Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

A MORC-driven transcriptional switch controls Toxoplasma developmental trajectories and sexual commitment

Abstract

Toxoplasma gondii has a complex life cycle that is typified by asexual development that takes place in vertebrates, and sexual reproduction, which occurs exclusively in felids and is therefore less studied. The developmental transitions rely on changes in the patterns of gene expression, and recent studies have assigned roles for chromatin shapers, including histone modifications, in establishing specific epigenetic programs for each given stage. Here, we identified the T. gondii microrchidia (MORC) protein as an upstream transcriptional repressor of sexual commitment. MORC, in a complex with Apetala 2 (AP2) transcription factors, was shown to recruit the histone deacetylase HDAC3, thereby impeding the accessibility of chromatin at the genes that are exclusively expressed during sexual stages. We found that MORC-depleted cells underwent marked transcriptional changes, resulting in the expression of a specific repertoire of genes, and revealing a shift from asexual proliferation to sexual differentiation. MORC acts as a master regulator that directs the hierarchical expression of secondary AP2 transcription factors, and these transcription factors potentially contribute to the unidirectionality of the life cycle. Thus, MORC plays a cardinal role in the T. gondii life cycle, and its conditional depletion offers a method to study the sexual development of the parasite in vitro, and is proposed as an alternative to the requirement of T. gondii infections in cats.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: MORC, a highly conserved ATPase protein, interacts with HDAC3 in T. gondii.
Fig. 2: Recruitment of HDAC3 to chromatin is mediated by MORC.
Fig. 3: Depletion of MORC phenocopies the inhibition of HDAC3 by inducing the expression of sexual-stage-specific genes.
Fig. 4: Degradation of MORC, similar to the inhibition of HDAC3, induces the expression of merozoite-restricted transcripts.
Fig. 5: MORC regulates developmental transitions at multiple checkpoints.
Fig. 6: MORC guides developmental trajectories recruiting downstream regulating pathways.

Similar content being viewed by others

Data availability

The data supporting the findings of this study are available from the corresponding author on reasonable request. The RNA-seq demultiplexed FASTQ files and gene-wise quantifications have been deposited in the NCBI Gene Expression Omnibus (GEO) and are accessible under accession number GSE136123. The ChIP–seq data have been deposited in the GEO under accession number GSE136060. The MS proteomics data have been deposited in the ProteomeXchange Consortium through the PRIDE partner repository with the dataset identifiers PXD016846 (MORC interactome) and PXD016845 (proteome-wide analyses). Source data for Figs. 1, 2 and 5 are presented with this paper.

References

  1. Dubey, J. P., Lindsay, D. S. & Speer, C. A. Structures of Toxoplasma gondii tachyzoites, bradyzoites, and sporozoites and biology and development of tissue cysts. Clin. Microbiol. Rev. 11, 267–299 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Balaji, S. Discovery of the principal specific transcription factors of Apicomplexa and their implication for the evolution of the AP2-integrase DNA binding domains. Nucleic Acids Res. 33, 3994–4006 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Kim, K. The epigenome, cell cycle, and development in Toxoplasma. Annu. Rev. Microbiol. 72, 479–499 (2018).

    CAS  PubMed  Google Scholar 

  4. Saksouk, N. et al. Histone-modifying complexes regulate gene expression pertinent to the differentiation of the protozoan parasite Toxoplasma gondii. Mol. Cell. Biol. 25, 10301–10314 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Bougdour, A. et al. Drug inhibition of HDAC3 and epigenetic control of differentiation in Apicomplexa parasites. J. Exp. Med. 206, 953–966 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Moissiard, G. et al. MORC family ATPases required for heterochromatin condensation and gene silencing. Science 336, 1448–1451 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Lorković, Z. J. MORC proteins and epigenetic regulation. Plant Signal. Behav. 7, 1561–1565 (2012).

    PubMed  PubMed Central  Google Scholar 

  8. Harris, C. J. et al. Arabidopsis AtMORC4 and AtMORC7 form nuclear bodies and repress a large number of protein-coding genes. PLoS Genet. 12, e1005998 (2016).

