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:

Reversible 2′-OH acylation enhances RNA stability

Abstract

The presence of a hydroxyl group at the 2′-position in its ribose makes RNA susceptible to hydrolysis. Stabilization of RNAs for storage, transport and biological application thus remains a serious challenge, particularly for larger RNAs that are not accessible by chemical synthesis. Here we present reversible 2′-OH acylation as a general strategy to preserve RNA of any length or origin. High-yield polyacylation of 2′-hydroxyls (‘cloaking’) by readily accessible acylimidazole reagents effectively shields RNAs from both thermal and enzymatic degradation. Subsequent treatment with water-soluble nucleophilic reagents removes acylation adducts quantitatively (‘uncloaking’) and recovers a remarkably broad range of RNA functions, including reverse transcription, translation and gene editing. Furthermore, we show that certain α-dimethylamino- and α-alkoxy- acyl adducts are spontaneously removed in human cells, restoring messenger RNA translation with extended functional half-lives. These findings support the potential of reversible 2′-acylation as a simple and general molecular solution for enhancing RNA stability and provide mechanistic insights for stabilizing RNA regardless of length or origin.

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: RNA polyacylation (cloaking) inhibits thermal degradation of RNA.
Fig. 2: Structurally diverse acylimidazole reagents enhance RNA stability in solution.
Fig. 3: RNA cloaking suppresses enzymatic RNA degradation by RNases and biofluids.
Fig. 4: Nucleophilic reagents remove 2′-polyacylation to uncloak RNA.
Fig. 5: Nucleophile-promoted RNA uncloaking restores RNA functions.
Fig. 6: Spontaneous RNA uncloaking restores mRNA translation with extended functional half-lives in human cells.

Similar content being viewed by others

Data availability

All relevant data supporting the findings of this study are available within the article and supplementary information. The characterization data of all organic compounds are provided within the supplementary information. All reagents generated in this study are available from the corresponding author upon reasonable request. Data used for this paper are also available via figshare at https://figshare.com/articles/dataset/NCHEM-22030547_source_data/19555132. Source data are provided with this paper.

References

  1. Fabre, A. L., Colotte, M., Luis, A., Tuffet, S. & Bonnet, J. An efficient method for long-term room temperature storage of RNA. Eur. J. Hum. Genet. 22, 379–385 (2014).

    CAS  PubMed  Google Scholar 

  2. Wayment-Steele, H. K. et al. Theoretical basis for stabilizing messenger RNA through secondary structure design. Nucleic Acids Res. 49, 10604–10617 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Kellerman, D. L., York, D. M., Piccirilli, J. A. & Harris, M. E. Altered (transition) states: mechanisms of solution and enzyme catalyzed RNA 2’-O-transphosphorylation. Curr. Opin. Chem. Biol. 21, 96–102 (2014).

    CAS  PubMed  Google Scholar 

  4. Guo, F. et al. Effect of ribose conformation on RNA cleavage via internal transesterification. J. Am. Chem. Soc. 140, 11893–11897 (2018).

    CAS  PubMed  Google Scholar 

  5. Breslow, R. & Chapman, W. H. Jr On the mechanism of action of ribonuclease A: relevance of enzymatic studies with a p-nitrophenylphosphate ester and a thiophosphate ester. Proc. Natl Acad. Sci. USA 93, 10018–10021 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Lee, J. B., Hong, J., Bonner, D. K., Poon, Z. & Hammond, P. T. Self-assembled RNA interference microsponges for efficient siRNA delivery. Nat. Mater. 11, 316–322 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Leppek, K. et al. Combinatorial optimization of mRNA structure, stability, and translation for RNA-based therapeutics. Nat. Commun. 13, 1536 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Hendel, A. et al. Chemically modified guide RNAs enhance CRISPR-Cas genome editing in human primary cells. Nat. Biotechnol. 33, 985–989 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Miller, P. S. et al. Nonionic nucleic acid analogues. Synthesis and characterization of dideoxyribonucleoside methylphosphonates. Biochemistry 18, 5134–5143 (1979).

    CAS  PubMed  Google Scholar 

  10. Wu, S. Y. et al. 2′-OMe-phosphorodithioate-modified siRNAs show increased loading into the RISC complex and enhanced anti-tumour activity. Nat. Commun. 5, 3459 (2014).

