1932

Abstract

Pattern recognition receptors (PRRs) survey intra- and extracellular spaces for pathogen-associated molecular patterns (PAMPs) within microbial products of infection. Recognition and binding to cognate PAMP ligand by specific PRRs initiates signaling cascades that culminate in a coordinated intracellular innate immune response designed to control infection. In particular, our immune system has evolved specialized PRRs to discriminate viral nucleic acid from host. These are critical sensors of viral RNA to trigger innate immunity in the vertebrate host. Different families of PRRs of virus infection have been defined and reveal a diversity of PAMP specificity for wide viral pathogen coverage to recognize and extinguish virus infection. In this review, we discuss recent insights in pathogen recognition by the RIG-I-like receptors, related RNA helicases, Toll-like receptors, and other RNA sensor PRRs, to present emerging themes in innate immune signaling during virus infection.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-immunol-042617-053309
2018-04-26
2024-06-27
Loading full text...

Full text loading...

/deliver/fulltext/immunol/36/1/annurev-immunol-042617-053309.html?itemId=/content/journals/10.1146/annurev-immunol-042617-053309&mimeType=html&fmt=ahah

Literature Cited

  1. Kell AM, Gale M Jr. 1.  2015. RIG-I in RNA virus recognition. Virology 479–480110–21 [Google Scholar]
  2. Loo YM, Gale M Jr. 2.  2011. Immune signaling by RIG-I-like receptors. Immunity 34680–92 [Google Scholar]
  3. Luo D, Kohlway A, Pyle AM. 3.  2013. Duplex RNA activated ATPases (DRAs): platforms for RNA sensing, signaling and processing. RNA Biol 10111–20 [Google Scholar]
  4. Saito T, Hirai R, Loo YM, Owen D, Johnson CL. 4.  et al. 2007. Regulation of innate antiviral defenses through a shared repressor domain in RIG-I and LGP2. PNAS 104582–87 [Google Scholar]
  5. Cui S, Eisenacher K, Kirchhofer A, Brzozka K, Lammens A. 5.  et al. 2008. The C-terminal regulatory domain is the RNA 5′-triphosphate sensor of RIG-I. Mol. Cell 29169–79 [Google Scholar]
  6. Li X, Lu C, Stewart M, Xu H, Strong RK. 6.  et al. 2009. Structural basis of double-stranded RNA recognition by the RIG-I like receptor MDA5. Arch. Biochem. Biophys. 48823–33 [Google Scholar]
  7. Kageyama M, Takahasi K, Narita R, Hirai R, Yoneyama M. 7.  et al. 2011. 55 Amino acid linker between helicase and carboxyl terminal domains of RIG-I functions as a critical repression domain and determines inter-domain conformation. Biochem. Biophys. Res. Commun. 41575–81 [Google Scholar]
  8. Takahasi K, Kumeta H, Tsuduki N, Narita R, Shigemoto T. 8.  et al. 2009. Solution structures of cytosolic RNA sensor MDA5 and LGP2 C-terminal domains: identification of the RNA recognition loop in RIG-I-like receptors. J. Biol. Chem. 28417465–74 [Google Scholar]
  9. Pippig DA, Hellmuth JC, Cui S, Kirchhofer A, Lammens K. 9.  et al. 2009. The regulatory domain of the RIG-I family ATPase LGP2 senses double-stranded RNA. Nucleic Acids Res 372014–25 [Google Scholar]
  10. Yoneyama M, Kikuchi M, Matsumoto K, Imaizumi T, Miyagishi M. 10.  et al. 2005. Shared and unique functions of the DExD/H-box helicases RIG-I, MDA5, and LGP2 in antiviral innate immunity. J. Immunol. 1752851–58 [Google Scholar]
  11. Murali A, Li X, Ranjith-Kumar CT, Bhardwaj K, Holzenburg A. 11.  et al. 2008. Structure and function of LGP2, a DEX(D/H) helicase that regulates the innate immunity response. J. Biol. Chem. 28315825–33 [Google Scholar]
  12. Bruns AM, Leser GP, Lamb RA, Horvath CM. 12.  2014. The innate immune sensor LGP2 activates antiviral signaling by regulating MDA5-RNA interaction and filament assembly. Mol. Cell 55771–81 [Google Scholar]
  13. Komuro A, Horvath CM. 13.  2006. RNA- and virus-independent inhibition of antiviral signaling by RNA helicase LGP2. J. Virol. 8012332–42 [Google Scholar]
  14. Rothenfusser S, Goutagny N, DiPerna G, Gong M, Monks BG. 14.  et al. 2005. The RNA helicase Lgp2 inhibits TLR-independent sensing of viral replication by retinoic acid-inducible gene-I. J. Immunol. 1755260–68 [Google Scholar]
  15. Venkataraman T, Valdes M, Elsby R, Kakuta S, Caceres G. 15.  et al. 2007. Loss of DExD/H box RNA helicase LGP2 manifests disparate antiviral responses. J. Immunol. 1786444–55 [Google Scholar]
  16. Satoh T, Kato H, Kumagai Y, Yoneyama M, Sato S. 16.  et al. 2010. LGP2 is a positive regulator of RIG-I- and MDA5-mediated antiviral responses. PNAS 1071512–17 [Google Scholar]
  17. Cai X, Xu H, Chen ZJ. 17.  2017. Prion-like polymerization in immunity and inflammation. Cold Spring Harb. Perspect. Biol. 9a023580 [Google Scholar]
  18. Schmitz ML, Kracht M, Saul VV. 18.  2014. The intricate interplay between RNA viruses and NF-κB. Biochim. Biophys. Acta 18432754–64 [Google Scholar]
  19. Horner SM, Liu HM, Park HS, Briley J, Gale M Jr. 19.  2011. Mitochondrial-associated endoplasmic reticulum membranes (MAM) form innate immune synapses and are targeted by hepatitis C virus. PNAS 10814590–95 [Google Scholar]
  20. Bender S, Reuter A, Eberle F, Einhorn E, Binder M, Bartenschlager R. 20.  2015. Activation of type I and III interferon response by mitochondrial and peroxisomal MAVS and inhibition by hepatitis C virus. PLOS Pathog 11e1005264 [Google Scholar]
  21. Seth RB, Sun L, Ea CK, Chen ZJ. 21.  2005. Identification and characterization of MAVS, a mitochondrial antiviral signaling protein that activates NF-κB and IRF 3. Cell 122669–82 [Google Scholar]
