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:

Functional and structural basis of extreme conservation in vertebrate 5′ untranslated regions

Abstract

The lack of knowledge about extreme conservation in genomes remains a major gap in our understanding of the evolution of gene regulation. Here, we reveal an unexpected role of extremely conserved 5′ untranslated regions (UTRs) in noncanonical translational regulation that is linked to the emergence of essential developmental features in vertebrate species. Endogenous deletion of conserved elements within these 5′ UTRs decreased gene expression, and extremely conserved 5′ UTRs possess cis-regulatory elements that promote cell-type-specific regulation of translation. We further developed in-cell mutate-and-map (icM2), a new methodology that maps RNA structure inside cells. Using icM2, we determined that an extremely conserved 5′ UTR encodes multiple alternative structures and that each single nucleotide within the conserved element maintains the balance of alternative structures important to control the dynamic range of protein expression. These results explain how extreme sequence conservation can lead to RNA-level biological functions encoded in the untranslated regions of vertebrate genomes.

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: Hyperconserved 5′ UTRs in vertebrate genomes.
Fig. 2: Hyperconserved 5′ UTRs impact translation efficiency.
Fig. 3: Noncanonical translation enhancer in hyperconserved 5′ UTRs.
Fig. 4: Cellular remodeling hyperconserved 5′ UTR RNA structures.
Fig. 5: icM2 reveals structured elements in the hyperconserved Csde1 5′ UTR.
Fig. 6: Csde1 5′ UTR tunes translation efficiency by encoding multiple alternative structures that are actively maintained by RNA helicases.

Similar content being viewed by others

Data availability

Raw sequencing data (related to Figs. 4, 5 and 6) are deposited to GEO with accession code GSE155656. Processed reactivity data have been deposited in the RNA Mapping Database (RMDB) with accession codes CSDE1_DMS_0000 and CSDE1_DMS_0001. Sources for publicly available data are described in the Methods.

Code availability

All software used to analyze the study data are listed in the Methods and in the Nature Research Reporting Summary and are publicly available. All codes used to analyze icM2 data are available through a Github repository: github.com/barnalab/icm2p.

References

  1. Dermitzakis, E. T., Reymond, A. & Antonarakis, S. E. Conserved non-genic sequences – an unexpected feature of mammalian genomes. Nat. Rev. Genet. 6, 151–157 (2005).

    Article  CAS  PubMed  Google Scholar 

  2. Harmston, N., Baresic, A. & Lenhard, B. The mystery of extreme non-coding conservation. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 368, 20130021 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  3. Halligan, D. L. et al. Positive and negative selection in murine ultraconserved noncoding elements. Mol. Biol. Evol. 28, 2651–2660 (2011).

    Article  CAS  PubMed  Google Scholar 

  4. Bejerano, G. et al. Ultraconserved elements in the human genome. Science 304, 1321–1325 (2004).

    Article  CAS  PubMed  Google Scholar 

  5. Dimitrieva, S. & Bucher, P. Genomic context analysis reveals dense interaction network between vertebrate ultraconserved non-coding elements. Bioinformatics 28, i395–i401 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Boffelli, D., Nobrega, M. A. & Rubin, E. M. Comparative genomics at the vertebrate extremes. Nat. Rev. Genet. 5, 456–465 (2004).

    Article  CAS  PubMed  Google Scholar 

  7. Lindblad-Toh, K. et al. Genome sequence, comparative analysis and haplotype structure of the domestic dog. Nature 438, 803–819 (2005).

    Article  CAS  PubMed  Google Scholar 

  8. Sandelin, A. et al. Arrays of ultraconserved non-coding regions span the loci of key developmental genes in vertebrate genomes. BMC Genomics 5, 99 (2004).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. de la Calle-Mustienes, E. et al. A functional survey of the enhancer activity of conserved non-coding sequences from vertebrate Iroquois cluster gene deserts. Genome Res. 15, 1061–1072 (2005).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  10. Sakuraba, Y. et al. Identification and characterization of new long conserved noncoding sequences in vertebrates. Mamm. Genome 19, 703–712 (2008).

