterns, (v) hypothesize that low levels of WNT signaling in the mammary gland mesenchyme (due to relatively high expression of WNT signaling inhibitors) are responsible for mammary specific branching, (vi) show that hyperactivation of WNT/CTNNB1 signaling in the mesenchyme indeed induces hyperbranching, (vii) identify Eda and Igf1 as putative mediators and paracrine signaling factors that regulate branching of the mammary epithelium upon secretion from the mesenchyme downstream of WNT/CTNNB1 signaling and (viii) show that mammary gland branching is impaired in Igfr1 null embryos.

      Major comments:

      1. Overall, this is a solid study that is well controlled and technically of high quality. The materials and methods should allow follow up and replication by others and the transcriptomic data have been made available via NCBI GEO. I think the authors convincingly demonstrate points (i), (iii), (iv) and (vi) and (viii). I have some questions regarding (ii), (v) and (vii) and (viii) that I will pose below.
      2. Re: (ii): The authors try to study the link between basal cell fate and branching. They use position of the cells (which they describe clearly and which is a good choice), since they cannot use specific markers due to the fact that the basal and luminal linages have not yet segregated at this point. This part of the manuscript is not the most straightforward to follow. The most obvious experiment would have been to focus on the location of the cells and their associated cell cycle profile - but the authors themselves have just recently published a pre-print (their REF #54, now also out in JCB) that is an in-depth study of the link between cell proliferation + cell motility and branching, but this only becomes apparent in the discussion. In that sense, Fig2 of the current manuscript is less novel, although it is nice to see that it holds up in a slightly different analysis. Instead of focusing on the cell cycle markers, the authors turn to a K14-Eda mouse model - which shows precocious branching and a temporary reduction in K8 expression. They also analyze Eda-KO embryos. Quite frankly, I find the authors' reasoning difficult to follow here and I cannot deduce how these experiments really address the question at hand (i.e. how lineage and cell proliferation affect branching), so I hope they can rewrite this section of the paper to make the arguments more clear and easy to follow for the reader who, at this point, knows little about Eda. For example, the authors present the argument that K14-Eda mice show a transient reduction in K8 expression - but we don't know if that also really means a (temporary?) change in (future?) luminal cell fate. In fact, since Eda later also makes an appearance as a candidate factor to be secreted by the mesenchyme together with Igf1, I wonder if their K14-Eda data would not be better suited to underscore that point instead and if the authors should perhaps eliminate this section altogether and just refer to their prior work in REF #45. If the authors think the current data add something more, than they need to be more explicit about this (and then also introduce the link to REF #45 in the results section).
      3. Re: (v): Do the authors have any WNT/CTNNB1 target genes that they can include in their transcriptomics analysis to show that the WNT/CTNNB1 signaling levels are indeed lower in the mammary mesenchyme? Axin2 comes to mind, but there are some other negative feedback targets that are often induced across tissues, e.g. Rnf43 and/or Znrf3 and/or Sp5?E.g. to include in FIg6E?
      4. Re: (vii) and (viii): The authors convincingly show the phenotype of the Igfr1 KO mice, but I hope the authors concur that an epithelial only Igfr1 KO (or alternatively a mesenchymal only Igf1 KO, or epithelial/mesenchymal recombination experiments with WT vs IGFR1 null or IGF1 null tissue, or experiments with small molecule inhibitors of IGF1/IGFR1 signaling) would have given more solid mechanistic evidence regarding the presumed paracrine effect of IGF1 signaling. I am not asking the authors to perform another mouse experiment or even generate or use these conditional strains, but if the authors agree, then I do think this would merit some attention in the discussion section. See also my comments regarding Eda in point 1.

      Minor comments:

      • A few minor spelling/grammar errors, including a couple of "the"s missing (first line of the abstract, and also preceding "Majority" in line 148.
      • Line 517-518: please also include the details for the Eda mice.
      • 1f spelling error: separation

      Referees cross-commenting

      Having read all three review reports I think they are pretty much in agreement, with shared questions about the inclusion/meaning/discussion of the lineage specification data and also agreement about the overall technical solidity of the data and this approach.

      I gather that reviewer #2 asks for more controls than myself or reviewer #3 and while I think all of their points are valid, in principle, I don't think all of these are required. I should add that I am inclined to trust the authors on their ability to separate mesenchyme and epithelium as they have been developing and optimising this system over many years.

      Significance

      General assessment:

      This is a carefully executed study in which an impressive amount of (combinatorial) embryonic mammary tissue explant experiments are combined with quantitative imaging and transcriptomics analysis.

      The main limitations of the work lie in the fact that the investigation of a potential link between branching and the cell cycle is not entirely novel, as the authors themselves recently published an nice pre-print (now also out in JCB) describing similar analyses. In addition, the mechanistic link between WNT/CTNNB1 signaling in the mesenchyme and the paracrine signaling activities of the presumed downstream effectors EDA and IGF, while plausible, is not yet complete. The work also does not yet addresses what exactly the branching identity is that is bestowed upon the mammary epithelium between E13.5 and E16.5 and how this then becomes an intrinsic (epigenetic?) feature of the mammary gland.

      Advance:

      This work provides more insight into the embryonic branching of the mammary gland - a stage of mammary gland development that is still poorly understood and that is, in general, understudied. In part, the work confirms prior work in the literature (their REF #19) regarding mammary and salivary gland tissue recombination experiments. It supplements this with a more elaborate time series of heterochronic and heterologous epithelium/mesenchyme explant cultures, using genetically engineered (and fluorescently labeled) mouse tissues to allow better and quantitative imaging. The transcriptomic analysis of different mesenchyme populations is also informative and allows the researchers to propose a putative mechanism for why the mammary gland branches differently from the salivary gland. The advance is both technical and functional, as well as conceptual, with some advance in terms of mechanism.

      Audience: This works should appeal to mammary gland biologists interested in the molecular and cellular mechanisms of (early) mammary gland development, as well as to a broader community of developmental biologists studying branching morphogenesis in tissues such as lung, kidney and salivary gland.

      My expertise:

      WNT signaling and mammary gland biology, at the intersection of developmental, stem cell and cancer biology

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      The authors do not wish to provide a response at this time.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Summary:

      In this study, Hwang et al. develop an inducible Cas9 hiPSC line and perform with it a pooled CRISPR knockout screen using a custom sgRNA library to identify novel genes involved in human primordial germ cell like cell (hPGCLC) differentiation in vitro. Thereby they find the AKT coactivator TCL1A to be important in the proliferation/survival of hPGCLCs after specification.

      Specific Comments:

      1.) p.7-p.8: "Using the MAGeCK algorithm (Li et al, 2014) to call hits on merged replicates, 25 genes scored as significantly depleted from the AG+ population at p < 0.05. Among the top hits was SOX17, and near-hits included TFAP2C, both of which are well-known drivers of the hPGCLC state (Fig. S2E)."

      When looking at Table S3, it appears that only 2 genes were significantly depleted (P < 0.05) in replicate 1, 23 in replicate 2 and 10 genes when rep 1 and rep 2 were analyzed together. The essential germ cell genes SOX17 and TFAP2C were not significantly depleted in replicate 1 and only TFAP2C but not SOX17 was depleted significantly in replicate 2. Also the main hits discussed in this paper, METTL7 and TCL1A were not significantly depleted in replicate 1 and only METTL7 but not TCL1A was significantly depleted in replicate 2. This indicates that replicate 1 might not have been robust enough to reliably detect depleted genes and that TCL1A was not among the significant hits. A potential explanation could be that not enough cells have been used to ensure a sufficient representation of sgRNAs to provide significant results in a depletion screen. Ideally the screen would need to be repeated to provide another informative replicate, or the authors should at least correct the sentences above and openly state that their hits are only based on one replicate of the screen and that the list of their hits might therefore not be fully reliable. Also the statement on page 8 that "the screen was both technically and biologically successful" might need to be toned down.

      2.) p.8 and Fig. 2E: The authors do not clearly describe, what are the 25 top hits in PGCLC(+) and PGCLC(-) cells and how they were chosen (Score, p-value or fold change), which they compared in the gene set enrichment analysis to the RNA-Seq data.

      3.) Fig. S3I-K: The authors mention in the text on p. 9 a significant reduction in hPGCLC induction efficiency for both TCL1A and METTL7A KO cells, but they do not provide statistics and do not mention, how many biological replicates have been used. As hiPSCs generally show a high clone to clone variability in hPGCLC induction efficiency, results from a single KO clone can not be considered as a reliable result. The authors should provide results from additional wt and KO clones (they are showing in S3E multiple for each gene) to ensure reliable effects, especially for METTL7A KO cells, where the reduction in PGCLC induction efficiency is more modest and might not be significant (Fig. S3I). Another way of validating the phenotype would be to use individual control, TCL1A and METTL7A sgRNAS from the screen and compare induction efficiencies with or without DOX-induced Cas9 expression.

      4.) Fig. 3F, Supp Tables S4, S5, S6: It is not clearly described, what was the criteria to define DEGs for the GO term analysis of TCL1A KO cells, as more genes have been used than the relatively few significant DEGs reported in Table S4. Furthermore, only FDRs or otherwise adjusted p-values (not raw p-values as done in the figure and tables) should be used to determine significantly enriched GO terms. Also no representative gene names are displayed in Fig.3F, as stated in the figure legend.

      5.) Fig.4: p-AKT and p-mTOR signaling are represented as the mechanism, by which TCLA1 KO affects hPGCLC maintenance/proliferation. It is not clear from the presented data, what is meant with biological triplicates (different germ cell inductions, different subclones) mentioned in the figure legend. As some of the effects observed are quite small (e.g. p-mTOR differences in 4E, cell cycle differences in 4J), biological replicates with a different KO clone should be performed (see point 3). Otherwise it is hard to judge how robust the data and conclusions derived are.

      CROSS-CONSULTATION COMMENTS

      Apart from my points regarding the weakness in statistical confidence I agree with the other reviewers that it needs to be shown whether the effect of TCL1A KO is based on a general proliferation defect of the entire aggregation body or if the effect is really hPGCLC-specific.

      Significance

      The study is the first CRISPR screen performed during hPGCLC differentiation and provides a proof of principle for a useful tool to allow dissection of gene regulatory networks during human germ cell development. Overall this is a technical advancement and an ambitious study and will generate interest in the human germ cell field. Generally it is easy to follow but could be improved in the description of some of the methodology (points 2+4). Overall the study suffers from weaknesses in the statistical robustness, as the screen itself did not provide many significant hits (major point 1) and the follow up was only performed using a single KO clone (points 3+5). Therefore adding replicates would be necessary to strengthen the confidence in the drawn conclusions.

      Reviewers' relevant expertise: in vitro germ cell differentiation, pluripotency, CRISPR screening

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      The authors conducted a custom CRISPR screening including 422 coding-genes in hPGC-like cells (hPGCLCs) to identify genes important to hPGCLC. Based on the screen, they found two candidates, TCL1A and METTL7A, that regulate hPGCLC specification. They concluded that TCL1A, an AKT coactivator, is critical for hPGCLC specification through regulating AKT-mTOR signaling. Unfortunately, we found that the evidences for the key conclusions are not quite convincing. Reasons are as below:

      1. The results to demonstrate the key role of TCL1A on hPGCLC specification is not convincing. Fig. 3B, C & D. the cell number per aggregate is also significantly reduced in TCL1A KO (2843/20.7% =13734) compared to that in WT cells (545/16.4%=3323). Despite that, the percentage of AG+ cells per aggregate is significantly, but not dramatically decreased in TCL1A KO (20.7%) vs WT (16.4%) cells. Thus, the effect of TCL1A KO may not be specific on the AG+ cells, but on the whole aggregate.
      2. The overall effect of METTL7A KO on hPGCLC development is too moderate to conclude it as a key regulator for hPGCLC specification.

      Significance

      a CRISPR screening for key regulator for human germline cell apecification was not reported before.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      In this manuscript, the authors performed CRISPR/Cas9-based screening to identify genes involved in human germ cell development. Using human PGCLC system, the authors found that sgRNAs targeting TCL1A and METTL7A genes were enriched in non-PGCLC population. Gene-disruption of each gene resulted in a significant reduction in PGCLC differentiation. Moreover, TCL1A-knockout PGCLCs failed to proliferation during PGCLC induction, possibly due to attenuation of AKT signaling. Further analyses showed that protein synthesis and cell-cycle progression, especially in S-phase, were impaired in TCL1A-knockout PGCLCs. Thus, the authors provided remarks on successful identification of genes functionally important for human PGCLC differentiation and the importance of translational control in human germ line.

      This manuscript demonstrates a genetic screening for genes involved in hPGCLC differentiation. Although the number of genes targeted by sgRNAs were limited (422 genes), it is still valuable to show such genetic screening can be applied to the hPGCLC system. Overall, statements and data in the manuscript are convincing, except for following points, and the novelty is sufficient for publication. It would be further improved, considering following points with additional experiment, if feasible.

      1. A major concern in this manuscript is whether TCL1A function is specifically involved in hPGC development or generally important for other cell type. As AKT signaling plays multiple roles on many cell contexts, this is important to verify the author's conclusion. For example, is there any defect in proliferation of TCL1A-knockout iPS lines? The authors should quantify the doubling rate of TCL1A-knockout iPS cells, the number of iMeLCs yielded, and the cell number included in the aggregates. Looking at Figure S3K and K, the total number of the cells in the aggregates seems lower in TCL1A-knockout aggregates than in WT.

      2. Related to the comment above, the author should add a statement describing expression pattern and level of TCL1A and METTL7A in tissue. Are they preferentially expressed in the germ line, or generally expressed in a broad range of tissue?

      3. The quantification of pAKT and p-mTOR is vague. The authors should quantify in a different way. Although the author claimed that Western blot analysis was not able to detect pAKT and p-mTOR in PGCLCs, there are a number of reports that detect these proteins. As an advantage of PGCLC system is to handle a large number of cells in culture, the authors should perform rigorously the quantification.

      4. In the abstract, there is a statement "demonstrate the importance of translational control in human reproduction". Isn't this too general? Is it a new finding? With critical examination for cell-specificity of the translational control by TCL1A as described above, this should be refined.