    PubMed  PubMed Central  Google Scholar 

  9. Weiser, N. E. et al. MORC-1 integrates nuclear RNAi and transgenerational chromatin architecture to promote germline immortality. Dev. Cell 41, 408–423 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Pastor, W. A. et al. MORC1 represses transposable elements in the mouse male germline. Nat. Commun. 5, 5795 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Tchasovnikarova, I. A. et al. Hyperactivation of HUSH complex function by Charcot–Marie–Tooth disease mutation in MORC2. Nat. Genet. 49, 1035–1044 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Kim, H. et al. The gene-silencing protein MORC-1 topologically entraps DNA and forms multimeric assemblies to cause DNA compaction. Mol. Cell 75, 700–710 (2019).

    CAS  PubMed  Google Scholar 

  13. Inoue, N. New gene family defined by MORC, a nuclear protein required for mouse spermatogenesis. Hum. Mol. Genet. 8, 1201–1207 (1999).

    CAS  PubMed  Google Scholar 

  14. Iyer, L. M., Abhiman, S. & Aravind, L. MutL homologs in restriction-modification systems and the origin of eukaryotic MORC ATPases. Biol. Direct 3, 8 (2008).

    PubMed  PubMed Central  Google Scholar 

  15. Iyer, L. M., Anantharaman, V., Wolf, M. Y. & Aravind, L. Comparative genomics of transcription factors and chromatin proteins in parasitic protists and other eukaryotes. Int. J. Parasitol. 38, 1–31 (2008).

    CAS  PubMed  Google Scholar 

  16. Andrews, F. H. et al. Multivalent chromatin engagement and inter-domain crosstalk regulate MORC3 ATPase. Cell Rep. 16, 3195–3207 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Sindikubwabo, F. et al. Modifications at K31 on the lateral surface of histone H4 contribute to genome structure and expression in apicomplexan parasites. eLife 6, e29391 (2017).

    PubMed  PubMed Central  Google Scholar 

  18. Sidik, S. M. et al. A genome-wide CRISPR screen in Toxoplasma identifies essential apicomplexan genes. Cell 166, 1423–1435 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Brown, K. M., Long, S. & Sibley, L. D. Plasma membrane association by N-acylation governs PKG function in Toxoplasma gondii. mBio 8, e00375-17 (2017).

    PubMed  PubMed Central  Google Scholar 

  20. Pittman, K. J., Aliota, M. T. & Knoll, L. J. Dual transcriptional profiling of mice and Toxoplasma gondii during acute and chronic infection. BMC Genom. 15, 806 (2014).

    Google Scholar 

  21. Hehl, A. B. et al. Asexual expansion of Toxoplasma gondii merozoites is distinct from tachyzoites and entails expression of non-overlapping gene families to attach, invade, and replicate within feline enterocytes. BMC Genom. 16, 66 (2015).

    Google Scholar 

  22. Ramakrishnan, C. et al. An experimental genetically attenuated live vaccine to prevent transmission of Toxoplasma gondii by cats. Sci. Rep. 9, 1474 (2019).

    PubMed  PubMed Central  Google Scholar 

  23. Fritz, H. M. et al. Transcriptomic analysis of Toxoplasma development reveals many novel functions and structures specific to sporozoites and oocysts. PLoS ONE 7, e29998 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Behnke, M. S., Zhang, T. P., Dubey, J. P. & Sibley, L. D. Toxoplasma gondii merozoite gene expression analysis with comparison to the life cycle discloses a unique expression state during enteric development. BMC Genom. 15, 350 (2014).

    Google Scholar 

  25. Ramakrishnan, C., Walker, R. A., Eichenberger, R. M., Hehl, A. B. & Smith, N. C. The merozoite-specific protein, TgGRA11B, identified as a component of the Toxoplasma gondii parasitophorous vacuole in a tachyzoite expression model. Int. J. Parasitol. 47, 597–600 (2017).

    CAS  PubMed  Google Scholar 

  26. Smith, E. F. PF16 encodes a protein with armadillo repeats and localizes to a single microtubule of the central apparatus in Chlamydomonas flagella. J. Cell Biol. 132, 359–370 (1996).

    CAS  PubMed  Google Scholar 

  27. Sapiro, R. et al. Male infertility, impaired sperm motility, and hydrocephalus in mice deficient in sperm-associated antigen 6. Mol. Cell. Biol. 22, 6298–6305 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Straschil, U. et al. The armadillo repeat protein PF16 is essential for flagellar structure and function in plasmodium male gametes. PLoS ONE 5, e12901 (2010).

