1932

Abstract

Understanding the factors that shape viral evolution is critical for developing effective antiviral strategies, accurately predicting viral evolution, and preventing pandemics. One fundamental determinant of viral evolution is the interplay between viral protein biophysics and the host machineries that regulate protein folding and quality control. Most adaptive mutations in viruses are biophysically deleterious, resulting in a viral protein product with folding defects. In cells, protein folding is assisted by a dynamic system of chaperones and quality control processes known as the proteostasis network. Host proteostasis networks can determine the fates of viral proteins with biophysical defects, either by assisting with folding or by targeting them for degradation. In this review, we discuss and analyze new discoveries revealing that host proteostasis factors can profoundly shape the sequence space accessible to evolving viral proteins. We also discuss the many opportunities for research progress proffered by the proteostasis perspective on viral evolution and adaptation.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-virology-100220-112120
2023-09-29
2024-04-25
Loading full text...

Full text loading...

/deliver/fulltext/virology/10/1/annurev-virology-100220-112120.html?itemId=/content/journals/10.1146/annurev-virology-100220-112120&mimeType=html&fmt=ahah

Literature Cited

  1. 1.
    Elena SF, Lenski RE. 2003. Evolution experiments with microorganisms: the dynamics and genetic bases of adaptation. Nat. Rev. Genet. 4:457–69
    [Google Scholar]
  2. 2.
    Sanjuán R, Nebot MR, Chirico N, Mansky LM, Belshaw R. 2010. Viral mutation rates. J. Virol. 84:9733–48
    [Google Scholar]
  3. 3.
    Peck KM, Lauring AS. 2018. Complexities of viral mutation rates. J. Virol. 92:e01031–17
    [Google Scholar]
  4. 4.
    Worobey M, Holmes EC. 1999. Evolutionary aspects of recombination in RNA viruses. J. Gen. Virol. 80:Part 102535–43
    [Google Scholar]
  5. 5.
    Lowen AC. 2017. Constraints, drivers, and implications of influenza A virus reassortment. Annu. Rev. Virol. 4:105–21
    [Google Scholar]
  6. 6.
    Montville R, Froissart R, Remold SK, Tenaillon O, Turner PE. 2005. Evolution of mutational robustness in an RNA virus. PLOS Biol. 3:e381
    [Google Scholar]
  7. 7.
    Lauring AS, Frydman J, Andino R. 2013. The role of mutational robustness in RNA virus evolution. Nat. Rev. Microbiol. 11:327–36
    [Google Scholar]
  8. 8.
    Wylie CS, Shakhnovich EI. 2011. A biophysical protein folding model accounts for most mutational fitness effects in viruses. PNAS 108:9916–21
    [Google Scholar]
  9. 9.
    Klaips CL, Jayaraj GG, Hartl FU. 2018. Pathways of cellular proteostasis in aging and disease. J. Cell Biol. 217:51–63
    [Google Scholar]
  10. 10.
    Mogk A, Huber D, Bukau B. 2011. Integrating protein homeostasis strategies in prokaryotes. Cold Spring Harb. Perspect. Biol. 3:a004366
    [Google Scholar]
  11. 11.
    Aviner R, Frydman J. 2020. Proteostasis in viral infection: unfolding the complex virus–chaperone interplay. Cold Spring Harb. Perspect. Biol. 12:a034090
    [Google Scholar]
  12. 12.
    Wan Q, Song D, Li H, He ML. 2020. Stress proteins: the biological functions in virus infection, present and challenges for target-based antiviral drug development. Signal Transduct. Target. Ther. 5:125
    [Google Scholar]
  13. 13.
    de Visser JAGM, Hermisson J, Wagner GP, Ancel Meyers L, Bagheri-Chaichian H et al. 2003. Perspective: evolution and detection of genetic robustness. Evolution 57:1959–72
    [Google Scholar]
  14. 14.
    Sanjuán R. 2012. From molecular genetics to phylodynamics: evolutionary relevance of mutation rates across viruses. PLOS Pathog. 8:e1002685
    [Google Scholar]
  15. 15.
    Duffy S, Shackelton LA, Holmes EC. 2008. Rates of evolutionary change in viruses: patterns and determinants. Nat. Rev. Genet. 9:267–76
    [Google Scholar]
  16. 16.
    Keightley PD, Lynch M. 2003. Toward a realistic model of mutations affecting fitness. Evolution 57:683–85
    [Google Scholar]
  17. 17.
    Orr HA. 2003. The distribution of fitness effects among beneficial mutations. Genetics 163:1519–26
    [Google Scholar]
  18. 18.
