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.

  • Letter
  • Published:

Genome-wide probing of RNA structure reveals active unfolding of mRNA structures in vivo

Subjects

Abstract

RNA has a dual role as an informational molecule and a direct effector of biological tasks. The latter function is enabled by RNA’s ability to adopt complex secondary and tertiary folds and thus has motivated extensive computational1,2 and experimental3,4,5,6,7,8 efforts for determining RNA structures. Existing approaches for evaluating RNA structure have been largely limited to in vitro systems, yet the thermodynamic forces which drive RNA folding in vitro may not be sufficient to predict stable RNA structures in vivo5. Indeed, the presence of RNA-binding proteins and ATP-dependent helicases can influence which structures are present inside cells. Here we present an approach for globally monitoring RNA structure in native conditions in vivo with single-nucleotide precision. This method is based on in vivo modification with dimethyl sulphate (DMS), which reacts with unpaired adenine and cytosine residues9, followed by deep sequencing to monitor modifications. Our data from yeast and mammalian cells are in excellent agreement with known messenger RNA structures and with the high-resolution crystal structure of the Saccharomyces cerevisiae ribosome10. Comparison between in vivo and in vitro data reveals that in rapidly dividing cells there are vastly fewer structured mRNA regions in vivo than in vitro. Even thermostable RNA structures are often denatured in cells, highlighting the importance of cellular processes in regulating RNA structure. Indeed, analysis of mRNA structure under ATP-depleted conditions in yeast shows that energy-dependent processes strongly contribute to the predominantly unfolded state of mRNAs inside cells. Our studies broadly enable the functional analysis of physiological RNA structures and reveal that, in contrast to the Anfinsen view of protein folding whereby the structure formed is the most thermodynamically favourable, thermodynamics have an incomplete role in determining mRNA structure in vivo.

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

Figure 1: Using dimethyl sulphate for RNA structure probing by deep sequencing.
Figure 2: Comparison of DMS-seq data to known RNA structures.
Figure 3: Identification of structured mRNA regions reveals far less structure in vivo than in vitro.
Figure 4: Factors affecting the difference between mRNA structure in vivo and in vitro.

Similar content being viewed by others

Accession codes

Accessions

Gene Expression Omnibus

Data deposits

All data are deposited in Gene Expression Omnibus (accession number GSE45803).

References

  1. Gruber, A. R., Neuböck, R., Hofacker, I. L. & Washietl, S. The RNAz web server: prediction of thermodynamically stable and evolutionarily conserved RNA structures. Nucleic Acids Res. 35, W335–W338 (2007)

    Article  PubMed  PubMed Central  Google Scholar 

  2. Ouyang, Z., Snyder, M. P. & Chang, H. Y. SeqFold: Genome-scale reconstruction of RNA secondary structure integrating high-throughput sequencing data. Genome Res. 23, 377–387 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Kertesz, M. et al. Genome-wide measurement of RNA secondary structure in yeast. Nature 467, 103–107 (2010)

    Article  ADS  CAS  PubMed  Google Scholar 

  4. Underwood, J. G. et al. FragSeq: transcriptome-wide RNA structure probing using high-throughput sequencing. Nature Methods 7, 995–1001 (2010)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Spitale, R. C. et al. RNA SHAPE analysis in living cells. Nature Chem. Biol. 9, 18–20 (2013)

    Article  CAS  Google Scholar 

  6. Lucks, J. B. et al. Multiplexed RNA structure characterization with selective 2′-hydroxyl acylation analyzed by primer extension sequencing (SHAPE-Seq). Proc. Natl Acad. Sci. USA 108, 11063–11068 (2011)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  7. Deigan, K. E., Li, T. W., Mathews, D. H. & Weeks, K. M. Accurate SHAPE-directed RNA structure determination. Proc. Natl Acad. Sci. USA 106, 97–102 (2009)

    Article  ADS  CAS  PubMed  Google Scholar 

  8. Wan, Y. et al. Genome-wide measurement of RNA folding energies. Mol. Cell 48, 169–181 (2012)

    Article  PubMed  PubMed Central  Google Scholar 

  9. Wells, S. E., Hughes, J. M., Igel, A. H. & Ares, M., Jr Use of dimethyl sulfate to probe RNA structure in vivo. Methods Enzymol. 318, 479–493 (2000)

    Article  CAS  PubMed  Google Scholar 

  10. Ben-Shem, A. et al. The structure of the eukaryotic ribosome at 3.0 Å resolution. Science 334, 1524–1529 (2011)

    Article  ADS  CAS  PubMed  Google Scholar 

  11. Ziehler, W. A. & Engelke, D. R. Probing RNA structure with chemical reagents and enzymes. Curr. Protoc. Nucleic Acid Chem. 6, Unit 6.1 (2001)