      Significance

      As authors performed CRISPR/Cas9-based screening to identify genes involved in human germ cell development using human PGCLC system, and then successfully isolated TCL1A and METTL7A genes that are previously known as drives for PGCLC induction. Therefore, from both technological and scientific viewpoints, there is a significant advance shown in this manuscript.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Response to Reviewer comments

      Reviewer #1 (Evidence, reproducibility and clarity (Required)):

      __Summary __

      The manuscript by Parker et al addresses the important question of how different organisms have evolved pre-messenger RNA systems that are either more or less complex. This question underlies the evolution of complex organisms and the genome adaptation of simple organisms to their specific environments, so is an important question to answer. This manuscript now provides the underlying molecular mechanisms of how 5' splice site sequence preference may have evolved which is both an interesting and exciting advance for the field.

      We thank the reviewer for these kind comments.

      __Major comments __

      __This manuscript builds on the previous work from this group where they identified the role of adenosine N6 methylation (m6A) of the U6 small nuclear RNA (snRNA) of the spliceosome by METTL16 as being important for 5' splice site selection. This work led to the speculation that loss of a METTL16 ortholog, or potentially other splicing factors, in some species could contribute to an evolutionary change in 5' splice site sequence preference. Here the authors now use the power of phylogenetics, interspecies association mapping and the available spliceosome structures to provide convincing conclusions that 5' splice site sequence preferences in the extensive number of organisms examined correlate with the presence of the U6 snRNA methyltransferase METTL16 and the splicing factor SNRNP27K. __

      __An analysis of METTL16 conservation was first carried out by comparing the METTL16 methyltransferase domain (MTD) in 29 diverse eukaryotic species. All the METTL16 orthologs were found to have either one or two globular domains. Three domain types were identified and compared in detail. What was not clear from this analysis was the functional significance of orthologs having either one or two domains. __

      We identified several species, including Drosophila melanogaster, whose METTL16 orthologs do not contain a VCR domain. However, in this study we do not draw specific conclusions about the functional significance of orthologs having different domain topologies.

      __In addition, while this analysis provides important new information on the domain structure of METTL16 orthologs, especially where these domains had not been identified previously, the link between this section of the results and the following sections is not that apparent. __

      We agree that there is a significant difference in approach between the first section of the Results and the following sections. However, we are keen to keep this part of the manuscript because it provides an orthogonal line of evidence suggesting that the ancestral role of METTL16 in eukaryotes is specifically the methylation of U6 snRNA.

      __Next novel bioinformatics pipelines were developed to compare both introns and orthologous groupings of protein coding genes between 227 Sacchromycotina genomes as well as 13 well-annotated eukaryote genomes. First, the 5' splice site sequence preference was compared and clearly indicates that the +4 position has the greatest variation in preferences within the Sacchromycotina. The ability to now compare a large number of genomes has provided novel information on the evolution of the 5' splice site sequence and the conclusion that there is more complexity to the 5' splice site in fungi that previously recognized. While it is apparent why only the 5' splice site signal was investigated here, with its relationship to the U6 snRNA and METTL16, it seems a shame the other splice site sequences were not analyzed using this novel pipeline. In any case, the complexity of the 5' splice site +4 position now allows, for the first time, interesting interspecies association studies. __

      We have now included the variance plots for 3’SS motifs (analogous to the 5’SS variance plots shown in Figure 2B) as Figure 2 supplementary figure 4A, and a traitgram for 3’SS -3C to U ratio as Figure 2 supplementary figure 4B. We have included a short section of text in the Results section to describe these additional findings.

      __With ____the 5' splice site +4 variation identified, the next step was to determine the underlying molecular mechanisms that dictate the evolution of the various sequence preferences. Some obvious players here are the U1 and U6 snRNAs which directly interact with the 5' splice site during splicing. However, no association was found between these snRNAs and the 5' splice site +4 sequence. __

      __The powerful interspecies association mapping was then used to determine whether the presence or absence of METTL16 ortholog or a splicing factor correlated with the 5' splice site +4 sequence variation. Interestingly, a clear association was found between METTL16 and the 5' splice site +4 position; METTL16 presence was associated with +4A at the 5' splice site and METTL16 absence was associated with +4U at the 5' splice site. This is an exciting and significant finding. __

      We thank the reviewer for these comments on the importance of this study.

      __Interestingly, the next most significant association with the 5' splice site +4 position was with SNRNP27K. This result makes sense as in the cryo-EM structure of the pre-B spliceosome complex the C-terminal domain of SNRNP27K is found near the region of the U6 snRNA that will interact with the 5' splices site. Absence of SNRNP27K was associated with an increased preference for +4U at the 5' splice site. Now the real power of the interspecies association mapping was demonstrated by investigating whether any association could be determined specifically within the C-terminus of SNRNP27K. Significantly, the methionine 141 position in SNRNP27K was found to be associated with the +4 position of the 5' splice site. This finding fits nicely with previous studies where mutation of M141 caused a shift in 5' splice site selection away from +4A 5' splice sites, to 5' splice sites without +4A. What is not clear is whether M141 is conserved or invariant between all the species that were compared? __

      M141 is not completely conserved across the species that were compared for the SNRNP27K C-terminus analysis. We did not test positions with very strong sequence conservation, because without variation in both the genotype and phenotype it is not possible to test for an association. We have rephrased the relevant Results and Methods sections to make this point clearer. In addition, we have incorporated a sequence logo to illustrate the degree of conservation of each position in the SNRNP27K C-terminal domain as Figure 5 -figure supplement 1A. Finally, we have included an additional box-plot to illustrate the finding that species which have lost SNRNP27K or have only lost the Methionine equivalent to human SNRNP27K position 141, show a similar preference for +4U at 5’ SSs. This is now included as Figure 5 - figure supplement 1B.

      Overall, this result reveals the power of the interspecies association approach and provides interesting and exciting information on the molecular determinants of 5' splice site evolution.

      We are grateful to the reviewer for these comments.

      __The final analysis was to investigate the interaction potentials of the U5 and U6 snRNAs with the 5' splice site in the Sacchromycotina genomes and try to relate this to species with fewer introns and less alternative splicing. Species with low intron numbers and low splicing complexity were revealed to have weaker U5 and U6 anti-correlation potentials and favor +4U at the 5' splice site. On the other hand, species with high intron number and presumably higher splicing complexity featured anti-correlated U5 and U6 snRNA interaction potentials and favored +4A 5' splice sites. This extensive analysis provides novel information on the interactions and splice site properties of species with simple and complex splicing. Again, I see why there is emphasis on the 5' splice site here but a similar analysis with the U2 snRNA and the branch site could also be informative. __

      We absolutely agree that inter-species association mapping could be applied to other splicing signal phenotypes including 3’ splice sites and intron branchpoints. Accordingly, we raise this subject in the final section of the Discussion. However, branchpoint sequences are challenging to predict with genomic data. Because preliminary analyses suggest independent variation in these other splicing signal phenotypes, we feel a separate focused study is required to properly explain (and substantiate) even the analytical approaches involved. We hope the reviewer would agree that incorporating U2 snRNA and branchpoint variation analyses into this manuscript as well, could detract from the clarity of the conceptual advances that we make here.

      __Minor comments __

      __Should the Title include SNRNP27K? __

      We have included SNRNP27K into the revised title.

      Should the title specify that it is the evolution of only the 5’ splice site sequence preference being studied here?

      Because apostrophes in titles can compromise some scholarly online search engines (https://insights.uksg.org/articles/10.1629/uksg.534), we would prefer not to include 5’ in the title.

      Include information on intron number and 5’ splice site interaction potential of U5 and U6 snRNA in the Summary?

      We thank the reviewer for this suggestion. We have updated the Summary to include our findings on U5 and U6 interaction potential in species with reduced intron number.

      __Figure 1C is not referred to in the text? __

      We apologise for this oversight. We have added references to figure 1C in the appropriate Results section.

      Page 8, line 5 – better to say “splicing signal phenotypes”.

      We have amended this statement on Page 8 and at other places in the text where related phrasing was made.

      __What are the other points on Figure 3B? What is the next point below SNRNP27K? Is it U2A’? __

      The other points on Figure 3B represent Orthofinder orthogroups which contain human orthologs that are known components of the spliceosome. The list of spliceosomal components was taken from Sales-Lee et al. 2021. The third most significant point is indeed the orthogroup containing the human ortholog of U2A’. As we state in the text, however, the correlation of U2A’ with the 5’SS+4 A to U ratio phenotype is no longer significant once METTL16 presence/absence is controlled for, indicating that the correlation of U2A’ with the +4A phenotype is likely explained by similarity in the patterns of gene loss of U2A’ and METTL16.

      __The second paragraph of the Discussion is vague and lacks a reference. “we could also identify an association with a methionine residue in the conserved C-terminal domain of SNRNP27K orthologs.” There are a few methionines in the C-terminus, which one? Please reference the statement “transcriptome analysis of C. elegans SNRP-27 M141T mutants..” __

      We apologise for the lower quality of writing in this section of the Discussion. We have updated the text, made the statements about the SNRNP27K C-terminus less ambiguous, and added the relevant citations as appropriate.

      Reviewer #1 (Significance (Required)):

      Overall, this is a well written and clearly presented study that provides some key molecular information on the splicing factors involved in the evolution of 5’ splice sites and shows the power of interspecies association studies. Some important conceptual principles have now been defined for the field going forward.

      With thank the reviewer for this kind comment on the importance of this work.

      __The question remains as to whether METTL16 and SNRNP27K are the sole determinants of 5’ splice site preference evolution at +4? __

      We cannot say for certain that METTL16 and/or SNRNP27K determine the 5’SS +4 phenotype – only that they are correlated with it. In our response to reviewer 3, and in a new Discussion section, we have detailed some of the scenarios that could explain these correlations. We also cannot rule out whether there are changes in the presence/absence (or domain/sequence-level changes) of other, untested proteins that correlate with the 5’SS +4 phenotype and we allude to this in the final section of the Discussion.

      One splicing factor that immediately comes to mind is Prp8 where there is extensive evidence for involvement in splice site selection and is clearly in the right location throughout splicing to be involved. This question should at least be discussed but Prp8 would also be a very interesting candidate for the interspecies association mapping.

      Prp8 is a core component of spliceosomes and is conserved throughout the Saccharomycotina. For this reason, we were unable to associate splicing phenotypes with Prp8 presence or absence variation at the level of orthogroups. However, we revisited this question posed by the reviewer. Our experience with inter-species association mapping, so far, indicates it works well with orthogroup presence/absence or when straightforward amino acid substitutions can be detected in conserved and hence alignable protein sequence domains. We analysed the conserved U6 snRNA-interacting region of the Prp8 linker domain, which maps close to the 5’ splice site in cryo-EM models, using the profile HMM PF10596 available from Pfam. We found that the majority of this domain was extremely highly conserved with variation in only a few species and positions. The strongest correlation with the +4A to U ratio phenotype was at position 58, which is conserved as a Glycine in all but 8 species (6 Dipodascaceae, 2 CUG-Ser1), that also tend to have a stronger preference for +4A. However, examination of the species contributing to this result (and to similar results at other positions) indicated that in the 6 Dipodascaceae species, this change is part of a larger deletion or replacement that makes the whole linker region align poorly to the model. Hence, the G58 position itself may not be specifically important for the +4 phenotype. Although the wholesale loss or replacement of the U6 snRNA-interacting region in these species is potentially interesting, these larger scale structural changes in a small number of species are difficult to interpret. Therefore, to maintain the focus of the manuscript and the clear links to METTL16 and SNRNP27K that have orthogonal support, we have decided not to add these results to the manuscript but present them here (Figure not available on biorXiv commenting window).

      Also, as mentioned previously, only the 5’ splice site was investigated here and the manuscript could become a more substantial piece of work if the other splice sites were included in some way.

      We agree that it will be exciting to apply this approach to other splicing signal phenotypes and in other phylogenetic clades with emerging tree-of-life-scale genomics data. We have included variation in 3’ splice sites in the revised manuscript. As the first of its kind, this study should pioneer a wider use of this approach, by us and others, to understand the mechanisms and functions of molecular interactions not only in splicing but in other areas of biology too.

      __The obvious audience here are those directly in the splicing field but the overall principles are relevant for evolutionary biologists and those studying organismal complexity. __

      We thank the reviewer for recognising the broad importance of this work.

      My expertise is in yeast and human splicing mechanisms. I do not have the expertise to critically evaluate the bioinformatic pipelines but they were clearly explained and presented.

      Reviewer #2 (Evidence, reproducibility and clarity (Required)):

      In their manuscript, Parker et al. investigate the evolutionary patterns of splice site preference, focusing on the A/U ratio at position A+4 on the 5´ splice site. Building upon prior studies in S. pombe and A. thaliana, the authors establish a strong correlation between this preference and the co-evolution of the METTL16 U6 snRNA methyltransferase. Furthermore, through inter-species association mapping, they identify the involvement of the splicing factor SNRNP27K in altered A/U ratios and highlight the significance of the residue Met-141 in SNRNP27K for this function. Overall, the paper effectively presents impactful new findings on the evolution of METTL16, U6 snRNA, and splicing.

      We thank the reviewer for these kind comments on the importance of our study.

      The computational analyses employed in this study are situated outside our field of expertise, preventing us from offering a comprehensive evaluation of the methodology’s appropriateness and rigor. Nonetheless, the identification of METTL16 through the authors’ methods, which aligns with previous research in S. pombe and A. thaliana, lends support to the validity of their approach. Notably, the close proximity between SNRNP27K and the methylated A43 residue in U6 snRNA within the spliceosome, particularly near Met-141, is an impressive finding. Previous studies have shown that a mutation at position M141T affects splicing at +4A introns, thus providing robust validation for their methods.

      We thank the reviewer for these kind comments on our work.

      The data presented in this study furnish crucial insights into the role of METTL16, U6 snRNA methylation, and splice site recognition. The authors expand upon recent observations that the “vertebrate conserved region” exists in non-vertebrates, despite the absence of primary sequence homology. These results will serve as a valuable guide for future molecular investigations into U6 snRNA methylation and its mechanisms in splicing. Furthermore, the implications of this paper extend to human evolution, as the plasticity in splicing is an essential factor in the evolution of developmental complexity.

      We thank the reviewer for these kind comments.