    Shen X, Song S, Li C, Zhang J. 2022. Synonymous mutations in representative yeast genes are mostly strongly non-neutral. Nature 606:725–31
    [Google Scholar]
  19. 19.
    Sanjuán R. 2010. Mutational fitness effects in RNA and single-stranded DNA viruses: common patterns revealed by site-directed mutagenesis studies. Philos. Trans. R. Soc. B 365:1975–82
    [Google Scholar]
  20. 20.
    Steinhauer DA, Domingo E, Holland JJ. 1992. Lack of evidence for proofreading mechanisms associated with an RNA virus polymerase. Gene 122:281–88
    [Google Scholar]
  21. 21.
    Holmes EC. 2003. Error thresholds and the constraints to RNA virus evolution. Trends Microbiol. 11:543–46
    [Google Scholar]
  22. 22.
    Tokuriki N, Stricher F, Schymkowitz J, Serrano L, Tawfik DS. 2007. The stability effects of protein mutations appear to be universally distributed. J. Mol. Biol. 369:1318–32
    [Google Scholar]
  23. 23.
    DePristo MA, Weinreich DM, Hartl DL. 2005. Missense meanderings in sequence space: a biophysical view of protein evolution. Nat. Rev. Genet. 6:678–87
    [Google Scholar]
  24. 24.
    Bloom JD, Labthavikul ST, Otey CR, Arnold FH. 2006. Protein stability promotes evolvability. PNAS 103:5869–74
    [Google Scholar]
  25. 25.
    Tokuriki N, Tawfik DS. 2009. Stability effects of mutations and protein evolvability. Curr. Opin. Struct. Biol. 19:596–604
    [Google Scholar]
  26. 26.
    Zeldovich KB, Chen P, Shakhnovich EI. 2007. Protein stability imposes limits on organism complexity and speed of molecular evolution. PNAS 104:16152–57
    [Google Scholar]
  27. 27.
    Tokuriki N, Stricher F, Serrano L, Tawfik DS. 2008. How protein stability and new functions trade off. PLOS Comput. Biol. 4:e1000002
    [Google Scholar]
  28. 28.
    Lafforgue G, Michon T, Charon J. 2022. Analysis of the contribution of intrinsic disorder in shaping potyvirus genetic diversity. Viruses 14:1959
    [Google Scholar]
  29. 29.
    Lefeuvre P, Lett JM, Reynaud B, Martin DP. 2007. Avoidance of protein fold disruption in natural virus recombinants. PLOS Pathog. 3:e181
    [Google Scholar]
  30. 30.
    Strobel HM, Horwitz EK, Meyer JR. 2022. Viral protein instability enhances host-range evolvability. PLOS Genet. 18:e1010030
    [Google Scholar]
  31. 31.
    Olabode AS, Kandathil SM, Lovell SC, Robertson DL. 2017. Adaptive HIV-1 evolutionary trajectories are constrained by protein stability. Virus Evol. 3:vex019
    [Google Scholar]
  32. 32.
    Gong LI, Suchard MA, Bloom JD. 2013. Stability-mediated epistasis constrains the evolution of an influenza protein. eLife 2:e00631
    [Google Scholar]
  33. 33.
    Klein EY, Blumenkrantz D, Serohijos A, Shakhnovich E, Choi JM et al. 2018. Stability of the influenza virus hemagglutinin protein correlates with evolutionary dynamics. mSphere 3:e00554–17
    [Google Scholar]
  34. 34.
    Rotem A, Serohijos AWR, Chang CB, Wolfe JT, Fischer AE et al. 2018. Evolution on the biophysical fitness landscape of an RNA virus. Mol. Biol. Evol. 35:2390–400
    [Google Scholar]
  35. 35.
    Labbadia J, Morimoto RI. 2015. The biology of proteostasis in aging and disease. Annu. Rev. Biochem. 84:435–64
    [Google Scholar]
  36. 36.
    Sebastian RM, Shoulders MD. 2020. Chemical biology framework to illuminate proteostasis. Annu. Rev. Biochem. 89:529–55
    [Google Scholar]
  37. 37.
    Powers ET, Morimoto RI, Dillin A, Kelly JW, Balch WE. 2009. Biological and chemical approaches to diseases of proteostasis deficiency. Annu. Rev. Biochem. 78:959–91
    [Google Scholar]
  38. 38.
    Hartl FU, Bracher A, Hayer-Hartl M. 2011. Molecular chaperones in protein folding and proteostasis. Nature 475:324–32
    [Google Scholar]
  39. 39.