    PubMed  PubMed Central  Google Scholar 

  29. Clark, T. HAP2/GCS1: mounting evidence of our true biological EVE? PLoS Biol. 16, e3000007 (2018).

    PubMed  PubMed Central  Google Scholar 

  30. Mori, T., Hirai, M., Kuroiwa, T. & Miyagishima, S. The functional domain of GCS1-based gamete fusion resides in the amino terminus in plant and parasite species. PLoS ONE 5, e15957 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Liu, Y. et al. The conserved plant sterility gene HAP2 functions after attachment of fusogenic membranes in Chlamydomonas and Plasmodium gametes. Genes Dev. 22, 1051–1068 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Angrisano, F. et al. Targeting the conserved fusion loop of HAP2 inhibits the transmission of Plasmodium berghei and falciparum. Cell Rep. 21, 2868–2878 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Gondim, L. F. P. et al. Characterization of an IgG monoclonal antibody targeted to both tissue cyst and sporocyst walls of Toxoplasma gondii. Exp. Parasitol. 163, 46–56 (2016).

    CAS  PubMed  Google Scholar 

  34. Radke, J. R. et al. Identification of a sporozoite-specific member of the Toxoplasma SAG superfamily via genetic complementation: T. gondii sporozoite developmental antigens. Mol. Microbiol. 52, 93–105 (2004).

    CAS  PubMed  Google Scholar 

  35. Tomita, T. et al. The Toxoplasma gondii cyst wall protein CST1 is critical for cyst wall integrity and promotes bradyzoite persistence. PLoS Pathog. 9, e1003823 (2013).

    PubMed  PubMed Central  Google Scholar 

  36. Radke, J. B. et al. ApiAP2 transcription factor restricts development of the Toxoplasma tissue cyst. Proc. Natl Acad. Sci. USA 110, 6871–6876 (2013).

    CAS  PubMed  Google Scholar 

  37. Poran, A. et al. Single-cell RNA sequencing reveals a signature of sexual commitment in malaria parasites. Nature 551, 95–99 (2017).

    PubMed  PubMed Central  Google Scholar 

  38. Jiang, C. & Pugh, B. F. Nucleosome positioning and gene regulation: advances through genomics. Nat. Rev. Genet. 10, 161–172 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Radman-Livaja, M. & Rando, O. J. Nucleosome positioning: how is it established, and why does it matter? Dev. Biol. 339, 258–266 (2010).

    CAS  PubMed  Google Scholar 

  40. Teif, V. B. et al. Genome-wide nucleosome positioning during embryonic stem cell development. Nat. Struct. Mol. Biol. 19, 1185–1192 (2012).

    CAS  PubMed  Google Scholar 

  41. Li, D.-Q. et al. MORC2 signaling integrates phosphorylation-dependent, ATPase-coupled chromatin remodeling during the DNA damage response. Cell Rep. 2, 1657–1669 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Wadman, M. Closure of U.S. Toxoplasma lab draws ire. Science 364, 109 (2019).

    CAS  PubMed  Google Scholar 

  43. Bougdour, A. et al. Host cell subversion by Toxoplasma GRA16, an exported dense granule protein that targets the host cell nucleus and alters gene expression. Cell Host Microbe 13, 489–500 (2013).

    CAS  PubMed  Google Scholar 

  44. Sangaré, L. O. et al. Unconventional endosome-like compartment and retromer complex in Toxoplasma gondii govern parasite integrity and host infection. Nat. Commun. 7, 11191 (2016).

    PubMed  PubMed Central  Google Scholar 

  45. Salvetti, A. et al. Nuclear functions of nucleolin through global proteomics and interactomic approaches. J. Proteome Res. 15, 1659–1669 (2016).

    CAS  PubMed  Google Scholar 

  46. Gajria, B. et al. ToxoDB: an integrated Toxoplasma gondii database resource. Nucleic Acids Res. 36, D553–D556 (2007).

    PubMed  PubMed Central  Google Scholar 

  47. Tyanova, S., Temu, T. & Cox, J. The MaxQuant computational platform for mass spectrometry-based shotgun proteomics. Nat. Protoc. 11, 2301–2319 (2016).