      Minor suggestions for improvement:

      1. __ Given the significance of the interaction between U6 snRNA and the intron for understanding the data, it would be beneficial to include a figure illustrating the RNA-RNA base-pairing interactions between U6 snRNA and the 5´ splice site. This addition is particularly important if the paper is intended for publication in a journal with a general readership.__  We thank the reviewer for this excellent suggestion. We have included this as Figure 3A.

      __ Similarly, the section on U1 snRNA would be more comprehensible with the inclusion of U1 RNA-RNA intron diagrams and improved descriptions of both the figures and the assay. Despite being negative data in the supplement, clarifying this section is essential. As currently written, it is challenging to follow.__ 

      We agree that this section is difficult to follow. We have updated the text to improve the readability and included a figure of U1 snRNA:5’SS basepairing as Figure 3 – figure supplement 1A.

      __ Whenever possible, consider increasing the figure and font sizes to enhance readability for readers.__ 

      We agree that some of the more complex figures can be difficult to read when embedded into a Word document/pdf. We hope that providing high-resolution figures for reading online will mitigate this.

      __ In the text, there is no reference to Figure 1C.__ 

      We apologise for this oversight. We have resolved this issue with the appropriate references in the Results text.

      __ In Figure 5B, the y-axis in the top panel is labelled “species,” but the legend only mentions U5/6p as the y-axis. Please revise the legend to include the appropriate information.__ 

      We apologise for the confusion caused by our poorly written legend for this plot. We have updated the legend so that the text clearly refers to either the scatter plot or the marginal histograms.

      Reviewer #2 (Significance (Required)):

      The data presented in this study furnish crucial insights into the role of METTL16, U6 snRNA methylation, and splice site recognition. The authors expand upon recent observations that the “vertebrate conserved region” exists in non-vertebrates, despite the absence of primary sequence homology. These results will serve as a valuable guide for future molecular investigations into U6 snRNA methylation and its mechanisms in splicing. Furthermore, the implications of this paper extend to human evolution, as the plasticity in splicing is an essential factor in the evolution of developmental complexity.

      We are grateful to the reviewer for these kind comments on the importance of this work.

      Reviewer #3 (Evidence, reproducibility and clarity (Required)):

      In this manuscript, Parker et al present a nice exploration of the evolutionary and mechanistic relationships between 5′ splice site consensus sequences, intron numbers and METTL16/SNRNP27K. By performing inter-species association mapping in Saccharomycotina species, they found that a T in position +4 is strongly associated with the absence of METTL16 (and/or in some cases SNRNP27K or mutations in it). They also provide solid structural modelling data in support of this association.

      In general, I think this is a very nice manuscript. I only have a few comments, which could be addressed by rewording specific parts and/or improving the current figures.

      We are grateful to the reviewer for the kind comments on this work.

      1) As the authors acknowledge, a key issue that cannot be fully resolved in this study is causality between the different events investigated. Overall, the authors are careful about this, but there are some exceptions that should be corrected. Probably the most important is in the abstract, where they write: “We conclude that variation in concerted processes of 5’ splice site selection by U6 snRNA is crucial to evolutionary change in splicing complexity”. I suggest they write something more open (and correct), such as: “We conclude that variation in concerted processes of 5’ splice site selection by U6 snRNA is associated with evolutionary changes in splicing complexity”. Similarly, other plausible scenarios should be discussed in the corresponding Discussion section.

      We agree with the reviewer that it is not possible to infer the causal relationship between METTL16 absence and 5’SS+4 preference change from the current data. We, therefore, apologise for failing to be more careful in the Summary and Introduction. We have reworded these statements to better reflect what we can currently say about the evolutionary relationship between METTL16 and 5’SS sequence preference.

      The correlation between METTL16 absence and 5'SS+4 sequence preference change could most likely be explained by one of several scenarios: (a) sudden loss of METTL16 causes a rapid necessity to change 5'SS sequence preferences. This is unlikely as such rapid change without widespread corresponding 5'SS changes would likely impose a high fitness cost. (b) Changes in 5'SS sequence preference occur first, driven by some other selective pressure, until there is no longer a benefit to retaining the METTL16 gene. (c) Gradual changes in the expression or catalytic efficiency of METTL16 reduce the stoichiometry of U6 snRNA m6A modification, which permits gradual change in 5'SS+4 sequence preference until complete loss of the METTL16 no longer imposes a major fitness cost. As we suggest in the Discussion, future work could examine this question by determining whether the METTL16 orthologs found in Zygosaccharomyces and Eremothecium species, which have altered their 5'SS+4 preference to a U, are expressed and functional. We have updated the Discussion to include a new section that addresses these scenarios.

      2) I do not agree with the statement that "The extent of alternative splicing is the best genomic predictor of developmental complexity". To start with, there are many ways to quantify "extent of alternative splicing" and there are also different types of alternative splicing that might have different prevalence and biological impact. Then, this claim is usually related with exon skipping, which is tightly linked with intron length, and that is likely a better prediction of complexity (yet clearly not causative). My concern is: to what extent has this claim been formally and properly assessed by comparing splicing prevalence with other genomic features, such as intergenic region length, intron length, or average distance between enhancer-promoter interactions (arguably the most relevant predictor, in light of many other studies)? Moreover, I found it a bit misleading to frame the work presented in this study as directly related with developmental (or even splicing) complexity. The work is very interesting on its own, and I doubt their findings on +4 position preference in Saccharomycotina has anything to do with developmental complexity (as the Abstract and Introduction seem to imply).

      On reflection, we agree with the reviewer. Some of our framing of the text isn’t balanced with other studies on the scaling of alternative splicing with developmental complexity. We have edited the Summary and Introduction sections accordingly and cited other references that broaden the consideration of this subject. We are grateful to the reviewer for this suggestion because the changes we make improve the focus of the manuscript since our findings relate more to splicing simplification than to an understanding of increased developmental complexity.

      __3) I found Figure 2 and its associated supplementary figure very difficult to follow. I suggest the authors try to improve it and make it clearer. Also, other trees summarizing the results might be helpful. __

      We apologise for the complexity of these figures. We opted to show phylogenetic trees with phenotypes plotted on the y axis, rather than simply trait histograms or box-plots, because the underlying structure of the tree is important for demonstrating that multiple independent changes in the 5’SS phenotype have occurred in the Saccharomycotina. We have tried to improve the comprehensibility of the figures in the following ways: (a) We have added 5’SS sequence motifs to the x-axis of figure 2B to make what the plot represents clearer, (b) as suggested by the reviewer, we have created a pruned tree showing the 5’SS motifs of a selection of Saccharomycotina species, which demonstrates that the changes in 5’SS+4 position preferences seen in S. cerevisiae and C. albicans are likely to be a result of convergent evolution. We have added this tree as Figure 2 - figure supplement 3.

      __4) I also found the Results section corresponding to Figure 5B a bit confusing. I would argue (as I think the authors do) that there are two main patterns here: below 500 introns, there is no association, while above 500 introns there is an increasingly negative association (correlation). I think it would help to more explicitly distinguishing these two patterns. Then, for the intron-poor species: is the correlation (or lack of) for species with a T or an A in position +4 different? __

      We do indeed think that there are two patterns here, as indicated by the reviewer. In the previous version of the manuscript, we separated species into those having an overall preference for A at the +4 position, and those having +4U. By showing regression lines for these two classes, rather than for the general relationship between intron number and U5/6rho, we somewhat imply that the switch in +4 base preference might be causing the loss of correlation between U5/6rho and intron number. However, since essentially all species with a 5'SS +4U preference are intron poor, it seems more likely that these trends are the result of a loss of the negative correlation between intron number and U5/6rho in intron poor species, as suggested by the reviewer. To address this issue, we have replaced the regression lines on Figure 6B with a single loess (locally estimated scatterplot smoothing) regression line for all species and updated the text to make it clearer that we think loss of U5/6rho and +4A preference are separate traits of intron poor species. Although this is not exactly what the reviewer requested, we hope that it satisfies their issue with the analysis.

      __Reviewer #3 (Significance (Required)): __

      __This is a very interesting study that sheds light on an intriguing evolutionary pattern: the change in consensus sequence at position +4 of the 5' splice site. This topic is relevant since it is closely associated with intron loss and splicing efficiency and evolution. __

      We thank the reviewer for the kind and constructive comments on this study.

    2. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1 (Evidence, reproducibility and clarity):

      Summary

      The manuscript by Parker et al addresses the important question of how different organisms have evolved pre-messenger RNA systems that are either more or less complex. This question underlies the evolution of complex organisms and the genome adaptation of simple organisms to their specific environments, so is an important question to answer. This manuscript now provides the underlying molecular mechanisms of how 5' splice site sequence preference may have evolved which is both an interesting and exciting advance for the field.

      We thank the reviewer for these kind comments.

      Major comments

      This manuscript builds on the previous work from this group where they identified the role of adenosine N6 methylation (m6A) of the U6 small nuclear RNA (snRNA) of the spliceosome by METTL16 as being important for 5' splice site selection. This work led to the speculation that loss of a METTL16 ortholog, or potentially other splicing factors, in some species could contribute to an evolutionary change in 5' splice site sequence preference. Here the authors now use the power of phylogenetics, interspecies association mapping and the available spliceosome structures to provide convincing conclusions that 5' splice site sequence preferences in the extensive number of organisms examined correlate with the presence of the U6 snRNA methyltransferase METTL16 and the splicing factor SNRNP27K. 

      An analysis of METTL16 conservation was first carried out by comparing the METTL16 methyltransferase domain (MTD) in 29 diverse eukaryotic species. All the METTL16 orthologs were found to have either one or two globular domains. Three domain types were identified and compared in detail. What was not clear from this analysis was the functional significance of orthologs having either one or two domains.

      We identified several species, including Drosophila melanogaster, whose METTL16 orthologs do not contain a VCR domain. However, in this study we do not draw specific conclusions about the functional significance of orthologs having different domain topologies.

      In addition, while this analysis provides important new information on the domain structure of METTL16 orthologs, especially where these domains had not been identified previously, the link between this section of the results and the following sections is not that apparent.

      We agree that there is a significant difference in approach between the first section of the Results and the following sections. However, we are keen to keep this part of the manuscript because it provides an orthogonal line of evidence suggesting that the ancestral role of METTL16 in eukaryotes is specifically the methylation of U6 snRNA.

      Next novel bioinformatics pipelines were developed to compare both introns and orthologous groupings of protein coding genes between 227 Sacchromycotina genomes as well as 13 well-annotated eukaryote genomes. First, the 5' splice site sequence preference was compared and clearly indicates that the +4 position has the greatest variation in preferences within the Sacchromycotina. The ability to now compare a large number of genomes has provided novel information on the evolution of the 5' splice site sequence and the conclusion that there is more complexity to the 5' splice site in fungi that previously recognized. While it is apparent why only the 5' splice site signal was investigated here, with its relationship to the U6 snRNA and METTL16, it seems a shame the other splice site sequences were not analyzed using this novel pipeline. In any case, the complexity of the 5' splice site +4 position now allows, for the first time, interesting interspecies association studies.

      We have now included the variance plots for 3’SS motifs (analogous to the 5’SS variance plots shown in Figure 2B) as Figure 2 supplementary figure 4A, and a traitgram for 3’SS -3C to U ratio as Figure 2 supplementary figure 4B. We have included a short section of text in the Results section to describe these additional findings.

      With the 5' splice site +4 variation identified, the next step was to determine the underlying molecular mechanisms that dictate the evolution of the various sequence preferences. Some obvious players here are the U1 and U6 snRNAs which directly interact with the 5' splice site during splicing. However, no association was found between these snRNAs and the 5' splice site +4 sequence. 

      The powerful interspecies association mapping was then used to determine whether the presence or absence of METTL16 ortholog or a splicing factor correlated with the 5' splice site +4 sequence variation. Interestingly, a clear association was found between METTL16 and the 5' splice site +4 position; METTL16 presence was associated with +4A at the 5' splice site and METTL16 absence was associated with +4U at the 5' splice site. This is an exciting and significant finding.

      We thank the reviewer for these comments on the importance of this study.

      Interestingly, the next most significant association with the 5' splice site +4 position was with SNRNP27K. This result makes sense as in the cryo-EM structure of the pre-B spliceosome complex the C-terminal domain of SNRNP27K is found near the region of the U6 snRNA that will interact with the 5' splices site. Absence of SNRNP27K was associated with an increased preference for +4U at the 5' splice site. Now the real power of the interspecies association mapping was demonstrated by investigating whether any association could be determined specifically within the C-terminus of SNRNP27K. Significantly, the methionine 141 position in SNRNP27K was found to be associated with the +4 position of the 5' splice site. This finding fits nicely with previous studies where mutation of M141 caused a shift in 5' splice site selection away from +4A 5' splice sites, to 5' splice sites without +4A. What is not clear is whether M141 is conserved or invariant between all the species that were compared?

      M141 is not completely conserved across the species that were compared for the SNRNP27K C-terminus analysis. We did not test positions with very strong sequence conservation, because without variation in both the genotype and phenotype it is not possible to test for an association. We have rephrased the relevant Results and Methods sections to make this point clearer. In addition, we have incorporated a sequence logo to illustrate the degree of conservation of each position in the SNRNP27K C-terminal domain as Figure 5 -figure supplement 1A. Finally, we have included an additional box-plot to illustrate the finding that species which have lost SNRNP27K or have only lost the Methionine equivalent to human SNRNP27K position 141, show a similar preference for +4U at 5’ SSs. This is now included as Figure 5 - figure supplement 1B.

      Overall, this result reveals the power of the interspecies association approach and provides interesting and exciting information on the molecular determinants of 5' splice site evolution.

      We are grateful to the reviewer for these comments.

      The final analysis was to investigate the interaction potentials of the U5 and U6 snRNAs with the 5' splice site in the Sacchromycotina genomes and try to relate this to species with fewer introns and less alternative splicing. Species with low intron numbers and low splicing complexity were revealed to have weaker U5 and U6 anti-correlation potentials and favor +4U at the 5' splice site. On the other hand, species with high intron number and presumably higher splicing complexity featured anti-correlated U5 and U6 snRNA interaction potentials and favored +4A 5' splice sites. This extensive analysis provides novel information on the interactions and splice site properties of species with simple and complex splicing. Again, I see why there is emphasis on the 5' splice site here but a similar analysis with the U2 snRNA and the branch site could also be informative.