    Chen B, Retzlaff M, Roos T, Frydman J. 2011. Cellular strategies of protein quality control. Cold Spring Harb. Perspect. Biol. 3:a004374
    [Google Scholar]
  40. 40.
    Sun Z, Brodsky JL. 2019. Protein quality control in the secretory pathway. J. Cell Biol. 218:3171–87
    [Google Scholar]
  41. 41.
    Jayaraj GG, Hipp MS, Hartl FU. 2020. Functional modules of the proteostasis network. Cold Spring Harb. Perspect. Biol. 12:a033951
    [Google Scholar]
  42. 42.
    Hipp MS, Kasturi P, Hartl FU. 2019. The proteostasis network and its decline in ageing. Nat. Rev. Mol. Cell Biol. 20:421–35
    [Google Scholar]
  43. 43.
    Wong MY, DiChiara AS, Suen PH, Chen K, Doan ND, Shoulders MD. 2018. Adapting secretory proteostasis and function through the unfolded protein response. Curr. Top. Microbiol. Immunol. 414:1–25
    [Google Scholar]
  44. 44.
    Braakman I, van Anken E. 2000. Folding of viral envelope glycoproteins in the endoplasmic reticulum. Traffic 1:533–39
    [Google Scholar]
  45. 45.
    Åkerfelt M, Morimoto RI, Sistonen L. 2010. Heat shock factors: integrators of cell stress, development and lifespan. Nat. Rev. Mol. Cell Biol. 11:545–55
    [Google Scholar]
  46. 46.
    Walter P, Ron D 2011. The unfolded protein response: from stress pathway to homeostatic regulation. Science 334:1081–86
    [Google Scholar]
  47. 47.
    Shoulders MD, Ryno LM, Genereux JC, Moresco JJ, Tu PG et al. 2013. Stress-independent activation of XBP1s and/or ATF6 reveals three functionally diverse ER proteostasis environments. Cell Rep. 3:1279–92
    [Google Scholar]
  48. 48.
    Wong MY, Chen K, Antonopoulos A, Kasper BT, Dewal MB et al. 2018. XBP1s activation can globally remodel N-glycan structure distribution patterns. PNAS 115:E10089–98
    [Google Scholar]
  49. 49.
    Sullivan CS, Pipas JM. 2001. The virus–chaperone connection. Virology 287:1–8
    [Google Scholar]
  50. 50.
    Wang Y, Jin F, Wang R, Li F, Wu Y et al. 2017. HSP90: a promising broad-spectrum antiviral drug target. Arch. Virol. 162:3269–82
    [Google Scholar]
  51. 51.
    Gao G, Luo H. 2006. The ubiquitin–proteasome pathway in viral infections. Can. J. Physiol. Pharmacol. 84:5–14
    [Google Scholar]
  52. 52.
    Morito D, Nagata K. 2015. Pathogenic hijacking of ER-associated degradation: Is ERAD flexible?. Mol. Cell 59:335–44
    [Google Scholar]
  53. 53.
    Aviner R, Li KH, Frydman J, Andino R. 2021. Cotranslational prolyl hydroxylation is essential for flavivirus biogenesis. Nature 596:558–64
    [Google Scholar]
  54. 54.
    Choi Y, Bowman JW, Jung JU. 2018. Autophagy during viral infection—a double-edged sword. Nat. Rev. Microbiol. 16:341–54
    [Google Scholar]
  55. 55.
    Tabata K, Arakawa M, Ishida K, Kobayashi M, Nara A et al. 2021. Endoplasmic reticulum-associated degradation controls virus protein homeostasis, which is required for flavivirus propagation. J. Virol. 95:e0223420
    [Google Scholar]
  56. 56.
    Fujita M, Akari H, Sakurai A, Yoshida A, Chiba T et al. 2004. Expression of HIV-1 accessory protein Vif is controlled uniquely to be low and optimal by proteasome degradation. Microbes Infect. 6:791–98
    [Google Scholar]
  57. 57.
    de Groot RJ, Rümenapf T, Kuhn RJ, Strauss EG, Strauss JH. 1991. Sindbis virus RNA polymerase is degraded by the N-end rule pathway. PNAS 88:8967–71
    [Google Scholar]
  58. 58.
    Suzuki R, Moriishi K, Fukuda K, Shirakura M, Ishii K et al. 2009. Proteasomal turnover of hepatitis C virus core protein is regulated by two distinct mechanisms: a ubiquitin-dependent mechanism and a ubiquitin-independent but PA28γ-dependent mechanism. J. Virol. 83:2389–92
    [Google Scholar]
  59. 59.
    Gao L, Tu H, Shi ST, Lee KJ, Asanaka M et al. 2003. Interaction with a ubiquitin-like protein enhances the ubiquitination and degradation of hepatitis C virus RNA-dependent RNA polymerase. J. Virol. 77:4149–59
    [Google Scholar]
  60. 60.
    Griffin SDC, Beales LP, Clarke DS, Worsfold O, Evans SD et al. 2003. The p7 protein of hepatitis C virus forms an ion channel that is blocked by the antiviral drug, Amantadine. FEBS Lett. 535:34–38
    [Google Scholar]
  61. 61.
    Franck N, Le Seyec J, Guguen-Guillouzo C, Erdtmann L 2005. Hepatitis C virus NS2 protein is phosphorylated by the protein kinase CK2 and targeted for degradation to the proteasome. J. Virol. 79:2700–8
    [Google Scholar]
  62. 62.
    Gladding RL, Haas AL, Gronros DL, Lawson TG. 1997. Evaluation of the susceptibility of the 3C proteases of hepatitis A virus and poliovirus to degradation by the ubiquitin-mediated proteolytic system. Biochem. Biophys. Res. Commun. 238:119–25
    [Google Scholar]
  63. 63.
    Tamura T, Yamashita T, Segawa H, Taira H. 2002. N-linked oligosaccharide chains of Sendai virus fusion protein determine the interaction with endoplasmic reticulum molecular chaperones. FEBS Lett. 513:153–58
    [Google Scholar]
  64. 64.
    Phillips AM, Ponomarenko AI, Chen K, Ashenberg O, Miao J et al. 2018. Destabilized adaptive influenza variants critical for innate immune system escape are potentiated by host chaperones. PLOS Biol. 16:e3000008
    [Google Scholar]
  65. 65.
    Phillips AM, Gonzalez LO, Nekongo EE, Ponomarenko AI, McHugh SM et al. 2017. Host proteostasis modulates influenza evolution. eLife 6:e28652
    [Google Scholar]
  66. 66.
    Phillips AM, Doud MB, Gonzalez LO, Butty VL, Lin YS et al. 2018. Enhanced ER proteostasis and temperature differentially impact the mutational tolerance of influenza hemagglutinin. eLife 7:e38795
    [Google Scholar]
  67. 67.
    Yoon J, Nekongo EE, Patrick JE, Hui T, Phillips AM et al. 2022. The endoplasmic reticulum proteostasis network profoundly shapes the protein sequence space accessible to HIV envelope. PLOS Biol. 20:e3001569
    [Google Scholar]
  68. 68.
    Geller R, Pechmann S, Acevedo A, Andino R, Frydman J. 2018. Hsp90 shapes protein and RNA evolution to balance trade-offs between protein stability and aggregation. Nat. Commun. 9:1781
    [Google Scholar]
  69. 69.
    Queitsch C, Sangster TA, Lindquist S. 2002. Hsp90 as a capacitor of phenotypic variation. Nature 417:618–24
    [Google Scholar]
  70. 70.
    Cowen LE, Lindquist S. 2005. Hsp90 potentiates the rapid evolution of new traits: drug resistance in diverse fungi. Science 309:2185–89
    [Google Scholar]
  71. 71.
    Sangster TA, Salathia N, Lee HN, Watanabe E, Schellenberg K et al. 2008. Hsp90-buffered genetic variation is common in Arabidopsis thaliana. PNAS 105:2969–74
    [Google Scholar]
  72. 72.
    Rohner N, Jarosz DF, Kowalko JE, Yoshizawa M, Jeffery WR et al. 2013. Cryptic variation in morphological evolution: Hsp90 as a capacitor for loss of eyes in cavefish. Science 342:1372–75
    [Google Scholar]
  73. 73.
    Sollars V, Lu X, Xiao L, Wang X, Garfinkel MD, Ruden DM 2003. Evidence for an epigenetic mechanism by which Hsp90 acts as a capacitor for morphological evolution. Nat. Genet. 33:70–74
    [Google Scholar]
  74. 74.
    Rutherford SL, Lindquist S. 1998. Hsp90 as a capacitor for morphological evolution. Nature 396:336–42
    [Google Scholar]
  75. 75.
    Aguilar-Rodríguez J, Sabater-Muñoz B, Montagud-Martínez R, Berlanga V, Alvarez-Ponce D et al. 2016. The molecular chaperone DnaK is a source of mutational robustness. Genome Biol. Evol. 8:2979–91
    [Google Scholar]
  76. 76.