    CAS  PubMed  Google Scholar 

  48. Wieczorek, S. et al. DAPAR & ProStaR: software to perform statistical analyses in quantitative discovery proteomics. Bioinformatics 33, 135–136 (2017).

    CAS  PubMed  Google Scholar 

  49. Chen, K. et al. DANPOS: dynamic analysis of nucleosome position and occupancy by sequencing. Genome Res. 23, 341–351 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank the developers of the ToxoDB.org Genome Resource. ToxoDB and EuPathDB are part of the National Institutes of Health/National Institutes of Allergy and Infectious Diseases (NIH/NIAID)-funded Bioinformatics Resource Center. We also thank G. Communie for his help computing Danpos software. This work was supported by the Laboratoire d’Excellence (LabEx) ParaFrap (grant no. ANR-11-LABX-0024), the Agence Nationale pour la Recherche (Project HostQuest, grant no. ANR-18-CE15-0023 and ProFI grant no. ANR-10-INSB-08-01), the European Research Council (ERC Consolidator grant no. 614880 Hosting TOXO to M.-A.H.) and Fondation pour la Recherche Médicale (FRM grant no. FDT201904008364 to D.C.F.).

Author information

Authors and Affiliations

Authors

Contributions

M.-A.H. supervised the research. M.-A.H. and D.C.F. generated the genetic tools, performed and analysed all of the genome-wide studies (RNA-seq, ChIP–seq and MNase assay), conducted most of biochemical experiments and the immunofluorescence assays. C.D., D.C. and F.S. generated biochemical reagents. L.B. and P.-J.D.B. performed MS analysis. Y.C. guided the MS experiments. C.S. performed structural modelling and analysed the MNase assay. A.B. computed and analysed the RNA-seq data. P.O. and M.B. analysed the ChIP–seq data. M.-A.H. and D.C.F. wrote the paper.

Corresponding author

Correspondence to Mohamed-Ali Hakimi.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Domain architectures and expression levels during the life cycle of MORC and partners.

a, Representative domain architectures of T. gondii MORC and partners identified by mass spectrometry-based proteomic are shown. Domains were predicted by SMART and PFAM: ELM2 (Egl-27 and MTA1 homology 2), PHD (plant homeodomain), and AP2 (APETALA2) b, Heatmap representation of MORC and partners gene expression in different life cycle stages (source: ToxoDB). Gene expression values were mean log2 transformed and median centered for clustering. Transcriptomic data from tachyzoite, merozoite, longitudinal studies on enteroepithelial stages (EES1 to EES5), immature (day 0), maturing (day 4) and mature (day 10) stages of oocyst development and cysts from chronically infected mice were used. c, Mapping of domains and identified phosphorylation and ubiquitination sites detected by mass-spectrometry (source: ToxoDB).

Extended Data Fig. 2 mAID-based MORC inducible KD system successfully ablated its expression on the protein post-translational level in both type I (RH) and type II (Pru) strains.

a, Auxin-inducible degradation system for controlling protein stability in T. gondii. We first engineered RH and Pru strains ectopically expressing the plant auxin receptor called transport inhibitor response 1 (TIR1). We chose the UPRT locus to integrate TIR1 under the control of a promoter allowing a mild expression of the protein tagged to Ty. We utilized a mini-AID (mAID) tagging LIC system for conditional MORC depletion. The resulting cell lines are referred to as MORC-KD hereafter. Conditional depletion of MORC-mAID-HA is reliant on auxin (IAA), TIR1, and the proteasome. b, Depletion of MORC-mAID-HA, upon addition of IAA for 16 hours, was measured by IFA in cells infected with Pru MORC KD. Fixed and permeabilized parasites were probed with MORC-mAID-HA (red) and GAP45 (green). MORC-depleted cells displayed strong inhibition of proliferation when compared to untreated cells. Yet, this growth defect phenotype was completely reversed upon IAA washout, indicating that MORC depletion while impeding the cell cycle progression does not kill the parasites. Graph on the right, in situ quantification of nuclear MORC-mAID-HA using IFA. The horizontal bars represent the mean ± s.d. of the nuclear MORC intensity from three independent experiments (n = 50 nuclei per dot). The p-values were calculated using one-way ANOVA. Scale bar, 10 μm. c, in situ quantification (related to Fig. 2c) of nuclear MORC-mAID-HA using IFA in RH MORC KD as described above. d, Smoothed and background-subtracted tag density profiles are displayed over representative regions of Chr. V (top) and X (bottom). The ChIP-seq profiles were obtained with antibodies directed against pan-acetylated histone H4, HDAC3 and HA (MORC detection) from chromatin sampled from a Pru MORC KD strain left untreated (UT) and or treated with IAA for 24 hours. The experiment was repeated independently twice with similar results.

Extended Data Fig. 3 MORC protein depletion, and HDAC3 inhibition induces gene expression.

a, Volcano plots showing gene expression differences identified from comparison of UT versus FR235222 in Pru MORC KD (left graph) and of UT versus IAA (24 hours) in RH MORC KD (right graph) (n=8914 genes, Supplementary Table 3). The orange and green dots indicate the number of significantly up- and down-regulated genes, respectively, using adjusted p < 0.01 (Bonferroni-corrected) and ± 2-fold change as the cut-off threshold corresponding to each comparison. X-axis showing log2 fold change, Y-axis showing -log10(p value). b, Volcano plots illustrating changes in protein expression between UT versus FR235222 in Pru MORC KD (left graph, n=2566 proteins, Benjamini-Hochberg FDR = 1 %, p-value ≤ 0.00794) and of UT versus IAA (24 hours) in RH MORC KD (right graph, n=2139 proteins, Benjamini-Hochberg FDR = 1.01%, p-value ≤ 0.00501). Overexpressed proteins upon MORC depletion are indicated in red and under-expressed ones in blue. c, Pru strains were engineered to endogenously epitope tag with HA two MORC-regulated genes, TGME49_207210 and TGME49_216140. Expression was monitored following HDAC3 chemical inactivation by FR235222 and HA staining. Scale bars, 10 μm. Experiments were conducted more than three times and representative images are displayed.

Extended Data Fig. 4 MORC alongside HDAC3 represses the expression of sexual stages-specific genes.

a, A type II (ME49) strain was engineered to endogenously expressed a HA tag version of the merozoite gene TGME49_238915. Genetic inactivation of HDAC3, MORC and AP2-XII-1 promote TGME49_243940 protein expression (in red). The efficiency of genetic disruption in Cas9-expressing parasites was monitored by cas9-GFP expression (in green). Scale bar, 5 μm. b, Expression of TGME49_243940 was monitored following HDAC3 chemical inactivation by FR235222 and HA staining (in red). Scale bar, 10 μm. c, Heatmap showing hierarchical clustering analysis of selected MORC-regulated genes from cluster 1 through different strain/induction combinations, including the abundance of their transcripts in the various stages of development, namely tachyzoite, bradyzoite/cyst, merozoite, enteroepithelial stages (EES) and oocyst stages. The color scale bar indicates log2 fold changes. d, Expression of TGME49_316130 was monitored following HDAC3 chemical inactivation by FR235222 and HA staining (in red). Scale bar, 10 μm. Experiments in a, b and d were conducted more than three times and representative images are displayed.

Extended Data Fig. 5 MORC KD derepresses proteins involved in merozoite and fertilization.

a, Smoothed and background-subtracted tag density profiles are displayed over Chr. VI. The ChIP-seq profiles were obtained with antibodies directed against various histone marks, HDAC3 and HA (MORC detection) from chromatin sampled from a Pru MORC KD strain left untreated (UT) and or treated with IAA for 24 hours. RNA-seq data through different strain/induction combinations are shown in black. The y-axis depicts read density from ChIP-seq and RPKM values for RNA-seq data. The merozoite-specific gene TGME49_238915 is shown in dark blue. b, Pru strain was engineered to endogenously expressed HA epitope-tagged TGME49_238915. Expression was monitored following HDAC3 chemical inactivation by FR235222 and HA staining. Scale bar, 10 μm. Experiment was conducted more than three times and representative images are displayed. c, Density profiles are displayed on Chromosome V at the TgHAP2 (TGME49_285940) locus. d, Density profiles are displayed on Chromosome X around a merozoite-specific cluster of tandemly repeated genes (SRS48 family). (a, c-d) The experiment was repeated independently twice with similar results. Chromosomal positions are indicated on x-axis.

Extended Data Fig. 6 MORC regulates microgamete related genes including those coding for flagella components.

