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.

  • Review Article
  • Published:

Cellular functions of eukaryotic RNA helicases and their links to human diseases

Abstract

RNA helicases are highly conserved proteins that use nucleoside triphosphates to bind or remodel RNA, RNA–protein complexes or both. RNA helicases are classified into the DEAD-box, DEAH/RHA, Ski2-like, Upf1-like and RIG-I families, and are the largest class of enzymes active in eukaryotic RNA metabolism — virtually all aspects of gene expression and its regulation involve RNA helicases. Mutation and dysregulation of these enzymes have been linked to a multitude of diseases, including cancer and neurological disorders. In this Review, we discuss the regulation and functional mechanisms of RNA helicases and their roles in eukaryotic RNA metabolism, including in transcription regulation, pre-mRNA splicing, ribosome assembly, translation and RNA decay. We highlight intriguing models that link helicase structure, mechanisms of function (such as local strand unwinding, translocation, winching, RNA clamping and displacing RNA-binding proteins) and biological roles, including emerging connections between RNA helicases and cellular condensates formed through liquid–liquid phase separation. We also discuss associations of RNA helicases with human diseases and recent efforts towards the design of small-molecule inhibitors of these pivotal regulators of eukaryotic gene expression.

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: Characteristics of RNA helicases.
Fig. 2: Mechanisms of function of RNA helicases.
Fig. 3: Regulation of helicase functions.
Fig. 4: Roles of RNA helicases in pre-mRNA splicing.
Fig. 5: Roles of RNA helicases in translation and RNA decay.
Fig. 6: Functions of RNA helicases in ribosome assembly.
Fig. 7: RNA helicases in viral infections and RNP granule (dis)assembly.

Similar content being viewed by others

References

  1. Jankowsky, E. RNA helicases at work: binding and rearranging. Trends Biochem. Sci. 36, 19–29 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Fairman-Williams, M. E., Guenther, U.-P. & Jankowsky, E. SF1 and SF2 helicases: family matters. Curr. Opin. Struct. Biol. 20, 313–324 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Pyle, A. M. Translocation and unwinding mechanisms of RNA and DNA helicases. Annu. Rev. Biophys. 37, 317–336 (2008).

    Article  CAS  PubMed  Google Scholar 

  4. Feng, Z., Jia, B. & Zhang, M. Liquid–liquid phase separation in biology: specific stoichiometric molecular interactions vs promiscuous interactions mediated by disordered sequences. Biochemistry 60, 2397–2406 (2021).

    Article  CAS  PubMed  Google Scholar 

  5. Gorbalenya, A. E. & Koonin, E. V. Helicases: amino acid sequence comparisons and structure-function relationships. Curr. Opin. Struct. Biol. 3, 419–429 (1993).

    Article  CAS  Google Scholar 

  6. Linder, P. & Owttrim, G. W. Plant RNA helicases: linking aberrant and silencing RNA. Trends Plant. Sci. 14, 344–352 (2009).

    Article  CAS  PubMed  Google Scholar 

  7. Li, X. et al. Functions and mechanisms of RNA helicases in plants. J. Exp. Bot. 74, 2295–2310 (2009).

    Article  Google Scholar 

  8. Ariumi, Y. Host cellular RNA helicases regulate SARS-CoV-2 infection. J. Virol. 96, e0000222 (2022).

    Article  PubMed  Google Scholar 

  9. Ali, M. A. M. DEAD-box RNA helicases: the driving forces behind RNA metabolism at the crossroad of viral replication and antiviral innate immunity. Virus Res. 296, 198352 (2021).

    Article  CAS  PubMed  Google Scholar 

  10. Sloan, K. E. & Bohnsack, M. T. Unravelling the mechanisms of RNA helicase regulation. Trends Biochem. Sci. 43, 237–250 (2018).

    Article  CAS  PubMed  Google Scholar 

  11. Thoresen, D. et al. The molecular mechanism of RIG‐I activation and signaling. Immunol. Rev. 304, 154–168 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Putnam, A. A. & Jankowsky, E. DEAD-box helicases as integrators of RNA, nucleotide and protein binding. Biochim. Biophys. Acta Gene Regul. Mech. 1829, 884–893 (2013).

    Article  CAS  Google Scholar 

  13. De Bortoli, F., Espinosa, S. & Zhao, R. DEAH-Box RNA helicases in pre-mRNA splicing. Trends Biochem. Sci. 46, 225–238 (2021).

    Article  PubMed  Google Scholar 

  14. Donsbach, P. & Klostermeier, D. Regulation of RNA helicase activity: principles and examples. Biol. Chem. 402, 529–559 (2021).

    Article  CAS  PubMed  Google Scholar 

  15. Jankowsky, E. & Harris, M. E. Specificity and nonspecificity in RNA–protein interactions. Nat. Rev. Mol. Cell Biol. 16, 533–544 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Yadav, M. et al. The KH domain facilitates the substrate specificity and unwinding processivity of DDX43 helicase. J. Biol. Chem. 296, 100085 (2021).

    Article  CAS  PubMed  Google Scholar 

  17. Chen, M. C. et al. Structural basis of G-quadruplex unfolding by the DEAH/RHA helicase DHX36. Nature 558, 465–469 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Srinivasan, S., Liu, Z., Chuenchor, W., Xiao, T. S. & Jankowsky, E. Function of auxiliary domains of the DEAH/RHA helicase DHX36 in RNA remodeling. J. Mol. Biol. 432, 2217–2231 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Schult, P. & Paeschke, K. The DEAH helicase DHX36 and its role in G-quadruplex-dependent processes. Biol. Chem. 402, 581–591 (2021).

    Article  CAS  PubMed  Google Scholar 

  20. Lee, T. & Pelletier, J. The biology of DHX9 and its potential as a therapeutic target. Oncotarget 7, 42716–42739 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  21. Zhang, S. & Grosse, F. Multiple functions of nuclear DNA helicase II (RNA helicase A) in nucleic acid metabolism. Acta Biochim. Biophys. Sin. 36, 177–183 (2004).

    Article  CAS  PubMed  Google Scholar 

  22. Sharma, D., Putnam, A. A. & Jankowsky, E. Biochemical differences and similarities between the DEAD-Box helicase orthologs DDX3X and Ded1p. J. Mol. Biol. 429, 3730–3742 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Ozgur, S. et al. The conformational plasticity of eukaryotic RNA-dependent ATPases. FEBS J. 282, 850–863 (2015).

    Article  CAS  PubMed  Google Scholar 

  24. Gilman, B., Tijerina, P. & Russell, R. Distinct RNA-unwinding mechanisms of DEAD-box and DEAH-box RNA helicase proteins in remodeling structured RNAs and RNPs. Biochem. Soc. Trans. 45, 1313–1321 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Patrick, E. M., Srinivasan, S., Jankowsky, E. & Comstock, M. J. The RNA helicase Mtr4p is a duplex-sensing translocase. Nat. Chem. Biol. 13, 99–104 (2017).

    Article  CAS  PubMed  Google Scholar 

  26. Semlow, D. R., Blanco, M. R., Walter, N. G. & Staley, J. P. Spliceosomal DEAH-Box ATPases remodel Pre-mRNA to activate alternative splice sites. Cell 164, 985–998 (2016). This study describes winching by DEAH/RHA helicases during pre-mRNA splicing.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Kanaan, J. et al. UPF1-like helicase grip on nucleic acids dictates processivity. Nat. Commun. 9, 3752 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Enders, M., Ficner, R. & Adio, S. Regulation of the DEAH/RHA helicase Prp43 by the G-patch factor Pfa1. Proc. Natl Acad. Sci. USA 119, e2203567119 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Yang, Q. & Jankowsky, E. The DEAD-box protein Ded1 unwinds RNA duplexes by a mode distinct from translocating helicases. Nat. Struct. Mol. Biol. 13, 981–986 (2006).

    Article  CAS  PubMed  Google Scholar 

  30. Yang, Q., Del Campo, M., Lambowitz, A. M. & Jankowsky, E. DEAD-box proteins unwind duplexes by local strand separation. Mol. Cell 28, 253–263 (2007).

    Article  CAS  PubMed  Google Scholar 

  31. Chen, Y. et al. DEAD-box proteins can completely separate an RNA duplex using a single ATP. Proc. Natl Acad. Sci. USA 105, 20203–20208 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Liu, F., Putnam, A. & Jankowsky, E. ATP hydrolysis is required for DEAD-box protein recycling but not for duplex unwinding. Proc. Natl Acad. Sci. USA 105, 20209–20214 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Russell, R., Jarmoskaite, I. & Lambowitz, A. M. Toward a molecular understanding of RNA remodeling by DEAD-box proteins. RNA Biol. 10, 44–55 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Cruz, V. E. et al. Sequence-specific remodeling of a topologically complex RNP substrate by Spb4. Nat. Struct. Mol. Biol. 29, 1228–1238 (2022). This study shows how local strand unwinding by a DEAD-box protein causes long-range remodelling within a pre-ribosomal particle.

    Article  CAS  PubMed  Google Scholar 

  35. Jankowsky, E., Gross, C. H., Shuman, S. & Pyle, A. M. Active disruption of an RNA-protein interaction by a DExH/D RNA helicase. Science 291, 121–125 (2001).

    Article  CAS  PubMed  Google Scholar 

  36. Fairman, M. E. et al. Protein displacement by DExH/D ‘RNA helicases’ without duplex unwinding. Science 304, 730–734 (2004).

    Article  CAS  PubMed  Google Scholar 

  37. Liu, F., Putnam, A. A. & Jankowsky, E. DEAD-box helicases form nucleotide-dependent, long-lived complexes with RNA. Biochemistry 53, 423–433 (2014).

    Article  CAS  PubMed  Google Scholar 

  38. Andersen, C. B. F. et al. Structure of the exon junction core complex with a trapped DEAD-box ATPase bound to RNA. Science 313, 1968–1972 (2006).

    Article  CAS  PubMed  Google Scholar 

  39. Ballut, L. et al. The exon junction core complex is locked onto RNA by inhibition of eIF4AIII ATPase activity. Nat. Struct. Mol. Biol. 12, 861–869 (2005).

    Article  CAS  PubMed  Google Scholar 

  40. Le Hir, H., Saulière, J. & Wang, Z. The exon junction complex as a node of post-transcriptional networks. Nat. Rev. Mol. Cell Biol. 17, 41–54 (2016).