    PubMed  Google Scholar 

  11. Korobeynikov, V. A., Lyashchenko, A. K., Blanco-Redondo, B., Jafar-Nejad, P. & Shneider, N. A. Antisense oligonucleotide silencing of FUS expression as a therapeutic approach in amyotrophic lateral sclerosis. Nat. Med. 28, 104–116 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Yin, H. et al. Structure-guided chemical modification of guide RNA enables potent non-viral in vivo genome editing. Nat. Biotechnol. 35, 1179–1187 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Guinea-Viniegra, J. et al. Targeting miR-21 to treat psoriasis. Sci. Transl. Med. 6, 225re221 (2014).

    Google Scholar 

  14. Wang, S. R. et al. Conditional control of RNA-guided nucleic acid cleavage and gene editing. Nat. Commun. 11, 91 (2020).

    PubMed  PubMed Central  Google Scholar 

  15. Ovodov, S. & Alakhov, Yu,B. mRNA acetylated at 2′-OH-groups of ribose residues is functionally active in the cell-free translation system from wheat embryos. FEBS Lett. 270, 111–114 (1990).

    CAS  Google Scholar 

  16. Goldsborough, S. Modified polynucleotides and uses thereof. US patent 2003/0039985 A1 (2003).

  17. Steen, K. A., Siegfried, N. A. & Weeks, K. M. Selective 2’-hydroxyl acylation analyzed by protection from exoribonuclease (RNase-detected SHAPE) for direct analysis of covalent adducts and of nucleotide flexibility in RNA. Nat. Protoc. 6, 1683–1694 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Fessler, A. B. et al. Water-soluble isatoic anhydrides: a platform for RNA-SHAPE analysis and protein bioconjugation. Bioconjug. Chem. 29, 3196–3202 (2018).

    CAS  PubMed  Google Scholar 

  19. Fessler, A. B., Fowler, A. J. & Ogle, C. A. Directly quantifiable biotinylation using a water-soluble isatoic anhydride platform. Bioconjug. Chem. 32, 904–908 (2021).

    CAS  PubMed  Google Scholar 

  20. Velema, W. A. & Kool, E. T. The chemistry and applications of RNA 2’-OH acylation. Nat. Rev. Chem. 4, 22–37 (2020).

    CAS  PubMed  Google Scholar 

  21. Spitale, R. C. et al. Structural imprints in vivo decode RNA regulatory mechanisms. Nature 519, 486–490 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Velema, W. A., Kietrys, A. M. & Kool, E. T. RNA control by photoreversible acylation. J. Am. Chem. Soc. 140, 3491–3495 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Park, H. S., Kietrys, A. M. & Kool, E. T. Simple alkanoyl acylating agents for reversible RNA functionalization and control. Chem. Commun. 55, 5135–5138 (2019).

    CAS  Google Scholar 

  24. Kadina, A., Kietrys, A. M. & Kool, E. T. RNA cloaking by reversible acylation. Angew. Chem. Int. Ed. Engl. 57, 3059–3063 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Habibian, M. et al. Reversible RNA acylation for control of CRISPR-Cas9 gene editing. Chem. Sci. 11, 1011–1016 (2019).

    PubMed  PubMed Central  Google Scholar 

  26. Li, Y. & Breaker, R. R. Kinetics of RNA degradation by specific base catalysis of transesterification involving the 2‘-hydroxyl group. J. Am. Chem. Soc. 121, 5364–5372 (1999).

    CAS  Google Scholar 

  27. Soukup, G. A. & Breaker, R. R. Relationship between internucleotide linkage geometry and the stability of RNA. RNA 5, 1308–1325 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Duffy, K., Arangundy-Franklin, S. & Holliger, P. Modified nucleic acids: replication, evolution, and next-generation therapeutics. BMC Biol. 18, 112 (2020).

    PubMed  PubMed Central  Google Scholar 

  29. Takata, J. et al. Vitamin K prodrugs: 1. Synthesis of amino acid esters of menahydroquinone-4 and enzymatic reconversion to an active form. Pharm. Res. 12, 18–23 (1995).

    CAS  PubMed  Google Scholar 

  30. Johnson, S. L. in Advances in Physical Organic Chemistry Vol. 5 (ed. Gold, V.) 237–330 (Academic Press, 1967).

  31. delCardayre, S. B. & Raines, R. T. Structural determinants of enzymatic processivity. Biochemistry 33, 6031–6037 (1994).

    CAS  PubMed  Google Scholar 

  32. Raines, R. T. Ribonuclease A. Chem. Rev. 98, 1045–1066 (1998).

    CAS  PubMed  Google Scholar 

  33. Garnett, E. R. & Raines, R. T. Emerging biological functions of ribonuclease 1 and angiogenin. Crit. Rev. Biochem. Mol. Biol. 57, 244–260 (2022).

    CAS  PubMed  Google Scholar 

  34. Pace, C. N., Heinemann, U., Hahn, U. & Saenger, W. Ribonuclease T1: structure, function, and stability. Angew. Chem. Int. Ed. Engl. 30, 343–360 (1991).

    Google Scholar 

  35. Tsui, N. B., Ng, E. K. & Lo, Y. M. Stability of endogenous and added RNA in blood specimens, serum, and plasma. Clin. Chem. 48, 1647–1653 (2002).

    CAS  PubMed  Google Scholar 

  36. Daou-Chabo, R. & Condon, C. RNase J1 endonuclease activity as a probe of RNA secondary structure. RNA 15, 1417–1425 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Wang, Z., Treder, K. & Miller, W. A. Structure of a viral cap-independent translation element that functions via high affinity binding to the eIF4E subunit of eIF4F. J. Biol. Chem. 284, 14189–14202 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Jencks, W. P. & Carriuolo, J. Reactivity of nucleophilic reagents toward esters. J. Am. Chem. Soc. 82, 1778–1786 (1960).

    CAS  Google Scholar 

  39. van der Helm, M. P., Klemm, B. & Eelkema, R. Organocatalysis in aqueous media. Nat. Rev. Chem. 3, 491–508 (2019).

    Google Scholar 

  40. Jarvinen, P., Oivanen, M. & Lonnberg, H. Interconversion and phosphoester hydrolysis of 2′,5′- and 3′,5′-dinucleoside monophosphates: kinetics and mechanisms. J. Org. Chem. 56, 5396–5401 (1991).

    CAS  Google Scholar 

  41. Xiao, L., Jun, Y. W. & Kool, E. T. DNA tiling enables precise acylation-based labeling and control of mRNA. Angew. Chem. Int. Ed. Engl. 60, 26798–26805 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Werber, M. M. & Shalitin, Y. The reaction of tertiary amino alcohols with active esters: I. Acylation and deacylation steps. Bioorg. Chem. 2, 202–220 (1973).

    CAS  Google Scholar 

  43. de Jersey, J., Fihelly, A. K. & Zerner, B. On the mechanism of the reaction of tris(hydroxymethyl)aminomethane with activated carbonyl compounds: a model for the serine proteinases. Bioorg. Chem. 9, 153–162 (1980).

    Google Scholar 

  44. Park, H. S., Jash, B., Xiao, L., Jun, Y. W. & Kool, E. T. Control of RNA with quinone methide reversible acylating reagents. Org. Biomol. Chem. 19, 8367–8376 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Kain, S. R. Green fluorescent protein (GFP): applications in cell-based assays for drug discovery. Drug Discov. Today 4, 304–312 (1999).

    CAS  PubMed  Google Scholar 

  46. Singh, A., Razooky, B. S., Dar, R. D. & Weinberger, L. S. Dynamics of protein noise can distinguish between alternate sources of gene-expression variability. Mol. Syst. Biol. 8, 607 (2012).

    PubMed  PubMed Central  Google Scholar 

  47. Li, X. et al. Generation of destabilized green fluorescent protein as a transcription reporter. J. Biol. Chem. 273, 34970–34975 (1998).

    CAS  PubMed  Google Scholar 

  48. LoPachin, R. M. & Gavin, T. Reactions of electrophiles with nucleophilic thiolate sites: relevance to pathophysiological mechanisms and remediation. Free Radic. Res. 50, 195–205 (2016).

    CAS  PubMed  Google Scholar 

  49. Hou, X., Zaks, T., Langer, R. & Dong, Y. Lipid nanoparticles for mRNA delivery. Nat. Rev. Mater. 6, 1078–1094 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Raj, A., Peskin, C. S., Tranchina, D., Vargas, D. Y. & Tyagi, S. Stochastic mRNA synthesis in mammalian cells. PLoS Biol. 4, e309 (2006).

    PubMed  PubMed Central  Google Scholar 

  51. Kerpedjiev, P., Hammer, S. & Hofacker, I. L. Forna (force-directed RNA): simple and effective online RNA secondary structure diagrams. Bioinformatics 31, 3377–3379 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Gu, C. et al. Chemical synthesis of stimuli-responsive guide RNA for conditional control of CRISPR-Cas9 gene editing. Chem. Sci. 12, 9934–9945 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Warren, L. et al. Highly efficient reprogramming to pluripotency and directed differentiation of human cells with synthetic modified mRNA. Cell Stem Cell 7, 618–630 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by a grant from the US National Institutes of Health GM127295 to E.T.K. The funder had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript. We thank K. Fukui and other staff at the Stanford PAN facility for performing BioAnalyzer QC. We thank Stanford Shared FACS Facility (SSFF) for the flow cytometry analysis. We thank T. McLaughlin in the Vincent Coates Foundation Mass Spectrometry Laboratory, Stanford University Mass Spectrometry (RRID:SCR_017801) for acquiring the HRMS data. We also thank T. Trinh for performing circular dichroism measurements and L. Zheng for the helpful discussion.

