1932

Abstract

Detection of double-stranded RNAs (dsRNAs) is a central mechanism of innate immune defense in many organisms. We here discuss several families of dsRNA-binding proteins involved in mammalian antiviral innate immunity. These include RIG-I-like receptors, protein kinase R, oligoadenylate synthases, adenosine deaminases acting on RNA, RNA interference systems, and other proteins containing dsRNA-binding domains and helicase domains. Studies suggest that their functions are highly interdependent and that their interdependence could offer keys to understanding the complex regulatory mechanisms for cellular dsRNA homeostasis and antiviral immunity. This review aims to highlight their interconnectivity, as well as their commonalities and differences in their dsRNA recognition mechanisms.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-immunol-042718-041356
2019-04-26
2024-04-25
Loading full text...

Full text loading...

/deliver/fulltext/immunol/37/1/annurev-immunol-042718-041356.html?itemId=/content/journals/10.1146/annurev-immunol-042718-041356&mimeType=html&fmt=ahah

Literature Cited

  1. 1.
    Neidle S 2008. Principles of Nucleic Acid Structure London: Academic
  2. 2.
    Gesteland RF, Cech TR, Atkins JF 1999. The RNA World Cold Spring Harbor, NY: Cold Spring Harb. Lab. Press. , 2nd ed..
  3. 3.
    Seeman NC, Rosenberg JM, Rich A 1976. Sequence-specific recognition of double helical nucleic acids by proteins. PNAS 73:804–8
    [Google Scholar]
  4. 4.
    Rohs R, Jin X, West SM, Joshi R, Honig B, Mann RS 2010. Origins of specificity in protein-DNA recognition. Annu. Rev. Biochem. 79:233–69
    [Google Scholar]
  5. 5.
    Masliah G, Barraud P, Allain FH 2013. RNA recognition by double-stranded RNA binding domains: a matter of shape and sequence. Cell Mol. Life Sci. 70:1875–95
    [Google Scholar]
  6. 6.
    Peisley A, Lin C, Bin W, Orme-Johnson M, Liu M et al. 2011. Cooperative assembly and dynamic disassembly of MDA5 filaments for viral dsRNA recognition. PNAS 108:21010–15
    [Google Scholar]
  7. 7.
    Berke IC, Modis Y 2012. MDA5 cooperatively forms dimers and ATP-sensitive filaments upon binding double-stranded RNA. EMBO J 7:1714–26
    [Google Scholar]
  8. 8.
    Ahmad S, Mu X, Yang F, Greenwald E, Park JW et al. 2018. Breaching self-tolerance to Alu duplex RNA underlies MDA5-mediated inflammation. Cell 172:797–810This paper describes a method to identify RNA ligand for MDA5 using protein conformation instead of affinity.
    [Google Scholar]
  9. 9.
    Mu X, Greenwald E, Ahmad S, Hur S 2018. An origin of the immunogenicity of in vitro transcribed RNA. Nucleic Acids Res 46:5239–49This paper describes a common dsRNA by-product of in vitro transcription.
    [Google Scholar]
  10. 10.
    Schlee M, Roth A, Hornung V, Hagmann CA, Wimmenauer V et al. 2009. Recognition of 5′ triphosphate by RIG-I helicase requires short blunt double-stranded RNA as contained in panhandle of negative-strand virus. Immunity 31:25–34
    [Google Scholar]
  11. 11.
    Heinicke LA, Wong CJ, Lary J, Nallagatla SR, Diegelman-Parente A et al. 2009. RNA dimerization promotes PKR dimerization and activation. J. Mol. Biol 390:319–38This paper describes a common phenomenon of hairpin RNA dimerization and its impact on PKR activity analysis.
    [Google Scholar]
  12. 12.
    Hull CM, Bevilacqua PC 2016. Discriminating self and non-self by RNA: roles for RNA structure, misfolding, and modification in regulating the innate immune sensor PKR. Acc. Chem. Res. 49:1242–49
    [Google Scholar]
  13. 13.
    Liu ZR, Wilkie AM, Clemens MJ, Smith CW 1996. Detection of double-stranded RNA-protein interactions by methylene blue-mediated photo-crosslinking. RNA 2:611–21
    [Google Scholar]
  14. 14.
    Ricci EP, Kucukural A, Cenik C, Mercier BC, Singh G et al. 2014. Staufen1 senses overall transcript secondary structure to regulate translation. Nat. Struct. Mol. Biol. 21:26–35
    [Google Scholar]
  15. 15.
    Sugimoto Y, Vigilante A, Darbo E, Zirra A, Militti C et al. 2015. hiCLIP reveals the in vivo atlas of mRNA secondary structures recognized by Staufen 1. Nature 519:491–94
    [Google Scholar]
  16. 16.
    Moy RH, Cole BS, Yasunaga A, Gold B, Shankarling G et al. 2014. Stem-loop recognition by DDX17 facilitates miRNA processing and antiviral defense. Cell 158:764–77
    [Google Scholar]
  17. 17.
    Wu B, Hur S 2015. How RIG-I like receptors activate MAVS. Curr. Opin. Virol. 12:91–98
    [Google Scholar]
  18. 18.
    Kato H, Takahasi K, Fujita T 2011. RIG-I-like receptors: cytoplasmic sensors for non-self RNA. Immunol. Rev. 243:91–98
    [Google Scholar]
  19. 19.
    Venkataraman T, Valdes M, Elsby R, Kakuta S, Caceres G et al. 2007. Loss of DExD/H box RNA helicase LGP2 manifests disparate antiviral responses. J. Immunol. 178:6444–55
    [Google Scholar]
  20. 20.
    Childs KS, Randall RE, Goodbourn S 2013. LGP2 plays a critical role in sensitizing mda-5 to activation by double-stranded RNA. PLOS ONE 8:e64202
    [Google Scholar]
  21. 21.
    Satoh T, Kato H, Kumagain Y, Yoneyama M, Sato S et al. 2010. LGP2 is a positive regulator of RIG-I– and MDA5-mediated antiviral responses. PNAS 107:1512–17
    [Google Scholar]
  22. 22.
    Parisien JP, Lenoir JJ, Mandhana R, Rodriguez KR, Qian K et al. 2018. RNA sensor LGP2 inhibits TRAF ubiquitin ligase to negatively regulate innate immune signaling. EMBO Rep 19:e45176
    [Google Scholar]
  23. 23.
    Peisley A, Jo M, Lin C, Wu B, Orme-Johnson M et al. 2012. Kinetic mechanism for viral dsRNA length discrimination by MDA5 filament. PNAS 109:E3340–49
    [Google Scholar]
  24. 24.
    Feng Q, Hato SV, Langereis MA, Zoll J, Virgen-Slane R et al. 2012. MDA5 detects the double-stranded RNA replicative form in picornavirus-infected cells. Cell Rep 29:1187–96
    [Google Scholar]
  25. 25.
    Kowalinski E, Lunardi T, McCarthy AA, Louber J, Brunel J et al. 2011. Structural basis for the activation of innate immune pattern-recognition receptor RIG-I by viral RNA. Cell 147:423–35
    [Google Scholar]
  26. 26.
    Luo D, Kohlway A, Vela A, Pyle AM 2012. Visualizing the determinants of viral RNA recognition by innate immune sensor RIG-I. Structure 20:1983–88
    [Google Scholar]
  27. 27.
    Jiang F, Ramanathan A, Miller MT, Tang G-Q, Gale M Jr et al. 2011. Structural basis of RNA recognition and activation by innate immune receptor RIG-I. Nature 479:423–27
    [Google Scholar]
  28. 28.
    Rawling DC, Fitzgerald ME, Pyle AM 2015. Establishing the role of ATP for the function of the RIG-I innate immune sensor. eLife 4:e03931
    [Google Scholar]
  29. 29.
    Sohn J, Hur S 2016. Filament assemblies in foreign nucleic acid sensors. Curr. Opin. Struct. Biol. 37:134–44
    [Google Scholar]
  30. 30.
    Peisley A, Wu B, Yao H, Walz T, Hur S 2013. RIG-I forms signaling-competent filaments in an ATP-dependent, ubiquitin-independent manner. Mol. Cell 51:573–83
    [Google Scholar]
  31. 31.
    Patel JR, Jain A, Chou YY, Baum A, Ha T, Garcia-Sastre A 2013. ATPase-driven oligomerization of RIG-I on RNA allows optimal activation of type-I interferon. EMBO Rep 14:780–87
    [Google Scholar]
  32. 32.
    Myong S, Cui S, Cornish PV, Kirchhofer A, Gack MU et al. 2009. Cytosolic viral sensor RIG-I is a 5′-triphosphate–dependent translocase on double-stranded RNA. Science 323:1070–74
    [Google Scholar]
  33. 33.
    Devarkar SC, Schweibenz B, Wang C, Marcotrigiano J, Patel SS 2018. RIG-I uses an ATPase-powered translocation-throttling mechanism for kinetic proofreading of RNAs and oligomerization. Mol. Cell 72:355–68.e354
    [Google Scholar]
  34. 34.
    Jiang X, Kinch LN, Brautigam CA, Chen X, Du F et al. 2012. Ubiquitin-induced oligomerization of the RNA sensors RIG-I and MDA5 activates antiviral innate immune response. Immunity 36:959–73
    [Google Scholar]
  35. 35.
    Peisley A, Wu B, Xu H, Chen ZJ, Hur S 2014. Structural basis for ubiquitin-mediated antiviral signal activation by RIG-I. Nature 509:110–14
    [Google Scholar]
  36. 36.
    Wu B, Peisley A, Tetrault D, Li Z, Egelman EH et al. 2014. Molecular imprinting as a signal activation mechanism of the viral RNA sensor RIG-I. Mol. Cell 55:511–23
    [Google Scholar]
  37. 37.
    Hou F, Sun L, Zheng H, Skaug B, Jiang QX, Chen ZJ 2011. MAVS forms functional prion-like aggregates to activate and propagate antiviral innate immune response. Cell 146:448–61
    [Google Scholar]
  38. 38.
    Triantafilou K, Vakakis E, Kar S, Richer E, Evans GL, Triantafilou M 2012. Visualisation of direct interaction of MDA5 and the dsRNA replicative intermediate form of positive strand RNA viruses. J. Cell Sci. 125:4761–69
    [Google Scholar]
  39. 39.
    Uchikawa E, Lethier M, Malet H, Brunel J, Gerlier D, Cusack S 2016. Structural analysis of dsRNA binding to anti-viral pattern recognition receptors LGP2 and MDA5. Mol. Cell 62:586–602
    [Google Scholar]
  40. 40.
    Li X, Ranjith-Kumar CT, Brooks MT, Dharmaiah S, Herr AB et al. 2009. The RIG-I-like receptor LGP2 recognizes the termini of double-stranded RNA. J. Biol. Chem. 284:13881–91
    [Google Scholar]
  41. 41.
    Bruns AM, Pollpeter D, Hadizadeh N, Myong S, Marko JF, Horvath CM 2013. ATP hydrolysis enhances RNA recognition and antiviral signal transduction by the innate immune sensor, laboratory of genetics and physiology 2 (LGP2). J. Biol. Chem. 288:938–46
    [Google Scholar]
  42. 42.
    Bruns AM, Leser GP, Lamb RA, Horvath CM 2014. The innate immune sensor LGP2 activates antiviral signaling by regulating MDA5-RNA interaction and filament assembly. Mol. Cell 55:771–81
    [Google Scholar]
  43. 43.
    Funabiki M, Kato H, Miyachi Y, Toki H, Motegi H et al. 2014. Autoimmune disorders associated with gain of function of the intracellular sensor MDA5. Immunity 40:199–212
    [Google Scholar]
  44. 44.
    Rice GI, del Toro Duany Y, Jenkinson EM, Forte GM, Anderson BH et al. 2014. Gain-of-function mutations in IFIH1 cause a spectrum of human disease phenotypes associated with upregulated type I interferon signaling. Nat. Genet. 46:503–9
    [Google Scholar]
  45. 45.
    Van Eyck L, De Somer L, Pombal D, Bornschein S, Frans G et al. 2015. IFIH1 mutation causes systemic lupus erythematosus with selective IgA deficiency. Arthritis Rheumatol 67:1592–97
    [Google Scholar]
  46. 46.
    Bursztejn AC, Briggs TA, del Toro Duany Y, Anderson BH, O'Sullivan J et al. 2015. Unusual cutaneous features associated with a heterozygous gain-of-function mutation in IFIH1: overlap between Aicardi-Goutières and Singleton-Merten syndromes. Br. J. Dermatol. 173:1505–13
    [Google Scholar]
  47. 47.
    Oda H, Nakagawa A, Abe J, Awaya T, Funabiki M et al. 2014. Aicardi-Goutières syndrome is caused by IFIH1 mutations. Am. J. Hum. Genet. 95:121–25
    [Google Scholar]
  48. 48.
    Chung H, Calis JJA, Wu X, Sun T, Yu Y et al. 2018. Human ADAR1 prevents endogenous RNA from triggering translational shutdown. Cell 172:811–24.e14
    [Google Scholar]
  49. 49.
    Chiappinelli KB, Strissel PL, Desrichard A, Li H, Henke C et al. 2015. Inhibiting DNA methylation causes an interferon response in cancer via dsRNA including endogenous retroviruses. Cell 162:974–86
    [Google Scholar]
  50. 50.
    Roulois D, Loo Yau H, Singhania R, Wang Y, Danesh A et al. 2015. DNA-demethylating agents target colorectal cancer cells by inducing viral mimicry by endogenous transcripts. Cell 162:961–73
    [Google Scholar]
  51. 51.
    Dhir A, Dhir S, Borowski LS, Jimenez L, Teitell M et al. 2018. Mitochondrial double-stranded RNA triggers antiviral signalling in humans. Nature 560:238–42
    [Google Scholar]
  52. 52.
    Jang M-A, Kim EK, Now H, Nguyen NTH, Kim W-J et al. 2015. Mutations in DDX58, which encodes RIG-I, cause atypical Singleton-Merten syndrome. Am. J. Hum. Genet. 96:266–74
    [Google Scholar]
  53. 53.
    Ferreira CR, Crow YJ, Gahl WA, Goldbach-Mansky R, Hur S et al. 2018. DDX58 and classic Singleton-Merten syndrome. J. Clin. Immnol. In press
    [Google Scholar]
  54. 54.
    Ranoa DR, Parekh AD, Pitroda SP, Huang X, Darga T et al. 2016. Cancer therapies activate RIG-I-like receptor pathway through endogenous non-coding RNAs. Oncotarget 7:26496–515
    [Google Scholar]
  55. 55.
    Nabet BY, Qiu Y, Shabason JE, Wu TJ, Yoon T et al. 2017. Exosome RNA unshielding couples stromal activation to pattern recognition receptor signaling in cancer. Cell 170:352–66.e13
    [Google Scholar]
  56. 56.
    Boelens MC, Wu TJ, Nabet BY, Xu B, Qiu Y et al. 2014. Exosome transfer from stromal to breast cancer cells regulates therapy resistance pathways. Cell 159:499–513
    [Google Scholar]
  57. 57.
    Chiang JJ, Sparrer KMJ, van Gent M, Lassig C, Huang T et al. 2018. Viral unmasking of cellular 5S rRNA pseudogene transcripts induces RIG-I-mediated immunity. Nat. Immunol. 19:53–62
    [Google Scholar]
  58. 58.
    Malathi K, Dong B, Gale M Jr, Silverman RH 2007. Small self-RNA generated by RNase L amplifies antiviral innate immunity. Nature 448:816–19
    [Google Scholar]
  59. 59.
    Malathi K, Saito T, Crochet N, Barton DJ, Gale M Jr, Silverman RH 2010. RNase L releases a small RNA from HCV RNA that refolds into a potent PAMP. RNA 16:2108–19
    [Google Scholar]
  60. 60.
    Eckard SC, Rice GI, Fabre A, Badens C, Gray EE et al. 2014. The SKIV2L RNA exosome limits activation of the RIG-I-like receptors. Nat. Immunol. 15:839–45
    [Google Scholar]
  61. 61.
    Yao H, Dittmann M, Peisley A, Hoffmann H-H, Gilmore RH et al. 2015. ATP-dependent effector-like functions of RIG-I-like receptors. Mol. Cell 58:541–48
    [Google Scholar]
  62. 62.
    Sato S, Li K, Kameyama K, Hayashi T, Ishida Y et al. 2015. The RNA sensor RIG-I dually functions as an innate sensor and direct antiviral factor for hepatitis B virus. Immunity 42:123–32
    [Google Scholar]
  63. 63.
    Weber M, Sediri H, Felgenhauer U, Binzen I, Banfer S et al. 2015. Influenza virus adaptation PB2–627K modulates nucleocapsid inhibition by the pathogen sensor RIG-I. Cell Host Microbe 17:309–19
    [Google Scholar]
  64. 64.
    Benitez AA, Panis M, Xue J, Varble A, Shim JV et al. 2015. In vivo RNAi screening identifies MDA5 as a significant contributor to the cellular defense against influenza A virus. Cell Rep 11:1714–26
    [Google Scholar]
  65. 65.
    van der Veen AG, Maillard PV, Schmidt JM, Lee SA, Deddouche-Grass S et al. 2018. The RIG-I-like receptor LGP2 inhibits Dicer-dependent processing of long double-stranded RNA and blocks RNA interference in mammalian cells. EMBO J 37:e97479
    [Google Scholar]
  66. 66.
    Takahashi T, Nakano Y, Onomoto K, Murakami F, Komori C et al. 2018. LGP2 virus sensor regulates gene expression network mediated by TRBP-bound microRNAs. Nucleic Acids Res 46:9134–47
    [Google Scholar]
  67. 67.
    Pfeller CK, Li Z, George CX, Samuel CE 2011. Protein kinase PKR and RNA adenosine deaminase ADAR1: new roles for old players as modulators of the interferon response. Curr. Opin. Immunol. 23:573–82
    [Google Scholar]
  68. 68.
    Nanduri S, Carpick BW, Yang Y, Williams BR, Qin J 1998. Structure of the double-stranded RNA-binding domain of the protein kinase PKR reveals the molecular basis of its dsRNA-mediated activation. EMBO J 17:5458–65
    [Google Scholar]
  69. 69.
    Dey M, Cao C, Dar AC, Tamura T, Ozato K et al. 2005. Mechanistic link between PKR dimerization, autophosphorylation, and eIF2α substrate recognition. Cell 122:901–13
    [Google Scholar]
  70. 70.
    Gleghorn ML, Maquat LE 2014. ‘Black sheep’ that don't leave the double-stranded RNA-binding domain fold. Trends Biochem. Sci. 39:328–40This review provides a comprehensive overview of dsRBD structures and functions.
    [Google Scholar]
  71. 71.
    Bevilacqua PC, Cech TR 1996. Minor-groove recognition of double-stranded RNA by the double-stranded RNA-binding domain from the RNA-activated protein kinase PKR. Biochemistry 35:9983–94
    [Google Scholar]
  72. 72.
    Kim I, Liu CW, Puglisi JD 2006. Specific recognition of HIV TAR RNA by the dsRNA binding domains (dsRBD1–dsRBD2) of PKR. J. Mol. Biol. 358:430–42
    [Google Scholar]
  73. 73.
    Ucci JW, Kobayashi Y, Choi G, Alexandrescu AT, Cole JL 2007. Mechanism of interaction of the double-stranded RNA (dsRNA) binding domain of protein kinase R with short dsRNA sequences. Biochemistry 46:55–65
    [Google Scholar]
  74. 74.
    Husain B, Mukerji I, Cole JL 2012. Analysis of high-affinity binding of protein kinase R to double-stranded RNA. Biochemistry 51:8764–70
    [Google Scholar]
  75. 75.
    Gelev V, Aktas H, Marintchev A, Ito T, Frueh D et al. 2006. Mapping of the auto-inhibitory interactions of protein kinase R by nuclear magnetic resonance. J. Mol. Biol. 364:352–63
    [Google Scholar]
  76. 76.
    Nanduri S, Rahman F, Williams BR, Qin J 2000. A dynamically tuned double-stranded RNA binding mechanism for the activation of antiviral kinase PKR. EMBO J 19:5567–74
    [Google Scholar]
  77. 77.
    McKenna SA, Lindhout DA, Kim I, Liu CW, Gelev VM et al. 2007. Molecular framework for the activation of RNA-dependent protein kinase. J. Biol. Chem. 282:11474–86
    [Google Scholar]
  78. 78.
    Dey M, Mann BR, Anshu A, Mannan MA 2014. Activation of protein kinase PKR requires dimerization-induced cis-phosphorylation within the activation loop. J. Biol. Chem. 289:5747–57
    [Google Scholar]
  79. 79.
    Husain B, Hesler S, Cole JL 2015. Regulation of PKR by RNA: formation of active and inactive dimers. Biochemistry 54:6663–72
    [Google Scholar]
  80. 80.
    Dar AC, Dever TE, Sicheri F 2005. Higher-order substrate recognition of eIF2α by the RNA-dependent protein kinase PKR. Cell 122:887–900
    [Google Scholar]
  81. 81.
    Li FZ, Li SW, Wang Z, Shen YQ, Zhang TC, Yang X 2013. Structure of the kinase domain of human RNA-dependent protein kinase with K296R mutation reveals a face-to-face dimer. Chinese Sci. Bull. 58:998–1002
    [Google Scholar]
  82. 82.
    Kim Y, Lee JH, Park JE, Cho J, Yi H, Kim VN 2014. PKR is activated by cellular dsRNAs during mitosis and acts as a mitotic regulator. Genes Dev 28:1310–22
    [Google Scholar]
  83. 83.
    Youssef OA, Safran SA, Nakamura T, Nix DA, Hotamisligil GS, Bass BL 2015. Potential role for snoRNAs in PKR activation during metabolic stress. PNAS 112:5023–28
    [Google Scholar]
  84. 84.
    Kim Y, Park J, Kim S, Kim M, Kang MG et al. 2018. PKR senses nuclear and mitochondrial signals by interacting with endogenous double-stranded RNAs. Mol. Cell 81:1051–63.e6
    [Google Scholar]
  85. 85.
    Nallagatla SR, Hwang J, Toroney R, Zheng X, Cameron CE, Bevilacqua PC 2007. 5′-Triphosphate-dependent activation of PKR by RNAs with short stem-loops. Science 318:1455–58
    [Google Scholar]
  86. 86.
    Schulz O, Pichlmair A, Rehwinkel J, Rogers NC, Scheuner D et al. 2010. Protein kinase R contributes to immunity against specific viruses by regulating interferon mRNA integrity. Cell Host Microbe 7:354–61
    [Google Scholar]
  87. 87.
    Pham AM, Santa Maria FG, Lahiri T, Friedman E, Marie IJ, Levy DE 2016. PKR transduces MDA5-dependent signals for type I IFN induction. PLOS Pathog 12:e1005489
    [Google Scholar]
  88. 88.
    Oh SW, Onomoto K, Wakimoto M, Onoguchi K, Ishidate F et al. 2016. Leader-containing uncapped viral transcript activates RIG-I in antiviral stress granules. PLOS Pathog 12:e1005444
    [Google Scholar]
  89. 89.
    Zamanian-Daryoush M, Mogensen TH, DiDonato JA, Williams BR 2000. NF-κB activation by double-stranded-RNA-activated protein kinase (PKR) is mediated through NF-κB-inducing kinase and IκB kinase. Mol. Cell. Biol. 20:1278–90
    [Google Scholar]
  90. 90.
    Sadler AJ, Williams BR 2007. Structure and function of the protein kinase R. Curr. Top. Microbiol. Immunol. 316:253–92
    [Google Scholar]
  91. 91.
    Ishii T, Kwon H, Hiscott J, Mosialos G, Koromilas AE 2001. Activation of the I kappa B alpha kinase (IKK) complex by double-stranded RNA-binding defective and catalytic inactive mutants of the interferon-inducible protein kinase PKR. Oncogene 20:1900–12
    [Google Scholar]
  92. 92.
    Onomoto K, Jogi M, Yoo JS, Narita R, Morimoto S et al. 2013. Critical role of an antiviral stress granule containing RIG-I and PKR in viral detection and innate immunity. PLOS ONE 7:e43031
    [Google Scholar]
  93. 93.
    Langereis MA, Feng Q, van Kuppeveld FJ 2013. MDA5 localizes to stress granules, but this localization is not required for the induction of type I interferon. J. Virol. 87:6314–25
    [Google Scholar]
  94. 94.
    McAllister CS, Taghavi N, Samuel CE 2012. Protein kinase PKR amplification of interferon β induction occurs through initiation factor eIF-2α-mediated translational control. J. Biol. Chem. 287:36384–92
    [Google Scholar]
  95. 95.
    Dalet A, Arguello RJ, Combes A, Spinelli L, Jaeger S et al. 2017. Protein synthesis inhibition and GADD34 control IFN-β heterogeneous expression in response to dsRNA. EMBO J 36:761–82
    [Google Scholar]
  96. 96.
    Kristiansen H, Gad HH, Eskildsen-Larsen S, Despres P, Hartmann R 2011. The oligoadenylate synthetase family: an ancient protein family with multiple antiviral activities. J. Interferon Cytokine Res. 31:41–47
    [Google Scholar]
  97. 97.
    Hovanessian AG, Justesen J 2007. The human 2′-5′ oligoadenylate synthetase family: unique interferon-inducible enzymes catalyzing 2′-5′ instead of 3′-5′ phosphodiester bond formation. Biochimie 89:779–88
    [Google Scholar]
  98. 98.
    Han Y, Donovan J, Rath S, Whitney G, Chitrakar A, Korennykh A 2014. Structure of human RNase L reveals the basis for regulated RNA decay in the IFN response. Science 343:1244–48
    [Google Scholar]
  99. 99.
    Donovan J, Dufner M, Korennykh A 2013. Structural basis for cytosolic double-stranded RNA surveillance by human oligoadenylate synthetase 1. PNAS 110:1652–57
    [Google Scholar]
  100. 100.
    Hartmann R, Justesen J, Sarkar SN, Sen GC, Yee VC 2003. Crystal structure of the 2′-specific and double-stranded RNA-activated interferon-induced antiviral protein 2′-5′-oligoadenylate synthetase. Mol. Cell 12:1173–85
    [Google Scholar]
  101. 101.
    Hornung V, Hartmann R, Ablasser A, Hopfner KP 2014. OAS proteins and cGAS: unifying concepts in sensing and responding to cytosolic nucleic acids. Nat. Rev. Immunol. 14:521–28
    [Google Scholar]
  102. 102.
    Zhou W, Whiteley AT, de Oliveira Mann CC, Morehouse BR, Nowak RP et al. 2018. Structure of the human cGAS-DNA complex reveals enhanced control of immune surveillance. Cell 174:300–11.e11
    [Google Scholar]
  103. 103.
    Donovan J, Whitney G, Rath S, Korennykh A 2015. Structural mechanism of sensing long dsRNA via a noncatalytic domain in human oligoadenylate synthetase 3. PNAS 112:3949–54
    [Google Scholar]
  104. 104.
    Sarkar SN, Ghosh A, Wang HW, Sung SS, Sen GC 1999. The nature of the catalytic domain of 2′-5′-oligoadenylate synthetases. J. Biol. Chem. 274:25535–42
    [Google Scholar]
  105. 105.
    Ilson DH, Torrence PF, Vilcek J 1986. Two molecular weight forms of human 2′,5′-oligoadenylate synthetase have different activation requirements. J. Interferon Res. 6:5–12
    [Google Scholar]
  106. 106.
    Ibsen MS, Gad HH, Thavachelvam K, Boesen T, Despres P, Hartmann R 2014. The 2′-5′-oligoadenylate synthetase 3 enzyme potently synthesizes the 2′-5′-oligoadenylates required for RNase L activation. J. Virol. 88:14222–31
    [Google Scholar]
  107. 107.
    Li Y, Banerjee S, Wang Y, Goldstein SA, Dong B et al. 2016. Activation of RNase L is dependent on OAS3 expression during infection with diverse human viruses. PNAS 113:2241–46
    [Google Scholar]
  108. 108.
    Schoggins JW, Wilson SJ, Panis M, Murphy MY, Jones CT et al. 2011. A diverse range of gene products are effectors of the type I interferon antiviral response. Nature 472:481–85
    [Google Scholar]
  109. 109.
    Ibsen MS, Gad HH, Andersen LL, Hornung V, Julkunen I et al. 2015. Structural and functional analysis reveals that human OASL binds dsRNA to enhance RIG-I signaling. Nucleic Acids Res 43:5236–48
    [Google Scholar]
  110. 110.
    Hartmann R, Olsen HS, Widder S, Jorgensen R, Justesen J 1998. p59 OASL, a 2′-5′ oligoadenylate synthetase like protein: a novel human gene related to the 2′-5′ oligoadenylate synthetase family. Nucleic Acids Res 26:4121–28
    [Google Scholar]
  111. 111.
    Marques J, Anwar J, Eskildsen-Larsen S, Rebouillat D, Paludan SR et al. 2008. The p59 oligoadenylate synthetase-like protein possesses antiviral activity that requires the C-terminal ubiquitin-like domain. J. Gen. Virol. 89:2767–72
    [Google Scholar]
  112. 112.
    Zhu J, Zhang Y, Ghosh A, Cuevas RA, Forero A et al. 2014. Antiviral activity of human OASL protein is mediated by enhancing signaling of the RIG-I RNA sensor. Immunity 40:936–48
    [Google Scholar]
  113. 113.
    Lee MS, Kim B, Oh GT, Kim YJ 2013. OASL1 inhibits translation of the type I interferon-regulating transcription factor IRF7. Nat. Immunol. 14:346–55
    [Google Scholar]
  114. 114.
    Al-Ahmadi W, Al-Haj L, Al-Mohanna FA, Silverman RH, Khabar KS 2009. RNase L downmodulation of the RNA-binding protein, HuR, and cellular growth. Oncogene 28:1782–91
    [Google Scholar]
  115. 115.
    Brennan-Laun SE, Li XL, Ezelle HJ, Venkataraman T, Blackshear PJ et al. 2014. RNase L attenuates mitogen-stimulated gene expression via transcriptional and post-transcriptional mechanisms to limit the proliferative response. J. Biol. Chem. 289:33629–43
    [Google Scholar]
  116. 116.
    Fabre O, Salehzada T, Lambert K, Boo Seok Y, Zhou A et al. 2012. RNase L controls terminal adipocyte differentiation, lipids storage and insulin sensitivity via CHOP10 mRNA regulation. Cell Death Differ 19:1470–81
    [Google Scholar]
  117. 117.
    Li Y, Banerjee S, Goldstein SA, Dong B, Gaughan C et al. 2017. Ribonuclease L mediates the cell-lethal phenotype of double-stranded RNA editing enzyme ADAR1 deficiency in a human cell line. eLife 6:e25687
    [Google Scholar]
  118. 118.
    Oakes SR, Gallego-Ortega D, Stanford PM, Junankar S, Au WWY et al. 2017. A mutation in the viral sensor 2′-5′-oligoadenylate synthetase 2 causes failure of lactation. PLOS Genet 13:e1007072
    [Google Scholar]
  119. 119.
    Hartmann R, Norby PL, Martensen PM, Jorgensen P, James MC et al. 1998. Activation of 2′-5′ oligoadenylate synthetase by single-stranded and double-stranded RNA aptamers. J. Biol. Chem. 273:3236–46
    [Google Scholar]
  120. 120.
    Walkley CR, Li JB 2017. Rewriting the transcriptome: adenosine-to-inosine RNA editing by ADARs. Genome Biol 18:205
    [Google Scholar]
  121. 121.
    Slotkin W, Nishikura K 2013. Adenosine-to-inosine RNA editing and human disease. Genome Med 5:105
    [Google Scholar]
  122. 122.
    George CX, John L, Samuel CE 2014. An RNA editor, adenosine deaminase acting on double-stranded RNA (ADAR1). J. Interferon Cytokine Res. 34:437–46
    [Google Scholar]
  123. 123.
    Stefl R, Oberstrass FC, Hood JL, Jourdan M, Zimmermann M et al. 2010. The solution structure of the ADAR2 dsRBM-RNA complex reveals a sequence-specific readout of the minor groove. Cell 143:225–37
    [Google Scholar]
  124. 124.
    Placido D, Brown BA 2nd, Lowenhaupt K, Rich A, Athanasiadis A 2007. A left-handed RNA double helix bound by the Z alpha domain of the RNA-editing enzyme ADAR1. Structure 15:395–404
    [Google Scholar]
  125. 125.
    Chen CX, Cho DS, Wang Q, Lai F, Carter KC, Nishikura K 2000. A third member of the RNA-specific adenosine deaminase gene family, ADAR3, contains both single- and double-stranded RNA binding domains. RNA 6:755–67
    [Google Scholar]
  126. 126.
    Thomas JM, Beal PA 2017. How do ADARs bind RNA? New protein-RNA structures illuminate substrate recognition by the RNA editing ADARs. BioEssays 39:1600187
    [Google Scholar]
  127. 127.
    Matthews MM, Thomas JM, Zheng Y, Tran K, Phelps KJ et al. 2016. Structures of human ADAR2 bound to dsRNA reveal base-flipping mechanism and basis for site selectivity. Nat. Struct. Mol. Biol. 23:426–33
    [Google Scholar]
  128. 128.
    Klimasauskas S, Kumar S, Roberts RJ, Cheng X 1994. HhaI methyltransferase flips its target base out of the DNA helix. Cell 76:357–69
    [Google Scholar]
  129. 129.
    Tan MH, Li Q, Shanmugam R, Piskol R, Kohler J et al. 2017. Dynamic landscape and regulation of RNA editing in mammals. Nature 550:249–54
    [Google Scholar]
  130. 130.
    Hundley HA, Bass BL 2010. ADAR editing in double-stranded UTRs and other noncoding RNA sequences. Trends Biochem. Sci. 35:377–83
    [Google Scholar]
  131. 131.
    Oakes E, Anderson A, Cohen-Gadol A, Hundley HA 2017. Adenosine deaminase that acts on RNA 3 (ADAR3) binding to glutamate receptor subunit B pre-mRNA inhibits RNA editing in glioblastoma. J. Biol. Chem. 292:4326–35
    [Google Scholar]
  132. 132.
    Higuchi M, Maas S, Single FN, Hartner J, Rozov A et al. 2000. Point mutation in an AMPA receptor gene rescues lethality in mice deficient in the RNA-editing enzyme ADAR2. Nature 406:78–81
    [Google Scholar]
  133. 133.
    Hartner JC, Schmittwolf C, Kispert A, Muller AM, Higuchi M, Seeburg PH 2004. Liver disintegration in the mouse embryo caused by deficiency in the RNA-editing enzyme ADAR1. J. Biol. Chem. 279:4894–902
    [Google Scholar]
  134. 134.