    Article  CAS  PubMed  Google Scholar 

  41. Karbstein, K. Attacking a DEAD problem: the role of DEAD-box ATPases in ribosome assembly and beyond. Methods Enzymol. 673, 19–38 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Vergani-Junior, C. A., Tonon-da-Silva, G., Inan, M. D. & Mori, M. A. DICER: structure, function, and regulation. Biophys. Rev. 13, 1081–1090 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Rawling, D. C. & Pyle, A. M. Parts, assembly and operation of the RIG-I family of motors. Curr. Opin. Struct. Biol. 25, 25–33 (2014).

    Article  CAS  PubMed  Google Scholar 

  44. Putnam, A. A. et al. Division of labor in an oligomer of the DEAD-Box RNA helicase Ded1p. Mol. Cell 59, 541–552 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Ma, W. K. & Tran, E. J. Measuring helicase inhibition of the DEAD-box protein Dbp2 by Yra1. Methods Mol. Biol. 1259, 183–197 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Rössler, O. G., Straka, A. & Stahl, H. Rearrangement of structured RNA via branch migration structures catalysed by the highly related DEAD-box proteins p68 and p72. Nucleic Acids Res. 29, 2088–2096 (2001).

    Article  PubMed  PubMed Central  Google Scholar 

  47. Yang, Q. & Jankowsky, E. ATP- and ADP-dependent modulation of RNA unwinding and strand annealing activities by the DEAD-box protein DED1. Biochemistry 44, 13591–13601 (2005).

    Article  CAS  PubMed  Google Scholar 

  48. Jarmoskaite, I. & Russell, R. RNA helicase proteins as chaperones and remodelers. Annu. Rev. Biochem. 83, 697–725 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Hondele, M. et al. DEAD-box ATPases are global regulators of phase-separated organelles. Nature 573, 144–148 (2019). This study reveals dynamic regulation of RNP granules by DEAD-box proteins.

    Article  CAS  PubMed  Google Scholar 

  50. Hondele, M., Heinrich, S., De Los Rios, P. & Weis, K. Membraneless organelles: phasing out of equilibrium. Emerg. Top. Life Sci. 4, 331–342 (2020).

    PubMed  Google Scholar 

  51. Fullam, A., Gu, L., Höhn, Y. & Schröder, M. DDX3 directly facilitates IKKα activation and regulates downstream signalling pathways. Biochem. J. 475, 3595–3607 (2018).

    Article  CAS  PubMed  Google Scholar 

  52. Dolde, C. et al. A CK1 FRET biosensor reveals that DDX3X is an essential activator of CK1ε. J. Cell Sci. 131, jcs207316 (2018).

    PubMed  PubMed Central  Google Scholar 

  53. Tanaka, K. et al. Knockdown of DEAD-box RNA helicase DDX5 selectively attenuates serine 311 phosphorylation of NF-κB p65 subunit and expression level of anti-apoptotic factor Bcl-2. Cell Signal. 65, 109428 (2020).

    Article  CAS  PubMed  Google Scholar 

  54. Collins, R. et al. The DEXD/H-box RNA helicase DDX19 is regulated by an α-helical switch. J. Biol. Chem. 284, 10296–10300 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Absmeier, E. et al. The large N-terminal region of the Brr2 RNA helicase guides productive spliceosome activation. Genes Dev. 29, 2576–2587 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Floor, S. N., Condon, K. J., Sharma, D., Jankowsky, E. & Doudna, J. A. Autoinhibitory interdomain interactions and subfamily-specific extensions redefine the catalytic core of the human DEAD-box protein DDX3. J. Biol. Chem. 291, 2412–2421 (2016).

    Article  CAS  PubMed  Google Scholar 

  57. Wojtas, M. N. et al. Regulation of m6A transcripts by the 3′ → 5′ RNA helicase YTHDC2 is essential for a successful meiotic program in the mammalian germline. Mol. Cell 68, 374–387.e12 (2017).

    Article  CAS  PubMed  Google Scholar 

  58. Kretschmer, J. et al. The m6A reader protein YTHDC2 interacts with the small ribosomal subunit and the 5′-3′ exoribonuclease XRN1. RNA 24, 1339–1350 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Mao, Y. et al. m6A in mRNA coding regions promotes translation via the RNA helicase-containing YTHDC2. Nat. Commun. 10, 5332 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  60. Luo, D. et al. Structural insights into RNA recognition by RIG-I. Cell 147, 409–422 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Wu, B. et al. Structural basis for dsRNA recognition, filament formation, and antiviral signal activation by MDA5. Cell 152, 276–289 (2013).

    Article  CAS  PubMed  Google Scholar 

  62. Rehwinkel, J. & Gack, M. U. RIG-I-like receptors: their regulation and roles in RNA sensing. Nat. Rev. Immunol. 20, 537–551 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Wilkinson, M. E., Charenton, C. & Nagai, K. RNA splicing by the spliceosome. Annu. Rev. Biochem. 89, 359–388 (2020).

    Article  CAS  PubMed  Google Scholar 

  64. Song, C., Hotz-Wagenblatt, A., Voit, R. & Grummt, I. SIRT7 and the DEAD-box helicase DDX21 cooperate to resolve genomic R loops and safeguard genome stability. Genes Dev. 31, 1370–1381 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Saito, M., Iestamantavicius, V., Hess, D. & Matthias, P. Monitoring acetylation of the RNA helicase DDX3X, a protein critical for formation of stress granules. Methods Mol. Biol. 2209, 217–234 (2021).

    Article  CAS  PubMed  Google Scholar 

  66. Kato, K. et al. Structural analysis of RIG-I-like receptors reveals ancient rules of engagement between diverse RNA helicases and TRIM ubiquitin ligases. Mol. Cell 81, 599–613.e8 (2021).

    Article  CAS  PubMed  Google Scholar 

  67. Kim, Y. K. & Maquat, L. E. UPFront and center in RNA decay: UPF1 in nonsense-mediated mRNA decay and beyond. RNA 25, 407–422 (2019). This study provides an excellent overview of the roles of UPF1 in NMD and in related RNA decay pathways.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Ponting, C. P. Novel eIF4G domain homologues linking mRNA translation with nonsense-mediated mRNA decay. Trends Biochem. Sci. 25, 423–426 (2000).

    Article  CAS  PubMed  Google Scholar 

  69. Aravind, L. & Koonin, E. V. G-patch: a new conserved domain in eukaryotic RNA-processing proteins and type D retroviral polyproteins. Trends Biochem. Sci. 24, 342–344 (1999).

    Article  CAS  PubMed  Google Scholar 

  70. Bohnsack, K. E., Ficner, R., Bohnsack, M. T. & Jonas, S. Regulation of DEAH-box RNA helicases by G-patch proteins. Biol. Chem. 402, 561–579 (2021).

    Article  CAS  PubMed  Google Scholar 

  71. Hamann, F. et al. Structural analysis of the intrinsically disordered splicing factor Spp2 and its binding to the DEAH-box ATPase Prp2. Proc. Natl Acad. Sci. USA 117, 2948–2956 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Studer, M. K., Ivanović, L., Weber, M. E., Marti, S. & Jonas, S. Structural basis for DEAH-helicase activation by G-patch proteins. Proc. Natl Acad. Sci. USA 117, 7159–7170 (2020). This study demonstrates the mechanism of DEAH/RHA helicase regulation by G-patch proteins.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Schütz, P. et al. Crystal structure of the yeast eIF4A-eIF4G complex: an RNA-helicase controlled by protein-protein interactions. Proc. Natl Acad. Sci. USA 105, 9564–9569 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  74. Hilbert, M., Kebbel, F., Gubaev, A. & Klostermeier, D. eIF4G stimulates the activity of the DEAD box protein eIF4A by a conformational guidance mechanism. Nucleic Acids Res. 39, 2260–2270 (2011).

    Article  CAS  PubMed  Google Scholar 

  75. Mathys, H. et al. Structural and biochemical insights to the role of the CCR4-NOT complex and DDX6 ATPase in microRNA repression. Mol. Cell 54, 751–765 (2014).

    Article  CAS  PubMed  Google Scholar 

  76. Alexandrov, A., Colognori, D. & Steitz, J. A. Human eIF4AIII interacts with an eIF4G-like partner, NOM1, revealing an evolutionarily conserved function outside the exon junction complex. Genes Dev. 25, 1078–1090 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Davila Gallesio, J., Hackert, P., Bohnsack, K. E. & Bohnsack, M. T. Sgd1 is an MIF4G domain-containing cofactor of the RNA helicase Fal1 and associates with the 5′ domain of the 18S rRNA sequence. RNA Biol. 17, 539–553 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Mugler, C. F. et al. ATPase activity of the DEAD-box protein Dhh1 controls processing body formation. eLife 5, e18746 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  79. Raisch, T., Sandmeir, F., Weichenrieder, O., Valkov, E. & Izaurralde, E. Structural and biochemical analysis of a NOT1 MIF4G-like domain of the CCR4-NOT complex. J. Struct. Biol. 204, 388–395 (2018).

    Article  CAS  PubMed  Google Scholar 

  80. Ozgur, S. et al. Structure of a human 4E-T/DDX6/CNOT1 complex reveals the different interplay of DDX6-binding proteins with the CCR4-NOT complex. Cell Rep. 13, 703–711 (2015).

    Article  CAS  PubMed  Google Scholar 

  81. Buchwald, G., Schüssler, S., Basquin, C., Le Hir, H. & Conti, E. Crystal structure of the human eIF4AIII-CWC22 complex shows how a DEAD-box protein is inhibited by a MIF4G domain. Proc. Natl Acad. Sci. USA 110, E4611–E4618 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Sharif, H. et al. Structural analysis of the yeast Dhh1-Pat1 complex reveals how Dhh1 engages Pat1, Edc3 and RNA in mutually exclusive interactions. Nucleic Acids Res. 41, 8377–8390 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Noble, C. G. & Song, H. MLN51 stimulates the RNA-helicase activity of eIF4AIII. PLoS ONE 2, e303 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  84. Weis, K. & Hondele, M. The role of DEAD-Box ATPases in gene expression and the regulation of RNA-protein condensates. Annu. Rev. Biochem. 91, 197–219 (2022).

    Article  PubMed  Google Scholar 

  85. Woodward, L. A., Mabin, J. W., Gangras, P. & Singh, G. The exon junction complex: a lifelong guardian of mRNA fate. Wiley Interdiscip. Rev. RNA 8, 1411 (2017).

    Article  Google Scholar 

  86. Weick, E.-M. & Lima, C. D. RNA helicases are hubs that orchestrate exosome-dependent 3′−5′ decay. Curr. Opin. Struct. Biol. 67, 86–94 (2021).