    Article  CAS  PubMed  Google Scholar 

  11. Dermitzakis, E. T. et al. Comparison of human chromosome 21 conserved nongenic sequences (CNGs) with the mouse and dog genomes shows that their selective constraint is independent of their genic environment. Genome Res. 14, 852–859 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Katzman, S. et al. Human genome ultraconserved elements are ultraselected. Science 317, 915 (2007).

    Article  CAS  PubMed  Google Scholar 

  13. Pennacchio, L. A. et al. In vivo enhancer analysis of human conserved non-coding sequences. Nature 444, 499–502 (2006).

    Article  CAS  PubMed  Google Scholar 

  14. Visel, A. et al. Ultraconservation identifies a small subset of extremely constrained developmental enhancers. Nat. Genet. 40, 158–160 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Visel, A. et al. A high-resolution enhancer atlas of the developing telencephalon. Cell 152, 895–908 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Ahituv, N. et al. Deletion of ultraconserved elements yields viable mice. PLoS Biol. 5, e234 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. McLean, C. & Bejerano, G. Dispensability of mammalian DNA. Genome Res. 18, 1743–1751 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Dickel, D. E. et al. Ultraconserved enhancers are required for normal development. Cell 172, 491–499.e15 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Osterwalder, M. et al. Enhancer redundancy provides phenotypic robustness in mammalian development. Nature 554, 239–243 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Lareau, L. F., Inada, M., Green, R. E., Wengrod, J. C. & Brenner, S. E. Unproductive splicing of SR genes associated with highly conserved and ultraconserved DNA elements. Nature 446, 926–929 (2007).

    Article  CAS  PubMed  Google Scholar 

  21. Ni, J. Z. et al. Ultraconserved elements are associated with homeostatic control of splicing regulators by alternative splicing and nonsense-mediated decay. Genes Dev. 21, 708–718 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Thomas, J. D. et al. RNA isoform screens uncover the essentiality and tumor-suppressor activity of ultraconserved poison exons. Nat. Genet. 52, 84–94 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Calin, G. A. et al. Ultraconserved regions encoding ncRNAs are altered in human leukemias and carcinomas. Cancer Cell 12, 215–229 (2007).

    Article  CAS  PubMed  Google Scholar 

  24. Liz, J. et al. Regulation of pri-miRNA processing by a long noncoding RNA transcribed from an ultraconserved region. Mol. Cell 55, 138–147 (2014).

    Article  CAS  PubMed  Google Scholar 

  25. Xue, S. et al. RNA regulons in Hox 5′ UTRs confer ribosome specificity to gene regulation. Nature 517, 33–38 (2015).

    Article  CAS  PubMed  Google Scholar 

  26. Siepel, A. et al. Evolutionarily conserved elements in vertebrate, insect, worm, and yeast genomes. Genome Res. 15, 1034–1050 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Jiang, L. et al. A quantitative proteome map of the human body. Cell 183, 269–283.e19 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. GTEx Consortium. The GTEx Consortium atlas of genetic regulatory effects across human tissues. Science 369, 1318–1330 (2020).

  29. Steri, M., Idda, M. L., Whalen, M. B. & Orrù, V. Genetic variants in mRNA untranslated regions. Wiley Interdiscip. Rev. RNA 9, e1474 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Blanco-Suarez, E., Liu, T.-F., Kopelevich, A. & Allen, N. J. Astrocyte-secreted chordin-like 1 drives synapse maturation and limits plasticity by increasing synaptic GluA2 AMPA receptors. Neuron 100, 1116–1132.e13 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Sakuta, H. et al. Ventroptin: a BMP-4 antagonist expressed in a double-gradient pattern in the retina. Science 293, 111–115 (2001).

    Article  CAS  PubMed  Google Scholar 

  32. Webb, T. R. et al. X-linked megalocornea caused by mutations in CHRDL1 identifies an essential role for ventroptin in anterior segment development. Am. J. Hum. Genet. 90, 247–259 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Gandal, M. J. et al. Shared molecular neuropathology across major psychiatric disorders parallels polygenic overlap. Science 359, 693–697 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Liu, T. et al. Chordin-like 1 improves osteogenesis of bone marrow mesenchymal stem cells through enhancing BMP4-SMAD pathway. Front. Endocrinol. 10, 360 (2019).

    Article  Google Scholar 

  35. Pei, Y.-F. et al. Hypermethylation of the CHRDL1 promoter induces proliferation and metastasis by activating Akt and Erk in gastric cancer. Oncotarget 8, 23155–23166 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Osório, C. et al. Growth differentiation factor 5 is a key physiological regulator of dendrite growth during development. Development 140, 4751–4762 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. O’Keeffe, G. W. et al. Region-specific role of growth differentiation factor-5 in the establishment of sympathetic innervation. Neural Dev. 11, 4 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Wu, H., Li, J., Xu, D., Zhang, Q. & Cui, T. Growth differentiation factor 5 improves neurogenesis and functional recovery in adult mouse hippocampus following traumatic brain injury. Front. Neurol. 9, 592 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Buxton, P., Edwards, C., Archer, C. W. & Francis-West, P. Growth/differentiation factor-5 (GDF-5) and skeletal development. J. Bone Joint Surg. Am. 83-A, S23–S30 (2001).

    Google Scholar 

  40. Panganiban, G. & Rubenstein, J. L. R. Developmental functions of the Distal-less/Dlx homeobox genes. Development 129, 4371–4386 (2002).

    Article  CAS  PubMed  Google Scholar 

  41. Depew, M. J., Simpson, C. A., Morasso, M. & Rubenstein, J. L. R. Reassessing the Dlx code: the genetic regulation of branchial arch skeletal pattern and development. J. Anat. 207, 501–561 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Polleux, F., Morrow, T. & Ghosh, A. Semaphorin 3A is a chemoattractant for cortical apical dendrites. Nature 404, 567–573 (2000).

    Article  CAS  PubMed  Google Scholar 

  43. Serini, G. et al. Class 3 semaphorins control vascular morphogenesis by inhibiting integrin function. Nature 424, 391–397 (2003).

    Article  CAS  PubMed  Google Scholar 

  44. Shelly, M. et al. Semaphorin 3A regulates neuronal polarization by suppressing axon formation and promoting dendrite growth. Neuron 71, 433–446 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Polleux, F., Giger, R. J., Ginty, D. D., Kolodkin, A. L. & Ghosh, A. Patterning of cortical efferent projections by semaphorin–neuropilin interactions. Science 282, 1904–1906 (1998).

    Article  CAS  PubMed  Google Scholar 

  46. Good, P. F. et al. A role for semaphorin 3A signaling in the degeneration of hippocampal neurons during Alzheimer’s disease. J. Neurochem. 91, 716–736 (2004).

    Article  CAS  PubMed  Google Scholar 

  47. Galan-Caridad, J. M. et al. Zfx controls the self-renewal of embryonic and hematopoietic stem cells. Cell 129, 345–357 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Lee, A. S. Y., Kranzusch, P. J. & Cate, J. H. D. eIF3 targets cell-proliferation messenger RNAs for translational activation or repression. Nature 522, 111–114 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Gilbert, W. V., Zhou, K., Butler, T. K. & Doudna, J. A. Cap-independent translation is required for starvation-induced differentiation in yeast. Science 317, 1224–1227 (2007).

    Article  CAS  PubMed  Google Scholar 

  50. Martin, F. et al. Cap-assisted internal initiation of translation of histone H4. Mol. Cell 41, 197–209 (2011).

    Article  CAS  PubMed  Google Scholar 

  51. Legnini, I. et al. Circ-ZNF609 is a circular RNA that can be translated and functions in myogenesis. Mol. Cell 66, 22–37.e9 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Pamudurti, N. R. et al. Translation of circRNAs. Mol. Cell 66, 9–21.e7 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Leppek, K. et al. Gene- and species-specific Hox mRNA translation by ribosome expansion segments. Mol. Cell 80, 980–995.e13 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Hershey, J. W. B., Sonenberg, N. & Mathews, M. B. Principles of translational control: an overview. Cold Spring Harb. Perspect. Biol. 4, a011528. (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Weingarten-Gabbay, S. et al. Comparative genetics. Systematic discovery of cap-independent translation sequences in human and viral genomes. Science 351, aad4939 (2016).