Heatmap representation of RNA-seq data portraying the number of genes involved in microgamete biology, alongside their levels of expression upon MORC depletion/HDAC3 inhibition through different strain/induction combinations. The abundance of their transcripts in the various stages of development, namely tachyzoite, bradyzoite/cyst, merozoite, enteroepithelial stages (EES) and oocyst stages was displayed. The color scale bar indicates log2 fold changes. The genes were divided in sets grouping together genes involved in axonemal cytoskeleton, and genes that harbor potential domains involved in intraflagellar transportation.

Extended Data Fig. 7 MORC depletion induces the expression of genes involved in oocyst wall formation and coding for sporozoite-specific markers.

a, Density profiles are displayed on Chromosome VIIb at the SporoSAG/SRS28 (TGME49_258550) locus, encoding for the hallmark surface antigen in sporozoite. The ChIP-seq profiles were obtained with antibodies directed against various histone PTMs, HDAC3 and HA (MORC detection) from chromatin sampled from a Pru MORC KD strain left UT and or treated with IAA for 24 hours. RNA-seq data through different strain/induction combinations are shown in black. The y-axis depicts read density from ChIP-seq and RPKM values for RNA-seq data. b, Density profiles are displayed on Chromosome XI at the TgHowp1 (TGME49_316890) locus. TgHOWP1 is a representative example of proteins expressed in the early stages of wall formation as revealed by the accumulation of its transcripts in both late developmental EES5 and in immature (D0) oocyst (ToxoDB data). c, Density profiles are displayed on Chromosome VIIb at the TGME49_262110 locus. Expression of TGME49_262110 was monitored following HDAC3 chemical inactivation by FR235222 and HA staining (in green). Experiments was conducted more than three times and representative images are displayed. Scale bar, 5 μm. (a-c) The experiment was repeated independently twice with similar results. Chromosomal positions are indicated on x-axis.

Extended Data Fig. 8 MORC depletion derepresses partly bradyzoite-specific genes.

a, Heatmap showing hierarchical clustering analysis of selected bradyzoite-specific and MORC-regulated genes from cluster 1 through different strain/induction combinations. The abundance of their transcripts in the various stages of development, namely tachyzoite, bradyzoite/cyst, merozoite, enteroepithelial stages (EES) and oocyst stages was displayed. The color scale bar indicates log2 fold changes. b, Density profiles are displayed on Chromosome VIIb at the Bag1 (TGME49_259020) locus, encoding for the hallmark surface antigen in bradyzoite. The ChIP-seq profiles were obtained with antibodies directed against various histone PTMs, HDAC3 and HA (MORC detection) from chromatin sampled from a Pru MORC KD strain left UT and or treated with IAA for 24 hours. RNA-seq data through different strain/induction combinations are shown in black. The y-axis depicts read density from ChIP-seq and RPKM values for RNA-seq data. The experiment was repeated independently twice with similar results. Chromosomal positions are indicated on x-axis.

Extended Data Fig. 9 MORC depletion causes indirect repression of tachyzoites specific genes.

a-b, Density profiles are displayed at the tachyzoite-specific ROP16 (a) and RON2 (b) genes. The ChIP-seq profiles were obtained with antibodies directed against various histone marks, HDAC3 and HA (MORC detection) from chromatin sampled from a Pru MORC KD strain left UT and or treated with IAA for 24 hours. RNA-seq data through different strain/induction combinations are shown in black. The y-axis depicts read density from ChIP-seq and RPKM values for RNA-seq data. The experiment was repeated independently twice with similar results. Chromosomal positions are indicated on x-axis.

Extended Data Fig. 10 MORC regulates the expression of AP2IV-3, a Plasmodium falciparum AP2-G homologous protein.

a, Density profiles are displayed at the merozoite-specific AP2IV-3 gene. The ChIP-seq profiles were obtained with antibodies directed against various histone marks, HDAC3 and HA (MORC detection) from chromatin sampled from a Pru MORC KD strain left UT and or treated with IAA for 24 hours. RNA-seq data through different strain/induction combinations are shown in black. The y-axis depicts read density from ChIP-seq and RPKM values for RNA-seq data. The experiment was repeated independently twice with similar results. Chromosomal positions are indicated on x-axis. b, Representative domain architectures of T. gondii AP2IV-3 and P. falciparum AP2-G, displaying the approximate AP2 domain position within the protein sequence. c, Protein alignment of the AP2IV-3 and AP2-G respective AP2 domains, with the amino acid homology shown in red.