    Article  CAS  PubMed  Google Scholar 

  87. Skourti-Stathaki, K., Proudfoot, N. J. & Gromak, N. Human senataxin resolves RNA/DNA hybrids formed at transcriptional pause sites to promote Xrn2-dependent termination. Mol. Cell 42, 794–805 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Arndt, K. M. & Reines, D. Termination of transcription of short noncoding RNAs by RNA polymerase II. Annu. Rev. Biochem. 84, 381–404 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Rivosecchi, J. et al. Senataxin homologue Sen1 is required for efficient termination of RNA polymerase III transcription. EMBO J. 38, e101955 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  90. Jurga, M., Abugable, A. A., Goldman, A. S. H. & El-Khamisy, S. F. USP11 controls R-loops by regulating senataxin proteostasis. Nat. Commun. 12, 5156 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Abraham, K. J. et al. Nucleolar RNA polymerase II drives ribosome biogenesis. Nature 585, 298–302 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Porrua, O. & Libri, D. Transcription termination and the control of the transcriptome: why, where and how to stop. Nat. Rev. Mol. Cell Biol. 16, 190–202 (2015).

    Article  CAS  PubMed  Google Scholar 

  93. Xing, Z., Ma, W. K. & Tran, E. J. The DDX5/Dbp2 subfamily of DEAD-box RNA helicases. Wiley Interdiscip. Rev. RNA 10, e1519 (2019).

    Article  PubMed  Google Scholar 

  94. Ma, W. K. et al. Recruitment, duplex unwinding and protein-mediated inhibition of the dead-box RNA helicase Dbp2 at actively transcribed chromatin. J. Mol. Biol. 428, 1091–1106 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Cloutier, S. C. et al. Regulated formation of lncRNA-DNA hybrids enables faster transcriptional induction and environmental adaptation. Mol. Cell 61, 393–404 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Tedeschi, F. A., Cloutier, S. C., Tran, E. J. & Jankowsky, E. The DEAD-box protein Dbp2p is linked to noncoding RNAs, the helicase Sen1p, and R-loops. RNA 24, 1693–1705 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Cloutier, S. C., Ma, W. K., Nguyen, L. T. & Tran, E. J. The DEAD-box RNA helicase Dbp2 connects RNA quality control with repression of aberrant transcription. J. Biol. Chem. 287, 26155–26166 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Sessa, G. et al. BRCA2 promotes DNA-RNA hybrid resolution by DDX5 helicase at DNA breaks to facilitate their repair‡. EMBO J. 40, e106018 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Yu, Z. et al. DDX5 resolves R-loops at DNA double-strand breaks to promote DNA repair and avoid chromosomal deletions. NAR Cancer 2, zcaa028 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  100. Giraud, G., Terrone, S. & Bourgeois, C. F. Functions of DEAD box RNA helicases DDX5 and DDX17 in chromatin organization and transcriptional regulation. BMB Rep. 51, 613–622 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Fuller-Pace, F. V. The DEAD box proteins DDX5 (p68) and DDX17 (p72): multi-tasking transcriptional regulators. Biochim. Biophys. Acta 1829, 756–763 (2013).

    Article  CAS  PubMed  Google Scholar 

  102. Jones, K. et al. Reduction of toxic RNAs in myotonic dystrophies type 1 and type 2 by the RNA helicase p68/DDX5. Proc. Natl Acad. Sci. USA 112, 8041–8045 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Mazurek, A. et al. Acquired dependence of acute myeloid leukemia on the DEAD-box RNA helicase DDX5. Cell Rep. 7, 1887–1899 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Calo, E. et al. RNA helicase DDX21 coordinates transcription and ribosomal RNA processing. Nature 518, 249–253 (2015). This study describes RNA helicase-mediated transcription regulation.

    Article  CAS  PubMed  Google Scholar 

  105. Xing, Y.-H. et al. SLERT regulates DDX21 rings associated with Pol I transcription. Cell 169, 664–678.e16 (2017).

    Article  CAS  PubMed  Google Scholar 

  106. Gonzalez, G. A. & Montminy, M. R. Cyclic AMP stimulates somatostatin gene transcription by phosphorylation of CREB at serine 133. Cell 59, 675–680 (1989).

    Article  CAS  PubMed  Google Scholar 

  107. Nakajima, T. et al. RNA helicase A mediates association of CBP with RNA polymerase II. Cell 90, 1107–1112 (1997).

    Article  CAS  PubMed  Google Scholar 

  108. Anderson, S. F., Schlegel, B. P., Nakajima, T., Wolpin, E. S. & Parvin, J. D. BRCA1 protein is linked to the RNA polymerase II holoenzyme complex via RNA helicase A. Nat. Genet. 19, 254–256 (1998).

    Article  CAS  PubMed  Google Scholar 

  109. Tang, W. et al. RNA helicase A acts as a bridging factor linking nuclear β-actin with RNA polymerase II. Biochem. J. 420, 421–428 (2009).

    Article  CAS  PubMed  Google Scholar 

  110. Huo, L. et al. RNA helicase A is a DNA-binding partner for EGFR-mediated transcriptional activation in the nucleus. Proc. Natl Acad. Sci. USA 107, 16125–16130 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Yuan, W. et al. TDRD3 promotes DHX9 chromatin recruitment and R-loop resolution. Nucleic Acids Res. 49, 8573–8591 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Rajendran, R. R. et al. Regulation of nuclear receptor transcriptional activity by a novel DEAD box RNA helicase (DP97). J. Biol. Chem. 278, 4628–4638 (2003).

    Article  CAS  PubMed  Google Scholar 

  113. De, I., Schmitzová, J. & Pena, V. The organization and contribution of helicases to RNA splicing. Wiley Interdiscip. Rev. RNA 7, 259–274 (2016).

    Article  CAS  PubMed  Google Scholar 

  114. Cordin, O. & Beggs, J. D. RNA helicases in splicing. RNA Biol. 10, 83–95 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Ruby, S. W., Chang, T. H. & Abelson, J. Four yeast spliceosomal proteins (PRP5, PRP9, PRP11, and PRP21) interact to promote U2 snRNP binding to pre-mRNA. Genes Dev. 7, 1909–1925 (1993).

    Article  CAS  PubMed  Google Scholar 

  116. Kistler, A. L. & Guthrie, C. Deletion of MUD2, the yeast homolog of U2AF65, can bypass the requirement for sub2, an essential spliceosomal ATPase. Genes Dev. 15, 42–49 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Shen, H. et al. Distinct activities of the DExD/H-box splicing factor hUAP56 facilitate stepwise assembly of the spliceosome. Genes Dev. 22, 1796–1803 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Liang, W.-W. & Cheng, S.-C. A novel mechanism for Prp5 function in prespliceosome formation and proofreading the branch site sequence. Genes Dev. 29, 81–93 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  119. Beier, D. H. et al. Dynamics of the DEAD-box ATPase Prp5 RecA-like domains provide a conformational switch during spliceosome assembly. Nucleic Acids Res. 47, 10842–10851 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Perriman, R. J. & Ares, M. Rearrangement of competing U2 RNA helices within the spliceosome promotes multiple steps in splicing. Genes Dev. 21, 811–820 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Staley, J. P. & Guthrie, C. An RNA switch at the 5′ splice site requires ATP and the DEAD box protein Prp28p. Mol. Cell 3, 55–64 (1999).

    Article  CAS  PubMed  Google Scholar 

  122. Yeh, F.-L. et al. Activation of Prp28 ATPase by phosphorylated Npl3 at a critical step of spliceosome remodeling. Nat. Commun. 12, 3082 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Laggerbauer, B., Achsel, T. & Lührmann, R. The human U5-200kD DEXH-box protein unwinds U4/U6 RNA duplices in vitro. Proc. Natl Acad. Sci. USA 95, 4188–4192 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Raghunathan, P. L. & Guthrie, C. RNA unwinding in U4/U6 snRNPs requires ATP hydrolysis and the DEIH-box splicing factor Brr2. Curr. Biol. 8, 847–855 (1998).

    Article  CAS  PubMed  Google Scholar 

  125. Charenton, C., Wilkinson, M. E. & Nagai, K. Mechanism of 5′ splice site transfer for human spliceosome activation. Science 364, 362–367 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Theuser, M., Höbartner, C., Wahl, M. C. & Santos, K. F. Substrate-assisted mechanism of RNP disruption by the spliceosomal Brr2 RNA helicase. Proc. Natl Acad. Sci. USA 113, 7798–7803 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Liu, H.-L. & Cheng, S.-C. The interaction of Prp2 with a defined region of the intron is required for the first splicing reaction. Mol. Cell Biol. 32, 5056–5066 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Bao, P., Höbartner, C., Hartmuth, K. & Lührmann, R. Yeast Prp2 liberates the 5′ splice site and the branch site adenosine for catalysis of pre-mRNA splicing. RNA 23, 1770–1779 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Bai, R. et al. Mechanism of spliceosome remodeling by the ATPase/helicase Prp2 and its coactivator Spp2. Science 371, eabe8863 (2021).

    Article  CAS  PubMed  Google Scholar 

  130. Bessonov, S., Anokhina, M., Will, C. L., Urlaub, H. & Lührmann, R. Isolation of an active step I spliceosome and composition of its RNP core. Nature 452, 846–850 (2008).

    Article  CAS  PubMed  Google Scholar 

  131. Agafonov, D. E. et al. Semiquantitative proteomic analysis of the human spliceosome via a novel two-dimensional gel electrophoresis method. Mol. Cell Biol. 31, 2667–2682 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. De, I. et al. The RNA helicase Aquarius exhibits structural adaptations mediating its recruitment to spliceosomes. Nat. Struct. Mol. Biol. 22, 138–144 (2015).

    Article  CAS  PubMed  Google Scholar 

  133. Sales-Lee, J. et al. Coupling of spliceosome complexity to intron diversity. Curr. Biol. 31, 4898–4910.e4 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Will, C. L. et al. Characterization of novel SF3b and 17S U2 snRNP proteins, including a human Prp5p homologue and an SF3b DEAD-box protein. EMBO J. 21, 4978–4988 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Barbosa, I. et al. Human CWC22 escorts the helicase eIF4AIII to spliceosomes and promotes exon junction complex assembly. Nat. Struct. Mol. Biol. 19, 983–990 (2012).

    Article  CAS  PubMed  Google Scholar 

  136. Zhang, Z. et al. Structural insights into how Prp5 proofreads the pre-mRNA branch site. Nature 596, 296–300 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Zhan, X., Yan, C., Zhang, X., Lei, J. & Shi, Y. Structures of the human pre-catalytic spliceosome and its precursor spliceosome. Cell Res. 28, 1129–1140 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Strittmatter, L. M. et al. psiCLIP reveals dynamic RNA binding by DEAH-box helicases before and after exon ligation. Nat. Commun. 12, 1488 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Toroney, R., Nielsen, K. H. & Staley, J. P. Termination of pre-mRNA splicing requires that the ATPase and RNA unwindase Prp43p acts on the catalytic snRNA U6. Genes Dev. 33, 1555–1574 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Semlow, D. R. & Staley, J. P. Staying on message: ensuring fidelity in pre-mRNA splicing. Trends Biochem. Sci. 37, 263–273 (2012).