  22. Loo YM, Owen DM, Li K, Erickson AK, Johnson CL. 22.  et al. 2006. Viral and therapeutic control of IFN-beta promoter stimulator 1 during hepatitis C virus infection. PNAS 1036001–6 [Google Scholar]
  23. Dixit E, Boulant S, Zhang Y, Lee AS, Odendall C. 23.  et al. 2010. Peroxisomes are signaling platforms for antiviral innate immunity. Cell 141668–81 [Google Scholar]
  24. Odendall C, Dixit E, Stavru F, Bierne H, Franz KM. 24.  et al. 2014. Diverse intracellular pathogens activate type III interferon expression from peroxisomes. Nat. Immunol. 15717–26 [Google Scholar]
  25. Rowland AA, Voeltz GK. 25.  2012. Endoplasmic reticulum-mitochondria contacts: function of the junction. Nat. Rev. Mol. Cell Biol. 13607–25 [Google Scholar]
  26. Hornung V, Ellegast J, Kim S, Brzozka K, Jung A. 26.  et al. 2006. 5′-Triphosphate RNA is the ligand for RIG-I. Science 314994–97 [Google Scholar]
  27. Franchi L, Eigenbrod T, Munoz-Planillo R, Ozkurede U, Kim YG. 27.  et al. 2014. Cytosolic double-stranded RNA activates the NLRP3 inflammasome via MAVS-induced membrane permeabilization and K+ efflux. J. Immunol. 1934214–22 [Google Scholar]
  28. Guo H, Callaway JB, Ting JP. 28.  2015. Inflammasomes: mechanism of action, role in disease, and therapeutics. Nat. Med. 21677–87 [Google Scholar]
  29. Poeck H, Bscheider M, Gross O, Finger K, Roth S. 29.  et al. 2010. Recognition of RNA virus by RIG-I results in activation of CARD9 and inflammasome signaling for interleukin 1 beta production. Nat. Immunol. 1163–69 [Google Scholar]
  30. Pothlichet J, Meunier I, Davis BK, Ting JP, Skamene E. 30.  et al. 2013. Type I IFN triggers RIG-I/TLR3/NLRP3-dependent inflammasome activation in influenza A virus infected cells. PLOS Pathog 9e1003256 [Google Scholar]
  31. Chattopadhyay S, Sen GC. 31.  2017. RIG-I-like receptor-induced IRF3 mediated pathway of apoptosis (RIPA): a new antiviral pathway. Protein Cell 8165–68 [Google Scholar]
  32. Saito T, Owen DM, Jiang F, Marcotrigiano J, Gale M Jr. 32.  2008. Innate immunity induced by composition-dependent RIG-I recognition of hepatitis C virus RNA. Nature 454523–27 [Google Scholar]
  33. Schnell G, Loo YM, Marcotrigiano J, Gale M Jr. 33.  2012. Uridine composition of the poly-U/UC tract of HCV RNA defines non-self recognition by RIG-I. PLOS Pathog 8e1002839 [Google Scholar]
  34. Runge S, Sparrer KM, Lassig C, Hembach K, Baum A. 34.  et al. 2014. In vivo ligands of MDA5 and RIG-I in measles virus-infected cells. PLOS Pathog 10e1004081 [Google Scholar]
  35. Sanchez David RY, Combredet C, Sismeiro O, Dillies MA, Jagla B. 35.  et al. 2016. Comparative analysis of viral RNA signatures on different RIG-I-like receptors. eLife 5e11275 [Google Scholar]
  36. Lu C, Xu H, Ranjith-Kumar CT, Brooks MT, Hou TY. 36.  et al. 2010. The structural basis of 5′ triphosphate double-stranded RNA recognition by RIG-I C-terminal domain. Structure 181032–43 [Google Scholar]
  37. Pichlmair A, Schulz O, Tan CP, Naslund TI, Liljestrom P. 37.  et al. 2006. RIG-I-mediated antiviral responses to single-stranded RNA bearing 5′-phosphates. Science 314997–1001 [Google Scholar]
  38. Schmidt A, Schwerd T, Hamm W, Hellmuth JC, Cui S. 38.  et al. 2009. 5′-Triphosphate RNA requires base-paired structures to activate antiviral signaling via RIG-I. PNAS 10612067–72 [Google Scholar]
  39. Wang Y, Ludwig J, Schuberth C, Goldeck M, Schlee M. 39.  et al. 2010. Structural and functional insights into 5′-ppp RNA pattern recognition by the innate immune receptor RIG-I. Nat. Struct. Mol. Biol. 17781–87 [Google Scholar]
  40. Banerjee AK, Shatkin AJ. 40.  1971. Guanosine-5′-diphosphate at the 5′ termini of reovirus RNA: evidence for a segmented genome within the virion. J. Mol. Biol. 61643–53 [Google Scholar]
  41. Goubau D, Schlee M, Deddouche S, Pruijssers AJ, Zillinger T. 41.  et al. 2014. Antiviral immunity via RIG-I-mediated recognition of RNA bearing 5′-diphosphates. Nature 514372–75 [Google Scholar]
  42. Abdullah Z, Schlee M, Roth S, Mraheil MA, Barchet W. 42.  et al. 2012. RIG-I detects infection with live Listeria by sensing secreted bacterial nucleic acids. EMBO J 314153–64 [Google Scholar]
  43. Hagmann CA, Herzner AM, Abdullah Z, Zillinger T, Jakobs C. 43.  et al. 2013. RIG-I detects triphosphorylated RNA of Listeria monocytogenes during infection in non-immune cells. PLOS ONE 8e62872 [Google Scholar]
  44. Schmolke M, Patel JR, de Castro E, Sanchez-Aparicio MT, Uccellini MB. 44.  et al. 2014. RIG-I detects mRNA of intracellular Salmonella enterica serovar Typhimurium during bacterial infection. mBio 5e01006–14 [Google Scholar]
  45. Malathi K, Dong B, Gale M Jr., Silverman RH. 45.  2007. Small self-RNA generated by RNase L amplifies antiviral innate immunity. Nature 448816–19 [Google Scholar]
  46. Malathi K, Saito T, Crochet N, Barton DJ, Gale M Jr., Silverman RH. 46.  2010. RNase L releases a small RNA from HCV RNA that refolds into a potent PAMP. RNA 162108–19 [Google Scholar]
  47. Ablasser A, Bauernfeind F, Hartmann G, Latz E, Fitzgerald KA, Hornung V. 47.  2009. RIG-I-dependent sensing of poly(dA:dT) through the induction of an RNA polymerase III-transcribed RNA intermediate. Nat. Immunol. 101065–72 [Google Scholar]
  48. Chiu YH, Macmillan JB, Chen ZJ. 48.  2009. RNA polymerase III detects cytosolic DNA and induces type I interferons through the RIG-I pathway. Cell 138576–91 [Google Scholar]