Supplementary information

Supplementary Information

Supplementary discussion, references and Figs. 1–7, and legends for Supplementary Tables 1–5.

Reporting Summary

Supplementary Table 1

MS-based characterization of the MORC interactome. Proteins eluted from Flag-tagged MORC co-immunoprecipitations performed from the RH and Pru strains were identified using MS-based proteomics. Toxoplasma proteins reproducibly identified from both strains with at least two peptides are presented. The identities of the proteins (accession number on ToxoDB, gene name and description) are indicated in columns A, B and C. Their molecular weights (MW) are noted in column D. The number of peptides identified for each protein is given in columns E and F for the RH and Pru strains, respectively. The abundance ranks of each protein in each interactome were deduced from extracted intensity-based absolute quantification values and are presented in columns G and H for the RH and Pru strains, respectively.

Supplementary Table 2

Description of T. gondii strains, plasmids, DNA synthesis and primers. List of T. gondii parasite lines as well as the plasmids used in this study. The primers and DNA synthesis construct used in this work are also shown in the table.

Supplementary Table 3

MORC- and HDAC3-regulated transcriptomes. Gene expression profiles in HFF showing that T. gondii genes are differentially regulated after MORC depletion or HDAC3 inhibition through different strain/induction combinations. RPKM values are shown for the indicated samples. The mean values for each gene/condition are indicated as well the replicate (REP) values. Differences in gene expression were identified from a comparison of (1) Pruku80 MORC KD left untreated (UT) or treated with IAA (24 h or 48 h); (2) RHku80 MORC KD left untreated or treated with IAA (7 h or 24 h); and (3) Pruku80 left unstimulated or treated FR235222 (FR) for 7 h or 18 h. Gene name (column A) and ToxoDB description (column B) are indicated.

Supplementary Table 4

MORC-regulated and stage-specific gene subsets as clustered in the study. A list of T. gondii genes related to Fig. 3b (columns A, B, C and D), Supplementary Fig. 3 (columns E and F) and Fig. 3g,h (columns G, H, I and J).

Supplementary Table 5

MS-based proteome-wide analyses following MORC KD or HDAC3 inhibition by FR235222. The proteomes from type I or type II strains infecting HFF cells were analysed by label-free quantitative proteomics (n = 8,322 predicted proteins). The quantification of proteins was based on razor + unique peptides (column E). Statistical significance was tested using limma. Differentially abundant proteins (marked by ‘1’ in column I) were defined by a log2-transformed fold change of ≥ 0.8 or ≤ −0.8 (column G) and P values (column H) were allowed to reach an FDR of ~1% according to the Benjamini–Hochberg estimator. Sheet 1: quantitative analysis of proteomes from Pru MORC KD parasites and Pru untreated (UT) parasites. P value cut-off = 0.00501. The red cells (column I) correspond to proteins that were enriched in MORC KD parasites compared with untreated parasites, whereas the blue cells (column I) correspond to proteins that were enriched in untreated parasites compared with MORC KD parasites. Sheet 2: quantitative analysis of proteomes from RH MORC KD parasites and RH untreated (UT) parasites. P value cut-off = 0.00631. The red cells (column I) correspond to proteins that were enriched in MORC KD parasites compared with untreated parasites, whereas the blue cells (column I) correspond to proteins that were enriched in untreated parasites compared with MORC KD parasites. Sheet 3: quantitative analysis of proteomes from Pru parasites that were either treated or not treated with FR235222. P value cut-off = 0.00794. The red cells (column I) correspond to proteins that were enriched in FR235222-treated parasites compared with untreated parasites, whereas the blue cells (column I) correspond to proteins that were enriched in untreated parasites compared with FR235222-treated parasites.

Source data

Source Data Fig. 1

Unprocessed western blots and a silver-stained gel.

Source Data Fig. 2

Unprocessed western blots.

Source Data Fig. 5

Unprocessed western blots.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Farhat, D.C., Swale, C., Dard, C. et al. A MORC-driven transcriptional switch controls Toxoplasma developmental trajectories and sexual commitment. Nat Microbiol 5, 570–583 (2020). https://doi.org/10.1038/s41564-020-0674-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-020-0674-4

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing