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:

N6-methyladenosine in poly(A) tails stabilize VSG transcripts

Abstract

RNA modifications are important regulators of gene expression1. In Trypanosoma brucei, transcription is polycistronic and thus most regulation happens post-transcriptionally2. N6-methyladenosine (m6A) has been detected in this parasite, but its function remains unknown3. Here we found that m6A is enriched in 342 transcripts using RNA immunoprecipitation, with an enrichment in transcripts encoding variant surface glycoproteins (VSGs). Approximately 50% of the m6A is located in the poly(A) tail of the actively expressed VSG transcripts. m6A residues are removed from the VSG poly(A) tail before deadenylation and mRNA degradation. Computational analysis revealed an association between m6A in the poly(A) tail and a 16-mer motif in the 3′ untranslated region of VSG genes. Using genetic tools, we show that the 16-mer motif acts as a cis-acting motif that is required for inclusion of m6A in the poly(A) tail. Removal of this motif from the 3′ untranslated region of VSG genes results in poly(A) tails lacking m6A, rapid deadenylation and mRNA degradation. To our knowledge, this is the first identification of an RNA modification in the poly(A) tail of any eukaryote, uncovering a post-transcriptional mechanism of gene regulation.

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: m6A is present in the poly(A) tail of VSG mRNA and other transcripts.
Fig. 2: m6A is removed from VSG mRNA before its degradation.
Fig. 3: Inclusion of m6A in the VSG poly(A) tail depends on de novo transcription.
Fig. 4: Conserved VSG 16-mer motif is required for inclusion of m6A in an adjacent poly(A) tail.
Fig. 5: VSG 16-mer motif inhibits CAF1 and poly(A) tail deadenylation.

Similar content being viewed by others

Data availability

RNA sequencing datasets of m6A-RIP experiments are deposited in the Sequence Read Archive under the accession code: PRJNA786734. The following panels have associated raw figures: 1b, c, e, h, 2b, c, 3c, d, f, g, i, k, 4b, e, f, 5b–d, f, Extended Data Figs. 2d, e, 3c, 4, 5a–c, 7. The following public databases were used; Trypanosome genome database (TryTrypDB): https://tritrypdb.org/tritrypdb/app; and RNA modifications database (MODOMICS): http://genesilico.pl/modomics/. All data are available on request from the corresponding author. Source data are provided with this paper.

References

  1. Zaccara, S., Ries, R. J. & Jaffrey, S. R. Reading, writing and erasing mRNA methylation. Nat. Rev. Mol. Cell Biol. 20, 608–624 (2019).

    Article  CAS  PubMed  Google Scholar 

  2. Horn, D. Antigenic variation in African trypanosomes. Mol. Biochem. Parasitol. 195, 123–129 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Liu, L. et al. Differential m6A methylomes between two major life stages allows potential regulations in Trypanosoma brucei. Biochem. Biophys. Res. Commun. 508, 1286–1290 (2019).

    Article  CAS  PubMed  Google Scholar 

  4. Franco, J. R., Simarro, P. P., Diarra, A., Ruiz-Postigo, J. A. & Jannin, J. G. The journey towards elimination of gambiense human African trypanosomiasis: not far, nor easy. Parasitology 141, 748–760 (2014).

    Article  CAS  PubMed  Google Scholar 

  5. Cross, G. A., Kim, H. S. & Wickstead, B. Capturing the variant surface glycoprotein repertoire (the VSGnome) of Trypanosoma brucei Lister 427. Mol. Biochem. Parasitol. 195, 59–73 (2014).

    Article  CAS  PubMed  Google Scholar 

  6. Cross, G. A. Identification, purification and properties of clone-specific glycoprotein antigens constituting the surface coat of Trypanosoma brucei. Parasitology 71, 393–417 (1975).

    Article  CAS  PubMed  Google Scholar 

  7. Nilsson, D. et al. Spliced leader trapping reveals widespread alternative splicing patterns in the highly dynamic transcriptome of Trypanosoma brucei. PLoS Pathog. 6, e1001037 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  8. Kraus, A. J., Brink, B. G. & Siegel, T. N. Efficient and specific oligo-based depletion of rRNA. Sci Rep. 9, 12281 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  9. Gunzl, A. et al. RNA polymerase I transcribes procyclin genes and variant surface glycoprotein gene expression sites in Trypanosoma brucei. Eukaryot. Cell 2, 542–551 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Fadda, A. et al. Transcriptome-wide analysis of trypanosome mRNA decay reveals complex degradation kinetics and suggests a role for co-transcriptional degradation in determining mRNA levels. Mol. Microbiol. 94, 307–326 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Berberof, M. et al. The 3′-terminal region of the mRNAs for VSG and procyclin can confer stage specificity to gene expression in Trypanosoma brucei. EMBO J. 14, 2925–2934 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Ridewood, S. et al. The role of genomic location and flanking 3′UTR in the generation of functional levels of variant surface glycoprotein in Trypanosoma brucei. Mol. Microbiol. 106, 614–634 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Roditi, I. et al. Procyclin gene expression and loss of the variant surface glycoprotein during differentiation of Trypanosoma brucei. J. Cell Biol. 108, 737–746 (1989).

    Article  CAS  PubMed  Google Scholar 

  14. Ehlers, B., Czichos, J. & Overath, P. RNA turnover in Trypanosoma brucei. Mol. Cell. Biol. 7, 1242–1249 (1987).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Matthews, K. R. The developmental cell biology of Trypanosoma brucei. J. Cell Sci. 118, 283–290 (2005).

    Article  CAS  PubMed  Google Scholar 

  16. Wei, C. M., Gershowitz, A. & Moss, B. Methylated nucleotides block 5′ terminus of HeLa cell messenger RNA. Cell 4, 379–386 (1975).

    Article  CAS  PubMed  Google Scholar 

  17. Perry, R. P., Kelley, D. E., Friderici, K. & Rottman, F. The methylated constituents of L cell messenger RNA: evidence for an unusual cluster at the 5′ terminus. Cell 4, 387–394 (1975).

    Article  CAS  PubMed  Google Scholar 

  18. Meyer, K. D. et al. Comprehensive analysis of mRNA methylation reveals enrichment in 3′ UTRs and near stop codons. Cell 149, 1635–1646 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Dominissini, D. et al. Topology of the human and mouse m6A RNA methylomes revealed by m6A-seq. Nature 485, 201–206 (2012).

    Article  CAS  ADS  PubMed  Google Scholar 

  20. Jia, G. et al. N6-methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO. Nat. Chem. Biol. 7, 885–887 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Zheng, G. et al. ALKBH5 is a mammalian RNA demethylase that impacts RNA metabolism and mouse fertility. Mol. Cell 49, 18–29 (2013).

    Article  CAS  PubMed  Google Scholar 

  22. Freistadt, M. S., Cross, G. A. & Robertson, H. D. Discontinuously synthesized mRNA from Trypanosoma brucei contains the highly methylated 5′ cap structure, m7GpppA*A*C(2′-O)mU*A. J. Biol. Chem. 263, 15071–15075 (1988).

    Article  CAS  PubMed  Google Scholar 

  23. Bangs, J. D., Crain, P. F., Hashizume, T., McCloskey, J. A. & Boothroyd, J. C. Mass spectrometry of mRNA cap 4 from trypanosomatids reveals two novel nucleosides. J. Biol. Chem. 267, 9805–9815 (1992).

    Article  CAS  PubMed  Google Scholar 

  24. Perry, K. L., Watkins, K. P. & Agabian, N. Trypanosome mRNAs have unusual “cap 4” structures acquired by addition of a spliced leader. Proc. Natl Acad. Sci. USA 84, 8190–8194 (1987).

    Article  CAS  ADS  PubMed  PubMed Central  Google Scholar 

  25. Hauenschild, R. et al. The reverse transcription signature of N-1-methyladenosine in RNA-seq is sequence dependent. Nucleic Acids Res. 43, 9950–9964 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Boccaletto, P. et al. MODOMICS: a database of RNA modification pathways. 2017 update. Nucleic Acids Res. 46, D303–D307 (2018).

    Article  CAS  PubMed  Google Scholar 

  27. Molinie, B. et al. m6A-LAIC-seq reveals the census and complexity of the m6A epitranscriptome. Nat. Methods 13, 692–698 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Xiang, J. F. et al. N6-methyladenosines modulate A-to-I RNA editing. Mol. Cell 69, 126–135.e6 (2018).

    Article  CAS  PubMed  Google Scholar 

  29. Clayton, C. E. Regulation of gene expression in trypanosomatids: living with polycistronic transcription. Open Biol. 9, 190072 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Schibler, U., Kelley, D. E. & Perry, R. P. Comparison of methylated sequences in messenger RNA and heterogeneous nuclear RNA from mouse L cells. J. Mol. Biol. 115, 695–714 (1977).

    Article  CAS  PubMed  Google Scholar 

  31. Linder, B. et al. Single-nucleotide-resolution mapping of m6A and m6Am throughout the transcriptome. Nat. Methods 12, 767–772 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Hoek, M., Zanders, T. & Cross, G. A. M. Trypanosoma brucei expression-site-associated-gene 8 protein interacts with a Pumilio family protein. Mol. Biochem. Parasitol. 120, 269–283 (2002).

    Article  CAS  PubMed  Google Scholar 

  33. Iyer, L. M., Zhang, D. & Aravind, L. Adenine methylation in eukaryotes: apprehending the complex evolutionary history and functional potential of an epigenetic modification. Bioessays 38, 27–40 (2016).

    Article  CAS  PubMed  Google Scholar 

  34. Domingo-Sananes, M. R., Szoor, B., Ferguson, M. A., Urbaniak, M. D. & Matthews, K. R. Molecular control of irreversible bistability during trypanosome developmental commitment. J. Cell Biol. 211, 455–468 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Melo do Nascimento, L. et al. Functional insights from a surface antigen mRNA-bound proteome. eLife 10, e68136 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Fadda, A., Färber, V., Droll, D. & Clayton, C. The roles of 3′-exoribonucleases and the exosome in trypanosome mRNA degradation. RNA 19, 937–947 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Schwede, A. et al. A role for Caf1 in mRNA deadenylation and decay in trypanosomes and human cells. Nucleic Acids Res. 36, 3374–3388 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Gallie, D. R. The role of the poly(A) binding protein in the assembly of the Cap-binding complex during translation initiation in plants. Translation (Austin) 2, e959378 (2014).

    Google Scholar 

  39. Garneau, N. L., Wilusz, J. & Wilusz, C. J. The highways and byways of mRNA decay. Nat. Rev. Mol. Cell Biol. 8, 113–126 (2007).

    Article  CAS  PubMed  Google Scholar 

  40. Chen, C. Y. & Shyu, A. B. Mechanisms of deadenylation-dependent decay. Wiley Interdiscip. Rev. RNA 2, 167–183 (2011).

    Article  CAS  PubMed  Google Scholar 

  41. Decker, C. J. & Parker, R. A turnover pathway for both stable and unstable mRNAs in yeast: evidence for a requirement for deadenylation. Genes Dev. 7, 1632–1643 (1993).

    Article  CAS  PubMed  Google Scholar 

  42. Tang, T. T. L., Stowell, J. A. W., Hill, C. H. & Passmore, L. A. The intrinsic structure of poly(A) RNA determines the specificity of Pan2 and Caf1 deadenylases. Nat. Struct. Mol. Biol. 26, 433–442 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We are grateful to support from the Howard Hughes Medical Institute International Early Career Scientist Program (55007419), a European Molecular Biology Organization Installation grant (2151) and La Caixa Foundation (HR20-00361). This work was also partially supported by the ONEIDA project (LISBOA-01-0145-FEDER-016417) co-funded by Fundos Europeus Estruturais e de Investimento (FEEI) from ‘Programa Operacional Regional Lisboa 2020’ and by national funds from Fundação para a Ciência e a Tecnologia (FCT). S.R.J. was supported by NIH (R35 NS111631). Researchers were funded by individual fellowships from FCT (PD/BD/105838/2014 to I.J.V., 2020.06827.BD to L.S., SFRH/BD/80718/2011 to F.A.-B., PD/BD/138891/2018 to A.T. and CEECIND/03322/2018 to L.M.F.); a Novartis Foundation for Biomedical-Biological research to J.P.d.M.; a Human Frontier Science Programme long-term postdoctoral fellowship to M.D.N. (LT000047/2019); a Marie Skłodowska-Curie Individual Standard European Fellowship to S.S.P. (grant no. 839960); the GlycoPar Marie Curie Initial Training Network (GA 608295) to J.A.R. We thank J. Thomas-Oates (University of York, Centre of Excellence in Mass Spectrometry, Department of Chemistry) for the mass spectrometry analysis. The York Centre of Excellence in Mass Spectrometry was created thanks to a major capital investment through Science City York, supported by Yorkshire Forward with funds from the Northern Way Initiative, and subsequent support from EPSRC (EP/K039660/1 and EP/M028127/1). We also thank A. Temudo, A. Nascimento and A. Lima for bioimaging assistance; the laboratories of A. Tomás and J. Kelly for providing RNA from Leishmania and T. cruzi, respectively; and A. Pena and members of the Figueiredo and Jaffrey laboratories for helpful discussions.

Author information

Authors and Affiliations

Authors

Contributions

I.J.V., J.P.d.M., L.S., M.D.N., A.T., S.S.P., A.H.M., E.B. and J.A.R. performed the experiments. I.J.V., J.P.d.M., L.S., M.D.N., A.T., S.S.P., A.H.M., E.B., J.A.R., F.A.-B., S.R.J. and L.M.F. planned the experiments and analysed the data. I.J.V., F.A.-B., J.A.R., S.R.J. and L.M.F. conceived the study. I.J.V., S.R.J. and L.M.F. wrote the manuscript, with contributions of all remaining authors.

Corresponding author

Correspondence to Luisa M. Figueiredo.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature thanks David Horn and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

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

Extended data figures and tables

Extended Data Fig. 1 Chemical structures of RNA modifications found in T. brucei.

a, 15 modifications were detected in poly(A)-enriched RNA (mRNA). b, 19 modifications were not detected in poly(A)-depleted RNA. All modifications were detected in total RNA. Structures were obtained from the database Modomics (http://genesilico.pl/modomics/).

Extended Data Fig. 2 Detection of m6A in T. brucei by mass-spectrometry.

a, b, Chromatograms obtained by LC-MS/MS analysis of a N6-methyladenosine standard and three RNA samples of T. brucei bloodstream form (BSF, a) or insect procyclic stage (PCF, b): total RNA, RNA enriched with poly(T)-beads (i.e., poly(A)-enriched RNA) and RNA that did not bind to polyT-beads (i.e., poly(A)-depleted RNA). c, Chromatogram obtained by LC-MS/MS analysis of a N1-methyladenosine standard with the 282->150 mass transition. The m1A peak is detected at 6.5 min. d, Standard curve of m6A. Increasing quantities of commercially synthesized m6A were loaded on the HPLC column and the area under the chromatogram peak was measured. e, Quantification of the m6A/A(%) in the mRNA in a second independent experiment. n = 5 mRNA samples. (see also Source Data of Extended Figures).

Source data

Extended Data Fig. 3 m6A detection in T. brucei by immunoblotting.

a, Specificity of anti-m6A antibody. Oligonucleotides containing either m6A (positive control), unmodified adenosine or m1A (negative controls) were manually spotted in the membrane, which was incubated with anti-m6A antibody. The antibody specifically recognized the oligos with m6A, while exhibiting low cross-reactivity to the oligos with only unmodified adenosine or containing m1A. b, m6A signal intensity in the immunoblot, measured by Image J, in the whole lane containing the poly(A)-enriched RNA of bloodstream forms. c, The intensity of the ~1.8 kb band was divided by the signal intensity of the entire lane. n = 5 biological replicates. d, m6A immunoblotting of RNA samples from two stages of T. brucei life cycle. Samples (from left to right): total RNA (Total), Poly(A)-enriched (A+) RNA and Poly(A)-depleted (A-) RNA from mammalian BSF and insect PCF. The last lane contains total mouse liver RNA (Mouse). 2 µg of total RNA, 2 µg of poly(A)-depleted RNA and 100 ng of poly(A)-enriched RNA was loaded per lane. rRNA was detected by staining RNA with methylene blue to confirm equal loading between total and poly(A)-depleted fractions. As expected rRNA is undetectable in the poly(A)-enriched fraction. (see also Supplementary Fig. 1 and Source Data of Extended Figures).

Source data

Extended Data Fig. 4 Poly(A) tail length, m6A and mRNA levels during VSG turnover.

Levels of m6A (immunoblot), length of VSG poly(A) tail (RNase H – northern blot) and levels of VSG mRNA (northern blot) after transcription halt by ActD. Signals were normalized to time point 0hr. The pattern observed is consistent with Fig. 2b. Two-way ANOVA with sidak correction for multiple test. (****P< 0.0001,*P = 0.0190 in mRNA vs m6A in 15 min, ***P = 0.0004 in poly(A) vs m6A in 15 min, ***P = 0.0001 in mRNA vs m6A in 120 min, *P = 0.0136 in poly(A) vs m6A in 240 min). n = 4 biological samples for mRNA and m6A levels, n = 3 biological samples for poly(A) tail length. Data are mean ± s.d. (see also Source Data of Extended Figures).

Source data

Extended Data Fig. 5 Subcellular distribution of m6A in bloodstream form parasites.

a, Proportion of m6A signal in nucleus and cytoplasm. Data are mean ± s.e.m. n = 4 independent experiments. b, Quantification of mean fluorescence intensity (MFI) levels of m6A in five independent replicates in three different conditions: untreated BSF, nuclease P1 (NP1)-treated BSF, and actinomycin D (ActD)-treated BSF. Raw MFIs were obtained, the average of the untreated BSF equalled to 100%, and all other values normalized to 100%. Data are mean ± s.e.m. c, Distribution of VSG2 mRNA in the nucleus and cytoplasm in single marker (SM) cell line (single VSG expression) and in the clones that express a second reporter VSG (6 clones of DE1 express VSG117 containing a WT 16-mer motif, 7 clones of DE2 express VSG117 containing a mutagenized 16-mer motif). VSG mRNA was quantified by FISH. Nucleus was delimited by Hoechst staining. Total signal was set as 100% and the nucleus and cytoplasm represented as percentage of total signal. Error bars represent s.d. d, Microscopic observation of nuclei purified after fractionation protocol. Nuclei were stained with Hoechst. The nuclear purification was compared with initial lysates and with the cytoplasmic fraction. Scale bars, 10 µm; n = 1 independent experiment. e, Western blot of subcellular fractions (total lysate, nuclear and cytoplasmic fractions) using antibodies against a nuclear protein (histone H2A; custom rabbit polyclonal 1:5000) and a cytoplasmic protein (β-tubulin; mouse monoclonal KMX-1 1:1000). For each sample, we loaded a protein equivalent to the same amount of cells. n = 3 independent experiments. (see also Supplementary Fig. 1 and Source Data of Extended Figures).

Source data

Extended Data Fig. 6 VSG double-expressor (DE) cell lines immunoblotting.

a, Overexposure of full immunoblot shown in Fig. 4c. Three independent 16-merWT clones and three independent 16-merMUT are shown (C1–C6). Note that with this exposure, most intense bands are saturated. The purpose of this high exposure is to observe the region of blot corresponding to the VSG117 transcript. No VSG117 band is observed in the 16-merMUT clones. It is also possible to observe a weak VSG2 band in the VSG2 single expressor lane and in the 16-merMUT clones, which correspond to incomplete RNase H digestion of VSG2 transcript. n = 3 independent clones for each genotype (C1–C6). b, Schematics of VSG double-expressor (DE) cell-lines DE3 and DE4. VSG8 was inserted in the active bloodstream expression site, which naturally contains VSG2 at the telomeric end. In DE3, VSG8 contains its endogenous 3’UTR with the conserved 16-mer motif (sequence in blue). In DE4, the 16-mer motif of VSG8 was scrambled (sequence in orange). c, m6A immunoblot of mRNA from DE3 and DE4 cell-lines, in which CAF1 was further depleted by RNAi by adding Tetracycline (Tet). RNase H digestion of VSG2 mRNA was used to resolve VSG2 and VSG8 transcripts. Two independent DE3 clones and two independent DE4 clones are shown (C1–C4), each with (+) or without (−) CAF1 downregulation. (see also Supplementary Fig. 1).

Extended Data Fig. 7 CAF1 depletion.

CAF1 transcript levels measured by RT-qPCR in CAF1 RNAi cell-line used in Fig. 5c–f. CAF1 downregulation was induced by adding tetracycline (Tet) to the medium. Unpaired two tailed t-test (**** P > 0.0001). Data are mean ± s.d. n = 3 independent clones.

Source data

Extended Data Table 1 Mass-spectrometry features of 34 nucleoside modifications found in T. brucei
Extended Data Table 2 List of oligonucleotides
Extended Data Table 3 Statistical parameters of time course experiments

Supplementary information

Supplementary Information

This file contains Supplementary Methods; Supplementary Fig. 1; Supplementary Table 1 legend and Supplementary References.

Reporting Summary

Peer Review File

Supplementary Table 1

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Viegas, I.J., de Macedo, J.P., Serra, L. et al. N6-methyladenosine in poly(A) tails stabilize VSG transcripts. Nature 604, 362–370 (2022). https://doi.org/10.1038/s41586-022-04544-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-04544-0

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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