      We absolutely agree that inter-species association mapping could be applied to other splicing signal phenotypes including 3’ splice sites and intron branchpoints. Accordingly, we raise this subject in the final section of the Discussion. However, branchpoint sequences are challenging to predict with genomic data. Because preliminary analyses suggest independent variation in these other splicing signal phenotypes, we feel a separate focused study is required to properly explain (and substantiate) even the analytical approaches involved. We hope the reviewer would agree that incorporating U2 snRNA and branchpoint variation analyses into this manuscript as well, could detract from the clarity of the conceptual advances that we make here.

      Minor comments

      Should the Title include SNRNP27K?

      There is certainly a case that the title should include SNRNP27K. Our aim was to make the title as short and informative as possible without too many acronyms that need explaining. Since the clearest correlation is with METTL16 and this has broader implications for understanding the role of this enzyme not only in splicing but in possibly modifying other RNA targets too, we think not including SNRNP27K is a suitable compromise. In addition, retaining the current title simplifies the tracking of the manuscript from pre-print through to journal publication.

      Should the title specify that it is the evolution of only the 5’ splice site sequence preference being studied here?

      Because apostrophes in titles can compromise some scholarly online search engines (https://insights.uksg.org/articles/10.1629/uksg.534), we would prefer not to include 5’ in the title.

      Include information on intron number and 5’ splice site interaction potential of U5 and U6 snRNA in the Summary?

      We thank the reviewer for this suggestion. We have updated the Summary to include our findings on U5 and U6 interaction potential in species with reduced intron number.

      Figure 1C is not referred to in the text?

      We apologise for this oversight. We have added references to figure 1C in the appropriate Results section.

      Page 8, line 5 – better to say “splicing signal phenotypes”.

      We have amended this statement on Page 8 and at other places in the text where related phrasing was made.

      What are the other points on Figure 3B? What is the next point below SNRNP27K? Is it U2A’? 

      The other points on Figure 3B represent Orthofinder orthogroups which contain human orthologs that are known components of the spliceosome. The list of spliceosomal components was taken from Sales-Lee et al. 2021. The third most significant point is indeed the orthogroup containing the human ortholog of U2A’. As we state in the text, however, the correlation of U2A’ with the 5’SS+4 A to U ratio phenotype is no longer significant once METTL16 presence/absence is controlled for, indicating that the correlation of U2A’ with the +4A phenotype is likely explained by similarity in the patterns of gene loss of U2A’ and METTL16.

      The second paragraph of the Discussion is vague and lacks a reference. “we could also identify an association with a methionine residue in the conserved C-terminal domain of SNRNP27K orthologs.” There are a few methionines in the C-terminus, which one? Please reference the statement “transcriptome analysis of C. elegans SNRP-27 M141T mutants..”

      We apologise for the lower quality of writing in this section of the Discussion. We have updated the text, made the statements about the SNRNP27K C-terminus less ambiguous, and added the relevant citations as appropriate.

      Reviewer #1 (Significance):

      Overall, this is a well written and clearly presented study that provides some key molecular information on the splicing factors involved in the evolution of 5’ splice sites and shows the power of interspecies association studies. Some important conceptual principles have now been defined for the field going forward.

      With thank the reviewer for this kind comment on the importance of this work.

      The question remains as to whether METTL16 and SNRNP27K are the sole determinants of 5’ splice site preference evolution at +4?

      We cannot say for certain that METTL16 and/or SNRNP27K determine the 5’SS +4 phenotype – only that they are correlated with it. In our response to reviewer 3, and in a new Discussion section, we have detailed some of the scenarios that could explain these correlations. We also cannot rule out whether there are changes in the presence/absence (or domain/sequence-level changes) of other, untested proteins that correlate with the 5’SS +4 phenotype and we allude to this in the final section of the Discussion.

      One splicing factor that immediately comes to mind is Prp8 where there is extensive evidence for involvement in splice site selection and is clearly in the right location throughout splicing to be involved. This question should at least be discussed but Prp8 would also be a very interesting candidate for the interspecies association mapping.

      Prp8 is a core component of spliceosomes and is conserved throughout the Saccharomycotina. For this reason, we were unable to associate splicing phenotypes with Prp8 presence or absence variation at the level of orthogroups. However, we revisited this question posed by the reviewer. Our experience with inter-species association mapping, so far, indicates it works well with orthogroup presence/absence or when straightforward amino acid substitutions can be detected in conserved and hence alignable protein sequence domains. We analysed the conserved U6 snRNA-interacting region of the Prp8 linker domain, which maps close to the 5’ splice site in cryo-EM models, using the profile HMM PF10596 available from Pfam. We found that the majority of this domain was extremely highly conserved with variation in only a few species and positions. The strongest correlation with the +4A to U ratio phenotype was at position 58, which is conserved as a Glycine in all but 8 species (6 Dipodascaceae, 2 CUG-Ser1), that also tend to have a stronger preference for +4A. However, examination of the species contributing to this result (and to similar results at other positions) indicated that in the 6 Dipodascaceae species, this change is part of a larger deletion or replacement that makes the whole linker region align poorly to the model. Hence, the G58 position itself may not be specifically important for the +4 phenotype. Although the wholesale loss or replacement of the U6 snRNA-interacting region in these species is potentially interesting, these larger scale structural changes in a small number of species are difficult to interpret. Therefore, to maintain the focus of the manuscript and the clear links to METTL16 and SNRNP27K that have orthogonal support, we have decided not to add these results to the manuscript but present them here (Figure not available on biorXiv commenting window).

      Also, as mentioned previously, only the 5’ splice site was investigated here and the manuscript could become a more substantial piece of work if the other splice sites were included in some way.

      We agree that it will be exciting to apply this approach to other splicing signal phenotypes and in other phylogenetic clades with emerging tree-of-life-scale genomics data. We have included variation in 3’ splice sites in the revised manuscript. As the first of its kind, this study should pioneer a wider use of this approach, by us and others, to understand the mechanisms and functions of molecular interactions not only in splicing but in other areas of biology too.

      The obvious audience here are those directly in the splicing field but the overall principles are relevant for evolutionary biologists and those studying organismal complexity.

      We thank the reviewer for recognising the broad importance of this work.

      My expertise is in yeast and human splicing mechanisms. I do not have the expertise to critically evaluate the bioinformatic pipelines but they were clearly explained and presented.

      Reviewer #2 (Evidence, reproducibility and clarity):

      In their manuscript, Parker et al. investigate the evolutionary patterns of splice site preference, focusing on the A/U ratio at position A+4 on the 5´ splice site. Building upon prior studies in S. pombe and A. thaliana, the authors establish a strong correlation between this preference and the co-evolution of the METTL16 U6 snRNA methyltransferase. Furthermore, through inter-species association mapping, they identify the involvement of the splicing factor SNRNP27K in altered A/U ratios and highlight the significance of the residue Met-141 in SNRNP27K for this function. Overall, the paper effectively presents impactful new findings on the evolution of METTL16, U6 snRNA, and splicing.

      We thank the reviewer for these kind comments on the importance of our study.

      The computational analyses employed in this study are situated outside our field of expertise, preventing us from offering a comprehensive evaluation of the methodology’s appropriateness and rigor. Nonetheless, the identification of METTL16 through the authors’ methods, which aligns with previous research in S. pombe and A. thaliana, lends support to the validity of their approach. Notably, the close proximity between SNRNP27K and the methylated A43 residue in U6 snRNA within the spliceosome, particularly near Met-141, is an impressive finding. Previous studies have shown that a mutation at position M141T affects splicing at +4A introns, thus providing robust validation for their methods.

      We thank the reviewer for these kind comments on our work.

      The data presented in this study furnish crucial insights into the role of METTL16, U6 snRNA methylation, and splice site recognition. The authors expand upon recent observations that the “vertebrate conserved region” exists in non-vertebrates, despite the absence of primary sequence homology. These results will serve as a valuable guide for future molecular investigations into U6 snRNA methylation and its mechanisms in splicing. Furthermore, the implications of this paper extend to human evolution, as the plasticity in splicing is an essential factor in the evolution of developmental complexity.

      We thank the reviewer for these kind comments.

      Minor suggestions for improvement:

      1. Given the significance of the interaction between U6 snRNA and the intron for understanding the data, it would be beneficial to include a figure illustrating the RNA-RNA base-pairing interactions between U6 snRNA and the 5´ splice site. This addition is particularly important if the paper is intended for publication in a journal with a general readership.

      We thank the reviewer for this excellent suggestion. We have included this as Figure 3A.

      1. Similarly, the section on U1 snRNA would be more comprehensible with the inclusion of U1 RNA-RNA intron diagrams and improved descriptions of both the figures and the assay. Despite being negative data in the supplement, clarifying this section is essential. As currently written, it is challenging to follow.

      We agree that this section is difficult to follow. We have updated the text to improve the readability and included a figure of U1 snRNA:5’SS basepairing as Figure 3 – figure supplement 1A.

      1. Whenever possible, consider increasing the figure and font sizes to enhance readability for readers.

      We agree that some of the more complex figures can be difficult to read when embedded into a Word document/pdf. We hope that providing high-resolution figures for reading online will mitigate this.

      1. In the text, there is no reference to Figure 1C.

      We apologise for this oversight. We have resolved this issue with the appropriate references in the Results text.

      1. In Figure 5B, the y-axis in the top panel is labelled “species,” but the legend only mentions U5/6p as the y-axis. Please revise the legend to include the appropriate information.

      We apologise for the confusion caused by our poorly written legend for this plot. We have updated the legend so that the text clearly refers to either the scatter plot or the marginal histograms.

      Reviewer #2 (Significance):

      The data presented in this study furnish crucial insights into the role of METTL16, U6 snRNA methylation, and splice site recognition. The authors expand upon recent observations that the “vertebrate conserved region” exists in non-vertebrates, despite the absence of primary sequence homology. These results will serve as a valuable guide for future molecular investigations into U6 snRNA methylation and its mechanisms in splicing. Furthermore, the implications of this paper extend to human evolution, as the plasticity in splicing is an essential factor in the evolution of developmental complexity.

      We are grateful to the reviewer for these kind comments on the importance of this work.

      Reviewer #3 (Evidence, reproducibility and clarity):

      In this manuscript, Parker et al present a nice exploration of the evolutionary and mechanistic relationships between 5′ splice site consensus sequences, intron numbers and METTL16/SNRNP27K. By performing inter-species association mapping in Saccharomycotina species, they found that a T in position +4 is strongly associated with the absence of METTL16 (and/or in some cases SNRNP27K or mutations in it). They also provide solid structural modelling data in support of this association.

      In general, I think this is a very nice manuscript. I only have a few comments, which could be addressed by rewording specific parts and/or improving the current figures.

      We are grateful to the reviewer for the kind comments on this work.

      1) As the authors acknowledge, a key issue that cannot be fully resolved in this study is causality between the different events investigated. Overall, the authors are careful about this, but there are some exceptions that should be corrected. Probably the most important is in the abstract, where they write: “We conclude that variation in concerted processes of 5’ splice site selection by U6 snRNA is crucial to evolutionary change in splicing complexity”. I suggest they write something more open (and correct), such as: “We conclude that variation in concerted processes of 5’ splice site selection by U6 snRNA is associated with evolutionary changes in splicing complexity”. Similarly, other plausible scenarios should be discussed in the corresponding Discussion section.

      We agree with the reviewer that it is not possible to infer the causal relationship between METTL16 absence and 5’SS+4 preference change from the current data. We, therefore, apologise for failing to be more careful in the Summary and Introduction. We have reworded these statements to better reflect what we can currently say about the evolutionary relationship between METTL16 and 5’SS sequence preference.

      The correlation between METTL16 absence and 5'SS+4 sequence preference change could most likely be explained by one of several scenarios: (a) sudden loss of METTL16 causes a rapid necessity to change 5'SS sequence preferences. This is unlikely as such rapid change without widespread corresponding 5'SS changes would likely impose a high fitness cost. (b) Changes in 5'SS sequence preference occur first, driven by some other selective pressure, until there is no longer a benefit to retaining the METTL16 gene. (c) Gradual changes in the expression or catalytic efficiency of METTL16 reduce the stoichiometry of U6 snRNA m6A modification, which permits gradual change in 5'SS+4 sequence preference until complete loss of the METTL16 no longer imposes a major fitness cost. As we suggest in the Discussion, future work could examine this question by determining whether the METTL16 orthologs found in Zygosaccharomyces and Eremothecium species, which have altered their 5'SS+4 preference to a U, are expressed and functional. We have updated the Discussion to include a new section that addresses these scenarios.

      2) I do not agree with the statement that "The extent of alternative splicing is the best genomic predictor of developmental complexity". To start with, there are many ways to quantify "extent of alternative splicing" and there are also different types of alternative splicing that might have different prevalence and biological impact. Then, this claim is usually related with exon skipping, which is tightly linked with intron length, and that is likely a better prediction of complexity (yet clearly not causative). My concern is: to what extent has this claim been formally and properly assessed by comparing splicing prevalence with other genomic features, such as intergenic region length, intron length, or average distance between enhancer-promoter interactions (arguably the most relevant predictor, in light of many other studies)? Moreover, I found it a bit misleading to frame the work presented in this study as directly related with developmental (or even splicing) complexity. The work is very interesting on its own, and I doubt their findings on +4 position preference in Saccharomycotina has anything to do with developmental complexity (as the Abstract and Introduction seem to imply).

      On reflection, we agree with the reviewer. Some of our framing of the text isn’t balanced with other studies on the scaling of alternative splicing with developmental complexity. We have edited the Summary and Introduction sections accordingly and cited other references that broaden the consideration of this subject. We are grateful to the reviewer for this suggestion because the changes we make improve the focus of the manuscript since our findings relate more to splicing simplification than to an understanding of increased developmental complexity.

      3) I found Figure 2 and its associated supplementary figure very difficult to follow. I suggest the authors try to improve it and make it clearer. Also, other trees summarizing the results might be helpful. 