    Article  PubMed  CAS  Google Scholar 

  56. Xiao, Z.-S., Simpson, L. G. & Quarles, L. D. IRES-dependent translational control of Cbfa1/Runx2 expression. J. Cell. Biochem. 88, 493–505 (2003).

    Article  CAS  PubMed  Google Scholar 

  57. Jang, G. M. et al. Structurally distinct elements mediate internal ribosome entry within the 5′-noncoding region of a voltage-gated potassium channel mRNA. J. Biol. Chem. 279, 47419–47430 (2004).

    Article  CAS  PubMed  Google Scholar 

  58. Holcik, M. & Sonenberg, N. Translational control in stress and apoptosis. Nat. Rev. Mol. Cell Biol. 6, 318–327 (2005).

    Article  CAS  PubMed  Google Scholar 

  59. El-Naggar, A. M. & Sorensen, P. H. Translational control of aberrant stress responses as a hallmark of cancer. J. Pathol. 244, 650–666 (2018).

    Article  CAS  PubMed  Google Scholar 

  60. Spriggs, K. A., Bushell, M. & Willis, A. E. Translational regulation of gene expression during conditions of cell stress. Mol. Cell 40, 228–237 (2010).

    Article  CAS  PubMed  Google Scholar 

  61. Washietl, S., Hofacker, I. L., Lukasser, M., Hüttenhofer, A. & Stadler, P. F. Mapping of conserved RNA secondary structures predicts thousands of functional noncoding RNAs in the human genome. Nat. Biotechnol. 23, 1383–1390 (2005).

    Article  CAS  PubMed  Google Scholar 

  62. Torarinsson, E. et al. Comparative genomics beyond sequence-based alignments: RNA structures in the ENCODE regions. Genome Res. 18, 242–251 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Parker, B. J. et al. New families of human regulatory RNA structures identified by comparative analysis of vertebrate genomes. Genome Res. 21, 1929–1943 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Smith, M. A., Gesell, T., Stadler, P. F. & Mattick, J. S. Widespread purifying selection on RNA structure in mammals. Nucleic Acids Res. 41, 8220–8236 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Eddy, S. R. Computational analysis of conserved RNA secondary structure in transcriptomes and genomes. Annu. Rev. Biophys. 43, 433–456 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Rivas, E., Clements, J. & Eddy, S. R. Estimating the power of sequence covariation for detecting conserved RNA structure. Bioinformatics 36, 3072–3076 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Homan, P. J. et al. Single-molecule correlated chemical probing of RNA. Proc. Natl Acad. Sci. USA 111, 13858–13863 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Zubradt, M. et al. DMS-MaPseq for genome-wide or targeted RNA structure probing in vivo. Nat. Methods 14, 75–82 (2017).

    Article  CAS  PubMed  Google Scholar 

  69. Mustoe, A. M., Lama, N. N., Irving, P. S., Olson, S. W. & Weeks, K. M. RNA base-pairing complexity in living cells visualized by correlated chemical probing. Proc. Natl Acad. Sci. USA 116, 24574–24582 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Beaudoin, J.-D. et al. Analyses of mRNA structure dynamics identify embryonic gene regulatory programs. Nat. Struct. Mol. Biol. 25, 677–686 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Patalano, S., Mihailovich, M., Belacortu, Y., Paricio, N. & Gebauer, F. Dual sex-specific functions of Drosophila Upstream of N-ras in the control of X chromosome dosage compensation. Development 136, 689–698 (2009).

    Article  CAS  PubMed  Google Scholar 

  72. Elatmani, H. et al. The RNA-binding protein Unr prevents mouse embryonic stem cells differentiation toward the primitive endoderm lineage. Stem Cells 29, 1504–1516 (2011).

    Article  CAS  PubMed  Google Scholar 

  73. Mitchell, S. A., Brown, E. C., Coldwell, M. J., Jackson, R. J. & Willis, A. E. Protein factor requirements of the Apaf-1 internal ribosome entry segment: roles of polypyrimidine tract binding protein and upstream of N-ras. Mol. Cell. Biol. 21, 3364–3374 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Schepens, B. et al. A role for hnRNP C1/C2 and Unr in internal initiation of translation during mitosis. EMBO J. 26, 158–169 (2007).

    Article  CAS  PubMed  Google Scholar 

  75. Guo, A.-X., Cui, J.-J., Wang, L.-Y. & Yin, J.-Y. The role of CSDE1 in translational reprogramming and human diseases. Cell Commun. Signal. 18, 14 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Moore, K. S. et al. Csde1 binds transcripts involved in protein homeostasis and controls their expression in an erythroid cell line. Sci. Rep. 8, 2628 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  77. Wurth, L. et al. UNR/CSDE1 drives a post-transcriptional program to promote melanoma invasion and metastasis. Cancer Cell 30, 694–707 (2016).

    Article  CAS  PubMed  Google Scholar 

  78. Horos, R. et al. Ribosomal deficiencies in Diamond-Blackfan anemia impair translation of transcripts essential for differentiation of murine and human erythroblasts. Blood 119, 262–272 (2012).

    Article  CAS  PubMed  Google Scholar 

  79. Guo, H. et al. Disruptive variants of CSDE1 associate with autism and interfere with neuronal development and synaptic transmission. Sci. Adv. 5, eaax2166 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Saltel, F. et al. Unr defines a novel class of nucleoplasmic reticulum involved in mRNA translation. J. Cell Sci. 130, 1796–1808 (2017).

    CAS  PubMed  Google Scholar 

  81. Sanders, S. J. et al. De novo mutations revealed by whole-exome sequencing are strongly associated with autism. Nature 485, 237–241 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Kladwang, W., VanLang, C. C., Cordero, P. & Das, R. A two-dimensional mutate-and-map strategy for non-coding RNA structure. Nat. Chem. 3, 954–962 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Cordero, P. & Das, R. Rich RNA structure landscapes revealed by mutate-and-map analysis. PLoS Comput. Biol. 11, e1004473 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  84. Bhaskaran, H. & Russell, R. Kinetic redistribution of native and misfolded RNAs by a DEAD-box chaperone. Nature 449, 1014–1018 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Alexa, A., Rahnenführer, J. & Lengauer, T. Improved scoring of functional groups from gene expression data by decorrelating GO graph structure. Bioinformatics 22, 1600–1607 (2006).

    Article  CAS  PubMed  Google Scholar 

  86. Motenko, H., Neuhauser, S. B., O’Keefe, M. & Richardson, J. E. MouseMine: a new data warehouse for MGI. Mamm. Genome 26, 325–330 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Concordet, J.-P. & Haeussler, M. CRISPOR: intuitive guide selection for CRISPR/Cas9 genome editing experiments and screens. Nucleic Acids Res. 46, W242–W245 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Yoon, A. et al. Impaired control of IRES-mediated translation in X-linked dyskeratosis congenita. Science 312, 902–906 (2006).

    Article  CAS  PubMed  Google Scholar 

  89. Robinson, M. D. & Oshlack, A. A scaling normalization method for differential expression analysis of RNA-seq data. Genome Biol. 11, R25 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  90. Law, C. W., Chen, Y., Shi, W. & Smyth, G. K. voom: precision weights unlock linear model analysis tools for RNA-seq read counts. Genome Biol. 15, R29 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  91. Ritchie, M. E. et al. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 43, e47 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  92. Lorenz, R. et al. ViennaRNA Package 2.0. Algorithms Mol. Biol. 6, 26 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank the members of the Barna laboratory for constructive criticism of the manuscript. This work was supported by New York Stem Cell Foundation grant NYSCF-R-I36 (M.B.), NIH grant 1R01HD086634 (M.B.), Alfred P. Sloan Research Fellowship (M.B.), Pew Scholars Award (M.B.), Mallinckrodt Foundation Award (M.B.), Benchmark Stanford Graduate Fellowship (G.W.B.) and Walter and Idun Berry Foundation (E.S.C.). M.B. is a New York Stem Cell Robertson Investigator.

Author information

Authors and Affiliations

Authors

Contributions

M.B., G.W.B. and E.S.C. conceived the project. M.B. supervised the project. L.J. and H.T. provided the GTEx data and critical feedback on its analysis. R.D. provided critical feedback on the development and analysis of icM2. E.S.C. carried out the large-scale reporter screens. G.W.B. performed all other experiments and data analysis. G.W.B. and M.B. wrote the manuscript in consultation with all authors.

Corresponding author

Correspondence to Maria Barna.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Genetics thanks Jean-Denis Beaudoin, Philip Bevilacqua and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

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

Extended data

Extended Data Fig. 1 Hyperconserved 5’UTRs in vertebrate genomes.

a, Left: heatmap of the positions of LOD ≥ 500 PhastCons elements in each h5UTR. Middle: heatmap of the relative positions (calculated in 100 bins across the h5UTRs) of the elements. Right: plot of average element overlap across the 100 bins to illustrate the positional preference. b, Histogram of the length of h5UTRs. Average length is 674nt. c, Histogram of the number of nucleotides overlap between LOD ≥ 500 PhastCons elements and h5UTRs. Average overlap is 389nt. d, Distributions of cross-tissue transcriptome-proteome correlations for all genes, genes with h5UTRs, or genes with variance-matched non-conserved 5’UTRs. Indicated p-values are from two-sided Wilcoxon rank sum tests for cross-tissue correlation values between h5UTR genes and all genes or between h5UTR genes and variance-matched non-conserved controls. e, Distributions of the number of annotated alternative 5’UTRs for all genes, genes with h5UTRs, or genes with size-matched non-conserved 5’UTRs. Indicated p-values are from two-sided Wilcoxon rank sum tests for the number of alternative 5’UTRs between h5UTR genes and all genes or between h5UTR genes and size-matched non-conserved controls. f, Scatter plot illustrating the lack of significant term enrichments for a size-matched set of non-conserved 5’ UTRs. X-axis and y-axis plots expected and the observed number of genes for each term. Blue dashed line indicates the minimum observed/expected ratio cutoff of 3. Green line indicates expected and observed counts where Fisher’s test p-value (pf) is estimated to have FDR = 0.05. Neighbor-weighted test p-value (pfw) ≤0.05 is further used as an additional cutoff. The final set of enriched terms passing filter is colored by pf and sized by pfw.

Extended Data Fig. 2 Non-canonical translation activation by hyperconserved 5’UTRs across cell types.

a, Density plots of non-canonical translation initiation activities from h5UTRs by bicistronic reporter assay. X-axis is the luciferase reporter activity ratios. Jittered dots mark individual reporter ratios for each h5UTR in each cell type. b, Summarized plot of ribosome load (sum of % mRNA times the ribosome number for each fraction) differential ratio between NSCs and ESCs calculated from polysome profiles for each gene shown in Extended Data Fig. 2c-l. Red indicates significant increase in NSCs and black indicates significant decrease (two-sided t-test p ≤ 0.05, n = 3, marked by asterisk). c-l, Endogenous polysome profiles of NSCs versus ESCs for genes with h5UTRs that show high non-canonical translation reporter activities in NSCs compared to ESCs. Distribution of mRNAs across sucrose gradient fractions are plotted. Y-axis plots the mean percent mRNA. Error bars indicate standard error. Asterisk indicates two-sided t-test p ≤ 0.05 for each fraction between the two cell types. n = 3 for each cell type. Indicated p-value (pf) is calculated by Fisher’s method across all fractions. Note that Extended Data Fig. 2c shows the profile of 18 S rRNA, which indicates lower global translation in NSCs compared to ESCs.

Extended Data Fig. 3 Non-canonical activation by hyperconserved 5’UTRs substantially contributes to translation.