    Google Scholar 

  12. Zaug, A. J. & Cech, T. R. Analysis of the structure of Tetrahymena nuclear RNAs in vivo: telomerase RNA, the self-splicing rRNA intron, and U2 snRNA. RNA 1, 363–374 (1995)

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Cordero, P., Kladwang, W., VanLang, C. C. & Das, R. Quantitative dimethyl sulfate mapping for automated RNA secondary structure inference. Biochemistry 51, 7037–7039 (2012)

    Article  CAS  PubMed  Google Scholar 

  14. Inoue, T. & Cech, T. R. Secondary structure of the circular form of the Tetrahymena rRNA intervening sequence: a technique for RNA structure analysis using chemical probes and reverse transcriptase. Proc. Natl Acad. Sci. USA 82, 648–652 (1985)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  15. Gonzalez, T. N., Sidrauski, C., Dörfler, S. & Walter, P. Mechanism of non-spliceosomal mRNA splicing in the unfolded protein response pathway. EMBO J. 18, 3119–3132 (1999)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Badis, G., Saveanu, C., Fromont-Racine, M. & Jacquier, A. Targeted mRNA degradation by deadenylation-independent decapping. Mol. Cell 15, 5–15 (2004)

    Article  CAS  PubMed  Google Scholar 

  17. Chartrand, P., Meng, X. H., Singer, R. H. & Long, R. M. Structural elements required for the localization of ASH1 mRNA and of a green fluorescent protein reporter particle in vivo. Curr. Biol. 9, 333–338 (1999)

    Article  CAS  PubMed  Google Scholar 

  18. Rüegsegger, U., Leber, J. H. & Walter, P. Block of HAC1 mRNA translation by long-range base pairing is released by cytoplasmic splicing upon induction of the unfolded protein response. Cell 107, 103–114 (2001)

    Article  PubMed  Google Scholar 

  19. Zuker, M. Mfold web server for nucleic acid folding and hybridization prediction. Nucleic Acids Res. 31, 3406–3415 (2003)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Wittebolle, L. et al. Initial community evenness favours functionality under selective stress. Nature 458, 623–626 (2009)

    Article  ADS  CAS  PubMed  Google Scholar 

  21. Ingolia, N. T., Ghaemmaghami, S., Newman, J. R. S. & Weissman, J. S. Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science 324, 218–223 (2009)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  22. Herschlag, D. RNA chaperones and the RNA folding problem. J. Biol. Chem. 270, 20871–20874 (1995)

    Article  CAS  PubMed  Google Scholar 

  23. Li, F. et al. Global analysis of RNA secondary structure in two metazoans. Cell Rep 1, 69–82 (2012)

    Article  CAS  PubMed  Google Scholar 

  24. Stade, K. et al. Exportin 1 (Crm1p) is an essential nuclear export factor. Cell 90, 1041–1050 (1997)

    Article  CAS  PubMed  Google Scholar 

  25. Kretz, M. et al. Control of somatic tissue differentiation by the long non-coding RNA TINCR. Nature 493, 231–235 (2013)

    Article  ADS  CAS  PubMed  Google Scholar 

  26. Memczak, S. et al. Circular RNAs are a large class of animal RNAs with regulatory potency. Nature (2013)

  27. Tan, X. et al. Tiling genomes of pathogenic viruses identifies potent antiviral shRNAs and reveals a role for secondary structure in shRNA efficacy. Proc. Natl Acad. Sci. USA 109, 869–874 (2012)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  28. Tang, J. & Breaker, R. R. Structural diversity of self-cleaving ribozymes. Proc. Natl Acad. Sci. USA 97, 5784–5789 (2000)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  29. Li, S. & Breaker, R. R. Eukaryotic TPP riboswitch regulation of alternative splicing involving long-distance base pairing. Nucleic Acids Res. 41, 3022–3031 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Meyer, M., Plass, M., Pérez-Valle, J., Eyras, E. & Vilardell, J. Deciphering 3′ss selection in the yeast genome reveals an RNA thermosensor that mediates alternative splicing. Mol. Cell 43, 1033–1039 (2011)

    Article  CAS  PubMed  Google Scholar 

  31. Kortmann, J., Sczodrok, S., Rinnenthal, J., Schwalbe, H. & Narberhaus, F. Translation on demand by a simple RNA-based thermosensor. Nucleic Acids Res. 39, 2855–2868 (2011)

    Article  CAS  PubMed  Google Scholar 

  32. Ingolia, N. T., Brar, G. A., Rouskin, S., McGeachy, A. M. & Weissman, J. S. The ribosome profiling strategy for monitoring translation in vivo by deep sequencing of ribosome-protected mRNA fragments. Nature Protocols 7, 1534–1550 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Ingolia, N. T., Ghaemmaghami, S., Newman, J. R. S. & Weissman, J. S. Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science 324, 218–223 (2009)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  34. Hastings, C., Mosteller, F., Tukey, J. W. & Winsor, C. P. Low moments for small samples: a comparative study of order statistics. Ann. Math. Stat. 18, 413–426 (1947)

    Article  MathSciNet  MATH  Google Scholar 

  35. Lee, B. & Richards, F. M. The interpretation of protein structures: estimation of static accessibility. J. Mol. Biol. 55, 379–400. IN3–IN4 (1971)

  36. Zuker, M. Mfold web server for nucleic acid folding and hybridization prediction. Nucleic Acids Res. 31, 3406–3415 (2003)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Nagalakshmi, U. et al. The transcriptional landscape of the yeast genome defined by RNA sequencing. Science 320, 1344–1349 (2008)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  38. Blanchette, M. et al. Aligning multiple genomic sequences with the threaded blockset aligner. Genome Res. 14, 708–715 (2004)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Bernhart, S. H., Hofacker, I. L., Will, S., Gruber, A. R. & Stadler, P. F. RNAalifold: improved consensus structure prediction for RNA alignments. BMC Bioinformatics 9, 474 (2008)

    Article  PubMed  PubMed Central  Google Scholar 

  40. Washietl, S., Hofacker, I. L. & Stadler, P. F. Fast and reliable prediction of noncoding RNAs. Proc. Natl Acad. Sci. USA 102, 2454–2459 (2005)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  41. Gruber, A. R., Findeiß, S., Washietl, S., Hofacker, I. L. & Stadler, P. F. Rnaz 2.0: improved noncoding RNA detection. Pac. Symp. Biocomput. 69–79 (2010)

  42. Brandman, O. et al. A ribosome-bound quality control complex triggers degradation of nascent peptides and signals translation stress. Cell 151, 1042–1054 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank R. Andino, M. Bassik, J. Doudna, J. Dunn, T. Faust, N. Stern-Ginossar, C. Gross, C. Guthrie, N. Ingolia, C. Jan, M. Kampmann, D. Koller, G.W. Li, S. Mortimer, E. Oh, C. Pop and members of the Weissman laboratory for discussions; J. Stewart-Ornstein and O. Brandman for plasmids; C. Chu, N. Ingolia and J. Lund for sequencing help. This research was supported by the Center for RNA Systems Biology (J.S.W.), the Howard Hughes Medical Institute (J.S.W.), and the National Science Foundation (M.Z.).

Author information

Authors and Affiliations

Authors

Contributions

S.R., M.Z. and J.S.W. designed the experiments. S.R. and M.Z. performed the experiments, and S.R. analysed the data. S.W. and M.K. completed the phylogenetic analysis. S.R., M.Z. and J.S.W. drafted and revised the manuscript.

Corresponding author

Correspondence to Jonathan S. Weissman.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 Ribosomal RNA analysis.

a, Histogram of raw counts distribution for denatured and structured 18S rRNA. Log2 (raw counts) for A bases plotted for in vivo and denatured samples. b, ROC curve on the DMS signal for A and C bases from the 25S rRNA in denatured, in vitro, intact ribosomes, and in vivo samples. True positives are defined as bases that are both unpaired and solvent-accessible, and true negatives are defined as bases that are paired. c, DMS signal on all of the 18S rRNA A bases plotted from least to most reactive in the denatured or in vivo samples. The A bases that are false negatives relative to the crystal structure are coloured as black dots on both the denatured and in vivo samples.

Extended Data Figure 2 mRNA structure analysis with different window sizes.

a, b, Scatter plots of Gini index versus r values in replicate samples and for in vivo or in vitro samples relative to denatured sample for mRNA regions with an average of at least 15 counts per position, spanning the sequence of 25 A/C nucleotides (5,000 randomly selected regions are shown) (a) or 100A/C nucleotides (b). Shown are regions spanning validated secondary structures (red dots) and regions from the 18S rRNA (green dots).

Extended Data Figure 3 Agreement of DMS-seq with validated structures in mammalian K562 cells.

Raw DMS counts were normalized to the most reactive nucleotide in the given region. A and C bases were normalized separately. a, The DMS signal is colour coded proportional to intensity and plotted onto the secondary structure model of MSRB1 selenocysteine insertion element, nucleotide 1 corresponds to nucleotide 966 of the transcript. b, TPRC iron recognition element, nucleotide 1 corresponds to nucleotide 3901 of the transcript. c, XBP1 conserved non canonical intron recognized by Ire1, nucleotide 1 corresponds to nucleotide 520 of the transcript.

Extended Data Figure 4 Global mRNA analysis of human foreskin fibroblast cells.

Scatter plots of Gini index versus r values for in vivo or in vitro samples relative to denatured sample for mRNA regions spanning the sequence of 50 A/C nucleotides.

Extended Data Figure 5 Functional verification of novel 5′ UTR structures in vivo.

a, Putative 5′ UTR stems were manipulated in the context of a Venus reporter in vivo. b, PMA1 5′ UTR structure was mutated and compensated twice with Venus reporter, differing in number and character of bases mutated. Mutation location shown in red on schematic. Reported P values relative to wild-type Venus levels, calculated by two-sided t-test (*P < 0.01, **P < 0.001, ***P < 0.0001). For all graphs, Venus signal normalized to cell size before calculating fold change and data presented is from two biological and two technical replicates. Error bars represent s.e.m. c, Secondary structure of functional PMA1 5′ UTR stem, with compensatory mutations (arrows) found in Saccharomyces paradoxus, Saccharomyces mikatae, Saccharomyces kudriavzevii and Saccharomyces bayanus. Raw DMS signal shown below (position 1 = chromosome VII:482745). d, SFT2 5′ UTR structure was mutated and compensated three times in Venus reporter system, differing in number, character and location of bases mutated. Mutation location shown in red on schematic. Stem stability as predicated by mfold. Reported P values relative to wild-type Venus levels, also by two-sided t-test (*P < 0.01, **P < 0.001, ***P < 0.0001). Error bars represent standard deviation. e, Secondary structure of functional SFT2 5′ UTR stem. Position 1 = chromosome II:24023.

Extended Data Figure 6 Functional verification of novel PRC1 3′ UTR structure in vivo.

a, Putative 3′ UTR stems were manipulated in the context of a Venus reporter in vivo, followed by Venus quantification with flow cytometry. b, PRC1 3′ UTR structure was mutated and compensated in Venus reporter system. For all data, reported P values relative to wild-type Venus levels, calculated by two-sided t-test (*P < 0.01, **P < 0.001, ***P < 0.0001). Venus signal was normalized to cell size with fold change reported relative to Venus levels seen with the wild-type stem. All results shown are derived from four measurements: two biological and two technical replicates. Error bars show standard deviation. c, Secondary structure of functional PRC1 3′ UTR stem, shown with raw DMS signal for in vivo and denatured samples. Position 1 = chromosome XIII:863554. d, Weakly structured 3′ UTRs in vivo were tested for function as in b, but reveal little effect when mutated and no evidence for compensation.

Extended Data Figure 7 Global analysis of mRNA structure.

a, In vitro DMS-seq on RNA refolded in different Mg+2 concentrations. b, Metagene plot of the average DMS signal (normalized to denatured control) over 5′ UTR, coding, and 3′ UTR regions. c, Scatter plot of Gini index (calculated over the first 100 A/C bases) of in vivo messages (relative to denatured) versus translation efficiency.

Extended Data Figure 8 In vivo structures forming in ATP depleted conditions.

a, b, Raw DMS counts were normalized to the most reactive nucleotide in a given region. a, The DMS signal is colour coded proportional to intensity and plotted onto the mfold predicted secondary structure model of CBF5, nucleotide 1 corresponds to chromosome XII position 506479. b, TCP1, nucleotide 1 corresponds to chromosome IV position 887991.

Extended Data Figure 9 Analysis of mRNA structure at 10 °C.

a, Histogram of Gini index difference (calculated over 100A or Cs) between 10 °C and wild-type (30 °C) samples. b, Scatter plot of the Gini index differences in ATP depleted or 10 °C yeast relative to wild-type yeast calculated over 50 As or Cs.

Extended Data Table 1 Sequences of functional structure mutations

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Cite this article

Rouskin, S., Zubradt, M., Washietl, S. et al. Genome-wide probing of RNA structure reveals active unfolding of mRNA structures in vivo. Nature 505, 701–705 (2014). https://doi.org/10.1038/nature12894

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature12894

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: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research