    Fares MA, Ruiz-González MX, Moya A, Elena SF, Barrio E. 2002. Endosymbiotic bacteria: GroEL buffers against deleterious mutations. Nature 417:398
    [Google Scholar]
  77. 77.
    Tokuriki N, Tawfik DS. 2009. Chaperonin overexpression promotes genetic variation and enzyme evolution. Nature 459:668–73
    [Google Scholar]
  78. 78.
    Williams TA, Fares MA. 2010. The effect of chaperonin buffering on protein evolution. Genome Biol. Evol. 2:609–19
    [Google Scholar]
  79. 79.
    Sabater-Muñoz B, Prats-Escriche M, Montagud-Martínez R, López-Cerdán A, Toft C et al. 2015. Fitness trade-offs determine the role of the molecular chaperonin GroEL in buffering mutations. Mol. Biol. Evol. 32:2681–93
    [Google Scholar]
  80. 80.
    Iyengar BR, Wagner A. 2022. GroEL/S overexpression helps to purge deleterious mutations and reduce genetic diversity during adaptive protein evolution. Mol. Biol. Evol. 39:msac047
    [Google Scholar]
  81. 81.
    Alvarez-Ponce D, Aguilar-Rodríguez J, Fares MA. 2019. Molecular chaperones accelerate the evolution of their protein clients in yeast. Genome Biol. Evol. 11:2360–75
    [Google Scholar]
  82. 82.
    Çetinbaş M, Shakhnovich EI. 2013. Catalysis of protein folding by chaperones accelerates evolutionary dynamics in adapting cell populations. PLOS Comput. Biol. 9:e1003269
    [Google Scholar]
  83. 83.
    Bogumil D, Dagan T. 2010. Chaperonin-dependent accelerated substitution rates in prokaryontes. Genome Biol. Evol. 2:602–8
    [Google Scholar]
  84. 84.
    Pechmann S, Frydman J. 2014. Interplay between chaperones and protein disorder promotes the evolution of protein networks. PLOS Comput. Biol. 10:e1003674
    [Google Scholar]
  85. 85.
    Bershtein S, Mu W, Serohijos AWR, Zhou J, Shakhnovich EI. 2013. Protein quality control acts on folding intermediates to shape the effects of mutations on organismal fitness. Mol. Cell 49:133–44
    [Google Scholar]
  86. 86.
    Matange N. 2020. Highly contingent phenotypes of Lon protease deficiency in Escherichia coli upon antibiotic challenge. J. Bacteriol. 202:e00561–19
    [Google Scholar]
  87. 87.
    Thompson S, Zhang Y, Ingle C, Reynolds KA, Kortemme T. 2020. Altered expression of a quality control protease in E. coli reshapes the in vivo mutational landscape of a model enzyme. eLife 9:e53476
    [Google Scholar]
  88. 88.
    Guerrero RF, Scarpino SV, Rodrigues JV, Hartl DL, Ogbunugafor CB. 2019. Proteostasis environment shapes higher-order epistasis operating on antibiotic resistance. Genetics 212:565–75
    [Google Scholar]
  89. 89.
    Joshi P, Stoddart CA. 2011. Impaired infectivity of ritonavir-resistant HIV is rescued by heat shock protein 90AB1. J. Biol. Chem. 286:24581–92
    [Google Scholar]
  90. 90.
    Ceuppens JL, Baroja ML, Lorre K, Van Damme J, Billiau A. 1988. Human T cell activation with phytohemagglutinin. The function of IL-6 as an accessory signal. J. Immunol. 141:3868–74
    [Google Scholar]
  91. 91.
    Joshi P, Sloan B, Torbett BE, Stoddart CA. 2013. Heat shock protein 90AB1 and hyperthermia rescue infectivity of HIV with defective cores. Virology 436:162–72
    [Google Scholar]
  92. 92.
    Ryno LM, Genereux JC, Naito T, Morimoto RI, Powers ET et al. 2014. Characterizing the altered cellular proteome induced by the stress-independent activation of heat shock factor 1. ACS Chem. Biol. 9:1273–83
    [Google Scholar]
  93. 93.
    Shoulders MD, Ryno LM, Cooley CB, Kelly JW, Wiseman RL. 2013. Broadly applicable methodology for the rapid and dosable small molecule-mediated regulation of transcription factors in human cells. J. Am. Chem. Soc. 135:8129–32
    [Google Scholar]
  94. 94.
    Shimamura T, Perera SA, Foley KP, Sang J, Rodig SJ et al. 2012. Ganetespib (STA-9090), a nongeldanamycin HSP90 inhibitor, has potent antitumor activity in in vitro and in vivo models of non-small cell lung cancer. Clin. Cancer Res. 18:4973–85
    [Google Scholar]
  95. 95.
    Geller R, Vignuzzi M, Andino R, Frydman J. 2007. Evolutionary constraints on chaperone-mediated folding provide an antiviral approach refractory to development of drug resistance. Genes Dev. 21:195–205
    [Google Scholar]
  96. 96.
    Liu Y. 2020. A code within the genetic code: Codon usage regulates co-translational protein folding. Cell Commun. Signal. 18:145
    [Google Scholar]
  97. 97.
    Woo HJ, Reifman J. 2014. Quantitative modeling of virus evolutionary dynamics and adaptation in serial passages using empirically inferred fitness landscapes. J. Virol. 88:1039–50
    [Google Scholar]
  98. 98.
    Fowler DM, Fields S. 2014. Deep mutational scanning: a new style of protein science. Nat. Methods 11:801–7
    [Google Scholar]
  99. 99.
    Thyagarajan B, Bloom JD. 2014. The inherent mutational tolerance and antigenic evolvability of influenza hemagglutinin. eLife 3:03300
    [Google Scholar]
  100. 100.
    Doud MB, Bloom JD. 2016. Accurate measurement of the effects of all amino-acid mutations on influenza hemagglutinin. Viruses 8:155
    [Google Scholar]
  101. 101.
    Starr TN, Greaney AJ, Addetia A, Hannon WW, Choudhary MC et al. 2021. Prospective mapping of viral mutations that escape antibodies used to treat COVID-19. Science 371:850–54
    [Google Scholar]
  102. 102.
    Setoh YX, Amarilla AA, Peng NYG, Griffiths RE, Carrera J et al. 2019. Determinants of Zika virus host tropism uncovered by deep mutational scanning. Nat. Microbiol. 4:876–87
    [Google Scholar]
  103. 103.
    Ashenberg O, Padmakumar J, Doud MB, Bloom JD. 2017. Deep mutational scanning identifies sites in influenza nucleoprotein that affect viral inhibition by MxA. PLOS Pathog. 13:e1006288
    [Google Scholar]
  104. 104.
    Mänz B, Dornfeld D, Götz V, Zell R, Zimmermann P et al. 2013. Pandemic influenza A viruses escape from restriction by human MxA through adaptive mutations in the nucleoprotein. PLOS Pathog. 9:e1003279
    [Google Scholar]
  105. 105.
    Moore CL, Dewal MB, Nekongo EE, Santiago S, Lu NB et al. 2016. Transportable, chemical genetic methodology for the small molecule-mediated inhibition of heat shock factor 1. ACS Chem. Biol. 11:200–10
    [Google Scholar]
  106. 106.
    Dewal MB, DiChiara AS, Antonopoulos A, Taylor RJ, Harmon CJ et al. 2015. XBP1s links the unfolded protein response to the molecular architecture of mature N-glycans. Chem. Biol. 22:1301–12
    [Google Scholar]
  107. 107.
    Sala AJ, Bott LC, Morimoto RI. 2017. Shaping proteostasis at the cellular, tissue, and organismal level. J. Cell Biol. 216:1231–41
    [Google Scholar]
  108. 108.
    Guisbert E, Czyz DM, Richter K, McMullen PD, Morimoto RI. 2013. Identification of a tissue-selective heat shock response regulatory network. PLOS Genet. 9:e1003466
    [Google Scholar]
  109. 109.
    Wang G, Cao P, Fan Y, Tan K. 2020. Emerging roles of HSF1 in cancer: cellular and molecular episodes. Biochim. Biophys. Acta Rev. Cancer 1874:188390
    [Google Scholar]
  110. 110.
    Gershenson A, Gierasch LM. 2011. Protein folding in the cell: challenges and progress. Curr. Opin. Struct. Biol. 21:32–41
    [Google Scholar]
  111. 111.
    Lauring AS, Acevedo A, Cooper SB, Andino R. 2012. Codon usage determines the mutational robustness, evolutionary capacity, and virulence of an RNA virus. Cell Host Microbe 12:623–32
    [Google Scholar]
  112. 112.
    Bitran A, Jacobs WM, Zhai X, Shakhnovich E. 2020. Cotranslational folding allows misfolding-prone proteins to circumvent deep kinetic traps. PNAS 117:1485–95
    [Google Scholar]
  113. 113.