  49. Leung DW, Amarasinghe GK. 49.  2016. When your cap matters: structural insights into self versus non-self recognition of 5′ RNA by immunomodulatory host proteins. Curr. Opin. Struct. Biol. 36133–41 [Google Scholar]
  50. Li X, Liu CX, Xue W, Zhang Y, Jiang S. 50.  et al. 2017. Coordinated circRNA biogenesis and function with NF90/NF110 in viral infection. Mol. Cell 67214–27.e7 [Google Scholar]
  51. Chen YG, Kim MV, Chen X, Batista PJ, Aoyama S. 51.  et al. 2017. Sensing self and foreign circular RNAs by intron identity. Mol. Cell 67228–38.e5 [Google Scholar]
  52. Zinder JC, Lima CD. 52.  2017. Targeting RNA for processing or destruction by the eukaryotic RNA exosome and its cofactors. Genes Dev 3188–100 [Google Scholar]
  53. Eckard SC, Rice GI, Fabre A, Badens C, Gray EE. 53.  et al. 2014. The SKIV2L RNA exosome limits activation of the RIG-I-like receptors. Nat. Immunol. 15839–45 [Google Scholar]
  54. Schuberth-Wagner C, Ludwig J, Bruder AK, Herzner AM, Zillinger T. 54.  et al. 2015. A conserved histidine in the RNA sensor RIG-I controls immune tolerance to N1-2′O-methylated self RNA. Immunity 4341–51 [Google Scholar]
  55. Devarkar SC, Wang C, Miller MT, Ramanathan A, Jiang F. 55.  et al. 2016. Structural basis for m7G recognition and 2′-O-methyl discrimination in capped RNAs by the innate immune receptor RIG-I. PNAS 113596–601 [Google Scholar]
  56. Durbin AF, Wang C, Marcotrigiano J, Gehrke L. 56.  2016. RNAs containing modified nucleotides fail to trigger RIG-I conformational changes for innate immune signaling. mBio 7e00833–16 [Google Scholar]
  57. Kato H, Takeuchi O, Mikamo-Satoh E, Hirai R, Kawai T. 57.  et al. 2008. Length-dependent recognition of double-stranded ribonucleic acids by retinoic acid-inducible gene-I and melanoma differentiation-associated gene 5. J. Exp. Med. 2051601–10 [Google Scholar]
  58. Pichlmair A, Schulz O, Tan CP, Rehwinkel J, Kato H. 58.  et al. 2009. Activation of MDA5 requires higher-order RNA structures generated during virus infection. J. Virol. 8310761–69 [Google Scholar]
  59. Feng Q, Hato SV, Langereis MA, Zoll J, Virgen-Slane R. 59.  et al. 2012. MDA5 detects the double-stranded RNA replicative form in picornavirus-infected cells. Cell Rep 21187–96 [Google Scholar]
  60. Triantafilou K, Vakakis E, Kar S, Richer E, Evans GL, Triantafilou M. 60.  2012. Visualisation of direct interaction of MDA5 and the dsRNA replicative intermediate form of positive strand RNA viruses. J. Cell Sci. 1254761–69 [Google Scholar]
  61. Deddouche S, Goubau D, Rehwinkel J, Chakravarty P, Begum S. 61.  et al. 2014. Identification of an LGP2-associated MDA5 agonist in picornavirus-infected cells. eLife 3e01535 [Google Scholar]
  62. Liehl P, Zuzarte-Luis V, Chan J, Zillinger T, Baptista F. 62.  et al. 2014. Host-cell sensors for Plasmodium activate innate immunity against liver-stage infection. Nat. Med. 2047–53 [Google Scholar]
  63. Nishikura K. 63.  2016. A-to-I editing of coding and non-coding RNAs by ADARs. Nat. Rev. Mol. Cell Biol. 1783–96 [Google Scholar]
  64. Liddicoat BJ, Piskol R, Chalk AM, Ramaswami G, Higuchi M. 64.  et al. 2015. RNA editing by ADAR1 prevents MDA5 sensing of endogenous dsRNA as nonself. Science 3491115–20 [Google Scholar]
  65. Mannion NM, Greenwood SM, Young R, Cox S, Brindle J. 65.  et al. 2014. The RNA-editing enzyme ADAR1 controls innate immune responses to RNA. Cell Rep 91482–94 [Google Scholar]
  66. Pestal K, Funk CC, Snyder JM, Price ND, Treuting PM, Stetson DB. 66.  2015. Isoforms of RNA-editing enzyme ADAR1 independently control nucleic acid sensor MDA5-driven autoimmunity and multi-organ development. Immunity 43933–44 [Google Scholar]
  67. Lumb JH, Li Q, Popov LM, Ding S, Keith MT. 67.  et al. 2017. DDX6 represses aberrant activation of interferon-stimulated genes. Cell Rep 20819–31 [Google Scholar]
  68. Zust R, Cervantes-Barragan L, Habjan M, Maier R, Neuman BW. 68.  et al. 2011. Ribose 2′-O-methylation provides a molecular signature for the distinction of self and non-self mRNA dependent on the RNA sensor Mda5. Nat. Immunol. 12137–43 [Google Scholar]
  69. Li X, Ranjith-Kumar CT, Brooks MT, Dharmaiah S, Herr AB. 69.  et al. 2009. The RIG-I-like receptor LGP2 recognizes the termini of double-stranded RNA. J. Biol. Chem. 28413881–91 [Google Scholar]
  70. Uchikawa E, Lethier M, Malet H, Brunel J, Gerlier D, Cusack S. 70.  2016. Structural analysis of dsRNA binding to anti-viral pattern recognition receptors LGP2 and MDA5. Mol. Cell 62586–602 [Google Scholar]
  71. Bamming D, Horvath CM. 71.  2009. Regulation of signal transduction by enzymatically inactive antiviral RNA helicase proteins MDA5, RIG-I, and LGP2. J. Biol. Chem. 2849700–12 [Google Scholar]
  72. Rodero MP, Crow YJ. 72.  2016. Type I interferon–mediated monogenic autoinflammation: the type I interferonopathies, a conceptual overview. J. Exp. Med. 2132527–38 [Google Scholar]
  73. Farrugia M, Baron B. 73.  2017. The role of Toll-like receptors in autoimmune diseases through failure of the self-recognition mechanism. Int. J. Inflamm. 20178391230 [Google Scholar]
  74. Kato H, Fujita T. 74.  2015. RIG-I-like receptors and autoimmune diseases. Curr. Opin. Immunol. 3740–45 [Google Scholar]
  75. Sun Z, Ren H, Liu Y, Teeling JL, Gu J. 75.  2011. Phosphorylation of RIG-I by casein kinase II inhibits its antiviral response. J. Virol. 851036–47 [Google Scholar]
  76. Gack MU, Nistal-Villan E, Inn KS, Garcia-Sastre A, Jung JU. 76.  2010. Phosphorylation-mediated negative regulation of RIG-I antiviral activity. J. Virol. 843220–29 [Google Scholar]
  77. Nistal-Villan E, Gack MU, Martinez-Delgado G, Maharaj NP, Inn KS. 77.  et al. 2010. Negative role of RIG-I serine 8 phosphorylation in the regulation of interferon-beta production. J. Biol. Chem. 28520252–61 [Google Scholar]
  78. Maharaj NP, Wies E, Stoll A, Gack MU. 78.  2012. Conventional protein kinase C-alpha (PKC-alpha) and PKC-beta negatively regulate RIG-I antiviral signal transduction. J. Virol. 861358–71 [Google Scholar]
  79. Wies E, Wang MK, Maharaj NP, Chen K, Zhou S. 79.  et al. 2013. Dephosphorylation of the RNA sensors RIG-I and MDA5 by the phosphatase PP1 is essential for innate immune signaling. Immunity 38437–49 [Google Scholar]
  80. Oshiumi H, Miyashita M, Inoue N, Okabe M, Matsumoto M, Seya T. 80.  2010. The ubiquitin ligase Riplet is essential for RIG-I-dependent innate immune responses to RNA virus infection. Cell Host Microbe 8496–509 [Google Scholar]
  81. Oshiumi H, Matsumoto M, Hatakeyama S, Seya T. 81.  2009. Riplet/RNF135, a RING finger protein, ubiquitinates RIG-I to promote interferon-beta induction during the early phase of viral infection. J. Biol. Chem. 284807–17 [Google Scholar]
  82. Oshiumi H, Miyashita M, Matsumoto M, Seya T. 82.  2013. A distinct role of Riplet-mediated K63-linked polyubiquitination of the RIG-I repressor domain in human antiviral innate immune responses. PLOS Pathog 9e1003533 [Google Scholar]
  83. Gack MU, Shin YC, Joo CH, Urano T, Liang C. 83.  et al. 2007. TRIM25 RING-finger E3 ubiquitin ligase is essential for RIG-I-mediated antiviral activity. Nature 446916–20 [Google Scholar]
  84. Gack MU, Kirchhofer A, Shin YC, Inn KS, Liang C. 84.  et al. 2008. Roles of RIG-I N-terminal tandem CARD and splice variant in TRIM25-mediated antiviral signal transduction. PNAS 10516743–48 [Google Scholar]
  85. Liu HM, Loo YM, Horner SM, Zornetzer GA, Katze MG, Gale M Jr. 85.  2012. The mitochondrial targeting chaperone 14-3-3ε regulates a RIG-I translocon that mediates membrane association and innate antiviral immunity. Cell Host Microbe 11528–37 [Google Scholar]
  86. Peisley A, Wu B, Yao H, Walz T, Hur S. 86.  2013. RIG-I forms signaling-competent filaments in an ATP-dependent, ubiquitin-independent manner. Mol. Cell 51573–83 [Google Scholar]
  87. Peisley A, Wu B, Xu H, Chen ZJ, Hur S. 87.  2014. Structural basis for ubiquitin-mediated antiviral signal activation by RIG-I. Nature 509110–14 [Google Scholar]
  88. Kuniyoshi K, Takeuchi O, Pandey S, Satoh T, Iwasaki H. 88.  et al. 2014. Pivotal role of RNA-binding E3 ubiquitin ligase MEX3C in RIG-I-mediated antiviral innate immunity. PNAS 1115646–51 [Google Scholar]
  89. Yan J, Li Q, Mao AP, Hu MM, Shu HB. 89.  2014. TRIM4 modulates type I interferon induction and cellular antiviral response by targeting RIG-I for K63-linked ubiquitination. J. Mol. Cell Biol. 6154–63 [Google Scholar]
  90. Zeng W, Sun L, Jiang X, Chen X, Hou F. 90.  et al. 2010. Reconstitution of the RIG-I pathway reveals a signaling role of unanchored polyubiquitin chains in innate immunity. Cell 141315–30 [Google Scholar]
  91. Shi Y, Yuan B, Zhu W, Zhang R, Li L. 91.  et al. 2017. Ube2D3 and Ube2N are essential for RIG-I-mediated MAVS aggregation in antiviral innate immunity. Nat. Commun. 815138 [Google Scholar]
  92. Lin W, Zhang J, Lin H, Li Z, Sun X. 92.  et al. 2016. Syndecan-4 negatively regulates antiviral signalling by mediating RIG-I deubiquitination via CYLD. Nat. Commun. 711848 [Google Scholar]
  93. Friedman CS, O'Donnell MA, Legarda-Addison D, Ng A, Cardenas WB. 93.  et al. 2008. The tumour suppressor CYLD is a negative regulator of RIG-I-mediated antiviral response. EMBO Rep 9930–36 [Google Scholar]
  94. Fan Y, Mao R, Yu Y, Liu S, Shi Z. 94.  et al. 2014. USP21 negatively regulates antiviral response by acting as a RIG-I deubiquitinase. J. Exp. Med. 211313–28 [Google Scholar]
  95. Cui J, Song Y, Li Y, Zhu Q, Tan P. 95.  et al. 2014. USP3 inhibits type I interferon signaling by deubiquitinating RIG-I-like receptors. Cell Res 24400–16 [Google Scholar]
  96. Arimoto K, Takahashi H, Hishiki T, Konishi H, Fujita T, Shimotohno K. 96.  2007. Negative regulation of the RIG-I signaling by the ubiquitin ligase RNF125. PNAS 1047500–5 [Google Scholar]
  97. Wang L, Zhao W, Zhang M, Wang P, Zhao K. 97.  et al. 2013. USP4 positively regulates RIG-I-mediated antiviral response through deubiquitination and stabilization of RIG-I. J. Virol. 874507–15 [Google Scholar]
  98. Liu HM, Jiang F, Loo YM, Hsu S, Hsiang TY, Marcotrigiano J, Gale M Jr. 98.  2016. Regulation of retinoic acid inducible gene-I (RIG-I) activation by the histone deacetylase 6. EBioMedicine 9195–206 [Google Scholar]
  99. Choi SJ, Lee HC, Kim JH, Park SY, Kim TH. 99.  et al. 2016. HDAC6 regulates cellular viral RNA sensing by deacetylation of RIG-I. EMBO J 35429–42 [Google Scholar]
  100. Kim MJ, Hwang SY, Imaizumi T, Yoo JY. 100.  2008. Negative feedback regulation of RIG-I-mediated antiviral signaling by interferon-induced ISG15 conjugation. J. Virol. 821474–83 [Google Scholar]
  101. Takashima K, Oshiumi H, Takaki H, Matsumoto M, Seya T. 101.  2015. RIOK3-mediated phosphorylation of MDA5 interferes with its assembly and attenuates the innate immune response. Cell Rep 11192–200 [Google Scholar]
  102. Takashima K, Oshiumi H, Seya T. 102.  2015. RIOK3 keeps MDA5 inactive. Oncotarget 630423–24 [Google Scholar]
  103. Jiang X, Kinch LN, Brautigam CA, Chen X, Du F. 103.  et al. 2012. Ubiquitin-induced oligomerization of the RNA sensors RIG-I and MDA5 activates antiviral innate immune response. Immunity 36959–73 [Google Scholar]
  104. Hu MM, Liao CY, Yang Q, Xie XQ, Shu HB. 104.  2017. Innate immunity to RNA virus is regulated by temporal and reversible sumoylation of RIG-I and MDA5. J. Exp. Med. 214973–89 [Google Scholar]
  105. Lang X, Tang T, Jin T, Ding C, Zhou R, Jiang W. 105.  2017. TRIM65-catalized ubiquitination is essential for MDA5-mediated antiviral innate immunity. J. Exp. Med. 214459–73 [Google Scholar]
  106. Baker PJ, De Nardo D, Moghaddas F, Tran LS, Bachem A. 106.  et al. 2017. Posttranslational modification as a critical determinant of cytoplasmic innate immune recognition. Physiol. Rev. 971165–209 [Google Scholar]
  107. Liu J, Qian C, Cao X. 107.  2016. Post-translational modification control of innate immunity. Immunity 4515–30 [Google Scholar]
  108. Peisley A, Lin C, Wu B, Orme-Johnson M, Liu M, Walz T, Hur S. 108.  2011. Cooperative assembly and dynamic disassembly of MDA5 filaments for viral dsRNA recognition. PNAS 10821010–15 [Google Scholar]
  109. Wu B, Peisley A, Richards C, Yao H, Zeng X. 109.  et al. 2013. Structural basis for dsRNA recognition, filament formation, and antiviral signal activation by MDA5. Cell 152276–89 [Google Scholar]
  110. Berke IC, Yu X, Modis Y, Egelman EH. 110.  2012. MDA5 assembles into a polar helical filament on dsRNA. PNAS 10918437–41 [Google Scholar]
  111. Wu B, Peisley A, Tetrault D, Li Z, Egelman EH. 111.  et al. 2014. Molecular imprinting as a signal-activation mechanism of the viral RNA sensor RIG-I. Mol. Cell 55511–23 [Google Scholar]
  112. Peisley A, Jo MH, Lin C, Wu B, Orme-Johnson M. 112.  et al. 2012. Kinetic mechanism for viral dsRNA length discrimination by MDA5 filaments. PNAS 109E3340–49 [Google Scholar]
  113. Motz C, Schuhmann KM, Kirchhofer A, Moldt M, Witte G. 113.  et al. 2013. Paramyxovirus V proteins disrupt the fold of the RNA sensor MDA5 to inhibit antiviral signaling. Science 339690–93 [Google Scholar]
  114. Gao D, Li W. 114.  2017. Structures and recognition modes of Toll-like receptors. Proteins 853–9 [Google Scholar]
  115. Leifer CA, Medvedev AE. 115.  2016. Molecular mechanisms of regulation of Toll-like receptor signaling. J. Leukoc. Biol. 100927–41 [Google Scholar]
  116. Satoh T, Akira S. 116.  2016. Toll-like receptor signaling and its inducible proteins. Microbiol. Spectr. 4MCHD–0040-2016 [Google Scholar]
  117. Kang S, Fernandes-Alnemri T, Rogers C, Mayes L, Wang Y. 117.  et al. 2015. Caspase-8 scaffolding function and MLKL regulate NLRP3 inflammasome activation downstream of TLR3. Nat. Commun. 67515 [Google Scholar]
  118. Li XD, Chen ZJ. 118.  2012. Sequence specific detection of bacterial 23S ribosomal RNA by TLR13. eLife 1e00102 [Google Scholar]
  119. Oldenburg M, Kruger A, Ferstl R, Kaufmann A, Nees G. 119.  et al. 2012. TLR13 recognizes bacterial 23S rRNA devoid of erythromycin resistance-forming modification. Science 3371111–15 [Google Scholar]
  120. Shi Z, Cai Z, Sanchez A, Zhang T, Wen S. 120.  et al. 2011. A novel Toll-like receptor that recognizes vesicular stomatitis virus. J. Biol. Chem. 2864517–24 [Google Scholar]
  121. Zhang Z, Ohto U, Shimizu T. 121.  2017. Toward a structural understanding of nucleic acid-sensing Toll-like receptors in the innate immune system. FEBS Lett 5913167–81 [Google Scholar]
  122. Bell JK, Botos I, Hall PR, Askins J, Shiloach J. 122.  et al. 2005. The molecular structure of the Toll-like receptor 3 ligand-binding domain. PNAS 10210976–80 [Google Scholar]
  123. Liu L, Botos I, Wang Y, Leonard JN, Shiloach J. 123.  et al. 2008. Structural basis of Toll-like receptor 3 signaling with double-stranded RNA. Science 320379–81 [Google Scholar]
  124. Choe J, Kelker MS, Wilson IA. 124.  2005. Crystal structure of human Toll-like receptor 3 (TLR3) ectodomain. Science 309581–85 [Google Scholar]
  125. Leonard JN, Ghirlando R, Askins J, Bell JK, Margulies DH. 125.  et al. 2008. The TLR3 signaling complex forms by cooperative receptor dimerization. PNAS 105258–63 [Google Scholar]
  126. Fukuda K, Watanabe T, Tokisue T, Tsujita T, Nishikawa S. 126.  et al. 2008. Modulation of double-stranded RNA recognition by the N-terminal histidine-rich region of the human Toll-like receptor 3. J. Biol. Chem. 28322787–94 [Google Scholar]
  127. Hardarson HS, Baker JS, Yang Z, Purevjav E, Huang CH. 127.  et al. 2007. Toll-like receptor 3 is an essential component of the innate stress response in virus-induced cardiac injury. Am. J. Physiol. Heart Circ. Physiol. 292H251–58 [Google Scholar]
  128. Daffis S, Samuel MA, Suthar MS, Gale M Jr., Diamond MS. 128.  2008. Toll-like receptor 3 has a protective role against West Nile virus infection. J. Virol. 8210349–58 [Google Scholar]
  129. Daffis S, Samuel MA, Suthar MS, Keller BC, Gale M Jr., Diamond MS. 129.  2008. Interferon regulatory factor IRF-7 induces the antiviral alpha interferon response and protects against lethal West Nile virus infection. J. Virol. 828465–75 [Google Scholar]
  130. Schulz O, Diebold SS, Chen M, Naslund TI, Nolte MA. 130.  et al. 2005. Toll-like receptor 3 promotes cross-priming to virus-infected cells. Nature 433887–92 [Google Scholar]
  131. Tabeta K, Georgel P, Janssen E, Du X, Hoebe K. 131.  et al. 2004. Toll-like receptors 9 and 3 as essential components of innate immune defense against mouse cytomegalovirus infection. PNAS 1013516–21 [Google Scholar]
  132. Wang T, Town T, Alexopoulou L, Anderson JF, Fikrig E, Flavell RA. 132.  2004. Toll-like receptor 3 mediates West Nile virus entry into the brain causing lethal encephalitis. Nat. Med. 101366–73 [Google Scholar]
  133. Tatematsu M, Nishikawa F, Seya T, Matsumoto M. 133.  2013. Toll-like receptor 3 recognizes incomplete stem structures in single-stranded viral RNA. Nat. Commun. 41833 [Google Scholar]
  134. Abe Y, Fujii K, Nagata N, Takeuchi O, Akira S. 134.  et al. 2012. The Toll-like receptor 3-mediated antiviral response is important for protection against poliovirus infection in poliovirus receptor transgenic mice. J. Virol. 86185–94 [Google Scholar]
  135. Kariko K, Ni H, Capodici J, Lamphier M, Weissman D. 135.  2004. mRNA is an endogenous ligand for Toll-like receptor 3. J. Biol. Chem. 27912542–50 [Google Scholar]
  136. Murakami Y, Fukui R, Motoi Y, Kanno A, Shibata T. 136.  et al. 2014. Roles of the cleaved N-terminal TLR3 fragment and cell surface TLR3 in double-stranded RNA sensing. J. Immunol. 1935208–17 [Google Scholar]
  137. Groskreutz DJ, Monick MM, Powers LS, Yarovinsky TO, Look DC, Hunninghake GW. 137.  2006. Respiratory syncytial virus induces TLR3 protein and protein kinase R, leading to increased double-stranded RNA responsiveness in airway epithelial cells. J. Immunol. 1761733–40 [Google Scholar]
  138. Matsumoto M, Kikkawa S, Kohase M, Miyake K, Seya T. 138.  2002. Establishment of a monoclonal antibody against human Toll-like receptor 3 that blocks double-stranded RNA-mediated signaling. Biochem. Biophys. Res. Commun. 2931364–69 [Google Scholar]
  139. Pohar J, Pirher N, Bencina M, Mancek-Keber M, Jerala R. 139.  2013. The role of UNC93B1 protein in surface localization of TLR3 receptor and in cell priming to nucleic acid agonists. J. Biol. Chem. 288442–54 [Google Scholar]
  140. Tanji H, Ohto U, Shibata T, Miyake K, Shimizu T. 140.  2013. Structural reorganization of the Toll-like receptor 8 dimer induced by agonistic ligands. Science 3391426–29 [Google Scholar]
  141. Tanji H, Ohto U, Shibata T, Taoka M, Yamauchi Y. 141.  et al. 2015. Toll-like receptor 8 senses degradation products of single-stranded RNA. Nat. Struct. Mol. Biol. 22109–15 [Google Scholar]
  142. Lund JM, Alexopoulou L, Sato A, Karow M, Adams NC. 142.  et al. 2004. Recognition of single-stranded RNA viruses by Toll-like receptor 7. PNAS 1015598–603 [Google Scholar]
  143. Diebold SS, Kaisho T, Hemmi H, Akira S, Reis e Sousa C. 143.  2004. Innate antiviral responses by means of TLR7-mediated recognition of single-stranded RNA. Science 3031529–31 [Google Scholar]
  144. Lee HK, Lund JM, Ramanathan B, Mizushima N, Iwasaki A. 144.  2007. Autophagy-dependent viral recognition by plasmacytoid dendritic cells. Science 3151398–401 [Google Scholar]
  145. Hornung V, Schlender J, Guenthner-Biller M, Rothenfusser S, Endres S. 145.  et al. 2004. Replication-dependent potent IFN-alpha induction in human plasmacytoid dendritic cells by a single-stranded RNA virus. J. Immunol. 1735935–43 [Google Scholar]
  146. Heil F, Hemmi H, Hochrein H, Ampenberger F, Kirschning C. 146.  et al. 2004. Species-specific recognition of single-stranded RNA via Toll-like receptor 7 and 8. Science 3031526–29 [Google Scholar]
  147. Awais M, Wang K, Lin X, Qian W, Zhang N. 147.  et al. 2017. TLR7 deficiency leads to TLR8 compensative regulation of immune response against JEV in mice. Front. Immunol. 8160 [Google Scholar]
  148. Mancuso G, Gambuzza M, Midiri A, Biondo C, Papasergi S. 148.  et al. 2009. Bacterial recognition by TLR7 in the lysosomes of conventional dendritic cells. Nat. Immunol. 10587–94 [Google Scholar]
  149. Eberle F, Sirin M, Binder M, Dalpke AH. 149.  2009. Bacterial RNA is recognized by different sets of immunoreceptors. Eur. J. Immunol. 392537–47 [Google Scholar]
  150. Gantier MP, Tong S, Behlke MA, Xu D, Phipps S. 150.  et al. 2008. TLR7 is involved in sequence-specific sensing of single-stranded RNAs in human macrophages. J. Immunol. 1802117–24 [Google Scholar]
  151. Cervantes JL, Dunham-Ems SM, La Vake CJ, Petzke MM, Sahay B. 151.  et al. 2011. Phagosomal signaling by Borrelia burgdorferi in human monocytes involves Toll-like receptor (TLR) 2 and TLR8 cooperativity and TLR8-mediated induction of IFN-beta. PNAS 1083683–88 [Google Scholar]
  152. Eigenbrod T, Pelka K, Latz E, Kreikemeyer B, Dalpke AH. 152.  2015. TLR8 senses bacterial RNA in human monocytes and plays a nonredundant role for recognition of Streptococcus pyogenes. . J. Immunol. 1951092–99 [Google Scholar]
  153. Choo MK, Sano Y, Kim C, Yasuda K, Li XD. 153.  et al. 2017. TLR sensing of bacterial spore-associated RNA triggers host immune responses with detrimental effects. J. Exp. Med. 2141297–311 [Google Scholar]
  154. Lehmann SM, Kruger C, Park B, Derkow K, Rosenberger K. 154.  et al. 2012. An unconventional role for miRNA: let-7 activates Toll-like receptor 7 and causes neurodegeneration. Nat. Neurosci. 15827–35 [Google Scholar]
  155. Gehrig S, Eberle ME, Botschen F, Rimbach K, Eberle F. 155.  et al. 2012. Identification of modifications in microbial, native tRNA that suppress immunostimulatory activity. J. Exp. Med. 209225–33 [Google Scholar]
  156. Jockel S, Nees G, Sommer R, Zhao Y, Cherkasov D. 156.  et al. 2012. The 2′-O-methylation status of a single guanosine controls transfer RNA-mediated Toll-like receptor 7 activation or inhibition. J. Exp. Med. 209235–41 [Google Scholar]
  157. Rimbach K, Kaiser S, Helm M, Dalpke AH, Eigenbrod T. 157.  2015. 2′-O-methylation within bacterial RNA acts as suppressor of TLR7/TLR8 activation in human innate immune cells. J. Innate Immun. 7482–93 [Google Scholar]
  158. Schmitt FCF, Freund I, Weigand MA, Helm M, Dalpke AH, Eigenbrod T. 158.  2017. Identification of an optimized 2′-O-methylated trinucleotide RNA motif inhibiting Toll-like receptors 7 and 8. RNA 231344–51 [Google Scholar]
  159. Gorden KK, Qiu XX, Binsfeld CC, Vasilakos JP, Alkan SS. 159.  2006. Cutting edge: activation of murine TLR8 by a combination of imidazoquinoline immune response modifiers and polyT oligodeoxynucleotides. J. Immunol. 1776584–87 [Google Scholar]
  160. Kandimalla ER, Struthers M, Bett AJ, Wisniewski T, Dubey SA. 160.  et al. 2011. Synthesis and immunological activities of novel Toll-like receptor 7 and 8 agonists. Cell Immunol 270126–34 [Google Scholar]
  161. Hayashi F, Means TK, Luster AD. 161.  2003. Toll-like receptors stimulate human neutrophil function. Blood 1022660–69 [Google Scholar]
  162. Kadowaki N, Ho S, Antonenko S, Malefyt RW, Kastelein RA. 162.  et al. 2001. Subsets of human dendritic cell precursors express different Toll-like receptors and respond to different microbial antigens. J. Exp. Med. 194863–69 [Google Scholar]
  163. Krug A, Towarowski A, Britsch S, Rothenfusser S, Hornung V. 163.  et al. 2001. Toll-like receptor expression reveals CpG DNA as a unique microbial stimulus for plasmacytoid dendritic cells which synergizes with CD40 ligand to induce high amounts of IL-12. Eur. J. Immunol. 313026–37 [Google Scholar]
  164. Hornung V, Rothenfusser S, Britsch S, Krug A, Jahrsdorfer B. 164.  et al. 2002. Quantitative expression of Toll-like receptor 1–10 mRNA in cellular subsets of human peripheral blood mononuclear cells and sensitivity to CpG oligodeoxynucleotides. J. Immunol. 1684531–37 [Google Scholar]
  165. Ito T, Amakawa R, Kaisho T, Hemmi H, Tajima K. 165.  et al. 2002. Interferon-alpha and interleukin-12 are induced differentially by Toll-like receptor 7 ligands in human blood dendritic cell subsets. J. Exp. Med. 1951507–12 [Google Scholar]
  166. Forsbach A, Nemorin JG, Montino C, Muller C, Samulowitz U. 166.  et al. 2008. Identification of RNA sequence motifs stimulating sequence-specific TLR8-dependent immune responses. J. Immunol. 1803729–38 [Google Scholar]
  167. Brinkmann MM, Spooner E, Hoebe K, Beutler B, Ploegh HL, Kim YM. 167.  2007. The interaction between the ER membrane protein UNC93B and TLR3, 7, and 9 is crucial for TLR signaling. J. Cell Biol. 177265–75 [Google Scholar]
  168. Tabeta K, Hoebe K, Janssen EM, Du X, Georgel P. 168.  et al. 2006. The Unc93b1 mutation 3d disrupts exogenous antigen presentation and signaling via Toll-like receptors 3, 7 and 9. Nat. Immunol. 7156–64 [Google Scholar]
  169. Fukui R, Saitoh S, Matsumoto F, Kozuka-Hata H, Oyama M. 169.  et al. 2009. Unc93B1 biases Toll-like receptor responses to nucleic acid in dendritic cells toward DNA- but against RNA-sensing. J. Exp. Med. 2061339–50 [Google Scholar]
  170. Fukui R, Saitoh S, Kanno A, Onji M, Shibata T. 170.  et al. 2011. Unc93B1 restricts systemic lethal inflammation by orchestrating Toll-like receptor 7 and 9 trafficking. Immunity 3569–81 [Google Scholar]
  171. Pelka K, Phulphagar K, Zimmermann J, Stahl R, Schmid-Burgk JL. 171.  et al. 2014. Cutting edge: The UNC93B1 tyrosine-based motif regulates trafficking and TLR responses via separate mechanisms. J. Immunol. 1933257–61 [Google Scholar]
  172. Tatematsu M, Funami K, Ishii N, Seya T, Obuse C, Matsumoto M. 172.  2015. LRRC59 regulates trafficking of nucleic acid-sensing TLRs from the endoplasmic reticulum via association with UNC93B1. J. Immunol. 1954933–42 [Google Scholar]
  173. Kono DH, Haraldsson MK, Lawson BR, Pollard KM, Koh YT. 173.  et al. 2009. Endosomal TLR signaling is required for anti-nucleic acid and rheumatoid factor autoantibodies in lupus. PNAS 10612061–66 [Google Scholar]
  174. Nakano S, Morimoto S, Suzuki S, Watanabe T, Amano H, Takasaki Y. 174.  2010. Up-regulation of the endoplasmic reticulum transmembrane protein UNC93B in the B cells of patients with active systemic lupus erythematosus. Rheumatology 49876–81 [Google Scholar]
  175. Pelka K, Shibata T, Miyake K, Latz E. 175.  2016. Nucleic acid-sensing TLRs and autoimmunity: novel insights from structural and cell biology. Immunol. Rev. 26960–75 [Google Scholar]
  176. Hipp MM, Shepherd D, Gileadi U, Aichinger MC, Kessler BM. 176.  et al. 2013. Processing of human Toll-like receptor 7 by furin-like proprotein convertases is required for its accumulation and activity in endosomes. Immunity 39711–21 [Google Scholar]
  177. Eigenbrod T, Pelka K, Latz E, Kreikemeyer B, Dalpke AH. 177.  2015. TLR8 senses bacterial RNA in human monocytes and plays a nonredundant role for recognition of Streptococcus pyogenes. . J. Immunol. 1951092–99 [Google Scholar]
  178. Ariumi Y. 178.  2014. Multiple functions of DDX3 RNA helicase in gene regulation, tumorigenesis, and viral infection. Front. Genet. 5423 [Google Scholar]
  179. Valiente-Echeverria F, Hermoso MA, Soto-Rifo R. 179.  2015. RNA helicase DDX3: at the crossroad of viral replication and antiviral immunity.. Rev. Med. Virol. 25286–99 [Google Scholar]
  180. Gringhuis SI, Hertoghs N, Kaptein TM, Zijlstra-Willems EM, Sarrami-Fooroshani R. 180.  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. 18225–35 [Google Scholar]
  181. Oshiumi H, Sakai K, Matsumoto M, Seya T. 181.  2010. DEAD/H box 3 (DDX3) helicase binds the RIG-I adaptor IPS-1 to up-regulate IFN-beta-inducing potential. Eur. J. Immunol. 40940–48 [Google Scholar]
  182. Zhang Z, Yuan B, Lu N, Facchinetti V, Liu YJ. 182.  2011. DHX9 pairs with IPS-1 to sense double-stranded RNA in myeloid dendritic cells. J. Immunol. 1874501–8 [Google Scholar]
  183. Aktas T, Avsar Ilik I, Maticzka D, Bhardwaj V, Pessoa Rodrigues C. 183.  et al. 2017. DHX9 suppresses RNA processing defects originating from the Alu invasion of the human genome. Nature 544115–19 [Google Scholar]
  184. Zhang Z, Kim T, Bao M, Facchinetti V, Jung SY. 184.  et al. 2011. DDX1, DDX21, and DHX36 helicases form a complex with the adaptor molecule TRIF to sense dsRNA in dendritic cells. Immunity 34866–78 [Google Scholar]
  185. Mitoma H, Hanabuchi S, Kim T, Bao M, Zhang Z. 185.  et al. 2013. The DHX33 RNA helicase senses cytosolic RNA and activates the NLRP3 inflammasome. Immunity 39123–35 [Google Scholar]
  186. Liu Y, Lu N, Yuan B, Weng L, Wang F. 186.  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. 1149–57 [Google Scholar]
  187. Subramanian N, Natarajan K, Clatworthy MR, Wang Z, Germain RN. 187.  2013. The adaptor MAVS promotes NLRP3 mitochondrial localization and inflammasome activation. Cell 153348–61 [Google Scholar]
  188. Chakrabarti A, Banerjee S, Franchi L, Loo YM, Gale M. 188.  Jr., et al. 2015. RNase L activates the NLRP3 inflammasome during viral infections. Cell Host Microbe 17466–77 [Google Scholar]
  189. Mosallanejad K, Sekine Y, Ishikura-Kinoshita S, Kumagai K, Nagano T. 189.  et al. 2014. The DEAH-box RNA helicase DHX15 activates NF-κB and MAPK signaling downstream of MAVS during antiviral responses. Sci. Signal. 7ra40 [Google Scholar]
  190. Lu H, Lu N, Weng L, Yuan B, Liu YJ, Zhang Z. 190.  2014. DHX15 senses double-stranded RNA in myeloid dendritic cells. J. Immunol. 1931364–72 [Google Scholar]
  191. Wang P, Zhu S, Yang L, Cui S, Pan W. 191.  et al. 2015. Nlrp6 regulates intestinal antiviral innate immunity. Science 350826–30 [Google Scholar]
  192. Miyashita M, Oshiumi H, Matsumoto M, Seya T. 192.  2011. DDX60, a DEXD/H box helicase, is a novel antiviral factor promoting RIG-I-like receptor-mediated signaling. Mol. Cell. Biol. 313802–19 [Google Scholar]
  193. Oshiumi H, Miyashita M, Okamoto M, Morioka Y, Okabe M. 193.  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 111193–207 [Google Scholar]
  194. Schoggins JW, Wilson SJ, Panis M, Murphy MY, Jones CT. 194.  et al. 2011. A diverse range of gene products are effectors of the type I interferon antiviral response. Nature 472481–85 [Google Scholar]
  195. Goubau D, van der Veen AG, Chakravarty P, Lin R, Rogers N. 195.  et al. 2015. Mouse superkiller-2-like helicase DDX60 is dispensable for type I IFN induction and immunity to multiple viruses. Eur. J. Immunol. 453386–403 [Google Scholar]
  196. Tremblay N, Baril M, Chatel-Chaix L, Es-Saad S, Park AY. 196.  et al. 2016. Spliceosome SNRNP200 promotes viral RNA sensing and IRF3 activation of antiviral response. PLOS Pathog 12e1005772 [Google Scholar]
  197. Luthra P, Sun D, Silverman RH, He B. 197.  2011. Activation of IFN-β expression by a viral mRNA through RNase L and MDA5. PNAS 1082118–23 [Google Scholar]
  198. Li LF, Yu J, Zhang Y, Yang Q, Li Y. 198.  et al. 2017. Interferon-inducible oligoadenylate synthetase-like protein acts as an antiviral effector against classical swine fever virus via the MDA5-mediated type I interferon-signaling pathway. J. Virol. 91e01514–16 [Google Scholar]
  199. Hull CM, Bevilacqua PC. 199.  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. 491242–49 [Google Scholar]
  200. Pham AM, Santa Maria FG, Lahiri T, Friedman E, Marie IJ. 200.  et al. 2016. PKR transduces MDA5-dependent signals for type I IFN induction. PLOS Pathog 12e1005489 [Google Scholar]
  201. Fensterl V, Sen GC. 201.  2015. Interferon-induced Ifit proteins: their role in viral pathogenesis. J. Virol. 892462–68 [Google Scholar]
  202. Yang P, An H, Liu X, Wen M, Zheng Y. 202.  et al. 2010. The cytosolic nucleic acid sensor LRRFIP1 mediates the production of type I interferon via a beta-catenin-dependent pathway. Nat. Immunol. 11487–94 [Google Scholar]
  203. Kuriakose T, Man SM, Malireddi RK, Karki R, Kesavardhana S. 203.  et al. 2016. ZBP1/DAI is an innate sensor of influenza virus triggering the NLRP3 inflammasome and programmed cell death pathways. Sci. Immunol. 1aag2045 [Google Scholar]
  204. Thapa RJ, Ingram JP, Ragan KB, Nogusa S, Boyd DF. 204.  et al. 2016. DAI senses influenza A virus genomic RNA and activates RIPK3-dependent cell death. Cell Host Microbe 20674–81 [Google Scholar]
  205. Ugrinova I, Pasheva E. 205.  2017. HMGB1 protein: a therapeutic target inside and outside the cell. Adv. Protein Chem. Struct. Biol. 10737–76 [Google Scholar]
/content/journals/10.1146/annurev-immunol-042617-053309
Loading
/content/journals/10.1146/annurev-immunol-042617-053309
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