    Article  CAS  PubMed  Google Scholar 

  141. Maciejewski, J. P., Padgett, R. A., Brown, A. L. & Müller-Tidow, C. DDX41-related myeloid neoplasia. Semin. Hematol. 54, 94–97 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  142. Maxwell, D. W., O’Keefe, R. T., Roy, S. & Hentges, K. E. The role of splicing factors in retinitis pigmentosa: links to cilia. Biochem. Soc. Trans. 49, 1221–1231 (2021).

    Article  CAS  PubMed  Google Scholar 

  143. Strässer, K. & Hurt, E. Splicing factor Sub2p is required for nuclear mRNA export through its interaction with Yra1p. Nature 413, 648–652 (2001).

    Article  PubMed  Google Scholar 

  144. Strässer, K. et al. TREX is a conserved complex coupling transcription with messenger RNA export. Nature 417, 304–308 (2002).

    Article  PubMed  Google Scholar 

  145. Xie, Y. et al. Cryo-EM structure of the yeast TREX complex and coordination with the SR-like protein Gbp2.eLife 10, e65699 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Ren, Y., Schmiege, P. & Blobel, G. Structural and biochemical analyses of the DEAD-box ATPase Sub2 in association with THO or Yra1. eLife 6, e20070 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  147. Strässer, K. & Hurt, E. Yra1p, a conserved nuclear RNA-binding protein, interacts directly with Mex67p and is required for mRNA export. EMBO J. 19, 410–420 (2000).

    Article  PubMed  PubMed Central  Google Scholar 

  148. Snay-Hodge, C. A., Colot, H. V., Goldstein, A. L. & Cole, C. N. Dbp5p/Rat8p is a yeast nuclear pore-associated DEAD-box protein essential for RNA export. EMBO J. 17, 2663–2676 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Tran, E. J., Zhou, Y., Corbett, A. H. & Wente, S. R. The DEAD-box protein Dbp5 controls mRNA export by triggering specific RNA:protein remodeling events. Mol. Cell 28, 850–859 (2007).

    Article  CAS  PubMed  Google Scholar 

  150. Weirich, C. S. et al. Activation of the DExD/H-box protein Dbp5 by the nuclear-pore protein Gle1 and its coactivator InsP6 is required for mRNA export. Nat. Cell Biol. 8, 668–676 (2006).

    Article  CAS  PubMed  Google Scholar 

  151. Alcázar-Román, A. R., Tran, E. J., Guo, S. & Wente, S. R. Inositol hexakisphosphate and Gle1 activate the DEAD-box protein Dbp5 for nuclear mRNA export. Nat. Cell Biol. 8, 711–716 (2006).

    Article  PubMed  Google Scholar 

  152. Lund, M. K. & Guthrie, C. The DEAD-box protein Dbp5p is required to dissociate Mex67p from exported mRNPs at the nuclear rim. Mol. Cell 20, 645–651 (2005).

    Article  CAS  PubMed  Google Scholar 

  153. Montpetit, B. et al. A conserved mechanism of DEAD-box ATPase activation by nucleoporins and InsP6 in mRNA export. Nature 472, 238–242 (2011).

    Article  CAS  PubMed  Google Scholar 

  154. Wu, H., Becker, D. & Krebber, H. Telomerase RNA TLC1 shuttling to the cytoplasm requires mRNA export factors and is important for telomere maintenance. Cell Rep. 8, 1630–1638 (2014).

    Article  CAS  PubMed  Google Scholar 

  155. Neumann, B., Wu, H., Hackmann, A. & Krebber, H. Nuclear Export of Pre-Ribosomal Subunits Requires Dbp5, but Not as an RNA-Helicase as for mRNA Export. PLoS ONE 11, e0149571 (2016).

    Article  PubMed  Google Scholar 

  156. Arul Nambi Rajan, A. & Montpetit, B. Emerging molecular functions and novel roles for the DEAD-box protein Dbp5/DDX19 in gene expression. Cell Mol. Life Sci. 78, 2019–2030 (2021).

    Article  CAS  PubMed  Google Scholar 

  157. Hinnebusch, A. G., Ivanov, I. P. & Sonenberg, N. Translational control by 5′-untranslated regions of eukaryotic mRNAs. Science 352, 1413–1416 (2016).

    Article  CAS  PubMed  Google Scholar 

  158. Merrick, W. C. eIF4F: a retrospective. J. Biol. Chem. 290, 24091–24099 (2015).

    Article  CAS  PubMed  Google Scholar 

  159. Gao, Z. et al. Coupling between the DEAD-box RNA helicases Ded1p and eIF4A. eLife 5, e16408 (2016).

    Article  PubMed  Google Scholar 

  160. Loh, P. G. et al. Structural basis for translational inhibition by the tumour suppressor Pdcd4. EMBO J. 28, 274–285 (2009).

    Article  CAS  PubMed  Google Scholar 

  161. Pelletier, J. & Sonenberg, N. The organizing principles of eukaryotic ribosome recruitment. Annu. Rev. Biochem. 88, 307–335 (2019).

    Article  CAS  PubMed  Google Scholar 

  162. Brito Querido, J. et al. Structure of a human 48S translational initiation complex. Science 369, 1220–1227 (2020).

    Article  CAS  PubMed  Google Scholar 

  163. Sokabe, M. & Fraser, C. S. A helicase-independent activity of eIF4A in promoting mRNA recruitment to the human ribosome. Proc. Natl Acad. Sci. USA 114, 6304–6309 (2017).

    Article  CAS  PubMed  Google Scholar 

  164. Soto-Rifo, R. et al. DEAD-box protein DDX3 associates with eIF4F to promote translation of selected mRNAs. EMBO J. 31, 3745–3756 (2012).

    Article  CAS  PubMed  Google Scholar 

  165. Venkataramanan, S., Gadek, M., Calviello, L., Wilkins, K. & Floor, S. N. DDX3X and DDX3Y are redundant in protein synthesis. RNA 27, 1577–1588 (2021).

    Article  CAS  PubMed  Google Scholar 

  166. Sen, N. D. et al. Functional interplay between DEAD-box RNA helicases Ded1 and Dbp1 in preinitiation complex attachment and scanning on structured mRNAs in vivo. Nucleic Acids Res. 47, 8785–8806 (2019).

    CAS  PubMed  Google Scholar 

  167. Guenther, U.-P. et al. The helicase Ded1p controls use of near-cognate translation initiation codons in 5′ UTRs. Nature 559, 130–134 (2018). This study reveals how an RNA helicase influences the selection of translation initiation sites.

    Article  CAS  PubMed  Google Scholar 

  168. Calviello, L. et al. DDX3 depletion represses translation of mRNAs with complex 5′ UTRs. Nucleic Acids Res. 49, 5336–5350 (2021).

    Article  CAS  PubMed  Google Scholar 

  169. Sen, N. D., Zhou, F., Ingolia, N. T. & Hinnebusch, A. G. Genome-wide analysis of translational efficiency reveals distinct but overlapping functions of yeast DEAD-box RNA helicases Ded1 and eIF4A. Genome Res. 25, 1196–1205 (2015).

    Article  CAS  PubMed  Google Scholar 

  170. Dang, C. V. Sex life, and death in MYC-driven lymphomagenesis. Mol. Cell 81, 3886–3887 (2021).

    Article  CAS  PubMed  Google Scholar 

  171. Gong, C. et al. Sequential inverse dysregulation of the RNA helicases DDX3X and DDX3Y facilitates MYC-driven lymphomagenesis. Mol. Cell 81, 4059–4075.e11 (2021).

    Article  CAS  PubMed  Google Scholar 

  172. Linsalata, A. E. et al. DDX3X and specific initiation factors modulate FMR1 repeat-associated non-AUG-initiated translation. EMBO Rep. 20, e47498 (2019).

    Article  PubMed  Google Scholar 

  173. Iserman, C. et al. Condensation of Ded1p promotes a translational switch from housekeeping to stress protein production. Cell 181, 818–831.e19 (2020).

    Article  CAS  PubMed  Google Scholar 

  174. Lennox, A. L. et al. Pathogenic DDX3X mutations Impair RNA metabolism and neurogenesis during fetal cortical development. Neuron 106, 404–420.e8 (2020).

    Article  CAS  PubMed  Google Scholar 

  175. He, Y. et al. A double-edged function of DDX3, as an oncogene or tumor suppressor, in cancer progression (Review). Oncol. Rep. 39, 883–892 (2018).

    CAS  PubMed  Google Scholar 

  176. Xu, C., Cao, Y. & Bao, J. Building RNA-protein germ granules: insights from the multifaceted functions of DEAD-box helicase Vasa/Ddx4 in germline development. Cell Mol. Life Sci. 79, 4 (2021).

    Article  PubMed  Google Scholar 

  177. Sweeney, T. R. et al. Functional role and ribosomal position of the unique N-terminal region of DHX29, a factor required for initiation on structured mammalian mRNAs. Nucleic Acids Res. 49, 12955–12969 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Hashem, Y. et al. Structure of the mammalian ribosomal 43S preinitiation complex bound to the scanning factor DHX29. Cell 153, 1108–1119 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Pisareva, V. P., Pisarev, A. V., Komar, A. A., Hellen, C. U. T. & Pestova, T. V. Translation initiation on mammalian mRNAs with structured 5′UTRs requires DExH-box protein DHX29. Cell 135, 1237–1250 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Murat, P. et al. RNA G-quadruplexes at upstream open reading frames cause DHX36- and DHX9-dependent translation of human mRNAs. Genome Biol. 19, 229 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  181. Chen, X. et al. Translational control by DHX36 binding to 5′UTR G-quadruplex is essential for muscle stem-cell regenerative functions. Nat. Commun. 12, 5043 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Sauer, M. et al. DHX36 prevents the accumulation of translationally inactive mRNAs with G4-structures in untranslated regions. Nat. Commun. 10, 2421 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  183. Zhang, Y., You, J., Wang, X. & Weber, J. The DHX33 RNA helicase promotes mRNA translation initiation. Mol. Cell Biol. 35, 2918–2931 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Li, Z. et al. RNA-binding protein DDX1 is responsible for fatty acid-mediated repression of insulin translation. Nucleic Acids Res. 46, 12052–12066 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Guenther, U.-P. et al. IGHMBP2 is a ribosome-associated helicase inactive in the neuromuscular disorder distal SMA type 1 (DSMA1). Hum. Mol. Genet. 18, 1288–1300 (2009).