    Geller R, Andino R, Frydman J. 2013. Hsp90 inhibitors exhibit resistance-free antiviral activity against respiratory syncytial virus. PLOS ONE 8:e56762
    [Google Scholar]
  114. 114.
    Taguwa S, Maringer K, Li X, Bernal-Rubio D, Rauch JN et al. 2015. Defining Hsp70 subnetworks in dengue virus replication reveals key vulnerability in flavivirus infection. Cell 163:1108–23
    [Google Scholar]
  115. 115.
    Taguwa S, Yeh M-T, Rainbolt TK, Nayak A, Shao H et al. 2019. Zika virus dependence on host Hsp70 provides a protective strategy against infection and disease. Cell Rep. 26:906–20.e3
    [Google Scholar]
  116. 116.
    Warfield KL, Schaaf KR, DeWald LE, Spurgers KB, Wang W et al. 2019. Lack of selective resistance of influenza A virus in presence of host-targeted antiviral, UV-4B. Sci. Rep. 9:7484
    [Google Scholar]
  117. 117.
    Coelmont L, Hanoulle X, Chatterji U, Berger C, Snoeck J et al. 2010. DEB025 (Alisporivir) inhibits hepatitis C virus replication by preventing a cyclophilin A induced cis-trans isomerisation in domain II of NS5A. PLOS ONE 5:e13687
    [Google Scholar]
  118. 118.
    Nekongo EE, Ponomarenko AI, Dewal MB, Butty VL, Browne EP, Shoulders MD. 2020. HSF1 activation can restrict HIV replication. ACS Infect. Dis. 6:1659–66
    [Google Scholar]
  119. 119.
    Hosfelt J, Richards A, Zheng M, Adura C, Nelson B et al. 2022. An allosteric inhibitor of bacterial Hsp70 chaperone potentiates antibiotics and mitigates resistance. Cell Chem. Biol. 29:854–69.e9
    [Google Scholar]
  120. 120.
    Piana S, Carloni P, Rothlisberger U. 2002. Drug resistance in HIV-1 protease: flexibility-assisted mechanism of compensatory mutations. Protein Sci. 11:2393–402
    [Google Scholar]
  121. 121.
    Bastys T, Gapsys V, Walter H, Heger E, Doncheva NT et al. 2020. Non-active site mutants of HIV-1 protease influence resistance and sensitisation towards protease inhibitors. Retrovirology 17:13
    [Google Scholar]
  122. 122.
    Baranovich T, Wong SS, Armstrong J, Marjuki H, Webby RJ et al. 2013. T-705 (favipiravir) induces lethal mutagenesis in influenza A H1N1 viruses in vitro. J. Virol. 87:3741–51
    [Google Scholar]
  123. 123.
    Arias A, Thorne L, Goodfellow I. 2014. Favipiravir elicits antiviral mutagenesis during virus replication in vivo. eLife 3:e03679
    [Google Scholar]
  124. 124.
    Harris KS, Brabant W, Styrchak S, Gall A, Daifuku R. 2005. KP-1212/1461, a nucleoside designed for the treatment of HIV by viral mutagenesis. Antiviral Res. 67:1–9
    [Google Scholar]
  125. 125.
    Dapp MJ, Clouser CL, Patterson S, Mansky LM. 2009. 5-Azacytidine can induce lethal mutagenesis in human immunodeficiency virus type 1. J. Virol. 83:11950–58
    [Google Scholar]
  126. 126.
    Dapp MJ, Heineman RH, Mansky LM. 2013. Interrelationship between HIV-1 fitness and mutation rate. J. Mol. Biol. 425:41–53
    [Google Scholar]
  127. 127.
    Crotty S, Cameron CE, Andino R. 2001. RNA virus error catastrophe: direct molecular test by using ribavirin. PNAS 98:6895–900
    [Google Scholar]
  128. 128.
    Biebricher CK, Eigen M. 2005. The error threshold. Virus Res. 107:117–27
    [Google Scholar]
  129. 129.
    Eigen M. 2002. Error catastrophe and antiviral strategy. PNAS 99:13374–76
    [Google Scholar]
  130. 130.
    Cuevas JM, González-Candelas F, Moya A, Sanjuán R. 2009. Effect of ribavirin on the mutation rate and spectrum of hepatitis C virus in vivo. J. Virol. 83:5760–64
    [Google Scholar]
  131. 131.
    Longdon B, Brockhurst MA, Russell CA, Welch JJ, Jiggins FM. 2014. The evolution and genetics of virus host shifts. PLOS Pathog. 10:e1004395
    [Google Scholar]
  132. 132.
    Nielsen KL, McLennan N, Masters M, Cowan NJ. 1999. A single-ring mitochondrial chaperonin (Hsp60-Hsp10) can substitute for GroEL-GroES in vivo. J. Bacteriol. 181:5871–75
    [Google Scholar]
  133. 133.
    Wang JD, Herman C, Tipton KA, Gross CA, Weissman JS. 2002. Directed evolution of substrate-optimized GroEL/S chaperonins. Cell 111:1027–39
    [Google Scholar]
  134. 134.
    Irving AT, Ahn M, Goh G, Anderson DE, Wang LF. 2021. Lessons from the host defences of bats, a unique viral reservoir. Nature 589:363–70
    [Google Scholar]
  135. 135.
    Chionh YT, Cui J, Koh J, Mendenhall IH, Ng JHJ et al. 2019. High basal heat-shock protein expression in bats confers resistance to cellular heat/oxidative stress. Cell Stress Chaperones 24:835–49
    [Google Scholar]
  136. 136.
    Rajkowitsch L, Chen D, Stampfl S, Semrad K, Waldsich C et al. 2007. RNA chaperones, RNA annealers and RNA helicases. RNA Biol. 4:118–30
    [Google Scholar]
  137. 137.
    Ganser LR, Kelly ML, Herschlag D, Al-Hashimi HM. 2019. The roles of structural dynamics in the cellular functions of RNAs. Nat. Rev. Mol. Cell Biol. 20:474–89
    [Google Scholar]
  138. 138.
    Jaafar ZA, Kieft JS. 2019. Viral RNA structure-based strategies to manipulate translation. Nat. Rev. Microbiol. 17:110–23
    [Google Scholar]
  139. 139.
    Bonilla SL, Sherlock ME, MacFadden A, Kieft JS. 2021. A viral RNA hijacks host machinery using dynamic conformational changes of a tRNA-like structure. Science 374:955–60
    [Google Scholar]
  140. 140.
    Zúñiga S, Sola I, Cruz JL, Enjuanes L. 2009. Role of RNA chaperones in virus replication. Virus Res. 139:253–66
    [Google Scholar]
  141. 141.
    Wasik BR, Turner PE. 2013. On the biological success of viruses. Annu. Rev. Microbiol. 67:519–41
    [Google Scholar]
  142. 142.
    Luo H. 2016. Interplay between the virus and the ubiquitin–proteasome system: molecular mechanism of viral pathogenesis. Curr. Opin. Virol. 17:1–10
    [Google Scholar]
  143. 143.
    Isaacson MK, Ploegh HL. 2009. Ubiquitination, ubiquitin-like modifiers, and deubiquitination in viral infection. Cell Host Microbe 5:559–70
    [Google Scholar]
  144. 144.
    Chiramel AI, Brady NR, Bartenschlager R. 2013. Divergent roles of autophagy in virus infection. Cells 2:83–104
    [Google Scholar]
  145. 145.
    Esvelt KM, Carlson JC, Liu DR. 2011. A system for the continuous directed evolution of biomolecules. Nature 472:499–503
    [Google Scholar]
  146. 146.
    Berman CM, Papa LJ III, Hendel SJ, Moore CL, Suen PH et al. 2018. An adaptable platform for directed evolution in human cells. J. Am. Chem. Soc. 140:18093–103
    [Google Scholar]
  147. 147.
    Ravikumar A, Arzumanyan GA, Obadi MKA, Javanpour AA, Liu CC. 2018. Scalable, continuous evolution of genes at mutation rates above genomic error thresholds. Cell 175:1946–57.e13
    [Google Scholar]
  148. 148.
    Moore CL, Papa LJ III, Shoulders MD 2018. A processive protein chimera introduces mutations across defined DNA regions in vivo. J. Am. Chem. Soc. 140:11560–64
    [Google Scholar]
  149. 149.
    Molina RS, Rix G, Mengiste AA, Álvarez B, Seo D et al. 2022. In vivo hypermutation and continuous evolution. Nat. Rev. Methods Primers 2:36
    [Google Scholar]
  150. 150.
    Hendel SJ, Shoulders MD. 2021. Directed evolution in mammalian cells. Nat. Methods 18:346–57
    [Google Scholar]
/content/journals/10.1146/annurev-virology-100220-112120
Loading
/content/journals/10.1146/annurev-virology-100220-112120
Loading

Data & Media loading...

Supplemental Material

Supplementary Data

  • 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