      We apologise for the complexity of these figures. We opted to show phylogenetic trees with phenotypes plotted on the y axis, rather than simply trait histograms or box-plots, because the underlying structure of the tree is important for demonstrating that multiple independent changes in the 5’SS phenotype have occurred in the Saccharomycotina. We have tried to improve the comprehensibility of the figures in the following ways: (a) We have added 5’SS sequence motifs to the x-axis of figure 2B to make what the plot represents clearer, (b) as suggested by the reviewer, we have created a pruned tree showing the 5’SS motifs of a selection of Saccharomycotina species, which demonstrates that the changes in 5’SS+4 position preferences seen in S. cerevisiae and C. albicans are likely to be a result of convergent evolution. We have added this tree as Figure 2 - figure supplement 3.

      4) I also found the Results section corresponding to Figure 5B a bit confusing. I would argue (as I think the authors do) that there are two main patterns here: below 500 introns, there is no association, while above 500 introns there is an increasingly negative association (correlation). I think it would help to more explicitly distinguishing these two patterns. Then, for the intron-poor species: is the correlation (or lack of) for species with a T or an A in position +4 different? 

      We do indeed think that there are two patterns here, as indicated by the reviewer. In the previous version of the manuscript, we separated species into those having an overall preference for A at the +4 position, and those having +4U. By showing regression lines for these two classes, rather than for the general relationship between intron number and U5/6rho, we somewhat imply that the switch in +4 base preference might be causing the loss of correlation between U5/6rho and intron number. However, since essentially all species with a 5'SS +4U preference are intron poor, it seems more likely that these trends are the result of a loss of the negative correlation between intron number and U5/6rho in intron poor species, as suggested by the reviewer. To address this issue, we have replaced the regression lines on Figure 6B with a single loess (locally estimated scatterplot smoothing) regression line for all species and updated the text to make it clearer that we think loss of U5/6rho and +4A preference are separate traits of intron poor species. Although this is not exactly what the reviewer requested, we hope that it satisfies their issue with the analysis.

      Reviewer #3 (Significance):

      This is a very interesting study that sheds light on an intriguing evolutionary pattern: the change in consensus sequence at position +4 of the 5' splice site. This topic is relevant since it is closely associated with intron loss and splicing efficiency and evolution. 

      We thank the reviewer for the kind and constructive comments on this study.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      Reviewer #1*. This is a good paper dealing with gap of our knowledge in understanding reason of ICB failures. Subject being difficult it is expected that the design and content of such experiment will be complex.But the authors forget practicality of readers attention and making paper apear interesting. They need to organise and may be classify the varied information in such a way that reader can find a rhythm in excavating data more easily. It appears confusing at time, so they may try to make it more simple. In this way they may concentrate more on methods and classify results too. A thorough revision is suggested, to make it consize. *

      __Authors’ answer: __We thank the Reviewer for his positive evaluation and constructive feedback. We appreciate the complexity of single-cell RNA-sequencing analyses. In order to simplify our manuscript, our revised manuscript now focuses on the transitional states of tumor-resident and circulating T cells found in ovarian cancer patients. Our study is timely as it is the first to report the developmental relationship of TILs in ovarian cancer. We substantially edited our manuscript to make it clear that our findings suggest a gradual acquisition of the exhaustion program initiated by effector-like cells (cluster CD8_GZMH) that eventually gives rise to more terminal states with features of tissue residency and chemotaxis (clusters CD8_CCL4, CD8_XCL1, and CD8_CXCL13). We also include new analyses revealing the presence and proportion of these T cell states in different cancer patients (New Fig. 4A-B), and how these T cell states associate with clinical responses to immune checkpoint blockade (ICB). We hope the Reviewer will find our revised manuscript easier to read.

      Reviewer #2. I think the first half of the article, in which the GZMH-CD8 cluster is considered to be in an intermediate state of transition to exhaustion, is interesting, and I feel that the single-cell seq and TCR data are well analyzed to make the point. On the other hand, I feel that the latter part of the paper may not be anything more than a hypothesis. In particular, the part claiming that it is related to prognosis or applicable to the prediction of the effect of ICB is insufficient, since their gene signature is not described in detail and the contents of the Figure are not mentioned in the manuscript. In the latter part, the effects of GPR184 and 25-HC, or the effects of IL21, would require experiments to verify (to verify whether the addition of chemokine or the inhibition of the receptor changes the specific CD8 population).

      Author’s answer: Thank you for discussing the limitation of the signature employed. We agree with the reviewer’s comment. Old Figure 5 has been removed from the revised manuscript.

      Reviewer #2. Minor point: In particular, there is little mention of Figure 5 in the text, making it difficult to understand.

      Author’s answer: Thank you for your comment. As we previously discussed, we have removed Figure 5 from the revised manuscript. The method used to generate the signature was found to be inappropriate.

      Reviewer #2. The latter part is difficult to understand. To begin with, it is already known that ovarian cancer does not contribute much to ICB, so what does it mean to analyze the CD8 population, which is known as a marker of ICB response in other carcinomas, as an indicator? Especially for clinicians like us, it is hard to imagine that the results will lead to clinical trials that will attempt to sort out the population that ICB is favored in.

      Author’s answer: Although immune checkpoint blockade has demonstrated limited effectiveness against ovarian cancer, subset analyses suggest superior efficacy for some patients and according to subtype. Combination anti-PD-1/CTLA-4 therapy for instance achieved response rates up to 31% (Zamarin et al., 2020), and superior benefit for single agent PD-1 blockade has been reported in clear cell ovarian cancer. Moreover, encouraging clinical results have recently been reported in studies exploring combinations with PARP and VEGF inhibitors. As example, interim analysis of the phase 3 DUO-O trial (NCT03737643) showed a statistically significant and clinically meaningful improvement in PFS in patients with newly diagnosed advanced ovarian cancer without a BRCA1/2 mutation (Harter et al., 2023).

      Our study aimed to better understand how ovarian tumor-infiltrating T cells acquire their exhaustion program after migrating from the periphery and whether these mechanisms are unique or shared amongst cancer types. Recent studies in other cancer types had shown the dynamics of T cells and demonstrated the clonal replacement of intratumoral T cells after ICB and emphasized the role of peripheral clones in this process (Wu et al., 2020; Yost et al., 2019). In lung cancer, it has been proposed a transitional state between precursor and terminally differentially cells (Gueguen et al., 2021). Our study demonstrates, for the first time in ovarian cancer, the presence of similar transitional states of CD8 T cells. Our revised manuscript also now includes new data revealing that pre-effector GZMK- and intermediary GZMH-expressing CD8 cells are better biomarkers of ICB response than terminally differentiated XCL1 and CXCL13 expressing CD8 T cells (New Figure 4). Altogether, our study provides important and novel insights on the development of tumor-infiltrating T cells in ovarian cancer patients, which may serve to better select ovarian cancer patients for ICB therapy.

      Reviewer #2. Since the first half of the study is very interesting, we feel that it is more important to confirm the mechanism of exhaustion from the blood via the intermediate (GZMH_CD8), including functional experiments. Also, as a clinician, we are very interested in the perspective of whether some of the fractions identified in this study are different in proportion in different patients and whether they correlate with the clinical course of the disease since the study only analyzed a sample of 5 patients.

      Author’s answer: We thank the reviewer for proposing to extend our analysis. As suggested, our revised manuscript now includes new analyses which reveals the different proportions of our identified T cells states in different cancer patients (New Figure 4). We further investigated whether these T cell states associate with clinical responses and observed that pre-effector GZMK- and intermediary GZMH-expressing CD8 T cells are better biomarkers of ICB response than terminally exhausted XCL1- and CXCL13-expressing CD8 T cells (New Figure 4).

      Reviewer #3. Question 1: Whether the distribution patterns of CD4+ and CD8+ T cell clusters in Figure 1B were comparable among the 5 patient samples? Whether the proportion of five types of clones in Figure 3C are comparable among the 5 patient samples?

      Author’s answer: Thank you for the question. We included the results to answer these questions in the supplementary material (fig. S1C-D). For each patient, we calculated the proportion of a cluster among T cells in the blood or tumor. As observed in the boxplot (fig. S1C), the proportion of some subsets were higher in certain patients, such as the higher proportion of CD8_GZMK in the tumor of patient p09454. A recent study classified patients’ tumors based on the spatial distribution of CD8 T cells and performed scRNA-seq to identified cell subsets enriched in the groups inflamed/infiltrated (characterized by the distribution of CD8 T cells within the tumor epithelium), excluded (infiltrating CD8 T cells are restricted to the tumor stroma) or desert (T cells are not present or have low frequency) (Hornburg et al., 2021). Interestingly, this subset of CD8_GZMK cells were enriched in desert tumors, suggesting that the difference we observed in our dataset might reflect the spatial distribution of CD8 T cells in patient p09454. Regarding the TCR-seq data, the frequency of the five types of clones was different among patients. To show this data, we included a barplot (fig. S2D), showing for example, a higher proportion of tumor-expanded clones in patient p10329.

      Reviewer #3. Question 2: In Figure S2C, only a very small number of cells in the CD8-GZMK K-22 population. Are these cells representative? Do they generally exist in multiple samples or only in one sample?

      Author’s answer: Thank you for your comment. The subcluster k_22 indeed has a smaller number of cells compared to other subclusters. Nevertheless, the K_22 cluster was found in every patient and in every healthy donor. To clarify, we edited our revised manuscript to include a statement that cluster k_22 was composed of fewer cells compared to other clusters.

      Reviewer #3. Question 3: In the Fig.S6 legend, the authors stated "Our results suggest the differentiation of cluster CD8-GZMK into the effector-like subset CD8-GZMH." However, there seems to be no corresponding analysis in the main text to support this conclusion.

      Author’s answer: We appreciate your attention to this statement. We agree the results of our study doesn’t sustain this statement and so we have excluded it in the revised manuscript.

      Reviewer #3____. Question 4: Is there more detailed clinical information that can be provided for the 5 patients included in the study? Per the methods all patients were receiving debulking surgery and were treatment naïve, but did they differ in stage, age, comorbidities, etc.?

      Author’s answer: Thank you for your comment on this. We have included a table with clinical information on the stage, age, and menopause status of the five patients.

      Reviewer #3. Question 5: Were any cells included for sequencing from adjacent 'normal' tissue uninvolved with tumor (these samples are from surgical debulking of primary tumors, which may include such areas of non-involved tissue.) While shared TCR clonotypes between blood and intratumoral T cells strongly suggests the tumor-resident populations are recruited from the blood, the degree of sharing with normal tissue-resident T cells would be of interest as well.

      Author’s answer: Thank you for your comment. Samples were provided for sc-RNA-seq after pathology review and validation of tumor histology. We did not perform sc-RNA-seq on normal adjacent tissue (NAT) We agree this would be interesting as a follow up study, since in other cancer types (renal, colon and lung) it has been demonstrated that T clones expanded in the tumor and NAT are also present in peripheral blood (Wu et al., 2020).

      Reviewer #3. Question 6: Very little is discussed about HGSOC itself in the main text (eg clinical background, prior literature on the composition of infiltrating immune populations and potential reasons for at best modest poor responses to IO) until the first sentence of the discussion. As the entirety of the new data produced in this study is from HGSOC tumors there should be more focus on this tumor type and conversation with the prior literature on it (mainly from prior studies on the immune environment of HGSOC). Further, how distinct do the authors suspect the cell populations found in their study to be to ovarian as opposed to other epithelial tumor types?

      Author’s answer: Thank you for the suggestion. We now included more background information on immunotherapy of HGSOC. Specifically, we added the following paragraph in our introduction: “In ovarian cancer, the presence of both T and B cells improves patients' survival (Nelson, 2015; Nielsen et al, 2012). They are usually organized in lymphoid aggregates ranging from a small group of cells to a well-organized TLS (Kroeger et al, 2016). Organized TLSs correlate with better survival, such as observed in patients treated with ICB. Although immunotherapy has demonstrated limited effectiveness against ovarian cancer, subsets of patients may thus benefit from ICB. In support of this, combination anti-PD-1/CTLA-4 therapy can achieve response rates above 30% (Zamarin et al., 2020), and encouraging clinical results have recently been reported when combining ICB with with PARP and VEGF inhibitors (Harter et al., 2023)”.

      Reviewer #3. Question 7: Were the signature genes used for analysis in figure 5 remove chosen in a formal, unbiased manner, or simply hand-picked as representative of the respective cell types? This information is not provided in the supplement.

      Author’s answer: Another reviewer has also expressed similar concerns. The genes selected to represent cell types were chosen manually, which we acknowledge is not the best method for defining a signature. As a result, we have decided to exclude Figure 5 from the manuscript under review. We believe an unbiased approach is more suitable for characterizing the cell network proposed in our study.

      Reviewer #3. Question 8: While the NicheNet analysis of potential interactions among lymphocyte populations raises some strong hypotheses, it would be interesting to extend the interaction analysis to all CD45+ populations, given the sequencing was done on CD45+ immune cells.

      Author’s answer: Thank you for suggesting analysis. We have included the results of cell interaction including all CD45+ cells (fig. S3). We observed CD40L as one of the top predicted ligands highly expressed in CD4_CXCL13 subset mediating a response in subsets of antigen-presenting cells, such as B cells (cluster B), plasma cells (cluster PC_2), and plasmacytoid dendritic cells (cluster pDC). Interestingly, this result also support the hypothesis of Tfh-like cells (cluster CD4_CXCL13) coordinating the action of intratumoral immune cells involved in the antitumor immune response.

      Reviewer #3. Question 9: A sample size of 5 patients is relatively small for current single cell RNAseq studies of human tumor patients.

      Author’s answer: We agree with the reviewer that a sample size of 5 patients is relatively small. Thus, to validate our results in other patients, we included in the reviewed manuscript the analysis of scRNA-seq of 47 patients across10 cancer types (dataset from (Zheng et al., 2021). As demonstrated in figure 3 and figure 5, we could identify subsets of CD8 and CD4 T cells from our ovarian cancer patients in those 10 cancer types dataset.

      Reviewer #3.____ Minor

      *1. In lines 96-97, "CD8-GZMB" was mentioned twice in the description. *

      2. In line 126, this section did not discuss residency markers, yet a conclusion about residency was made in this sentence.

      Author’s answer: We appreciate you bringing these errors to our attention. We fixed them in the updated version of the manuscript.

      References:

      Gueguen, P., Metoikidou, C., Dupic, T., Lawand, M., Goudot, C., Baulande, S., … Amigorena, S. (2021). Contribution of resident and circulating precursors to tumor-infiltrating CD8 T cell populations in lung cancer. Science Immunology, Vol. 6, p. eabd5778. doi:10.1126/sciimmunol.abd5778

      Harter, P., Trillsch, F., Okamoto, A., Reuss, A., Kim, J.-W., Rubio-Pérez, M. J., … Aghajanian, C. (2023). Durvalumab with paclitaxel/carboplatin (PC) and bevacizumab (bev), followed by maintenance durvalumab, bev, and olaparib in patients (pts) with newly diagnosed advanced ovarian cancer (AOC) without a tumor BRCA1/2 mutation (non-tBRCAm): Results from the randomized, placebo (pbo)-controlled phase III DUO-O trial. Journal of Clinical Orthodontics: JCO, 41(17_suppl), LBA5506–LBA5506.

      Hornburg, M., Desbois, M., Lu, S., Guan, Y., Lo, A. A., Kaufman, S., … Wang, Y. (2021). Single-cell dissection of cellular components and interactions shaping the tumor immune phenotypes in ovarian cancer. Cancer Cell. doi:10.1016/j.ccell.2021.04.004

      Wu, T. D., Madireddi, S., de Almeida, P. E., Banchereau, R., Chen, Y.-J. J., Chitre, A. S., … Grogan, J. L. (2020). Peripheral T cell expansion predicts tumour infiltration and clinical response. Nature. doi:10.1038/s41586-020-2056-8

      Yost, K. E., Satpathy, A. T., Wells, D. K., Qi, Y., Wang, C., Kageyama, R., … Chang, H. Y. (2019). Clonal replacement of tumor-specific T cells following PD-1 blockade. Nature Medicine. doi:10.1038/s41591-019-0522-3

      Zamarin, D., Burger, R. A., Sill, M. W., Powell, D. J., Jr, Lankes, H. A., Feldman, M. D., … Aghajanian, C. (2020). Randomized Phase II Trial of Nivolumab Versus Nivolumab and Ipilimumab for Recurrent or Persistent Ovarian Cancer: An NRG Oncology Study. Journal of Clinical Oncology: Official Journal of the American Society of Clinical Oncology, 38(16), 1814–1823.

      Zheng, L., Qin, S., Si, W., Wang, A., Xing, B., Gao, R., … Zhang, Z. (2021). Pan-cancer single-cell landscape of tumor-infiltrating T cells. Science, 374(6574), abe6474.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      Below are some questions as well as suggestions for revision and to strengthen the manuscript.

      Major:

      1. Whether the distribution patterns of CD4+ and CD8+ T cell clusters in Figure 1B were comparable among the 5 patient samples? Whether the proportion of five types of clones in Figure 3C are comparable among the 5 patient samples?
      2. In Figure S2C, only a very small number of cells in the CD8-GZMK K-22 population. Are these cells representative? Do they generally exist in multiple samples or only in one sample?
      3. In the Fig.S6 legend, the authors stated "Our results suggest the differentiation of cluster CD8-GZMK into the effector-like subset CD8-GZMH." However, there seems to be no corresponding analysis in the main text to support this conclusion.
      4. Is there more detailed clinical information that can be provided for the 5 patients included in the study? Per the methods all patients were receiving debulking surgery and were treatment naïve, but did they differ in stage, age, comorbidities, etc.?
      5. Were any cells included for sequencing from adjacent 'normal' tissue uninvolved with tumor (these samples are from surgical debulking of primary tumors, which may include such areas of non-involved tissue.) While shared TCR clonotypes between blood and intratumoral T cells strongly suggests the tumor-resident populations are recruited from the blood, the degree of sharing with normal tissue-resident T cells would be of interest as well.
      6. Very little is discussed about HGSOC itself in the main text (eg clinical background, prior literature on the composition of infiltrating immune populations and potential reasons for at best modest poor responses to IO) until the first sentence of the discussion. As the entirety of the new data produced in this study is from HGSOC tumors there should be more focus on this tumor type and conversation with the prior literature on it (mainly from prior studies on the immune environment of HGSOC). Further, how distinct do the authors suspect the cell populations found in their study to be to ovarian as opposed to other epithelial tumor types?
      7. Were the signature genes used for analysis in figure 5 chosen in a formal, unbiased manner, or simply hand-picked as representative of the respective cell types? This information is not provided in the supplement.
      8. While the NicheNet analysis of potential interactions among lymphocyte populations raises some strong hypotheses, it would be interesting to extend the interaction analysis to all CD45+ populations, given the sequencing was done on CD45+ immune cells.
      9. A sample size of 5 patients is relatively small for current single cell RNAseq studies of human tumor patients.

      Minor

      1. In lines 96-97, "CD8-GZMB" was mentioned twice in the description.
      2. In line 126, this section did not discuss residency markers, yet a conclusion about residency was made in this sentence.

      Significance

      In this manuscript titled "Predicting Developmental Relationships of Tumor-Resident and Circulating T Cells in Ovarian Cancer," Carneiro and colleagues employed single-cell transcriptomics and T cell receptor profiling of immune cells sorted from paired peripheral blood and tumor tissue in a small cohort of ovarian cancer patients to investigate the developmental relationships of T cell populations and their potential interactions. They identified a possible differentiation pathway involving GZMH-expressing CD8+ T cells that progresses towards tissue residency and exhaustion. The researchers suggested the effector function of intermediate GZMH-expressing CD8+ T cells could be sustained through interaction with CXCL13-expressing CD4+ Tfh-like cells. Moreover, they proposed that CD4+ Tfh-like cells could attract GPR183-expressing pre-effector GZMK-expressing CD8+ T cells and plasmacytoid dendritic cells via the production of 7α,25 dihydroxycholesterol (7α,25-HC). Ultimately, the study hypothesized that CD4+ Tfh-like cells expressing IL-21 among other molecules might enhance antitumor immunity against ovarian tumors by coordinating the actions of multiple immune populations. Strengths of the study include detailed, combined analysis of inferred developmental trajectories via shared TCR clonotypes across tissue as well as potential crosstalk between cellular populations, as well as association of signature genes with clinical outcomes. Weaknesses include a small number of patients and the dataset being limited only to single cell RNAseq and thus providing descriptive findings without functional validation or perturbation.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary: This study used single-cell transcriptomics and T cell receptor profiling to identify the developmental relationships of T cell populations in ovarian cancer patients. The researchers proposed a model of differentiation pathway that showed how an intermediate GZMH-expressing CD8 T cell subset progressively reinforces exhaustion and tissue residency programs towards terminally exhausted cells. Then they also focus on the nature of TPEX, dual-expanded clone, which is considered an important indicator for the efficacy of ICB, and argue that it is strongly related to GPR183, 25-OHC, and IL21. Based on the analysis of clinical samples, they argue that their proposed gene signature may also be prognostically relevant and predictive of ICB efficacy.

      Major comment: I think the first half of the article, in which the GZMH-CD8 cluster is considered to be in an intermediate state of transition to exhaustion, is interesting, and I feel that the single-cell seq and TCR data are well analyzed to make the point. On the other hand, I feel that the latter part of the paper may not be anything more than a hypothesis. In particular, the part claiming that it is related to prognosis or applicable to the prediction of the effect of ICB is insufficient, since their gene signature is not described in detail and the contents of the Figure are not mentioned in the manuscript. In the latter part, the effects of GPR184 and 25-HC, or the effects of IL21, would require experiments to verify (to verify whether the addition of chemokine or the inhibition of the receptor changes the specific CD8 population).

      Minor point: In particular, there is little mention of Figure 5 in the text, making it difficult to understand.

      Significance

      It is interesting to note that the authors simultaneously analyze immune cells in the blood and in the tumor, and examine in detail what is characteristic of the blood, what is characteristic of the tumor, and what is seen in both. And it is very interesting that they specifically proposes an intermediate group that is recruited from the blood to the tumor and is in the process of becoming exhausted. I am sure there are many studies on TILs and TLSs, but this study would be helpful to understand how they are concentrated locally (near the tumor) in comparison with immune cells in the blood as well.

      However, the latter part is difficult to understand. To begin with, it is already known that ovarian cancer does not contribute much to ICB, so what does it mean to analyze the CD8 population, which is known as a marker of ICB response in other carcinomas, as an indicator? Especially for clinicians like us, it is hard to imagine that the results will lead to clinical trials that will attempt to sort out the population that ICB is favored in.

      Since the first half of the study is very interesting, we feel that it is more important to confirm the mechanism of exhaustion from the blood via the intermediate state, including functional experiments. Also, as a clinician, we are very interested in the perspective of whether some of the fractions identified in this study are different in proportion in different patients and whether they correlate with the clinical course of the disease, since the study only analyzed a sample of 5 patients.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      This is a good paper dealing with gap of our knowledge in understanding reason of ICB failures. Subject being difficult it is expected that the design and content of such experiment will be complex.But the authors forget practicality of readers attention and making paper apear interesting.

      They need to organise and may be classify the varied information in such a way that reader can find a rhythm in excavating data more easily.

      It appears confusing at time, so they may try to make it more simple.

      In this way they may concentrate more on methods and classify results too.

      A thorough revision is suggested, to make it consize.

      Significance

      This is a good paper dealing with gap of our knowledge in understanding reason of ICB failures. Subject being difficult it is expected that the design and content of such experiment will be complex.But the authors forget practicality of readers attention and making paper apear interesting.

      They need to organise and may be classify the varied information in such a way that reader can find a rhythm in excavating data more easily.

      It appears confusing at time, so they may try to make it more simple.

      In this way they may concentrate more on methods and classify results too.

      A thorough revision is suggested, to make it consize.

    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      The authors do not wish to provide a response at this time.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      In the manuscript entitled "Aurora A mediated new phosphorylation of RAD51 is observed in Nuclear Speckles", the authors unveil the Serine S97 as a novel phosphorylation site of the RAD51 recombinase and that this phosphorylation is mediated by the Aurora A kinase using a set of in vitro and in cellulo experiments. The authors also describe this phosphorylation being in the nucleus specifically in nuclear speckles where mRNA maturation and splicing occurs suggesting a role of RAD51 in the latter. The confocal microscopy images provided to test this hypothesis are convincing. However, using confocal images as well, the authors claim that RAD51 phosphorylated at S97 foci do not colocalize with the DNA damage marker -H2AX, hence a function not related to DNA damage, however the data provided does not fully support this statement. In this study, Alaouid et al, utilize mutants of RAD51 that alter S97 phosphorylation to further study its function and provide data that support RAD51 as an RNA binding protein. Overall, the manuscript shows some interesting observations that are worth pursuing however the in vitro and in cellulo results are not aligned, lack some controls, and many points should be reconsidered.

      Major comments:

      • Are the key conclusions convincing?

      Not as stated.

      Fig. 1A. The authors conclude that pS97-RAD51 favors RAD51 strand invasion capacity using the D-loop assay. Indeed, the S97D phosphomimic increased the D-loop activity 2.5-fold compared to WT RAD51. However, the S97A mutant, which is the non-phosphorylated form also increased the D-loop activity by 2-fold compared to WT (figure 1C). So, the phosphorylation or the absence of it seem to promote strand invasion. So, what is the role of the phosphorylation? There is no discussion about this. Besides, no representative image of the D-loop assay is shown, this is very important as these experiments need to be run with the relevant controls to be meaningful.

      Fig. 1D. The polymerization rate of RAD51 is probably irrelevant for its function in the absence of DNA. What do they want to get at with this assay?

      In figure 2B, the authors conclude that RAD51 phosphorylation at S97 is dynamically regulated throughout the cell cycle. Indeed, the pS97-RAD51 is well observed in asynchronous cells, and the double thymidine block time course experiment followed by PI staining shows the oscillation of the pS97-RAD51 from G1 to G2/M stage. The authors quantified the ratio of pS97-RAD51/total RAD51 to conclude this. However, it would be more accurate to also divide the above over the intensity of the loading control (tubulin) because in figure 3A for example, they quantified the ratio of pS97-RAD51/tubulin but did not consider the levels of RAD51 in their quantifications.

      In figure 3B, the authors state that pS97-RAD51 is decreased after CPT treatment and that the pS97-RAD51 foci do not localize with the DNA damage marker -H2AX. The signal of gH2AX is already weird as it does not change from Ctrl to CPT conditions (especially in HCC1806 cells). A pre-extraction of soluble protein with CSK should be used to then look at the co-localization, with the pan-staining of the two signals is difficult to draw any conclusions of colocalization. Nevertheless, the signal of RAD51 seems equal in all conditions in the images shown and it does not seem to be reduced after CPT.

      In figure 4A, the authors show that Aurora A is responsible for the S97-RAD51 phosphorylation in cellulo. Indeed, the use of an Aurora A inhibitor reduces the pS97-RAD51 signal, however, this is only true in one cell line (HCC1806) but no effect was observed in HeLa cells. Is this effect cell-specific?

      The authors find that RAD51 binds both DNA and RNA and measure the affinities of the RAD51 bearing the S97D and S97A mutations. S97D shows the highest affinity for ssDNA and RNA in Fig. 7A, B, however the opposite is true for dsDNA in Fig 7C, D. All three forms of RAD51 bind RNA although with different affinities however no error bars are shown. The description of the results does not seem accurate. Importantly, these data should somehow correlate/be discussed with respect to the D-loop assay performed in Fig. 1. The authors conclude that the binding to RNA is reduced in S97D-RAD51 suggesting that the pRAD51 that they observe at nuclear speckles would be probably not associated with RNA at these nuclear speckles, right? this goes against their idea of this phosphorylated form being related to RNA splicing... - Should the authors qualify some of their claims as preliminary or speculative, or remove them altogether?

      The manuscript seems to be in early days and requires lots of editing, rewriting to relate the in vitro and in cell data and make a coherent story - Would additional experiments be essential to support the claims of the paper? Request additional experiments only where necessary for the paper as it is, and do not ask authors to open new lines of experimentation.

      The authors performed chromatin fractionation to determine the correct localization of the pS97-RAD51 and looked for the phosphorylated form by western blots. But then they confirmed the finding using immunofluorescence. I think it would be more convincing and consistent if the authors do a pre-extraction before the use the antibody because as such, they would be indeed confirming the localization of the protein they are looking at that is specifically in the nucleus.

      As well, in order to test the specificity of the pS97-RAD51 antibody they generated, a simple treatment of the lysates with phosphatases would be a good control for the specificity of their antibody These and the critics mentioned above need to be address. - Are the suggested experiments realistic in terms of time and resources? It would help if you could add an estimated cost and time investment for substantial experiments.

      This manuscript is not ready for submission - Are the data and the methods presented in such a way that they can be reproduced?

      Yes. However, the legends of the images are way too concise. - Are the experiments adequately replicated and statistical analysis adequate?

      In Fig. 2B, the authors performed a double thymidine block followed by a time course release to track cell cycle progression of the cells and phosphorylation of RAD51 at S97. They do not indicate the biological replicates they performed. There are no error bars in the estimated KD shown in Fig.7.

      Minor comments:

      • Specific experimental issues that are easily addressable.

      The authors conclude that the S97 is specifically phosphorylated by the Aurora A kinase. How? Have they looked at other documented kinases known to phosphorylate RAD51?

      In figure 6 the authors overexpress HA-tagged RAD51 proteins corresponding to WT, S97D and S97A mutants in cells and label them for immunofluorescence. Maybe it would be better to downregulate the endogenous RAD51 to discard possible combined effects.

      In figure - Are prior studies referenced appropriately?

      The authors show in their manuscript that RAD51 protein CAN interact with RNA in vitro, a finding not previously described to my knowledge. However, a recent study entitled "RAD51-dependent recruitment of TERRA lncRNA to telomeres through R-loops, Nature, 2020" provides in vitro data showing the binding of RAD51 to TERRA, a LncRNA, which I think would be worth mentioning their manuscript.

      The authors should mention previous contributions in the field especially when it comes to RAD51 in the HR pathway post DNA damage, which is quite documented and updated. For example, in this section of the introduction, "RAD51 is a recombinase protein implicated in the strand exchange mechanism during the DSB repair by the Homologous Recombination (HR) pathway. In the absence of DNA Damage (DD), RAD51 is predominantly cytoplasmic and translocates to the nucleus during the DNA Damage Response (DDR) to manage HR repair. As it needs the undamaged sister chromatid as a template, the HR repair pathway occurs mainly in the late S, G2 phases of the cell cycle. However, it has been documented that HR repair can also occur during G1 and early S phases, and in this case, the undamaged template used for the repair could be the homologous chromosome or an RNA transcript2". This statement is definitely worth more references.

      The same problem is recurrent in the rest of the introduction; therefore, it needs to be updated and better referenced. - Are the text and figures clear and accurate?

      The text needs a lot of editing to accurately describe the results, see for example: "The resulting KD evaluation shows that the S97D mutant had a dsDNA binding affinity lower to that of the WT (a KD of 2.26 μM for the S97D-RAD51 vs a KD of 0.38 μM for the WT RAD51). Concerning, the S97A mutant comparison to the WT RAD51, we observed modified association and dissociation curves that resulted in an identical affinity to dsDNA (a KD of 0.33 μM for the S97A-RAD51 vs a KD of 0.38 μM for the WT RAD51). We can conclude that in our in vitro conditions, the Ser97 phosphorylation has a high impact on RAD51 affinity for DNA by dividing its affinity by 5.8." Besides, the figures are of low quality and should be more carefully crafted and presented. Some experiments (such as the D-loop) are not represented in the figures.<br /> - Do you have suggestions that would help the authors improve the presentation of their data and conclusions?

      Using a different representation for the graphs would be a plus (also see previous comments)

      Referees cross-commenting

      I think the other reviewers and I have raised very important and complementary points that will help the authors improve the quality of the manuscript substantially.

      Significance

      • Describe the nature and significance of the advance (e.g. conceptual, technical, clinical) for the field.

      The discovery of a new phosphorylation site in RAD51 (S97) by Aurora A is potentially interesting for the field of the maintenance of genome stability as it could broaden the understanding of how such an important recombinase may be regulating the maintenance of genome integrity throughout the cell cycle. Also, the idea of RAD51 being involved in splicing and mRNA maturation seems very attractive and a very important conceptual advance. However, given the premature status of the text and the figures, the manuscript falls short to show convincing evidence. - Place the work in the context of the existing literature (provide references, where appropriate).

      Many works are highlighting the role of RNA binding proteins as an integral part of the DNA damage response. In addition, a wealth of evidence in the literature suggest that many DNA repair proteins are RNA binding proteins, and that RNA is an important player in the DDR. The possible finding that RAD51 interacts with RNA and localize to nuclear speckles possibly acting in splicing is very interesting and attractive. How is Aurora A involved in this, what is the trigger, and whether RAD51 is binding RNA at these sites is still unclear. - State what audience might be interested in and influenced by the reported findings.

      Labs working in genome integrity mechanisms and the crosstalk between transcription and DNA repair would be interested. - Define your field of expertise with a few keywords to help the authors contextualize your point of view.

      Genome Instability, homologous recombination

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      Summary

      Homologous recombination is central to stabilizing the genome where Rad51 recombinase plays a pivotal role. Authors found unexpected localization of Rad51 to nuclear speckles. This localization is associated with a novel phosphorylation site of Rad51 at Ser97, which is phosphorylated by Aurora A. Because nuclear speckles are where RNA maturation occurs, authors argue for the possible involvement of Rad51 in modulating splicing, a previously unsuspected role for this important recombinase.

      Major points:

      1. The discoveries made in this paper heavily rely on the Phospho-S97 specific antibody (PS97 antibody). The biggest concern of this reviewer is that the validation of this material is not rigid enough. The specificity of this antibody against PS97 is validated only by PS97 peptide competition. The outcome is not convincing either; the PS97 signal seems quite resistant to Si-RAD51 in Fig2A. Furthermore, in Fig2B, PSer97 signal seems rather constant throughout the cell cycle while Rad51 signal fluctuates.

      These observations make this reviewer wonder if the authors are really detecting the phosphorylation of Rad51 with this material (i.e., PSer97 antibody).

      I suggest the authors validate this antibody by doing the following experiments:

      1-1. Do phosphatase treatment to see if the western blotting signal depends on phospho-S97.

      1-2. Do competition experiments using the non-phospho peptide (i.e., the same polypeptides carrying a regular unmodified Ser at 97).

      1-3. Try western blotting using purified Rad51 proteins, one treated with AuroraA and another without the treatment.

      1-4. Do western blotting with cell extract from the cell line producing Rad51-S97A, S97D and compare with wild-type Rad51.

      1. P.10, line 4 The purity of the purified protein should be included (Rad51 and two other mutant proteins) by showing CBB-stained SDS-page gel.
      2. P. 10, line 7 (Fig1C). D-loop assay with Rad51 and its mutants. The actual data should be presented with the actual D-loop formation efficiency. Comparison with wild type value is not enough.
      3. Fig5AB There are lots of PSer97 signals that do not even overlap with DAPI (Fig5A) or Sc35 (Fig5B). How do authors explain that? Also, quantification needs to be done regarding colocalization between PSer97 and Sc35.
      4. Fig5D I do not know what to look for here. At least authors should employ proper negative controls such as siRad51 extract and WCE supplemented with PSer peptides.
      5. Fig6AB Quantification of the results needs to be presented. This reviewer is wondering if there is any explanation regarding the difference in the localization of overproduced HA-Rad51 between HeLa and HCC1806; HA-Rad51 goes into a nucleus in HeLa while it stays in the cytoplasm in HCC1806. Any explanation?

      Minor points:

      1. Please include line numbers.
      2. P.2, line 11 Could you cite the literature showing Rad51 is predominantly cytoplasmic?
      3. P.10, line 15 The authors are not measuring the polymerization rate here. The title is misleading.
      4. Fig2A What do NT and Si-Sc stand for? How come Pser97 signal is resistant to Si-RAD51?
      5. Fig2B P-Rad51/Rad51 ratio graph does not have error bars, making it difficult to assess its reproducibility.

      Referees cross-commenting

      I am pretty much in agreement with the comments/criticisms raised by the other two reviewers.

      Significance

      If Rad51 is indeed involved in RNA maturation, that will be a very novel and exciting discovery. The observations presented in this work, however, seem a bit too inconclusive to support the idea, at least, to this reviewer.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      This paper by Alaouid et al. describes the role of Aurora A-mediated RAD51 phosphorylation in RNA metabolism in the sub-nuclear organelle such as Nuclear Speckles (NS). By the combination of biochemistry and cell biology methods, the authors showed that Aurora A phosphorylates Ser97 of human RAD51 in vitro. The antibody against the phosphor-Ser97 RAD51 seems to recognize NS. Moreover, RAD51 binds not only DNA, but also RNA in vitro. These suggest the role of RAD51 in RNA processing in NS. Moreover, the authors analyzed the biochemical properties of a phosphor-mimetic version of RAD51 (RAD51-S97D) and a phosphor-defective version (RAD51-S97A) and the effect of the over-expression of these mutants in NS dynamics in a cell. The idea of the link to RAD51 in RNA processing in a specific nuclear organelle sounds very interesting and opens a new area of DNA damage response (DDR). However, this paper did not show any functional evidence on the link between RAD51 phosphorylation and RNA processing in NS. Moreover, there are lots of technical issues in the results reported in the manuscript. In some experiments, the authors have to check the reproducibility and be careful about statistics. More importantly, the specificity of the anti-PSer97-RAD51 antibody raised in this study was not properly evaluated in vivo, which makes it hard to interpret the results using this antibody such as western blotting and immuno-fluorescent analyses.

      Major comments:

      1. Because of the poor description, it is hard to evaluate the content. This manuscript needs a more detailed description of the results in a scientifically valid way.
      2. Please describes the basic biochemical activities (ssDNA (Figure, dsDNA binding, and ATP hydrolysis activities) of mutant RAD51 proteins used in this study; RAD51-S97A and RAD51-S97D proteins. See also minor comments in this respect.
      3. It is essential to check the localization signal of PS97 signal in cells with RAD51 depletion by siRNA. Alternatively, the authors used chicken DT40 RAD51 cKO cells (Sonoda et al. EMBO J. 1998) to check the specificity of the antibody (RAD51 is conserved between humans and chicken, and the antibody seems to work in DT40).
      4. Data not shown: Please show the data in Supplemental Figures or deposit it in a public database.
      5. Please cite original references to cite the previous results.
      6. Please make a single composite Figure.

      Minor comments:

      1. Page 4, the second paragraph, line 1: Please add the reference number of Chabot et al. (4).
      2. Page 4, last sentence, DNA/RNA binding activity: For the binding of RAD51 in the presence of ATP, Mg2+ ion or divalent ion is essential. However, there is no description on how much concentration of the divalent ion was used in the assay.
      3. Page 10, the first paragraph, line 4: Please show a Coomassie gel of purified RAD51, RAD51-S97A, and RAD51-S97D proteins.
      4. Figure 1C, D-loop assay: Please show the gel of the products in this assay. It would be nice to show the kinetics of the reactions by these RAD51 mutant proteins. Or the effect of a different RAD51 concentration was tested.
      5. Page 10, the third paragraph, line 1: Please explain what is "BS3"; how this chemical stabilizes the oligomer and the references related to the drug.
      6. P values: Please describe the method to calculate the value in Figure legends.
      7. Figure 1D: Since this assay cannot quantitatively measure the oligomerization status of RAD51, the authors' claims are not convincing. Electron microscopic observation, which is the best, and/or ultra-centrifugation or gel filtration would be recommended to see the difference in the oligomeric status of the RAD51.
      8. Figure 2A: This result is not convincing. Although siRNA for RAD51 largely decreases the amount of RAD51 in cell lysates (bottom, ~80%), a modest decrease of the signal is seen for Phospho-Ser97-RAD51 (top, ~50%)). The authors need to explain this discrepancy. More importantly, this phosphorylation is mediated by Aurora A kinase. It is important to show the signals detected with this antibody decrease in the treatment of the Aurora A inhibitor or siRNA for Aurora A subunits. The inhibition experiments shown in Figure 4A are not convincing because the effect of the inhibitor is very small.
      9. Figure 2B: How did the authors determine each stage of G1, S, G2, and M phases (bottom right graph)? There are no markers of the S phase in western blots such as Cyclin A. Moreover, FACS analysis would be recommended.
      10. Figure 2B, graphs: Please add error bars of quantification of the bands by doing multiple experiments to support the authors' claim on the increase of RAD51 S97 phosphorylation from G1/S to G2/M transition.
      11. Figure 2C top, fractionation assay: Please include a western blot of RAD51 as a control like the ones in the middle.
      12. Figure 2D top: Please include images of RAD51 as a control.
      13. Page 12, first paragraph, line 2: Please show representative images of immunostaining of different cell lines in the Supplementary Figure and quantify the size of the foci. Do show all the data in either main Figures or Supplemental Figures without "data not shown".
      14. Figure 3A: Please explain gammaH2AX blot in either text or legend.
      15. Page 12, the second paragraph, line 8, data not shown: Please show the data in Supplemental Figure.
      16. Page 12, the third paragraph, line 3: The authors need to explain what is the gammaH2AX to readers.
      17. Figure 3B and C: Please check BRCA1/2 or RPA32 (or other DNA repair center markers) localization rather than gammaH2AX for the marker for DNA damage focus. As shown in these figures, gammaH2AX signals spread over the DSB sites, make it difficult to check the colocalization. Why number and intensity of gammaH2AX signals are so different between B and C? In Figure 3C, did the authors use non-treated cells?
      18. Figure 3B, western blots: The top panel is over-exposed.
      19. Page 12 last paragraph-page 13 first paragraph: The short summary is not necessary. These sentences should be moved to Discussion.
      20. Figure 4A: Please include any positive marker for the Aurora-A inhibition such as histone H3S10-phosphorylation.
      21. Figure 4B: Did Aurora A overexpression induce any cell cycle arrest? If it induces G2/M arrest, this increased phosphorylation is simply due to the arrest (in Figure 2B, the authors showed an increase of the phosphorylation in G2/M phase).
      22. Figure 4B pSer97-RAD51/RAD51 ratio: This reviewer is not convincing the quantification. What is the dynamic range of this western? Do they try different cell lysate volumes to adjust constant RAD51 signals to compare the pSer97-RAD51 signals?
      23. Page 13, third paragraph, lines 2-3 and Figure 4B left graph: Is this statistically significant? Please show what statistical method was used here (show it in Legend).
      24. Figure 5B, PlaB treatment: The Images show a decreased focus number of PSer97-RAD51. This is more obvious than the formation of larger foci. The authors need a more precise description of the result in the text.
      25. Figure 5C: Please show the position of the full-length of RAD51 protein by an arrow. The position of RAD51 and pS97 are different-pS97 signal migrates faster than the RAD51 (opposed to the result in Figure 1A).
      26. Figure 5D, IE: What is "NIP"?
      27. Figure 5D, IP: Where is a band of Sc-35? In IP fraction (bottom), there is little band corresponding with the band in lysates. Three thick band are not specific.
      28. Figure 5E, page 14 last sentence, "improved this experiment": Without the quantification, the authors do not conclude this.
      29. Figure 5 experiments: It is not clear why the improvement of Rad51-IP by RNA treatment could explain the role of pSer97-RAD51 points out the RAD51-binding to RNA. Rather the opposite interpretation would be possible. If pSer97-RAD51 is tightly bound to an RNA-containing nuclear structure, the authors may try chromatin fractionation with RNAase treatment.
      30. Figure 6: Please quantify the number and size of Nuclear speckles in different conditions.
      31. Figure label of B and D: "B" and "D" should be placed on the graph for RNA binding.
      32. Page 15-16, DNA/RNA binding assay: Please indicate the length of DNA/RNA in the text. Moreover, it is well established that ATP analogs modulate RAD51-binding to DNA. It is important for the authors to check the effect of ATP and ADP et on DNA/RNA.
      33. Page 16, the second paragraph: In this paragraph, the authors mentioned about "ds"DNA rather than ssDNA described above. Which is true?
      34. Supplemental Figure 1: Please explain what the purple circle means. Moreover, how this result shows the phosphorylation of Ser97. The two spectra look very different. Do they have any other phosphorylation?

      Referees cross-commenting

      I also agree with the other two reviewers. My concern is that we need a re-review of the revised version. I am not familiar with how the Review Common works. Hope that the journal will take care of the re-reviewing after the authors address our concerns on this paper

      Significance

      This paper may offer a new idea in the biology of nuclei by providing a possible link between proteins involved in homologous recombination such as RAD51 and RNA processing in subnuclear compartments, which is regulated by Aurora A-phosphorylation.

      • Describe the nature and significance of the advance (e.g. conceptual, technical, clinical) for the field. This paper might provide a possible link of RAD51 protein involved in homologous recombination with RNA processing in subcompartments in nuclei.
      • Place the work in the context of the existing literature (provide references, where appropriate). The concept on the role of RAD51 in nuclear RNA processing sounds interesting.
      • State what the audience might be interested in and influenced by the reported findings. The results in the paper are of interest to researchers who work on DNA damage response and DNA repair as well as RNA metabolism.
      • Define your field of expertise with a few keywords to help the authors contextualize your point of view. Indicate if there are any parts of the paper that you do not have sufficient expertise to evaluate. I am a researcher on DNA damage response and DNA repair but is not working of RNA biology.
    1. Note: This rebuttal was posted by the corresponding author to Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Reply to the reviewers

      The authors do not wish to provide a response at this time.

    2. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #3

      Evidence, reproducibility and clarity

      This manuscript reports the interaction between poorly characterized FAM104 proteins to ubiquitin-dependent segregase VCP. VCP functions in protein and organelle quality control as well as in extraction of ubiquitylated proteins from chromatin to regulate DNA repair, replication and transcription. In addition, VCP mutations are causative for several human neurological disorders. The authors demonstrate that FAM104 proteins promote the nuclear localization of VCP and that their loss causes impaired growth and hypersensitivity to chemical inhibition of VCP. They show that FAM104 proteins bind to VCP directly via a non-canonical helical motif and model the interaction with AlphaFold Miner, which allowed the identification of critical amino acids that mediate the interaction, which was then validated in vivo and in vitro.

      The conclusions are supported with well-designed experiments and data of high quality, manuscript is written in a clear and precise way.

      Minor points

      • P3: The authors write that mutations in VCP are causative for cancer. This should be rephrased.
      • P3: I would suggest to add a reference to the new study that also shows that VCP is also exploited by bacteria rand not only viruses (
      • Could the authors better illustrate the difference between FAM104A and B, and provide some explanations of why A seems to interact better with VCP compared to B. Is it just matter of higher expression of FAM104A in the cells where the interaction has been tested?
      • The authors should quantify the IF results in Figure 4 and include the quantification in the main figure
      • UBXN2B interaction with FAM104A was found in HT affinity-MS (Huttlin et al) and Y2H (Luck et al) studies. Can the authors validate this interaction of UBXN2B with FAM104 proteins? This would help to understand whether FAM104 interacts mainly with nuclear adaptors.

      Significance

      The results presented in this manuscript will be of interest to the borad field of protein quality control and lay the foundation to study the functions of FAM104 proteins in chromatin-associated degradation.

    3. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #2

      Evidence, reproducibility and clarity

      In their manuscript ‚FAM104 proteins promote the nuclear localization of p97/VCP' Maria Körner, Susanne Meyer, and coauthors describe the identification of FAM104 proteins as cofactors of p97/VCP, a central factor in ubiquitin-mediated cellular proteolytic processes. Function as p97/VCP cofactors has not previously been clearly attributed to FAM104 proteins.

      The initial observation that FAM104 proteins interact directly with p97/VCP in yeast two-hybrid assays is confirmed by in vitro pulldown experiments with recombinantly purified proteins as well as in lysates of cultured mammalian cells. Using truncated proteins and structure predictions, the interaction interface of p97/VCP and FAM104 proteins is further narrowed down to single amino acids of a characteristic C-terminal alpha helix in FAM104. Overall, these interaction studies are technically sound and include meaningful control conditions to postulate FAM104 proteins as p97/VCP binders. Subsequent functional analyses using colocalization studies and cellular fractionation suggest that FAM104 proteins determine the nuclear/chromatin-associated fraction of p97/VCP. Based on this observation, the authors speculate that FAM104 proteins are of particular importance given the established nuclear/chromatin-associated processes involving p97/VCP activity. This hypothesis is supported by the observation that FAM104A knockout cells exhibit an impaired growth phenotype that is exacerbated in the presence of a p97/VCP inhibitor and in combination with a DNA damage trigger.

      Points of concern:

      1. The authors hypothesize that FAM104 proteins enhance the nuclear/chromatin-associated function of p97/VCP by sequestering it from the cytosol into nuclear/chromatin. In the corresponding experiments, overexpression of FAM104 species (Figures 4 and 5) in otherwise unperturbed cells is used. Because recruitment of p97/VCP to client proteins is thought to depend in large part on ubiquitylation, it is unclear how overexpression of FAM104 is sufficient to enhance nuclear/chromatin localization of VCP. Is nuclear/chromatin localization accompanied by changes in ubiquitylation and/or turnover of the corresponding proteins? In other words, does enhanced localization also correlate with increased activity, or could the enhanced nuclear/chromatin association also be explained by inhibited/captured p97/VCP?
      2. The authors link the function of FAM104 proteins in nuclear targeting of p97/VCP to the absence of a unique NLS peptide. Therefore, it would be interesting to determine whether the appearance of FAM104 proteins at the evolutionary level correlates with the strength/presence of NLS peptides in p97/VCP and/or its cofactors UFD1/NPL4/FAF1/UBXN3. Do FAM104 proteins compensate for the loss of NLS peptides in p97/cofactor complexes?
      3. Re 2) It remains unclear whether FAM104 proteins are responsible for the mere sequestration of p97/VCP in the nucleus or whether FAM104 proteins also contribute to process/client specificity in other ways. In this context, the authors could investigate a possible compensation of the reduced nuclear targeting of p97/VCP in FAM104 knock-out cells by fusion with an efficient cNLS peptide. Does this compensate for both nuclear/chromatin localization and growth/drug sensitivity?
      4. Re 3) How does overexpression of FAM104 alter drug sensitivity compared to knock-out cell lines (Figure 7)?
      5. Is there experimental evidence on how FAM104 proteins can bind p97/VCP to chromatin in this context and the proposed targeting of p97/VCP to the nucleus/chromatin? Does FAM104 mRNA/protein expression increase when p97/VCP-mediated processes are disrupted (e.g., in the presence of p97/VCP inhibition or DNA damage)? Are FAM104 protein levels stabilized under these conditions? Are FAM104 proteins differentially regulated (e.g., in terms of localization) under these conditions? Figure 3A suggests that FAM104 proteins may have a different function in relation to p97/VCP protein levels: FAM104A iso1/2 have lower p97/VCP protein levels than FAM104A iso5 and B iso3. The authors suggest that this is due to the solubility of p97/VCP. It should be clarified whether lower solubility equates to increased chromatin association.
      6. It remains unclear whether a FAM104-dependent shift in nuclear/chromatin-associated p97/VCP could also be a secondary compensatory effect versus functional impairment in FAM104 overexpression/depletion. The authors might include this in their discussion.

      Significance

      In summary, the author's conclusion that FAM104 proteins represent a previously underappreciated class of p97/VCP cofactors is well supported. Given the versatile and important role of p97/VCP and cellular protein homeostasis pathways, this finding is of interest to a broad audience. However, the functional role of FAM104 proteins in p97/VCP biology remains unclear. Therefore, the authors need to further elaborate the physiological contribution of FAM104 proteins to p97/VCP function in additional experiments. The suggestions are largely based on modifications of experiments already performed in this manuscript.

    4. Note: This preprint has been reviewed by subject experts for Review Commons. Content has not been altered except for formatting.

      Learn more at Review Commons


      Referee #1

      Evidence, reproducibility and clarity

      Summary:

      The manuscript submitted by Korner et al. presents data regarding the interaction of p97 with FAM104 protein family. This part of the manuscript is performed by Y2H as well as in-vitro and in-vivo pull-down assays. After molecular mapping and characterizations, the authors continue and address the role of FAM104-p97 interaction on nuclear localization of p97 as well cell health in respect to p97 dependent activity. Provide a short summary of the findings and key conclusions (including methodology and model system(s) where appropriate). Please place your comments about significance in section 2.

      Major comments:

      • Are the claims and the conclusions supported by the data or do they require additional experiments or analyses to support them?

      Claims concerning the mapping of FAM104 and p97 (Figures 1&2) are generally well concluded. Yet, minor issues concerning Fig2D (FAM104Aiso1 cdel26) as well as Figure 2E (p97-deltaN pull down) lack of interaction-are not supported by the presented data (both show weak interactions). Claims concerning nuclear/cytosol p97 distribution impact upon FAM104 manipulations (over-expression or KO) need to be further evaluated by additional methodologies. For example, the distribution impact using the FAM104 mutants in 4B should be evaluated by cell fractionation experiments (as performed in figure 5). Cell fractionation performed for FAM104A isoforms 1&2 should be performed on isoforms 5&3, the fact that they are expressed at lower levels has no impact, as the evaluation is on p97 and they were able to show in figure 3A an impact on p97 levels. Impact on distribution performed in Figure 6 using FAM104 KO cells should also include cell fractionation experiments in order to enable clear conclusion regarding FAM104 impact on p97 nuclear distribution.

      Also, statistics presented are somewhat problematic at several points. In figure S4C the ** difference between vector and deNLS mutant make no sense (I think they should have been non-significant). Figure 7 make no sense to compare WT and KO cells (in panels b&C) if their original growth was different. One should compare the differences in respect to the drug concentration in each cell type. Also, it may be useful for statistical purposes to evaluate cell numbers rather than growth% and this may enable to obtain better statistical significances. - Are the suggested experiments realistic in terms of time and resources? It would help if you could add an estimated time investment for substantial experiments.

      The suggested experiments are all in reasonable time frame. - Are the data and the methods presented in such a way that they can be reproduced?

      Y2H is not explained at all in methods, furthermore, it would be useful to present in a table the entire list of p97 interactome obtained in this screen. - Are the experiments adequately replicated and statistical analysis adequate?

      See comments above

      Minor comments:

      • Specific experimental issues that are easily addressable.
      • Are prior studies referenced appropriately?

      Previous reports regarding FAM104 interaction with p97 have been reported in two papers (PMID 32296183 sup. Table9 therein and PMID 32814053 S2 therein) this has not been stated at all. Furthermore, no data concerning previous knowledge of FAM104 is referred to in the introduction. - Are the text and figures clear and accurate?

      The text is written well and one can easily follow<br /> - Do you have suggestions that would help the authors improve the presentation of their data and conclusions?

      See major suggestion above

      Referee cross-commenting

      It seems reviewer #2 concerns are also situated close to our comments regarding nuclear function of FAM104 on p97 function. Reviewers 3 comment regarding UBXN2B possible tertiary complex with p97 and FAM104 should be attempted as it would help put p97 function in a slightly more specific context

      Significance

      Provide contextual information to readers (editors and researchers) about the novelty of the study, its value for the field and the communities that might be interested.

      The following aspects are important:

      • General assessment: provide a summary of the strengths and limitations of the study. What are the strongest and most important aspects? What aspects of the study should be improved or could be developed? The authors deal with a specific interaction of p97 with FAM104 protein family. While this interaction has been previously reported, their mapping of domains required for interaction is new. Conclusions regarding the additional binding partners of the FAM104-p97 complex would require additional double affinity and mass-spectrometry identifications (as well as possible substrates identification).
      • Advance: compare the study to the closest related results in the literature or highlight results reported for the first time to your knowledge; does the study extend the knowledge in the field and in which way? Describe the nature of the advance and the resulting insights (for example: conceptual, technical, clinical, mechanistic, functional,...). The study advances the repertoire of p97 adaptors and interacting domains.
      • Audience: describe the type of audience ("specialized", "broad", "basic research", "translational/clinical", etc...) that will be interested or influenced by this research; how will this research be used by others; will it be of interest beyond the specific field? The suitable audience would be p97 basic researchers
      • Please define your field of expertise with a few keywords to help the authors contextualize your point of view. Indicate if there are any parts of the paper that you do not have sufficient expertise to evaluate. p97 role in protein quality control and cellular homeostasis.