a, Scatter plot of luciferase activity versus RNA level ratios (mean from n = 3) observed for the bicistronic reporters of 90 h5UTRs measured in 10T1/2 cells. Dashed line marks the 10% FDR used in Fig. 3a. Spearman correlation indicated on top left. b, The effect of various truncations of the h5UTRs on non-canonical initiation and total translation efficiency (also see Fig. 3d). Left: positions of truncations. Dashed lines indicate truncations. Purple horizontal lines indicate uORFs; yellow and red lines indicate in-frame and out-of-frame uAUGs, respectively. Middle: non-canonical initiation efficiency. Right: total translation efficiency. X-axis indicates the mean of luciferase reporter ratios relative to the wild-type. Error bars indicate standard error. Dashed line marks the wild-type 5’UTR activity. Asterisk indicates two-sided t-test p ≤ 0.05 for each truncation versus the full-length. The numbers to the left of the bars indicate n and p-values. c, Comparison of translational activities between the full-length long, non-conserved 5’UTRs versus the only first 300nt truncation. 11 different pairs are tested. X-axis indicates the mean log2 luciferase reporter ratios of each truncation relative to its full-length wild-type. Error bars indicate standard error. Bars colored in red indicate significantly reduced translation in the shorter, truncated 300nt fragment; black indicates significant increase (two-sided t-test, paired n = 3, p ≤ 0.05, marked by asterisk). The numbers to the left of the bars indicate p-values. d, Violin plot of full-length/truncated reporter activity ratios (log2) from hyperconserved and non-conserved 5’UTRs. p indicates two-sided Wilcoxon rank sum test p-value. Box hinges: 25% quantile, median, 75% quantile, respectively from left to right. Whiskers: lower or upper hinge ±1.5*IQR. e, Scatter plot of change in translation efficiency between full-length and truncated h5UTRs shown in Fig. 3e versus change in uAUG density (change in number of AUGs / change in length between each pair of full-length and truncated h5UTRs). r indicates pearson’s correlation coefficient and p indicates two-tailed p-value.

Extended Data Fig. 4 Cellular remodeling of hyperconserved 5’UTR RNA structures.

a, Stacked bar plots showing proportions of significant (FDR ≤ 0.05) or not significant windows that overlap uAUG in black versus that do not overlap uAUG in red. OR indicates odds ratio for overlaps uAUG / does not overlap uAUG, and p indicates Fisher’s test p-value (one-sided, Ha = odds ratio>0). b, Stacked bar plots showing proportions of significant (FDR ≤ 0.05) or not significant windows that overlap uORF in black versus that do not overlap uORF in red. OR indicates odds ratio for overlaps uORF / does not overlap uORF, and p indicates Fisher’s test p-value (one-sided, Ha = odds ratio>0). c, Zoomed-in view of differential accessibilities along h5UTRs with one or more significantly different windows under ATP depletion. Top plot shows -log10 p-value for each window. Highlighted boxes mark significantly different windows, above the dashed line indicating 5% FDR. Middle plot shows differential accessibility on the y-axis, where greater than zero indicates increased accessibility upon ATP depletion and less than zero indicates decreased accessibility. Bottom plot shows differential accessibility for in vitro refolded RNA. Error bars in each plot show standard error, n = 3. The three profiled regions shown on the left side exhibit discordant profiles between accessibility changes observed in cells following ATP depletion and accessibility changes observed for in cell versus in vitro refolded RNA. The other three on the right side exhibit concordant profiles.

Extended Data Fig. 5 icM2 reveals structured elements in the hyperconserved Csde1 5’UTR.

a, Boxplot of average PhastCons scores in significant windows of ATP-dependent remodeling versus all windows shown in Fig. 4c. p indicates two-sided Wilcoxon rank sum test p-value. b, Same as Extended Data Fig. 5a, but showing the distribution of average PhyloP scores.

Extended Data Fig. 6 In-vitro M2 analysis of Csde1 5’UTR.

a, Heatmap of in-vitro M2 accessibility matrix for Csde1 5’UTR from position 190 to 386. For each row, the chemical mapping profile of a single-nucleotide variant of the RNA is plotted across the columns, where the colors indicate z-scaled accessibility change values from the wild-type RNA. 1D data from each mutant are vertically stacked to display a 2D matrix. White boxes mark perturbation signals that support the model shown in Extended Data Fig. 6b; color bars at the bottom indicate the nucleotide positions of the stems that match the same color in the model. b, The model for the in-vitro structure of Csde1 5’UTR from position 190 to 386. Also see Extended Data Fig. 6a.

Supplementary information

Supplementary Information

Supplementary Figs. 1–7 and Notes

Reporting Summary

Peer Review Information

Supplementary Tables

Supplementary Tables 1–9

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Byeon, G.W., Cenik, E.S., Jiang, L. et al. Functional and structural basis of extreme conservation in vertebrate 5′ untranslated regions. Nat Genet 53, 729–741 (2021). https://doi.org/10.1038/s41588-021-00830-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-021-00830-1

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