    Article  CAS  PubMed  Google Scholar 

  186. Lim, S. C., Bowler, M. W., Lai, T. F. & Song, H. The Ighmbp2 helicase structure reveals the molecular basis for disease-causing mutations in DMSA1. Nucleic Acids Res. 40, 11009–11022 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Radhakrishnan, A. et al. The DEAD-Box protein Dhh1p couples mRNA decay and translation by monitoring codon optimality. Cell 167, 122–132.e9 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Buschauer, R. et al. The Ccr4-Not complex monitors the translating ribosome for codon optimality. Science 368, eaay6912 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Hanson, G. & Coller, J. Codon optimality, bias and usage in translation and mRNA decay. Nat. Rev. Mol. Cell Biol. 19, 20–30 (2018).

    Article  CAS  PubMed  Google Scholar 

  190. Nott, A., Le Hir, H. & Moore, M. J. Splicing enhances translation in mammalian cells: an additional function of the exon junction complex. Genes Dev. 18, 210–222 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Tanabe, A. et al. RNA helicase YTHDC2 promotes cancer metastasis via the enhancement of the efficiency by which HIF-1α mRNA is translated. Cancer Lett. 376, 34–42 (2016).

    Article  CAS  PubMed  Google Scholar 

  192. Yuan, W. et al. The N6-methyladenosine reader protein YTHDC2 promotes gastric cancer progression via enhancing YAP mRNA translation. Transl. Oncol. 16, 101308 (2022).

    Article  CAS  PubMed  Google Scholar 

  193. Coller, J. M., Tucker, M., Sheth, U., Valencia-Sanchez, M. A. & Parker, R. The DEAD box helicase, Dhh1p, functions in mRNA decapping and interacts with both the decapping and deadenylase complexes. RNA 7, 1717–1727 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Wolin, S. L. & Maquat, L. E. Cellular RNA surveillance in health and disease. Science 366, 822–827 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Schneider, C. & Bohnsack, K. E. Caught in the act — visualizing ribonucleases during eukaryotic ribosome assembly. Wiley Interdiscip. Rev. RNA https://doi.org/10.1002/wrna.1766 (2022).

  196. Schmid, M. & Jensen, T. H. The nuclear RNA exosome and its cofactors. Adv. Exp. Med. Biol. 1203, 113–132 (2019).

    Article  CAS  PubMed  Google Scholar 

  197. Lingaraju, M. et al. The MTR4 helicase recruits nuclear adaptors of the human RNA exosome using distinct arch-interacting motifs. Nat. Commun. 10, 3393 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  198. Falk, S. et al. The molecular architecture of the TRAMP complex reveals the organization and interplay of its two catalytic activities. Mol. Cell 55, 856–867 (2014).

    Article  CAS  PubMed  Google Scholar 

  199. Puno, M. R. & Lima, C. D. Structural basis for MTR4-ZCCHC8 interactions that stimulate the MTR4 helicase in the nuclear exosome-targeting complex. Proc. Natl Acad. Sci. USA 115, E5506–E5515 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. Jia, H. et al. The RNA helicase Mtr4p modulates polyadenylation in the TRAMP complex. Cell 145, 890–901 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Weick, E.-M. et al. Helicase-dependent RNA decay illuminated by a Cryo-EM structure of a human nuclear RNA exosome-MTR4 complex. Cell 173, 1663–1677.e21 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Kögel, A., Keidel, A., Bonneau, F., Schäfer, I. B. & Conti, E. The human SKI complex regulates channeling of ribosome-bound RNA to the exosome via an intrinsic gatekeeping mechanism. Mol. Cell 82, 756–769.e8 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  203. Zinoviev, A., Ayupov, R. K., Abaeva, I. S., Hellen, C. U. T. & Pestova, T. V. Extraction of mRNA from stalled ribosomes by the Ski complex. Mol. Cell 77, 1340–1349.e6 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Fabre, A. et al. Novel mutations in TTC37 associated with tricho-hepato-enteric syndrome. Hum. Mutat. 32, 277–281 (2011).

    Article  CAS  PubMed  Google Scholar 

  205. Eckard, S. C. et al. The SKIV2L RNA exosome limits activation of the RIG-I-like receptors. Nat. Immunol. 15, 839–845 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  206. May, G. E. & McManus, C. J. High-throughput quantitation of yeast uORF regulatory impacts using FACS-uORF. Methods Mol. Biol. 2404, 331–351 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  207. Kurosaki, T., Popp, M. W. & Maquat, L. E. Quality and quantity control of gene expression by nonsense-mediated mRNA decay. Nat. Rev. Mol. Cell Biol. 20, 406–420 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Park, O. H. et al. Identification and molecular characterization of cellular factors required for glucocorticoid receptor-mediated mRNA decay. Genes Dev. 30, 2093–2105 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Elbarbary, R. A., Miyoshi, K., Hedaya, O., Myers, J. R. & Maquat, L. E. UPF1 helicase promotes TSN-mediated miRNA decay. Genes Dev. 31, 1483–1493 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Hug, N. & Cáceres, J. F. The RNA helicase DHX34 activates NMD by promoting a transition from the surveillance to the decay-inducing complex. Cell Rep. 8, 1845–1856 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Melero, R. et al. The RNA helicase DHX34 functions as a scaffold for SMG1-mediated UPF1 phosphorylation. Nat. Commun. 7, 10585 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  212. Nawaz, A., Shilikbay, T., Skariah, G. & Ceman, S. Unwinding the roles of RNA helicase MOV10. Wiley Interdiscip. Rev. RNA 13, e1682 (2022).

    Article  CAS  PubMed  Google Scholar 

  213. Hanet, A. et al. HELZ directly interacts with CCR4-NOT and causes decay of bound mRNAs. Life Sci. Alliance 2, e201900405 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  214. Klinge, S. & Woolford, J. L. Ribosome assembly coming into focus. Nat. Rev. Mol. Cell Biol. 20, 116–131 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Bohnsack, K. E. & Bohnsack, M. T. Uncovering the assembly pathway of human ribosomes and its emerging links to disease. EMBO J. 38, e100278 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  216. Martin, R., Straub, A. U., Doebele, C. & Bohnsack, M. T. DExD/H-box RNA helicases in ribosome biogenesis. RNA Biol. 10, 4–18 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Rodríguez-Galán, O., García-Gómez, J. J. & de la Cruz, J. Yeast and human RNA helicases involved in ribosome biogenesis: current status and perspectives. Biochim. Biophys. Acta 1829, 775–790 (2013).

    Article  PubMed  Google Scholar 

  218. Aquino, G. R. R. et al. The RNA helicase Dbp7 promotes domain V/VI compaction and stabilization of inter-domain interactions during early 60S assembly. Nat. Commun. 12, 6152 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  219. Brüning, L. et al. RNA helicases mediate structural transitions and compositional changes in pre-ribosomal complexes. Nat. Commun. 9, 5383 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  220. Martin, R. et al. A pre-ribosomal RNA interaction network involving snoRNAs and the Rok1 helicase. RNA 20, 1173–1182 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Bohnsack, M. T. et al. Prp43 bound at different sites on the pre-rRNA performs distinct functions in ribosome synthesis. Mol. Cell 36, 583–592 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Choudhury, P., Kretschmer, J., Hackert, P., Bohnsack, K. E. & Bohnsack, M. T. The DExD box ATPase DDX55 is recruited to domain IV of the 28S ribosomal RNA by its C-terminal region. RNA Biol. 18, 1124–1135 (2021).

    Article  CAS  PubMed  Google Scholar 

  223. Manikas, R.-G., Thomson, E., Thoms, M. & Hurt, E. The K+-dependent GTPase Nug1 is implicated in the association of the helicase Dbp10 to the immature peptidyl transferase centre during ribosome maturation. Nucleic Acids Res. 44, 1800–1812 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  224. Sanghai, Z. A. et al. Modular assembly of the nucleolar pre-60S ribosomal subunit. Nature 556, 126–129 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. Kater, L. et al. Visualizing the assembly pathway of nucleolar Pre-60S ribosomes. Cell 171, 1599–1610.e14 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  226. Zhou, D. et al. Cryo-EM structure of an early precursor of large ribosomal subunit reveals a half-assembled intermediate. Protein Cell 10, 120–130 (2019).

    Article  CAS  PubMed  Google Scholar 

  227. Cheng, J. et al. 90S pre-ribosome transformation into the primordial 40S subunit. Science 369, 1470–1476 (2020).

    Article  CAS  PubMed  Google Scholar 

  228. Du, Y. et al. Cryo-EM structure of 90S small ribosomal subunit precursors in transition states. Science 369, 1477–1481 (2020).

    Article  CAS  PubMed  Google Scholar 

  229. Bohnsack, M. T., Kos, M. & Tollervey, D. Quantitative analysis of snoRNA association with pre-ribosomes and release of snR30 by Rok1 helicase. EMBO Rep. 9, 1230–1236 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  230. Aquino, G. R. R. et al. RNA helicase-mediated regulation of snoRNP dynamics on pre-ribosomes and rRNA 2′-O-methylation. Nucleic Acids Res. 49, 4066–4084 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  231. Sardana, R. et al. The DEAH-box helicase Dhr1 dissociates U3 from the pre-rRNA to promote formation of the central pseudoknot. PLoS Biol. 13, e1002083 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  232. Jaafar, M. et al. Association of snR190 snoRNA chaperone with early pre-60S particles is regulated by the RNA helicase Dbp7 in yeast. Nat. Commun. 12, 6153 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  233. Khoshnevis, S. et al. The DEAD-box protein Rok1 orchestrates 40S and 60S ribosome assembly by promoting the release of Rrp5 from pre-40S ribosomes to allow for 60S maturation. PLoS Biol. 14, e1002480 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  234. Dembowski, J. A., Kuo, B. & Woolford, J. L. Has1 regulates consecutive maturation and processing steps for assembly of 60S ribosomal subunits. Nucleic Acids Res. 41, 7889–7904 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  235. Khreiss, A. et al. The DEAD-box protein Dbp6 is an ATPase and RNA annealase interacting with the peptidyl transferase center (PTC) of the ribosome. Nucleic Acids Res. 51, 744–764 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  236. Memet, I., Doebele, C., Sloan, K. E. & Bohnsack, M. T. The G-patch protein NF-κB-repressing factor mediates the recruitment of the exonuclease XRN2 and activation of the RNA helicase DHX15 in human ribosome biogenesis. Nucleic Acids Res. 45, 5359–5374 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  237. Boneberg, F. M. et al. Molecular mechanism of the RNA helicase DHX37 and its activation by UTP14A in ribosome biogenesis. RNA 25, 685–701 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  238. Choudhury, P., Hackert, P., Memet, I., Sloan, K. E. & Bohnsack, M. T. The human RNA helicase DHX37 is required for release of the U3 snoRNP from pre-ribosomal particles. RNA Biol. 16, 54–68 (2019).