    Hartner JC, Walkley CR, Lu J, Orkin SH 2008. ADAR1 is essential for the maintenance of hematopoiesis and suppression of interferon signaling. Nat. Immunol. 10:109–15
    [Google Scholar]
  135. 135.
    Hauenstein A, Zhang L, Wu H 2015. The hierarchical structural architecture of inflammasomes, supramolecular inflammatory machines. Curr. Opin. Struct. Biol. 31:75–83
    [Google Scholar]
  136. 136.
    Pestal K, Funk CC, Snyder JM, Price ND, Treuting PM, Stetson DB 2015. Isoforms of RNA-editing enzyme ADAR1 independently control nucleic acid sensor MDA5-driven autoimmunity and multi-organ development. Immunity 43:933–44
    [Google Scholar]
  137. 137.
    Liddicoat BJ, Piskol R, Chalk AM, Ramaswami G, Higuchi M et al. 2015. RNA editing by ADAR1 prevents MDA5 sensing of endogenous dsRNA as nonself. Science 349:1115–20
    [Google Scholar]
  138. 138.
    Mannion NM, Greenwood SM, Young R, Cox S, Brindle J et al. 2014. The RNA-editing enzyme ADAR1 controls innate immune responses to RNA. Cell Rep 9:1482–94
    [Google Scholar]
  139. 139.
    Rice GI, Kasher PR, Forte GM, Mannion NM, Greenwood SM et al. 2010. Mutations in ADAR1 cause Aicardi-Goutières syndrome associated with a type I interferon signature. Nat. Genet. 44:1243–48
    [Google Scholar]
  140. 140.
    Ward SV, George CX, Welch MJ, Liou LY, Hahm B et al. 2011. RNA editing enzyme adenosine deaminase is a restriction factor for controlling measles virus replication that also is required for embryogenesis. PNAS 108:331–36
    [Google Scholar]
  141. 141.
    Biswas N, Wang T, Ding M, Tumne A, Chen Y et al. 2012. ADAR1 is a novel multi targeted anti-HIV-1 cellular protein. Virology 422:265–77
    [Google Scholar]
  142. 142.
    Nishikura K 2016. A-to-I editing of coding and non-coding RNAs by ADARs. Nat. Rev. Mol. Cell Biol. 17:83–96
    [Google Scholar]
  143. 143.
    Yang W, Chendrimada TP, Wang Q, Higuchi M, Seeburg PH et al. 2006. Modulation of microRNA processing and expression through RNA editing by ADAR deaminases. Nat. Struct. Mol. Biol. 13:13–21
    [Google Scholar]
  144. 144.
    Kawahara Y, Zinshteyn B, Sethupathy P, Iizasa H, Hatzigeorgiou AG, Nishikura K 2007. Redirection of silencing targets by adenosine-to-inosine editing of miRNAs. Science 315:1137–40
    [Google Scholar]
  145. 145.
    Ota H, Sakurai M, Gupta R, Valente L, Wulff BE et al. 2013. ADAR1 forms a complex with Dicer to promote microRNA processing and RNA-induced gene silencing. Cell 153:575–89
    [Google Scholar]
  146. 146.
    Aktas T, Avsar Ilik I, Maticzka D, Bhardwaj V, Pessoa Rodrigues C et al. 2017. DHX9 suppresses RNA processing defects originating from the Alu invasion of the human genome. Nature 544:115–19
    [Google Scholar]
  147. 147.
    Elbarbary RA, Li W, Tian B, Maquat LE 2013. STAU1 binding 3′ UTR IRAlus complements nuclear retention to protect cells from PKR-mediated translational shutdown. Genes Dev 27:1495–510
    [Google Scholar]
  148. 148.
    Wilson RC, Doudna JA 2013. Molecular mechanisms of RNA interference. Annu. Rev. Biophys. 42:217–39
    [Google Scholar]
  149. 149.
    Sheu-Gruttadauria J, MacRae IJ 2017. Structural foundations of RNA silencing by Argonaute. J. Mol. Biol. 429:2619–39
    [Google Scholar]
  150. 150.
    Kim VN, Han J, Siomi MC 2009. Biogenesis of small RNAs in animals. Nat. Rev. Mol. Cell Biol. 10:126–39
    [Google Scholar]
  151. 151.
    Kwon SC, Nguyen TA, Choi YG, Jo MH, Hohng S et al. 2016. Structure of human DROSHA. Cell 164:81–90
    [Google Scholar]
  152. 152.
    Macrae IJ, Li F, Zhou K, Cande WZ, Doudna JA 2006. Structure of Dicer and mechanistic implications for RNAi. Cold Spring Harb. Symp. Quant. Biol. 71:73–80
    [Google Scholar]
  153. 153.
    Liu Z, Wang J, Cheng H, Ke X, Sun L et al. 2018. Cryo-EM structure of human Dicer and its complexes with a pre-miRNA substrate. Cell 173:1191–203 e12. Erratum. 2018 Cell 173:1549–50
    [Google Scholar]
  154. 154.
    Tian Y, Simanshu DK, Ma JB, Park JE, Heo I et al. 2014. A phosphate-binding pocket within the platform-PAZ-connector helix cassette of human Dicer. Mol. Cell 53:606–16
    [Google Scholar]
  155. 155.
    Sinha NK, Iwasa J, Shen PS, Bass BL 2018. Dicer uses distinct modules for recognizing dsRNA termini. Science 359:329–34
    [Google Scholar]
  156. 156.
    Burger K, Gullerova M 2015. Swiss army knives: non-canonical functions of nuclear Drosha and Dicer. Nat. Rev. Mol. Cell Biol. 16:417–30
    [Google Scholar]
  157. 157.
    Lee D, Shin C 2018. Emerging roles of DROSHA beyond primary microRNA processing. RNA Biol 15:186–93
    [Google Scholar]
  158. 158.
    Banez-Coronel M, Porta S, Kagerbauer B, Mateu-Huertas E, Pantano L et al. 2012. A pathogenic mechanism in Huntington's disease involves small CAG-repeated RNAs with neurotoxic activity. PLOS Genet 8:e1002481
    [Google Scholar]
  159. 159.
    Kaneko H, Dridi S, Tarallo V, Gelfand BD, Fowler BJ et al. 2011. DICER1 deficit induces Alu RNA toxicity in age-related macular degeneration. Nature 471:325–30
    [Google Scholar]
  160. 160.
    Tarallo V, Hirano Y, Gelfand BD, Dridi S, Kerur N et al. 2012. DICER1 loss and Alu RNA induce age-related macular degeneration via the NLRP3 inflammasome and MyD88. Cell 149:847–59
    [Google Scholar]
  161. 161.
    Ding SW 2010. RNA-based antiviral immunity. Nat. Rev. Immunol. 10:632–44
    [Google Scholar]
  162. 162.
    Seo GJ, Kincaid RP, Phanaksri T, Burke JM, Cox JE et al. 2013. Reciprocal inhibition between intracellular antiviral signaling and the RNAi machinery in mammalian cells. Cell Host Microbe 14:435–45
    [Google Scholar]
  163. 163.