Author information

Authors and Affiliations

Authors

Contributions

L.F. and E.T.K. conceived the project and designed the experiments. L.F. performed the experiments and data analysis. L.X. performed the Cas9-mediated DNA cleavage assay. Y.W.J. collected and analysed the confocal microscopy data. E.T.K. supervised the work. L.F. and E.T.K. wrote the paper, with input from all authors.

Corresponding author

Correspondence to Eric T. Kool.

Ethics declarations

Competing interests

Stanford University has filed a patent application (PCTUS2023/010686) based on the RNA cloaking and uncloaking reagents and methods described in this work, in which L.F. and E.T.K. are named as inventors. The other authors declare no competing interests.

Peer review

Peer review information

Nature Chemistry thanks Ashwani Sharma, Joanna Sztuba-Solinska, Wen Zhang and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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 Reaction conditions of RNA cloaking with acylimidazole reagents.

A heatmap shows reaction conditions for each acylimidazole reagent (1-7) required to introduce the desired level of RNA cloaking using a model 18nt tRF-3005 RNA. The X-axis indicates the number of 2′-acyl adducts per RNA molecule identified by MALDI-TOF. The color intensity of each cell represents the mass intensity of the acyl adduct peak (numbers at bottom show numbers of acyl adducts for the model RNA strand). See Supplementary Table 1 for quantification.

Source data

Extended Data Fig. 2 One-step synthesis of acylimidazoles for selectively 2′-OH acylation.

a, Schematic for one-step synthesis of acylimidazole reagents. b-c, Comparisons of RNA and DNA with the same sequence (tRF-3005-DNA and tRF-3005-DNA-Phos) after reacting with 1-7 show that reaction occurs entirely or almost entirely at 2′-OH groups rather than nucleobases or phosphodiester bonds. b, Number of acyl adducts on tRF-3005-DNA (18nt) upon treatment with acylimidazole reagents 1-7. tRF-3005-DNA contains an unmodified 3′-OH (Sequence: 5′-ATC CTG CCG ACT ACG CCA-3′-OH). c, Number of acyl adducts on tRF-3005-DNA-Phos (18nt) upon treatment with acylimidazole reagents 1, 5, 6, and 7. tRF-3005-DNA-Phos contains a blocked 3′-end by 3′-phosphorylation (Sequence: 5′-ATC CTG CCG ACT ACG CCA-3′-phosphate). The X-axis shows the number of acyl adducts per DNA identified by MALDI-TOF. The color intensity of each cell represents the mass intensity of the acyl adduct peak.

Source data

Extended Data Fig. 3 Sensitive acylimidazole-nucleophile pairs.

A summary of sensitive acylimidazole-nucleophile pairs that afford >50% hydrolysis of 2′-acyl adducts within 2 hours. The electrophilic centers are coloured orange.

Extended Data Fig. 4 Spontaneous RNA uncloaking restores mRNA translation in human cells.

a, Representative single-blind fluorescence micrographs of HeLa cells transfected with 1-cloaked eGFP-mRNA showing strongly inhibited mRNA translation after 24 or 48 h. Each condition was imaged at areas that contain more than 50 cells over two independent experiments. b, Time-course plot showing 7-cloaked eGFP-mRNA (~50% of 2′-hydroxyls acylated) translates in HeLa cells. Its translation kinetics were compared to unprotected or equimolarly 4-cloaked eGFP-mRNA. Data represent mean ± s.e.m., n = 3 per group from three biologically independent experiments.

Source data

Supplementary information

Supplementary Information

Supplementary Tables 1–4, Figs. 1–13, Synthesis protocol, Uncropped gels for supplementary information and NMR/HRMS data.

Reporting Summary

Source data

Source Data Fig. 1

Unprocessed gels.

Source Data Fig. 2

Statistical source data and unprocessed gels.

Source Data Fig. 3

Statistical source data and unprocessed gels.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fang, L., Xiao, L., Jun, Y.W. et al. Reversible 2′-OH acylation enhances RNA stability. Nat. Chem. 15, 1296–1305 (2023). https://doi.org/10.1038/s41557-023-01246-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41557-023-01246-6

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