    Article  PubMed  Google Scholar 

  239. Bohnsack, K. E., Kanwal, N. & Bohnsack, M. T. Prp43/DHX15 exemplify RNA helicase multifunctionality in the gene expression network. Nucleic Acids Res. 50, 9012–9022 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  240. Calo, E. et al. Tissue-selective effects of nucleolar stress and rDNA damage in developmental disorders. Nature 554, 112–117 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  241. Sloan, K. E. et al. The association of late-acting snoRNPs with human pre-ribosomal complexes requires the RNA helicase DDX21. Nucleic Acids Res. 43, 553–564 (2015).

    Article  CAS  PubMed  Google Scholar 

  242. Kim, D.-S. et al. Activation of PARP-1 by snoRNAs controls ribosome biogenesis and cell growth via the RNA helicase DDX21. Mol. Cell 75, 1270–1285.e14 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  243. Carthew, R. W. & Sontheimer, E. J. Origins and mechanisms of miRNAs and siRNAs. Cell 136, 642–655 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  244. Soifer, H. S. et al. A role for the Dicer helicase domain in the processing of thermodynamically unstable hairpin RNAs. Nucleic Acids Res. 36, 6511–6522 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  245. Cenik, E. S. et al. Phosphate and R2D2 restrict the substrate specificity of Dicer-2, an ATP-driven ribonuclease. Mol. Cell 42, 172–184 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  246. Welker, N. C. et al. Dicer’s helicase domain discriminates dsRNA termini to promote an altered reaction mode. Mol. Cell 41, 589–599 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  247. Suzuki, H. I. et al. Modulation of microRNA processing by p53. Nature 460, 529–533 (2009).

    Article  CAS  PubMed  Google Scholar 

  248. Davis, B. N., Hilyard, A. C., Lagna, G. & Hata, A. SMAD proteins control DROSHA-mediated microRNA maturation. Nature 454, 56–61 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  249. Gregory, R. I. et al. The Microprocessor complex mediates the genesis of microRNAs. Nature 432, 235–240 (2004).

    Article  CAS  PubMed  Google Scholar 

  250. Lambert, M.-P. et al. The RNA helicase DDX17 controls the transcriptional activity of REST and the expression of proneural microRNAs in neuronal differentiation. Nucleic Acids Res. 46, 7686–7700 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  251. Popow, J., Jurkin, J., Schleiffer, A. & Martinez, J. Analysis of orthologous groups reveals archease and DDX1 as tRNA splicing factors. Nature 511, 104–107 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  252. Kroupova, A. et al. Molecular architecture of the human tRNA ligase complex. eLife 10, e71656 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  253. Razew, M. et al. Structural analysis of mtEXO mitochondrial RNA degradosome reveals tight coupling of nuclease and helicase components. Nat. Commun. 9, 97 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  254. Golik, P., Szczepanek, T., Bartnik, E., Stepien, P. P. & Lazowska, J. The S. cerevisiae nuclear gene SUV3 encoding a putative RNA helicase is necessary for the stability of mitochondrial transcripts containing multiple introns. Curr. Genet. 28, 217–224 (1995).

    Article  CAS  PubMed  Google Scholar 

  255. Huang, H.-R. et al. The splicing of yeast mitochondrial group I and group II introns requires a DEAD-box protein with RNA chaperone function. Proc. Natl Acad. Sci. USA 102, 163–168 (2005).

    Article  CAS  PubMed  Google Scholar 

  256. Séraphin, B., Simon, M., Boulet, A. & Faye, G. Mitochondrial splicing requires a protein from a novel helicase family. Nature 337, 84–87 (1989).

    Article  PubMed  Google Scholar 

  257. Mallam, A. L., Del Campo, M., Gilman, B., Sidote, D. J. & Lambowitz, A. M. Structural basis for RNA-duplex recognition and unwinding by the DEAD-box helicase Mss116p. Nature 490, 121–125 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  258. De Silva, D. et al. The DEAD-box helicase Mss116 plays distinct roles in mitochondrial ribogenesis and mRNA-specific translation. Nucleic Acids Res. 45, 6628–6643 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  259. Markov, D. A., Wojtas, I. D., Tessitore, K., Henderson, S. & McAllister, W. T. Yeast DEAD box protein Mss116p is a transcription elongation factor that modulates the activity of mitochondrial RNA polymerase. Mol. Cell Biol. 34, 2360–2369 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  260. De Silva, D., Fontanesi, F. & Barrientos, A. The DEAD box protein Mrh4 functions in the assembly of the mitochondrial large ribosomal subunit. Cell Metab. 18, 712–725 (2013).

    Article  PubMed  Google Scholar 

  261. Tu, Y.-T. & Barrientos, A. The human mitochondrial DEAD-box protein DDX28 resides in RNA granules and functions in mitoribosome assembly. Cell Rep. 10, 854–864 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  262. Antonicka, H. & Shoubridge, E. A. Mitochondrial RNA granules are centers for posttranscriptional RNA processing and ribosome biogenesis. Cell Rep. 10, 920–932 (2015).

    Article  CAS  PubMed  Google Scholar 

  263. Bosco, B. et al. DHX30 coordinates cytoplasmic translation and mitochondrial function contributing to cancer cell survival. Cancers 13, 4412 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  264. Cruz-Zaragoza, L. D. et al. An in vitro system to silence mitochondrial gene expression. Cell 184, 5824–5837.e15 (2021).

    Article  CAS  PubMed  Google Scholar 

  265. Yoneyama, M. et al. Shared and unique functions of the DExD/H-box helicases RIG-I, MDA5, and LGP2 in antiviral innate immunity. J. Immunol. 175, 2851–2858 (2005).

    Article  CAS  PubMed  Google Scholar 

  266. Ren, X., Linehan, M. M., Iwasaki, A. & Pyle, A. M. RIG-I Selectively discriminates against 5′-monophosphate RNA. Cell Rep. 26, 2019–2027.e4 (2019).

    Article  CAS  PubMed  Google Scholar 

  267. Dickey, T. H., Song, B. & Pyle, A. M. RNA binding activates RIG-I by releasing an autorepressed signaling domain. Sci. Adv. 5, eaax3641 (2019). This study provides mechanistic insights into RIG-I-mediated recognition of viral RNA.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  268. Rawling, D. C., Fitzgerald, M. E. & Pyle, A. M. Establishing the role of ATP for the function of the RIG-I innate immune sensor. eLife 4, e09391 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  269. Devarkar, S. C., Schweibenz, B., Wang, C., Marcotrigiano, J. & Patel, S. S. RIG-I uses an ATPase-powered translocation-throttling mechanism for kinetic proofreading of RNAs and oligomerization. Mol. Cell 72, 355–368.e4 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  270. Lässig, C. et al. Unified mechanisms for self-RNA recognition by RIG-I Singleton-Merten syndrome variants. eLife 7, e38958 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  271. Chiang, J. J. et al. Viral unmasking of cellular 5S rRNA pseudogene transcripts induces RIG-I-mediated immunity. Nat. Immunol. 19, 53–62 (2018).

    Article  CAS  PubMed  Google Scholar 

  272. Oda, H. et al. Aicardi-Goutières syndrome is caused by IFIH1 mutations. Am. J. Hum. Genet. 95, 121–125 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  273. Rice, G. I. et al. Gain-of-function mutations in IFIH1 cause a spectrum of human disease phenotypes associated with upregulated type I interferon signaling. Nat. Genet. 46, 503–509 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  274. Duic, I. et al. Viral RNA recognition by LGP2 and MDA5, and activation of signaling through step-by-step conformational changes. Nucleic Acids Res. 48, 11664–11674 (2020).

    Article  CAS  PubMed  Google Scholar 

  275. Wang, Y. et al. Mitochondria-localised ZNFX1 functions as a dsRNA sensor to initiate antiviral responses through MAVS. Nat. Cell Biol. 21, 1346–1356 (2019).

    Article  CAS  PubMed  Google Scholar 

  276. Vance, R. E. & Cytosolic, D. N. A. Sensing: the field narrows. Immunity 45, 227–228 (2016).

    Article  CAS  PubMed  Google Scholar 

  277. Alberti, S. & Hyman, A. A. Biomolecular condensates at the nexus of cellular stress, protein aggregation disease and ageing. Nat. Rev. Mol. Cell Biol. 22, 196–213 (2021).

    Article  CAS  PubMed  Google Scholar 

  278. Alberti, S., Gladfelter, A. & Mittag, T. Considerations and challenges in studying liquid-liquid phase separation and biomolecular condensates. Cell 176, 419–434 (2019).

    Article  CAS  PubMed  Google Scholar 

  279. Lafontaine, D. L. J., Riback, J. A., Bascetin, R. & Brangwynne, C. P. The nucleolus as a multiphase liquid condensate. Nat. Rev. Mol. Cell Biol. 22, 165–182 (2021).

    Article  CAS  PubMed  Google Scholar 

  280. Boeynaems, S. et al. Protein phase separation: a new phase in cell biology. Trends Cell Biol. 28, 420–435 (2018).

    Article  CAS  PubMed  Google Scholar 

  281. Van Treeck, B. & Parker, R. Emerging roles for intermolecular RNA-RNA interactions in RNP assemblies. Cell 174, 791–802 (2018).

    Article  PubMed  Google Scholar 

  282. Jain, S. et al. ATPase-modulated stress granules contain a diverse proteome and substructure. Cell 164, 487–498 (2016).

    Article  CAS  PubMed  Google Scholar 

  283. Hubstenberger, A. et al. P-Body purification reveals the condensation of repressed mRNA regulons. Mol. Cell 68, 144–157.e5 (2017).

    Article  CAS  PubMed  Google Scholar 

  284. Sachdev, R. et al. Pat1 promotes processing body assembly by enhancing the phase separation of the DEAD-box ATPase Dhh1 and RNA. eLife 8, e41415 (2019).

    Article  PubMed  Google Scholar 

  285. Tauber, D. et al. Modulation of RNA condensation by the DEAD-box protein eIF4A. Cell 180, 411–426.e16 (2020). This study shows how eIF4A can antagonize the formation of RNP granules.

    Article  CAS  PubMed  Google Scholar 

  286. Tauber, D., Tauber, G. & Parker, R. Mechanisms and regulation of RNA condensation in RNP granule formation. Trends Biochem. Sci. 45, 764–778 (2020).