    Maillard PV, Ciaudo C, Marchais A, Li Y, Jay F et al. 2013. Antiviral RNA interference in mammalian cells. Science 342:235–38
    [Google Scholar]
  164. 164.
    Li Y, Lu J, Han Y, Fan X, Ding SW 2013. RNA interference functions as an antiviral immunity mechanism in mammals. Science 342:231–34
    [Google Scholar]
  165. 165.
    Li Y, Basavappa M, Lu J, Dong S, Cronkite DA et al. 2016. Induction and suppression of antiviral RNA interference by influenza A virus in mammalian cells. Nat. Microbiol. 2:16250
    [Google Scholar]
  166. 166.
    Aguado LC, Schmid S, May J, Sabin LR, Panis M et al. 2017. RNase III nucleases from diverse kingdoms serve as antiviral effectors. Nature 547:114–17
    [Google Scholar]
  167. 167.
    Heyam A, Lagos D, Plevin M 2015. Dissecting the roles of TRBP and PACT in double-stranded RNA recognition and processing of noncoding RNAs. Wiley Interdiscip. Rev. RNA 6:271–89
    [Google Scholar]
  168. 168.
    Jakob L, Treiber T, Treiber N, Gust A, Kramm K et al. 2016. Structural and functional insights into the fly microRNA biogenesis factor Loquacious. RNA 22:383–96
    [Google Scholar]
  169. 169.
    Heyam A, Coupland CE, Degut C, Haley RA, Baxter NJ et al. 2017. Conserved asymmetry underpins homodimerization of Dicer-associated double-stranded RNA-binding proteins. Nucleic Acids Res 45:12577–84
    [Google Scholar]
  170. 170.
    Wilson RC, Tambe A, Kidwell MA, Noland CL, Schneider CP, Doudna JA 2015. Dicer-TRBP complex formation ensures accurate mammalian microRNA biogenesis. Mol. Cell 57:397–407
    [Google Scholar]
  171. 171.
    Chukwurah E, Willingham V, Singh M, Castillo-Azofeifa D, Patel RC 2018. Contribution of the two dsRBM motifs to the double-stranded RNA binding and protein interactions of PACT. J. Cell Biochem. 119:3598–607
    [Google Scholar]
  172. 172.
    Patel RC, Sen GC 1998. PACT, a protein activator of the interferon-induced protein kinase, PKR. EMBO J 17:4379–90
    [Google Scholar]
  173. 173.
    Peters GA, Dickerman B, Sen GC 2009. Biochemical analysis of PKR activation by PACT. Biochemistry 48:7441–47
    [Google Scholar]
  174. 174.
    Singh M, Patel RC 2012. Increased interaction between PACT molecules in response to stress signals is required for PKR activation. J. Cell Biochem. 113:2754–64
    [Google Scholar]
  175. 175.
    Park H, Davies MV, Langland JO, Chang HW, Nam YS et al. 1994. TAR RNA-binding protein is an inhibitor of the interferon-induced protein kinase PKR. PNAS 91:4713–17
    [Google Scholar]
  176. 176.
    Chukwurah E, Patel RC 2018. Stress-induced TRBP phosphorylation enhances its interaction with PKR to regulate cellular survival. Sci. Rep. 8:1020
    [Google Scholar]
  177. 177.
    Laraki G, Clerzius G, Daher A, Melendez-Pena C, Daniels S, Gatignol A 2008. Interactions between the double-stranded RNA-binding proteins TRBP and PACT define the Medipal domain that mediates protein-protein interactions. RNA Biol 5:92–103
    [Google Scholar]
  178. 178.
    Vaughn LS, Bragg DC, Sharma N, Camargos S, Cardoso F, Patel RC 2015. Altered activation of protein kinase PKR and enhanced apoptosis in dystonia cells carrying a mutation in PKR activator protein PACT. J. Biol. Chem. 290:22543–57
    [Google Scholar]
  179. 179.
    Dickerman BK, White CL, Kessler PM, Sadler AJ, Williams BR, Sen GC 2015. The protein activator of protein kinase R, PACT/RAX, negatively regulates protein kinase R during mouse anterior pituitary development. FEBS J 282:4766–81
    [Google Scholar]
  180. 180.
    Clerzius G, Shaw E, Daher A, Burugu S, Gelinas JF et al. 2013. The PKR activator, PACT, becomes a PKR inhibitor during HIV-1 replication. Retrovirology 10:96
    [Google Scholar]
  181. 181.
    Meyer C, Garzia A, Mazzola M, Gerstberger S, Molina H, Tuschl T 2018. The TIA1 RNA-binding protein family regulates EIF2AK2-mediated stress response and cell cycle progression. Mol. Cell 69:622–35.e6
    [Google Scholar]
  182. 182.
    Daher A, Laraki G, Singh M, Melendez-Pena CE, Bannwarth S et al. 2009. TRBP control of PACT-induced phosphorylation of protein kinase R is reversed by stress. Mol. Cell Biol. 29:254–65
    [Google Scholar]
  183. 183.
    Singh M, Castillo D, Patel CV, Patel RC 2011. Stress-induced phosphorylation of PACT reduces its interaction with TRBP and leads to PKR activation. Biochemistry 50:4550–60
    [Google Scholar]
  184. 184.
    Lee HY, Zhou K, Smith AM, Noland CL, Doudna JA 2013. Differential roles of human Dicer-binding proteins TRBP and PACT in small RNA processing. Nucleic Acids Res 41:6568–76
    [Google Scholar]
  185. 185.
    Lee Y, Hur I, Park SY, Kim YK, Suh MR, Kim VN 2006. The role of PACT in the RNA silencing pathway. EMBO J 25:522–32
    [Google Scholar]
  186. 186.
    Kok KH, Ng MH, Ching YP, Jin DY 2007. Human TRBP and PACT directly interact with each other and associate with Dicer to facilitate the production of small interfering RNA. J. Biol. Chem. 282:17649–57
    [Google Scholar]
  187. 187.
    Fukunaga R, Han BW, Hung JH, Xu J, Weng Z, Zamore PD 2012. Dicer partner proteins tune the length of mature miRNAs in flies and mammals. Cell 151:533–46
    [Google Scholar]
  188. 188.
    Kim Y, Yeo J, Lee JH, Cho J, Seo D et al. 2014. Deletion of human tarbp2 reveals cellular microRNA targets and cell-cycle function of TRBP. Cell Rep 9:1061–74
    [Google Scholar]
  189. 189.
    Noland CL, Doudna JA 2013. Multiple sensors ensure guide strand selection in human RNAi pathways. RNA 19:639–48
    [Google Scholar]
  190. 190.
    Kok KH, Lui PY, Ng MH, Siu KL, Au SW, Jin DY 2011. The double-stranded RNA-binding protein PACT functions as a cellular activator of RIG-I to facilitate innate antiviral response. Cell Host Microbe 9:299–309
    [Google Scholar]
  191. 191.
    Ho TH, Kew C, Lui PY, Chan CP, Satoh T et al. 2016. PACT- and RIG-I-dependent activation of type I interferon production by a defective interfering RNA derived from measles virus vaccine. J. Virol. 90:1557–68
    [Google Scholar]
  192. 192.
    Luthra P, Ramanan P, Mire CE, Weisend C, Tsuda Y et al. 2013. Mutual antagonism between the Ebola virus VP35 protein and the RIG-I activator PACT determines infection outcome. Cell Host Microbe 14:74–84
    [Google Scholar]
  193. 193.
    Ding Z, Fang L, Yuan S, Zhao L, Wang X et al. 2017. The nucleocapsid proteins of mouse hepatitis virus and severe acute respiratory syndrome coronavirus share the same IFN-beta antagonizing mechanism: attenuation of PACT-mediated RIG-I/ MDA5 activation. Oncotarget 8:49655–70
    [Google Scholar]
  194. 194.
    Siu KL, Yeung ML, Kok KH, Yuen KS, Kew C et al. 2014. Middle East respiratory syndrome coronavirus 4a protein is a double-stranded RNA-binding protein that suppresses PACT-induced activation of RIG-I and MDA5 in the innate antiviral response. J. Virol. 88:4866–76
    [Google Scholar]
  195. 195.
    Komuro A, Homma Y, Negoro T, Barber GN, Horvath CM 2016. The TAR-RNA binding protein is required for immunoresponses triggered by Cardiovirus infection. Biochem. Biophys. Res. Commun. 480:187–93
    [Google Scholar]
  196. 196.
    Lui PY, Wong LR, Ho TH, Au SWN, Chan CP et al. 2017. PACT facilitates RNA-induced activation of MDA5 by promoting MDA5 oligomerization. J. Immunol. 199:1846–55
    [Google Scholar]
  197. 197.
    Miyamoto M, Komuro A 2017. PACT is required for MDA5-mediated immunoresponses triggered by Cardiovirus infection via interaction with LGP2. Biochem. Biophys. Res. Commun. 494:227–33
    [Google Scholar]
  198. 198.
    Singleton MR, Dillingham MS, Wigley DB 2007. Structure and mechanism of helicases and nucleic acid translocases. Annu. Rev. Biochem. 76:23–50
    [Google Scholar]
  199. 199.
    Putnam AA, Jankowsky E 2013. DEAD-box helicases as integrators of RNA, nucleotide and protein binding. Biochim. Biophys. Acta. 1829:884–93
    [Google Scholar]
  200. 200.
    Luo D, Kohlway A, Pyle AM 2013. Duplex RNA activated ATPases (DRAs): platforms for RNA sensing, signaling and processing. RNA Biol 10:111–20
    [Google Scholar]
  201. 201.
    Lee T, Pelletier J 2016. The biology of DHX9 and its potential as a therapeutic target. Oncotarget 7:42716–39
    [Google Scholar]
  202. 202.
    Brass AL, Dykxhoorn DM, Benita Y, Yan N, Engelman A et al. 2008. Identification of host proteins required for HIV infection through a functional genomic screen. Science 319:921–26
    [Google Scholar]
  203. 203.
    Zhang Z, Yuan B, Lu N, Facchinetti V, Liu YJ 2011. DHX9 pairs with IPS-1 to sense double-stranded RNA in myeloid dendritic cells. J. Immunol. 187:4501–8
    [Google Scholar]
  204. 204.
    Zhu S, Ding S, Wang P, Wei Z, Pan W et al. 2017. Nlrp9b inflammasome restricts rotavirus infection in intestinal epithelial cells. Nature 546:667–70
    [Google Scholar]
  205. 205.
    Miyashita M, Oshiumi H, Matsumoto M, Seya T 2011. DDX60, a DEXD/H box helicase, is a novel antiviral factor promoting RIG-I-like receptor-mediated signaling. Mol. Cell. Biol. 31:3802–19
    [Google Scholar]
  206. 206.
    Oshiumi H, Miyashita M, Okamoto M, Morioka Y, Okabe M et al. 2015. DDX60 is involved in RIG-I-dependent and independent antiviral responses, and its function is attenuated by virus-induced EGFR activation. Cell Rep 11:1193–207
    [Google Scholar]
  207. 207.
    Goubau D, van der Veen AG, Chakravarty P, Lin R, Rogers N et al. 2015. Mouse superkiller-2-like helicase DDX60 is dispensable for type I IFN induction and immunity to multiple viruses. Eur. J. Immunol. 45:3386–403
    [Google Scholar]
  208. 208.
    Wang P, Zhu S, Yang L, Cui S, Pan W et al. 2015. Nlrp6 regulates intestinal antiviral innate immunity. Science 350:826–30
    [Google Scholar]
  209. 209.
    Mitoma H, Hanabuchi S, Kim T, Bao M, Zhang Z et al. 2013. The DHX33 RNA helicase senses cytosolic RNA and activates the NLRP3 inflammasome. Immunity 39:123–35
    [Google Scholar]
  210. 210.
    Liu Y, Lu N, Yuan B, Weng L, Wang F et al. 2014. The interaction between the helicase DHX33 and IPS-1 as a novel pathway to sense double-stranded RNA and RNA viruses in myeloid dendritic cells. Cell Mol. Immunol. 11:49–57
    [Google Scholar]
  211. 211.
    Sugimoto N, Mitoma H, Kim T, Hanabuchi S, Liu YJ 2014. Helicase proteins DHX29 and RIG-I cosense cytosolic nucleic acids in the human airway system. PNAS 111:7747–52
    [Google Scholar]
  212. 212.
    Zhu Q, Tan P, Li Y, Lin M, Li C et al. 2018. DHX29 functions as an RNA co-sensor for MDA5-mediated EMCV-specific antiviral immunity. PLOS Pathog 14:e1006886
    [Google Scholar]
  213. 213.
    Gringhuis SI, Hertoghs N, Kaptein TM, Zijlstra-Willems EM, Sarrami-Forooshani R et al. 2017. HIV-1 blocks the signaling adaptor MAVS to evade antiviral host defense after sensing of abortive HIV-1 RNA by the host helicase DDX3. Nat. Immunol. 18:225–35
    [Google Scholar]
  214. 214.
    Gu L, Fullam A, McCormack N, Hohn Y, Schroder M 2017. DDX3 directly regulates TRAF3 ubiquitination and acts as a scaffold to co-ordinate assembly of signalling complexes downstream from MAVS. Biochem. J. 474:571–87
    [Google Scholar]
  215. 215.
    Zhang Z, Kim T, Bao M, Facchinetti V, Jung SY et al. 2011. DDX1, DDX21, and DHX36 helicases form a complex with the adaptor molecule TRIF to sense dsRNA in dendritic cells. Immunity 34:866–78
    [Google Scholar]
  216. 216.
    Sithole N, Williams C, Vaughan A, Lever A 2015. The roles of DEAD box helicases in the life cycle of HIV-1. Lancet 385:Suppl. 1S89
    [Google Scholar]
  217. 217.
    Bortz E, Westera L, Maamary J, Steel J, Albrecht RA et al. 2011. Host- and strain-specific regulation of influenza virus polymerase activity by interacting cellular proteins. mBio 2:e00151–11
    [Google Scholar]
/content/journals/10.1146/annurev-immunol-042718-041356
Loading
/content/journals/10.1146/annurev-immunol-042718-041356
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error