    Article  CAS  PubMed  Google Scholar 

  287. Banani, S. F., Lee, H. O., Hyman, A. A. & Rosen, M. K. Biomolecular condensates: organizers of cellular biochemistry. Nat. Rev. Mol. Cell Biol. 18, 285–298 (2017).

    Article  CAS  PubMed  Google Scholar 

  288. Valentin-Vega, Y. A. et al. Cancer-associated DDX3X mutations drive stress granule assembly and impair global translation. Sci. Rep. 6, 25996 (2016).

    Article  CAS  PubMed  Google Scholar 

  289. Lessel, D. et al. De novo missense mutations in DHX30 impair global translation and cause a neurodevelopmental disorder. Am. J. Hum. Genet. 101, 716–724 (2017).

    Article  CAS  PubMed  Google Scholar 

  290. Balak, C. et al. Rare de novo missense variants in RNA helicase DDX6 cause intellectual disability and dysmorphic features and lead to P-body defects and RNA dysregulation. Am. J. Hum. Genet. 105, 509–525 (2019).

    Article  CAS  PubMed  Google Scholar 

  291. Alberti, S. & Dormann, D. Liquid-liquid phase separation in disease. Annu. Rev. Genet. 53, 171–194 (2019).

    Article  CAS  PubMed  Google Scholar 

  292. Hubstenberger, A., Noble, S. L., Cameron, C. & Evans, T. C. Translation repressors, an RNA helicase, and developmental cues control RNP phase transitions during early development. Dev. Cell 27, 161–173 (2013).

    Article  CAS  PubMed  Google Scholar 

  293. Bordeleau, M.-E. et al. Functional characterization of IRESes by an inhibitor of the RNA helicase eIF4A. Nat. Chem. Biol. 2, 213–220 (2006).

    Article  CAS  PubMed  Google Scholar 

  294. Bordeleau, M.-E. et al. Stimulation of mammalian translation initiation factor eIF4A activity by a small molecule inhibitor of eukaryotic translation. Proc. Natl Acad. Sci. USA 102, 10460–10465 (2005).

    Article  CAS  PubMed  Google Scholar 

  295. Naineni, S. K. et al. Functional mimicry revealed by the crystal structure of an eIF4A:RNA complex bound to the interfacial inhibitor, desmethyl pateamine A. Cell Chem. Biol. 28, 825–834.e6 (2021).

    Article  CAS  PubMed  Google Scholar 

  296. Iwasaki, S. et al. The translation inhibitor rocaglamide targets a bimolecular cavity between eIF4A and polypurine RNA. Mol. Cell 73, 738–748.e9 (2019). This study demonstrates the molecular basis of eIF4A inhibition by rocaglamide.

    Article  CAS  PubMed  Google Scholar 

  297. Chu, J. et al. Amidino-rocaglates: a potent class of eIF4A inhibitors. Cell Chem. Biol. 26, 1586–1593.e3 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  298. Gerson-Gurwitz, A. et al. Zotatifin, an eIF4A-selective inhibitor, blocks tumor growth in receptor tyrosine kinase driven tumors. Front. Oncol. 11, 766298 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  299. Ernst, J. T. et al. Design of development candidate eFT226, a first in class inhibitor of eukaryotic initiation factor 4A RNA helicase. J. Med. Chem. 63, 5879–5955 (2020).

    Article  CAS  PubMed  Google Scholar 

  300. Chen, M. et al. Dual targeting of DDX3 and eIF4A by the translation inhibitor rocaglamide A. Cell Chem. Biol. 28, 475–486.e8 (2021).

    Article  CAS  PubMed  Google Scholar 

  301. Iwasaki, S., Floor, S. N. & Ingolia, N. T. Rocaglates convert DEAD-box protein eIF4A into a sequence-selective translational repressor. Nature 534, 558–561 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  302. Shen, L. & Pelletier, J. Selective targeting of the DEAD-box RNA helicase eukaryotic initiation factor (eIF) 4A by natural products. Nat. Prod. Rep. 37, 609–616 (2020).

    Article  CAS  PubMed  Google Scholar 

  303. Naineni, S. K. et al. A comparative study of small molecules targeting eIF4A. RNA 26, 541–549 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  304. Floor, S. N., Barkovich, K. J., Condon, K. J., Shokat, K. M. & Doudna, J. A. Analog sensitive chemical inhibition of the DEAD-box protein DDX3. Protein Sci. 25, 638–649 (2016).

    Article  CAS  PubMed  Google Scholar 

  305. Barkovich, K. J., Moore, M. K., Hu, Q. & Shokat, K. M. Chemical genetic inhibition of DEAD-box proteins using covalent complementarity. Nucleic Acids Res. 46, 8689–8699 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  306. Kost, G. C. et al. A novel anti-cancer agent, 1-(3,5-dimethoxyphenyl)-4-[(6-fluoro-2-methoxyquinoxalin-3-yl)aminocarbonyl] piperazine (RX-5902), interferes with β-catenin function through Y593 phospho-p68 RNA helicase. J. Cell. Biochem. 116, 1595–1601 (2015).

    Article  CAS  PubMed  Google Scholar 

  307. Iwatani-Yoshihara, M. et al. Discovery and characterization of a eukaryotic initiation factor 4A-3-selective inhibitor that suppresses nonsense-mediated mRNA decay. ACS Chem. Biol. 12, 1760–1768 (2017).

    Article  CAS  PubMed  Google Scholar 

  308. Yoneyama-Hirozane, M. et al. High-throughput screening to identify inhibitors of DEAD box helicase DDX41. SLAS Discov. 22, 1084–1092 (2017).

    Article  CAS  PubMed  Google Scholar 

  309. Marecki, J. C., Belachew, B., Gao, J. & Raney, K. D. RNA helicases required for viral propagation in humans. Enzymes 50, 335–367 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  310. Richard, P., Feng, S. & Manley, J. L. A SUMO-dependent interaction between Senataxin and the exosome, disrupted in the neurodegenerative disease AOA2, targets the exosome to sites of transcription-induced DNA damage. Genes Dev. 27, 2227–2232 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  311. Chen, Y.-Z. et al. DNA/RNA helicase gene mutations in a form of juvenile amyotrophic lateral sclerosis (ALS4). Am. J. Hum. Genet. 74, 1128–1135 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  312. Hirano, M. et al. Senataxin mutations and amyotrophic lateral sclerosis. Amyotroph. Lateral Scler. 12, 223–227 (2011).

    Article  CAS  PubMed  Google Scholar 

  313. Moreira, M.-C. et al. Senataxin, the ortholog of a yeast RNA helicase, is mutant in ataxia-ocular apraxia 2. Nat. Genet. 36, 225–227 (2004).

    Article  CAS  PubMed  Google Scholar 

  314. Chen, Y.-Z. et al. Senataxin, the yeast Sen1p orthologue: characterization of a unique protein in which recessive mutations cause ataxia and dominant mutations cause motor neuron disease. Neurobiol. Dis. 23, 97–108 (2006).

    Article  CAS  PubMed  Google Scholar 

  315. Alzahrani, F. et al. Recessive, deleterious variants in SMG8 expand the role of nonsense-mediated decay in developmental disorders in humans. Am. J. Hum. Genet. 107, 1178–1185 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  316. Sun, Y., Eshov, A., Zhou, J., Isiktas, A. U. & Guo, J. U. C9orf72 arginine-rich dipeptide repeats inhibit UPF1-mediated RNA decay via translational repression. Nat. Commun. 11, 3354 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  317. Yaojia Cheng, Y. X. et al. Aberrant expression of the UPF1 RNA surveillance gene disturbs keratinocyte homeostasis by stabilizing. Areg. Int. J. Mol. Med. 45, 1163–1175 (2020).

    PubMed  Google Scholar 

  318. Cottenie, E. et al. Truncating and missense mutations in IGHMBP2 cause Charcot-Marie Tooth disease type 2. Am. J. Hum. Genet. 95, 590–601 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  319. Grohmann, K. et al. Mutations in the gene encoding immunoglobulin mu-binding protein 2 cause spinal muscular atrophy with respiratory distress type 1. Nat. Genet. 29, 75–77 (2001).

    Article  CAS  PubMed  Google Scholar 

  320. Guenther, U. P. et al. Genomic rearrangements at the IGHMBP2 gene locus in two patients with SMARD1. Hum. Genet. 115, 319–326 (2004).

    Article  PubMed  Google Scholar 

  321. Guenther, U.-P. et al. Clinical and mutational profile in spinal muscular atrophy with respiratory distress (SMARD): defining novel phenotypes through hierarchical cluster analysis. Hum. Mutat. 28, 808–815 (2007).

    Article  CAS  PubMed  Google Scholar 

  322. Guenther, U.-P. et al. Clinical variability in distal spinal muscular atrophy type 1 (DSMA1): determination of steady-state IGHMBP2 protein levels in five patients with infantile and juvenile disease. J. Mol. Med. 87, 31–41 (2009).

    Article  CAS  PubMed  Google Scholar 

  323. Saladini, M. et al. Spinal muscular atrophy with respiratory distress type 1: clinical phenotypes, molecular pathogenesis and therapeutic insights. J. Cell Mol. Med. 24, 1169–1178 (2020).

    Article  CAS  PubMed  Google Scholar 

  324. Favaro, F. P. et al. A noncoding expansion in EIF4A3 causes Richieri-Costa-Pereira syndrome, a craniofacial disorder associated with limb defects. Am. J. Hum. Genet. 94, 120–128 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  325. Paine, I. et al. Paralog studies augment gene discovery: DDX and DHX genes. Am. J. Hum. Genet. 105, 302–316 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  326. Snijders Blok, L. et al. Mutations in DDX3X are a common cause of unexplained intellectual disability with gender-specific effects on Wnt signaling. Am. J. Hum. Genet. 97, 343–352 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  327. Johnson-Kerner, B. et al. in GeneReviews (eds. Adam, M. P. et al.) (Univ. Washington, 1993).

  328. Toriello, H. V. et al. Toriello-Carey syndrome: delineation and review. Am. J. Med. Genet. A 123A, 84–90 (2003).

    Article  PubMed  Google Scholar 

  329. Dikow, N. et al. DDX3X mutations in two girls with a phenotype overlapping Toriello-Carey syndrome. Am. J. Med. Genet. A 173, 1369–1373 (2017).

    Article  CAS  PubMed  Google Scholar 

  330. Zhang, L. et al. RNA helicase p68 inhibits the transcription and post-transcription of Pkd1 in ADPKD. Theranostics 10, 8281–8297 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  331. Laurent, F.-X. et al. New function for the RNA helicase p68/DDX5 as a modifier of MBNL1 activity on expanded CUG repeats. Nucleic Acids Res. 40, 3159–3171 (2012).

    Article  CAS  PubMed  Google Scholar 

  332. Burns, W. et al. Syndromic neurodevelopmental disorder associated with de novo variants in DDX23. Am. J. Med. Genet. A 185, 2863–2872 (2021).

    Article  CAS  PubMed  Google Scholar 

  333. Shamseldin, H. E. et al. Mutations in DDX59 implicate RNA helicase in the pathogenesis of orofaciodigital syndrome. Am. J. Hum. Genet. 93, 555–560 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  334. Latif, Z. et al. Confirmation of the role of DHX38 in the etiology of early-onset retinitis pigmentosa. Invest. Ophthalmol. Vis. Sci. 59, 4552–4557 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  335. da Silva, T. E. et al. Genetic evidence of the association of DEAH-box helicase 37 defects with 46,XY gonadal dysgenesis spectrum. J. Clin. Endocrinol. Metab. 104, 5923–5934 (2019).

    Article  PubMed  Google Scholar 

  336. McElreavey, K. et al. Pathogenic variants in the DEAH-box RNA helicase DHX37 are a frequent cause of 46,XY gonadal dysgenesis and 46,XY testicular regression syndrome. Genet. Med. 22, 150–159 (2020).

    Article  CAS  PubMed  Google Scholar 

  337. Karaca, E. et al. Genes that affect brain structure and function identified by rare variant analyses of Mendelian neurologic disease. Neuron 88, 499–513 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  338. Oved, J. H. et al. Human mutational constraint as a tool to understand biology of rare and emerging bone marrow failure syndromes. Blood Adv. 4, 5232–5245 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  339. Castillo, A. et al. 19q13.32 microdeletion syndrome: three new cases. Eur. J. Med. Genet. 57, 654–658 (2014).

    Article  PubMed  Google Scholar 

  340. Jang, M.-A. et al. Mutations in DDX58, which encodes RIG-I, cause atypical Singleton-Merten syndrome. Am. J. Hum. Genet. 96, 266–274 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  341. Rutsch, F. et al. A specific IFIH1 gain-of-function mutation causes Singleton-Merten syndrome. Am. J. Hum. Genet. 96, 275–282 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  342. Hu, H., Yang, H., Liu, Y. & Yan, B. Pathogenesis of anti-melanoma differentiation-associated gene 5 antibody-positive dermatomyositis: a concise review with an emphasis on type I interferon system. Front. Med. 8, 833114 (2021).

    Article  Google Scholar 

  343. Nombel, A., Fabien, N. & Coutant, F. Dermatomyositis with anti-MDA5 antibodies: bioclinical features, pathogenesis and emerging therapies. Front. Immunol. 12, 773352 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  344. Smyth, D. J. et al. A genome-wide association study of nonsynonymous SNPs identifies a type 1 diabetes locus in the interferon-induced helicase (IFIH1) region. Nat. Genet. 38, 617–619 (2006).

    Article  CAS  PubMed  Google Scholar 

  345. Schultz, K. A. P. et al. DICER1 and associated conditions: identification of at-risk individuals and recommended surveillance strategies. Clin. Cancer Res. 24, 2251–2261 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  346. de Kock, L., Wu, M. K. & Foulkes, W. D. Ten years of DICER1 mutations: provenance, distribution, and associated phenotypes. Hum. Mutat. 40, 1939–1953 (2019).

    Article  PubMed  Google Scholar 

  347. Hill, D. A. et al. DICER1 mutations in familial pleuropulmonary blastoma. Science 325, 965 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  348. Foulkes, W. D. et al. Extending the phenotypes associated with DICER1 mutations. Hum. Mutat. 32, 1381–1384 (2011).

    Article  CAS  PubMed  Google Scholar 

  349. Rio Frio, T. et al. DICER1 mutations in familial multinodular goiter with and without ovarian Sertoli-Leydig cell tumors. J. Am. Med. Assoc. 305, 68–77 (2011).

    Article  CAS  Google Scholar 

  350. Klein, S. et al. Expanding the phenotype of mutations in DICER1: mosaic missense mutations in the RNase IIIb domain of DICER1 cause GLOW syndrome. J. Med. Genet. 51, 294–302 (2014).

    Article  CAS  PubMed  Google Scholar 

  351. Fabre, A. et al. SKIV2L mutations cause syndromic diarrhea, or trichohepatoenteric syndrome. Am. J. Hum. Genet. 90, 689–692 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  352. Li, N., Mei, H., MacDonald, I. M., Jiao, X. & Hejtmancik, J. F. Mutations in ASCC3L1 on 2q11.2 are associated with autosomal dominant retinitis pigmentosa in a Chinese family. Invest. Ophthalmol. Vis. Sci. 51, 1036–1043 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  353. Zhao, C. et al. Autosomal-dominant retinitis pigmentosa caused by a mutation in SNRNP200, a gene required for unwinding of U4/U6 snRNAs. Am. J. Hum. Genet. 85, 617–627 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  354. Bowne, S. J. et al. Mutations in the small nuclear riboprotein 200 kDa gene (SNRNP200) cause 1.6% of autosomal dominant retinitis pigmentosa. Mol. Vis. 19, 2407–2417 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank J. D. Gallesio for help with preparation of figures. The authors acknowledge funding by the Deutsche Forschungsgemeinschaft (DFG) via SFB1565 (Projektnummer 469281184; P06 to M.T.B. and P12 to K.E.B.). Work in the Jankowsky group is funded by the NIH (GM118088–05).

Author information

Authors and Affiliations

Authors

Contributions

K.E.B., M.T.B. and E.J. conceptualized the article, and all authors contributed to preparing the manuscript, figures and tables.

Corresponding authors

Correspondence to Katherine E. Bohnsack, Eckhard Jankowsky or Markus T. Bohnsack.

Ethics declarations

Competing interests

E.J. is an employee and owns equity of Moderna. The other authors declare no competing interests.

Peer review

Peer review information

Nature Reviews Molecular Cell Biology thanks Katrin Karbstein, Vladimir Pena, and Stephen Floor with Jess Sheu-Gruttadauria, 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.

Supplementary information

Glossary

43S pre-initiation complex

(43S PIC). A ribonucleoprotein complex containing the small ribosomal subunit and translation initiation factors.

ATPase cycle

Cycle in which ATP bound by an enzyme is hydrolysed, ADP and Pi are released before ATP re-associates, and the steps are repeated.

Branch point

Nucleotide located within a short motif upstream of the 3′ splice site; the branch point nucleotide initiates a nucleophilic attack on the 5′ splice site to form an intron lariat.

Caspase activation and recruitment domains

(CARDs). Protein domains that mediate protein–protein interactions to activate signalling pathways in apoptosis, inflammation and innate immunity.

Codon optimality

The use of codons for optimal translation efficiency; a balance is required between the demand for different aminoacylated tRNAs by translating ribosomes and their availability in the cytoplasm.

Exon junction complex

(EJC). A protein complex that assembles 20 to 24 nucleotides upstream of exon–exon junctions during pre-mRNA splicing and remains bound to the mature mRNA to regulate export, translation and nonsense-mediated decay. The EJC is ejected during the first round of protein synthesis.

Exosome

A protein complex that degrades RNAs through endoribonucleolytic and 3′−5′ exoribonucleolytic activity.

Glucocorticoid receptor-mediated decay

mRNA degradation pathway in which the glucocorticoid receptor binds an mRNA transcript and, upon binding to glucocorticoid, recruits PNRC2 and UPF1 to initiate mRNA degradation.

Histone mRNA decay

Degradation pathway specific to histone mRNAs, which lack poly(A) tails. UPF1 is recruited to a conserved stem–loop structure in the 3′ untranslated region of histone mRNAs by a stem–loop binding protein and initiates histone mRNA degradation.

Intrinsically disordered regions

(IDRs). Protein regions without a defined secondary or tertiary structure.

Liquid–liquid phase separation

(LLPS). Phase separation (de-mixing) of protein, RNA and/or other macromolecules resulting in formation of dynamic liquid condensates, which can function as membraneless cellular compartments.

Long non-coding RNAs

A class of RNAs that are longer than 200 nucleotides and generally do not encode proteins.

Nonsense-mediated mRNA decay

(NMD). mRNA degradation pathway in which an mRNA with a premature termination codon is recognized and degraded.

Regnase 1-mediated decay

mRNA degradation pathway of immune homeostasis in which regnase 1 is recruited to a 3′ untranslated region decay element and, together with UPF1, initiates mRNA degradation.

R-loops

Nucleic acid structures consisting of a DNA–RNA hybrid and a displaced single strand of DNA; commonly associated with the function of RNA polymerases.

RNA chaperones

Proteins that facilitate RNA folding into appropriate secondary or tertiary structures by preventing assembly of aberrant structures, promoting formation of correct structures, displacing bound proteins and/or resolving misfolding.

RNA-induced silencing complexes

(RISCs). Heterogeneous ribonucleoprotein complexes, each containing a microRNA and Argonaute protein, that regulate mRNA expression at transcriptional and translational levels.

RNA remodelling

Dynamic alteration of the secondary or tertiary structure of RNA molecules.

Small nuclear RNA

(snRNA). A class of nuclear non-coding RNAs of 100–1200 nucleotides that form spliceosome complexes with partnering proteins.

Small nucleolar RNAs

(snoRNAs). A class of nucleolar non-coding RNAs of 60–300 nucleotides that guide methylation, pseudouridylation and acetylation of ribosomal (and other) RNAs, or that chaperone RNA folding.

Staufen-mediated decay

mRNA degradation pathway in which the double-stranded RNA-binding protein Staufen binds to a stem–loop structure in the mRNA 3′ untranslated region and recruits UPF1 to initiate mRNA degradation.

TRAMP

(TRF4–TRF45–AIR2–MTR4 polyadenylation). A protein complex that appends short polyadenine tails to the 3′ end of several non-coding RNAs.

TSN-mediated miRNA decay

MicroRNA (miRNA) degradation pathway in which the endoribonuclease translin (TSN) degrades a specific subset of miRNAs. UPF1 promotes dissociation of miRNAs from their targets, thereby rendering the miRNAs more susceptible to TSN-mediated miRNA decay.

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

Bohnsack, K.E., Yi, S., Venus, S. et al. Cellular functions of eukaryotic RNA helicases and their links to human diseases. Nat Rev Mol Cell Biol 24, 749–769 (2023). https://doi.org/10.1038/s41580-023-00628-5

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41580-023-00628-5

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer