1932

Abstract

Human coronavirus (HCoV) infection causes respiratory diseases with mild to severe outcomes. In the last 15 years, we have witnessed the emergence of two zoonotic, highly pathogenic HCoVs: severe acute respiratory syndrome coronavirus (SARS-CoV) and Middle East respiratory syndrome coronavirus (MERS-CoV). Replication of HCoV is regulated by a diversity of host factors and induces drastic alterations in cellular structure and physiology. Activation of critical signaling pathways during HCoV infection modulates the induction of antiviral immune response and contributes to the pathogenesis of HCoV. Recent studies have begun to reveal some fundamental aspects of the intricate HCoV-host interaction in mechanistic detail. In this review, we summarize the current knowledge of host factors co-opted and signaling pathways activated during HCoV infection, with an emphasis on HCoV-infection-induced stress response, autophagy, apoptosis, and innate immunity. The cross talk among these pathways, as well as the modulatory strategies utilized by HCoV, is also discussed.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-micro-020518-115759
2019-09-08
2024-04-16
Loading full text...

Full text loading...

/deliver/fulltext/micro/73/1/annurev-micro-020518-115759.html?itemId=/content/journals/10.1146/annurev-micro-020518-115759&mimeType=html&fmt=ahah

Literature Cited

  1. 1. 
    Angelini MM, Akhlaghpour M, Neuman BW, Buchmeier MJ 2013. Severe acute respiratory syndrome coronavirus nonstructural proteins 3, 4, and 6 induce double-membrane vesicles. mBio 4:4e00524–13
    [Google Scholar]
  2. 2. 
    Bailey CC, Zhong G, Huang I-C, Farzan M 2014. IFITM-family proteins: the cell's first line of antiviral defense. Annu. Rev. Virol. 1:261–83
    [Google Scholar]
  3. 3. 
    Bechill J, Chen Z, Brewer JW, Baker SC 2008. Coronavirus infection modulates the unfolded protein response and mediates sustained translational repression. J. Virol. 82:94492–501
    [Google Scholar]
  4. 4. 
    Belouzard S, Chu VC, Whittaker GR 2009. Activation of the SARS coronavirus spike protein via sequential proteolytic cleavage at two distinct sites. PNAS 106:145871–76
    [Google Scholar]
  5. 5. 
    Bertram S, Dijkman R, Habjan M, Heurich A, Gierer S et al. 2013. TMPRSS2 activates the human coronavirus 229E for cathepsin-independent host cell entry and is expressed in viral target cells in the respiratory epithelium. J. Virol. 87:116150–60
    [Google Scholar]
  6. 6. 
    Bertram S, Glowacka I, Müller MA, Lavender H, Gnirss K et al. 2011. Cleavage and activation of the severe acute respiratory syndrome coronavirus spike protein by human airway trypsin-like protease. J. Virol. 85:2413363–72
    [Google Scholar]
  7. 7. 
    Bosch BJ, Bartelink W, Rottier PJM 2008. Cathepsin L functionally cleaves the severe acute respiratory syndrome coronavirus class I fusion protein upstream of rather than adjacent to the fusion peptide. J. Virol. 82:178887–90
    [Google Scholar]
  8. 8. 
    Bulavin DV, Saito S, Hollander MC, Sakaguchi K, Anderson CW et al. 1999. Phosphorylation of human p53 by p38 kinase coordinates N-terminal phosphorylation and apoptosis in response to UV radiation. EMBO J 18:236845–54
    [Google Scholar]
  9. 9. 
    Castaño-Rodriguez C, Honrubia JM, Gutiérrez-Álvarez J, DeDiego ML, Nieto-Torres JL et al. 2018. Role of severe acute respiratory syndrome coronavirus viroporins E, 3a, and 8a in replication and pathogenesis. mBio 9:3e02325–17
    [Google Scholar]
  10. 10. 
    Chafekar A, Fielding BC. 2018. MERS-CoV: understanding the latest human coronavirus threat. Viruses 10:2E93
    [Google Scholar]
  11. 11. 
    Chan CP, Siu KL, Chin KT, Yuen KY, Zheng B, Jin DY 2006. Modulation of the unfolded protein response by the severe acute respiratory syndrome coronavirus spike protein. J. Virol. 80:189279–87
    [Google Scholar]
  12. 12. 
    Chang Y-J, Liu CY-Y, Chiang B-L, Chao Y-C, Chen C-C 2004. Induction of IL-8 release in lung cells via activator protein-1 by recombinant baculovirus displaying severe acute respiratory syndrome-coronavirus spike proteins: identification of two functional regions. J. Immunol. 173:127602–14
    [Google Scholar]
  13. 13. 
    Chen C-C, Krüger J, Sramala I, Hsu H-J, Henklein P et al. 2011. ORF8a of SARS-CoV forms an ion channel: experiments and molecular dynamics simulations. Biochim. Biophys. Acta 1808 2572–79
    [Google Scholar]
  14. 14. 
    Chen I-Y, Chang SC, Wu H-Y, Yu T-C, Wei W-C et al. 2010. Upregulation of the chemokine (C-C motif) ligand 2 via a severe acute respiratory syndrome coronavirus spike-ACE2 signaling pathway. J. Virol. 84:157703–12
    [Google Scholar]
  15. 15. 
    Chu H, Zhou J, Wong BH-Y, Li C, Chan JF-W et al. 2016. Middle East respiratory syndrome coronavirus efficiently infects human primary T lymphocytes and activates the extrinsic and intrinsic apoptosis pathways. J. Infect. Dis. 213:6904–14
    [Google Scholar]
  16. 16. 
    Clavarino G, Cláudio N, Dalet A, Terawaki S, Couderc T et al. 2012. Protein phosphatase 1 subunit Ppp1r15a/GADD34 regulates cytokine production in polyinosinic:polycytidylic acid-stimulated dendritic cells. PNAS 109:83006–11
    [Google Scholar]
  17. 17. 
    Corman VM, Ithete NL, Richards LR, Schoeman MC, Preiser W et al. 2014. Rooting the phylogenetic tree of Middle East respiratory syndrome coronavirus by characterization of a conspecific virus from an African bat. J. Virol. 88:1911297–303
    [Google Scholar]
  18. 18. 
    Cottam EM, Maier HJ, Manifava M, Vaux LC, Chandra-Schoenfelder P et al. 2011. Coronavirus nsp6 proteins generate autophagosomes from the endoplasmic reticulum via an omegasome intermediate. Autophagy 7:111335–47
    [Google Scholar]
  19. 19. 
    Cottam EM, Whelband MC, Wileman T 2014. Coronavirus NSP6 restricts autophagosome expansion. Autophagy 10:81426–41
    [Google Scholar]
  20. 20. 
    Credle JJ, Finer-Moore JS, Papa FR, Stroud RM, Walter P 2005. On the mechanism of sensing unfolded protein in the endoplasmic reticulum. PNAS 102:5218773–84
    [Google Scholar]
  21. 21. 
    Cruz JL, Sola I, Becares M, Alberca B, Plana J et al. 2011. Coronavirus gene 7 counteracts host defenses and modulates virus virulence. PLOS Pathog 7:6e1002090
    [Google Scholar]
  22. 22. 
    Cui J, Li F, Shi Z-L 2019. Origin and evolution of pathogenic coronaviruses. Nat. Rev. Microbiol. 17:3181–92
    [Google Scholar]
  23. 23. 
    Dalby KN, Morrice N, Caudwell FB, Avruch J, Cohen P 1998. Identification of regulatory phosphorylation sites in mitogen-activated protein kinase (MAPK)-activated protein kinase-1a/p90rsk that are inducible by MAPK. J. Biol. Chem. 273:31496–505
    [Google Scholar]
  24. 24. 
    de Groot RJ. 2006. Structure, function and evolution of the hemagglutinin-esterase proteins of corona- and toroviruses. Glycoconj. J. 23:1–259–72
    [Google Scholar]
  25. 25. 
    de Groot RJ, Baker SC, Baric RS, Brown CS, Drosten C et al. 2013. Middle East respiratory syndrome coronavirus (MERS-CoV): announcement of the Coronavirus Study Group. J. Virol. 87:147790–92
    [Google Scholar]
  26. 26. 
    de Haan CAM, de Wit M, Kuo L, Montalto-Morrison C, Haagmans BL et al. 2003. The glycosylation status of the murine hepatitis coronavirus M protein affects the interferogenic capacity of the virus in vitro and its ability to replicate in the liver but not the brain. Virology 312:2395–406
    [Google Scholar]
  27. 27. 
    DeDiego ML, Nieto-Torres JL, Jiménez-Guardeño JM, Regla-Nava JA, Álvarez E et al. 2011. Severe acute respiratory syndrome coronavirus envelope protein regulates cell stress response and apoptosis. PLOS Pathog 7:10e1002315
    [Google Scholar]
  28. 28. 
    Deng X, Hackbart M, Mettelman RC, O'Brien A, Mielech AM et al. 2017. Coronavirus nonstructural protein 15 mediates evasion of dsRNA sensors and limits apoptosis in macrophages. PNAS 114:21E4251–60This paper demonstrates how HCoV nsp15 mediates evasion of dsRNA sensing and suppresses apoptosis.
    [Google Scholar]
  29. 29. 
    Devaraj SG, Wang N, Chen Z, Chen Z, Tseng M et al. 2007. Regulation of IRF-3-dependent innate immunity by the papain-like protease domain of the severe acute respiratory syndrome coronavirus. J. Biol. Chem. 282:4432208–21
    [Google Scholar]
  30. 30. 
    Favreau DJ, Meessen-Pinard M, Desforges M, Talbot PJ 2012. Human coronavirus-induced neuronal programmed cell death is cyclophilin D dependent and potentially caspase dispensable. J. Virol. 86:181–93
    [Google Scholar]
  31. 31. 
    Fouchier RAM, Hartwig NG, Bestebroer TM, Niemeyer B, de Jong JC et al. 2004. A previously undescribed coronavirus associated with respiratory disease in humans. PNAS 101:166212–16
    [Google Scholar]
  32. 32. 
    Frödin M, Gammeltoft S. 1999. Role and regulation of 90 kDa ribosomal S6 kinase (RSK) in signal transduction. Mol. Cell Endocrinol. 151:1–265–77
    [Google Scholar]
  33. 33. 
    Fukushi M, Yoshinaka Y, Matsuoka Y, Hatakeyama S, Ishizaka Y et al. 2012. Monitoring S protein maturation in the endoplasmic reticulum by calnexin is important for the infectivity of severe acute respiratory syndrome coronavirus. J. Virol. 86:2111745–53
    [Google Scholar]
  34. 34. 
    Fung T, Liao Y, Liu DX 2016. Regulation of stress responses and translational control by coronavirus. Viruses 8:7184
    [Google Scholar]
  35. 35. 
    Fung T, Torres J, Liu D 2015. The emerging roles of viroporins in ER stress response and autophagy induction during virus infection. Viruses 7:62834–57
    [Google Scholar]
  36. 36. 
    Fung TS, Huang M, Liu DX 2014. Coronavirus-induced ER stress response and its involvement in regulation of coronavirus-host interactions. Virus Res 194:110–23
    [Google Scholar]
  37. 37. 
    Fung TS, Liao Y, Liu DX 2014. The ER stress sensor IRE1α protects cells from apoptosis induced by coronavirus infectious bronchitis virus. J. Virol. 88:2112752–64This paper demonstrates how the IRE1 branch of UPR promotes cell survival during CoV infection.
    [Google Scholar]
  38. 38. 
    Fung TS, Liu DX. 2014. Coronavirus infection, ER stress, apoptosis and innate immunity. Front. Microbiol. 5:296
    [Google Scholar]
  39. 39. 
    Fung TS, Liu DX. 2017. Activation of the c-Jun NH2-terminal kinase pathway by coronavirus infectious bronchitis virus promotes apoptosis independently of c-Jun. Cell Death Dis 8:123215This paper demonstrates how CoV infection activates JNK to modulate apoptosis induction.
    [Google Scholar]
  40. 40. 
    Fung TS, Liu DX. 2018. Post-translational modifications of coronavirus proteins: roles and function. Future Virol 13:6405–30
    [Google Scholar]
  41. 41. 
    Gorse GJ, O'Connor TZ, Hall SL, Vitale JN, Nichol KL 2009. Human coronavirus and acute respiratory illness in older adults with chronic obstructive pulmonary disease. J. Infect. Dis. 199:6847–57
    [Google Scholar]
  42. 42. 
    Graham RL, Donaldson EF, Baric RS 2013. A decade after SARS: strategies for controlling emerging coronaviruses. Nat. Rev. Microbiol. 11:12836–48
    [Google Scholar]
  43. 43. 
    Guo L, Yu H, Gu W, Luo X, Li R et al. 2016. Autophagy negatively regulates transmissible gastroenteritis virus replication. Sci. Rep. 6:23864
    [Google Scholar]
  44. 44. 
    Guo X, Zhang M, Zhang X, Tan X, Guo H et al. 2017. Porcine epidemic diarrhea virus induces autophagy to benefit its replication. Viruses 9:3E53
    [Google Scholar]
  45. 45. 
    Hamre D, Procknow JJ. 1966. A new virus isolated from the human respiratory tract. Exp. Biol. Med. 121:1190–93
    [Google Scholar]
  46. 46. 
    Han DP, Lohani M, Cho MW 2007. Specific asparagine-linked glycosylation sites are critical for DC-SIGN- and L-SIGN-mediated severe acute respiratory syndrome coronavirus entry. J. Virol. 81:2112029–39
    [Google Scholar]
  47. 47. 
    Hess J, Angel P, Schorpp-Kistner M 2004. AP-1 subunits: quarrel and harmony among siblings. J. Cell Sci. 117:255965–73
    [Google Scholar]
  48. 48. 
    Hogue BG, Machamer CE. 2008. Coronavirus structural proteins and virus assembly. Nidoviruses S Perlman, T Gallagher, EJ Snijder 179–200 Washington, DC: ASM
    [Google Scholar]
  49. 49. 
    Hollien J, Weissman JS. 2006. Decay of endoplasmic reticulum-localized mRNAs during the unfolded protein response. Science 313:5783104–7
    [Google Scholar]
  50. 50. 
    Hu B, Zeng L-P, Yang X-L, Ge X-Y, Zhang W et al. 2017. Discovery of a rich gene pool of bat SARS-related coronaviruses provides new insights into the origin of SARS coronavirus. PLOS Pathog 13:11e1006698
    [Google Scholar]
  51. 51. 
    Huang I-C, Bailey CC, Weyer JL, Radoshitzky SR, Becker MM et al. 2011. Distinct patterns of IFITM-mediated restriction of filoviruses, SARS coronavirus, and influenza A virus. PLOS Pathog 7:1e1001258
    [Google Scholar]
  52. 52. 
    Irigoyen N, Firth AE, Jones JD, Chung BY-W, Siddell SG, Brierley I 2016. High-resolution analysis of coronavirus gene expression by RNA sequencing and ribosome profiling. PLOS Pathog 12:2e1005473
    [Google Scholar]
  53. 53. 
    Jimenez-Guardeño JM, Nieto-Torres JL, DeDiego ML, Regla-Nava JA, Fernandez-Delgado R et al. 2014. The PDZ-binding motif of severe acute respiratory syndrome coronavirus envelope protein is a determinant of viral pathogenesis. PLOS Pathog 10:8e1004320This paper demonstrates how the PDZ-binding motif of the SARS-CoV E protein contributes to viral pathogenesis.
    [Google Scholar]
  54. 54. 
    Jimenez-Guardeño JM, Regla-Nava JA, Nieto-Torres JL, DeDiego ML, Castaño-Rodriguez C et al. 2015. Identification of the mechanisms causing reversion to virulence in an attenuated SARS-CoV for the design of a genetically stable vaccine. PLOS Pathog 11:10e1005215
    [Google Scholar]
  55. 55. 
    Keshet Y, Seger R. 2010. The MAP kinase signaling cascades: a system of hundreds of components regulates a diverse array of physiological functions. Methods Mol. Biol. 661:3–38
    [Google Scholar]
  56. 56. 
    Kindler E, Gil-Cruz C, Spanier J, Li Y, Wilhelm J et al. 2017. Early endonuclease-mediated evasion of RNA sensing ensures efficient coronavirus replication. PLOS Pathog 13:2e1006195
    [Google Scholar]
  57. 57. 
    Klumperman J, Locker JK, Meijer A, Horzinek MC, Geuze HJ, Rottier P 1994. Coronavirus M proteins accumulate in the Golgi complex beyond the site of virion budding. J. Virol. 68:106523–34
    [Google Scholar]
  58. 58. 
    Ko S, Gu MJ, Kim CG, Kye YC, Lim Y et al. 2017. Rapamycin-induced autophagy restricts porcine epidemic diarrhea virus infectivity in porcine intestinal epithelial cells. Antiviral Res 146:86–95
    [Google Scholar]
  59. 59. 
    Kono M, Tatsumi K, Imai AM, Saito K, Kuriyama T, Shirasawa H 2008. Inhibition of human coronavirus 229E infection in human epithelial lung cells (L132) by chloroquine: involvement of p38 MAPK and ERK. Antiviral Res 77:2150–52
    [Google Scholar]
  60. 60. 
    Kopecky-Bromberg SA, Martinez-Sobrido L, Palese P 2006. 7a protein of severe acute respiratory syndrome coronavirus inhibits cellular protein synthesis and activates p38 mitogen-activated protein kinase. J. Virol. 80:2785–93
    [Google Scholar]
  61. 61. 
    Krähling V, Stein DA, Spiegel M, Weber F, Mühlberger E 2009. Severe acute respiratory syndrome coronavirus triggers apoptosis via protein kinase R but is resistant to its antiviral activity. J. Virol. 83:52298–309
    [Google Scholar]
  62. 62. 
    Kwak H, Park MW, Jeong S 2011. Annexin A2 binds RNA and reduces the frameshifting efficiency of infectious bronchitis virus. PLOS ONE 6:8e24067
    [Google Scholar]
  63. 63. 
    Li S-W, Wang C-Y, Jou Y-J, Huang S-H, Hsiao L-H et al. 2016. SARS coronavirus papain-like protease inhibits the TLR7 signaling pathway through removing Lys63-linked polyubiquitination of TRAF3 and TRAF6. Int. J. Mol. Sci. 17:5E678
    [Google Scholar]
  64. 64. 
    Li W, Shi Z, Yu M, Ren W, Smith C et al. 2005. Bats are natural reservoirs of SARS-like coronaviruses. Science 310:5748676–79
    [Google Scholar]
  65. 65. 
    Li Y, Treffers EE, Napthine S, Tas A, Zhu L et al. 2014. Transactivation of programmed ribosomal frameshifting by a viral protein. PNAS 111:21E2172–81
    [Google Scholar]
  66. 66. 
    Liao Y, Fung TS, Huang M, Fang SG, Zhong Y, Liu DX 2013. Upregulation of CHOP/GADD153 during coronavirus infectious bronchitis virus infection modulates apoptosis by restricting activation of the extracellular signal-regulated kinase pathway. J. Virol. 87:148124–34
    [Google Scholar]
  67. 67. 
    Liao Y, Yuan Q, Torres J, Tam JP, Liu DX 2006. Biochemical and functional characterization of the membrane association and membrane permeabilizing activity of the severe acute respiratory syndrome coronavirus envelope protein. Virology 349:2264–75
    [Google Scholar]
  68. 68. 
    Lim KP, Liu DX. 2001. The missing link in coronavirus assembly retention of the avian coronavirus infectious bronchitis virus envelope protein in the pre-Golgi compartments and physical interaction between the envelope and membrane proteins. J. Biol. Chem. 276:2017515–23
    [Google Scholar]
  69. 69. 
    Lim Y, Ng Y, Tam J, Liu D 2016. Human coronaviruses: a review of virus-host interactions. Diseases 4:426
    [Google Scholar]
  70. 70. 
    Liu DX, Fung TS, Chong KK-L, Shukla A, Hilgenfeld R 2014. Accessory proteins of SARS-CoV and other coronaviruses. Antiviral Res 109:97–109
    [Google Scholar]
  71. 71. 
    Liu DX, Inglis SC. 1991. Association of the infectious bronchitis virus 3c protein with the virion envelope. Virology 185:2911–17
    [Google Scholar]
  72. 72. 
    Liu M, Yang Y, Gu C, Yue Y, Wu KK et al. 2007. Spike protein of SARS-CoV stimulates cyclooxygenase-2 expression via both calcium-dependent and calcium-independent protein kinase C pathways. FASEB J 21:71586–96
    [Google Scholar]
  73. 73. 
    Lu W, Zheng BJ, Xu K, Schwarz W, Du L et al. 2006. Severe acute respiratory syndrome-associated coronavirus 3a protein forms an ion channel and modulates virus release. PNAS 103:3312540–45
    [Google Scholar]
  74. 74. 
    Luo H, Chen Q, Chen J, Chen K, Shen X, Jiang H 2005. The nucleocapsid protein of SARS coronavirus has a high binding affinity to the human cellular heterogeneous nuclear ribonucleoprotein A1. FEBS Lett 579:122623–28
    [Google Scholar]
  75. 75. 
    Ma Y, Wang C, Xue M, Fu F, Zhang X et al. 2018. Coronavirus TGEV evades the type I interferon response through IRE1α-mediated manipulation of the miR-30a-5p/SOCS1/3 Axis. J. Virol. 92:e00728–18
    [Google Scholar]
  76. 76. 
    Madan V, García M, de J, Sanz MA, Carrasco L 2005. Viroporin activity of murine hepatitis virus E protein.. FEBS Lett 579:173607–12
    [Google Scholar]
  77. 77. 
    Maier HJ, Hawes PC, Cottam EM, Mantell J, Verkade P et al. 2013. Infectious bronchitis virus generates spherules from zippered endoplasmic reticulum membranes. mBio 4:5e00801–13
    [Google Scholar]
  78. 78. 
    Martinon F, Chen X, Lee A-H, Glimcher LH 2010. TLR activation of the transcription factor XBP1 regulates innate immune responses in macrophages. Nat. Immunol. 11:5411–18
    [Google Scholar]
  79. 79. 
    Masters PS. 2006. The molecular biology of coronaviruses. Adv. Virus Res. 66:193–292
    [Google Scholar]
  80. 80. 
    Matthews K, Schäfer A, Pham A, Frieman M 2014. The SARS coronavirus papain like protease can inhibit IRF3 at a post activation step that requires deubiquitination activity. Virol. J. 11:209
    [Google Scholar]
  81. 81. 
    Maurel M, Chevet E, Tavernier J, Gerlo S 2014. Getting RIDD of RNA: IRE1 in cell fate regulation. Trends Biochem. Sci. 39:5245–54
    [Google Scholar]
  82. 82. 
    Mesel-Lemoine M, Millet J, Vidalain P-O, Law H, Vabret A et al. 2012. A human coronavirus responsible for the common cold massively kills dendritic cells but not monocytes. J. Virol. 86:147577–87
    [Google Scholar]
  83. 83. 
    Mielech AM, Kilianski A, Baez-Santos YM, Mesecar AD, Baker SC 2014. MERS-CoV papain-like protease has deISGylating and deubiquitinating activities. Virology 450–451:64–70
    [Google Scholar]
  84. 84. 
    Millet JK, Whittaker GR. 2014. Host cell entry of Middle East respiratory syndrome coronavirus after two-step, furin-mediated activation of the spike protein. PNAS 111:4215214–19This study shows how host protease furin activates the spike protein of MERS-CoV.
    [Google Scholar]
  85. 85. 
    Mizutani T, Fukushi S, Murakami M, Hirano T, Saijo M et al. 2004. Tyrosine dephosphorylation of STAT3 in SARS coronavirus-infected Vero E6 cells. FEBS Lett 577:1–2187–92
    [Google Scholar]
  86. 86. 
    Mizutani T, Fukushi S, Saijo M, Kurane I, Morikawa S 2004. Phosphorylation of p38 MAPK and its downstream targets in SARS coronavirus-infected cells. Biochem. Biophys. Res. Commun. 319:41228–34
    [Google Scholar]
  87. 87. 
    Mizutani T, Fukushi S, Saijo M, Kurane I, Morikawa S 2005. JNK and PI3k/Akt signaling pathways are required for establishing persistent SARS-CoV infection in Vero E6 cells. Biochim. Biophys. Acta Mol. Basis Dis. 1741:14–10
    [Google Scholar]
  88. 88. 
    Mizutani T, Fukushi S, Saijo M, Kurane I, Morikawa S 2006. Regulation of p90RSK phosphorylation by SARS-CoV infection in Vero E6 cells. FEBS Lett 580:51417–24
    [Google Scholar]
  89. 89. 
    Monastyrska I, Ulasli M, Rottier PJM, Guan J-L, Reggiori F, de Haan CAM 2013. An autophagy-independent role for LC3 in equine arteritis virus replication. Autophagy 9:2164–74
    [Google Scholar]
  90. 90. 
    Nanda SK, Leibowitz JL. 2001. Mitochondrial aconitase binds to the 3′ untranslated region of the mouse hepatitis virus genome. J. Virol. 75:73352–62
    [Google Scholar]
  91. 91. 
    Nieto-Torres JL, DeDiego ML, Verdiá-Báguena C, Jimenez-Guardeño JM, Regla-Nava JA et al. 2014. Severe acute respiratory syndrome coronavirus envelope protein ion channel activity promotes virus fitness and pathogenesis. PLOS Pathog 10:5e1004077This paper demonstrates how ion channel activity of SARS-CoV E contributes to viral pathogenesis.
    [Google Scholar]
  92. 92. 
    Nieto-Torres JL, Verdiá-Báguena C, Jimenez-Guardeño JM, Regla-Nava JA, Castaño-Rodriguez C et al. 2015. Severe acute respiratory syndrome coronavirus E protein transports calcium ions and activates the NLRP3 inflammasome. Virology 485:330–39
    [Google Scholar]
  93. 93. 
    Ogata M, Hino S, Saito A, Morikawa K, Kondo S et al. 2006. Autophagy is activated for cell survival after endoplasmic reticulum stress. Mol. Cell. Biol. 26:249220–31
    [Google Scholar]
  94. 94. 
    Oostra M, de Haan CAM, Rottier PJM 2007. The 29-nucleotide deletion present in human but not in animal severe acute respiratory syndrome coronaviruses disrupts the functional expression of open reading frame 8. J. Virol. 81:2413876–88
    [Google Scholar]
  95. 95. 
    Padhan K, Minakshi R, Towheed MAB, Jameel S 2008. Severe acute respiratory syndrome coronavirus 3a protein activates the mitochondrial death pathway through p38 MAP kinase activation. J. Gen. Virol. 89:Part 81960–69
    [Google Scholar]
  96. 96. 
    Peiris J, Lai S, Poon L, Guan Y, Yam L et al. 2003. Coronavirus as a possible cause of severe acute respiratory syndrome. Lancet 361:93661319–25
    [Google Scholar]
  97. 97. 
    Pene F, Merlat A, Vabret A, Rozenberg F, Buzyn A et al. 2003. Coronavirus 229E-related pneumonia in immunocompromised patients. Clin. Infect. Dis. 37:7929–32
    [Google Scholar]
  98. 98. 
    Perlman S, Netland J. 2009. Coronaviruses post-SARS: update on replication and pathogenesis. Nat. Rev. Microbiol. 7:6439–50
    [Google Scholar]
  99. 99. 
    Prentice E, Jerome WG, Yoshimori T, Mizushima N, Denison MR 2004. Coronavirus replication complex formation utilizes components of cellular autophagy. J. Biol. Chem. 279:1110136–41
    [Google Scholar]
  100. 100. 
    Rabouw HH, Langereis MA, Knaap RCM, Dalebout TJ, Canton J et al. 2016. Middle East respiratory coronavirus accessory protein 4a inhibits PKR-mediated antiviral stress responses. PLOS Pathog 12:10e1005982
    [Google Scholar]
  101. 101. 
    Reggiori F, Monastyrska I, Verheije MH, Cali T, Ulasli M et al. 2010. Coronaviruses hijack the LC3-I-positive EDEMosomes, ER-derived vesicles exporting short-lived ERAD regulators, for replication. Cell Host Microbe 7:6500–8
    [Google Scholar]
  102. 102. 
    Ron D, Walter P. 2007. Signal integration in the endoplasmic reticulum unfolded protein response. Nat. Rev. Mol. Cell Biol. 8:7519–29
    [Google Scholar]
  103. 103. 
    Rüdiger A-T, Mayrhofer P, Ma-Lauer Y, Pohlentz G, Müthing J et al. 2016. Tubulins interact with porcine and human S proteins of the genus Alphacoronavirus and support successful assembly and release of infectious viral particles. Virology 497:185–97
    [Google Scholar]
  104. 104. 
    Schneider M, Ackermann K, Stuart M, Wex C, Protzer U et al. 2012. Severe acute respiratory syndrome coronavirus replication is severely impaired by MG132 due to proteasome-independent inhibition of M-calpain. J. Virol. 86:1810112–22
    [Google Scholar]
  105. 105. 
    Simmons G, Gosalia DN, Rennekamp AJ, Reeves JD, Diamond SL, Bates P 2005. Inhibitors of cathepsin L prevent severe acute respiratory syndrome coronavirus entry. PNAS 102:3311876–81
    [Google Scholar]
  106. 106. 
    Smeal T, Binetruy B, Mercola DA, Birrer M, Karin M 1991. Oncogenic and transcriptional cooperation with Ha-Ras requires phosphorylation of c-Jun on serines 63 and 73. Nature 354:6353494–96
    [Google Scholar]
  107. 107. 
    Song HC, Seo M-Y, Stadler K, Yoo BJ, Choo Q-L et al. 2004. Synthesis and characterization of a native, oligomeric form of recombinant severe acute respiratory syndrome coronavirus spike glycoprotein. J. Virol. 78:1910328–35
    [Google Scholar]
  108. 108. 
    Spagnolo JF, Hogue BG. 2000. Host protein interactions with the 3′ end of bovine coronavirus RNA and the requirement of the poly(A) tail for coronavirus defective genome replication. J. Virol. 74:115053–65
    [Google Scholar]
  109. 109. 
    Sui J, Li W, Murakami A, Tamin A, Matthews LJ et al. 2004. Potent neutralization of severe acute respiratory syndrome (SARS) coronavirus by a human mAb to S1 protein that blocks receptor association. PNAS 101:82536–41
    [Google Scholar]
  110. 110. 
    Sung SC, Chao CY, Jeng KS, Yang JY, Lai M 2009. The 8ab protein of SARS-CoV is a luminal ER membrane-associated protein and induces the activation of ATF6. Virology 387:2402–13
    [Google Scholar]
  111. 111. 
    Tan YW, Hong W, Liu DX 2012. Binding of the 5′-untranslated region of coronavirus RNA to zinc finger CCHC-type and RNA-binding motif 1 enhances viral replication and transcription. Nucleic Acids Res 40:115065–77This paper demonstrates how CoV 5′-UTR recruits host factors that regulate viral replication and transcription.
    [Google Scholar]
  112. 112. 
    Tan Y-X, Tan THP, Lee MJ-R, Tham P-Y, Gunalan V et al. 2007. Induction of apoptosis by the severe acute respiratory syndrome coronavirus 7a protein is dependent on its interaction with the Bcl-XL protein. J. Virol. 81:126346–55
    [Google Scholar]
  113. 113. 
    Tao X, Hill TE, Morimoto C, Peters CJ, Ksiazek TG, Tseng C-TK 2013. Bilateral entry and release of Middle East respiratory syndrome coronavirus induces profound apoptosis of human bronchial epithelial cells. J. Virol. 87:179953–58
    [Google Scholar]
  114. 114. 
    Taylor RC, Cullen SP, Martin SJ 2008. Apoptosis: controlled demolition at the cellular level. Nat. Rev. Mol. Cell Biol. 9:3231–41
    [Google Scholar]
  115. 115. 
    Teoh K-T, Siu Y-L, Chan W-L, Schlüter MA, Liu C-J et al. 2010. The SARS coronavirus E protein interacts with PALS1 and alters tight junction formation and epithelial morphogenesis. Mol. Biol. Cell 21:223838–52
    [Google Scholar]
  116. 116. 
    Tirosh B, Iwakoshi NN, Glimcher LH, Ploegh HL 2006. Rapid turnover of unspliced Xbp-1 as a factor that modulates the unfolded protein response. J. Biol. Chem. 281:95852–60
    [Google Scholar]
  117. 117. 
    To J, Surya W, Fung TS, Li Y, Verdià-Bàguena C et al. 2017. Channel-inactivating mutations and their revertant mutants in the envelope protein of infectious bronchitis virus. J. Virol. 91:5e02158–16This paper demonstrates that CoV E ion channel activity is required for efficient CoV release.
    [Google Scholar]
  118. 118. 
    Urano F, Wang X, Bertolotti A, Zhang Y, Chung P et al. 2000. Coupling of stress in the ER to activation of JNK protein kinases by transmembrane protein kinase IRE1. Science 287:5453664–66
    [Google Scholar]
  119. 119. 
    Verheije MH, Raaben M, Mari M, Te Lintelo EG, Reggiori F et al. 2008. Mouse hepatitis coronavirus RNA replication depends on GBF1-mediated ARF1 activation. PLOS Pathog 4:6e1000088
    [Google Scholar]
  120. 120. 
    Versteeg GA, Van De Nes PS, Bredenbeek PJ, Spaan WJM 2007. The coronavirus spike protein induces endoplasmic reticulum stress and upregulation of intracellular chemokine mRNA concentrations. J. Virol. 81:2010981–90
    [Google Scholar]
  121. 121. 
    Wang J, Fang S, Xiao H, Chen B, Tam JP, Liu DX 2009. Interaction of the coronavirus infectious bronchitis virus membrane protein with β-actin and its implication in virion assembly and budding. PLOS ONE 4:3e4908
    [Google Scholar]
  122. 122. 
    Wang K, Lu W, Chen J, Xie S, Shi H et al. 2012. PEDV ORF3 encodes an ion channel protein and regulates virus production. FEBS Lett 586:4384–91
    [Google Scholar]
  123. 123. 
    Wang X, Liao Y, Yap PL, Png KJ, Tam JP, Liu DX 2009. Inhibition of protein kinase R activation and upregulation of GADD34 expression play a synergistic role in facilitating coronavirus replication by maintaining de novo protein synthesis in virus-infected cells. J. Virol. 83:2312462–72
    [Google Scholar]
  124. 124. 
    Wang X, Ron D. 1996. Stress-induced phosphorylation and activation of the transcription factor CHOP (GADD153) by p38 MAP kinase. Science 272:52661347–49
    [Google Scholar]
  125. 125. 
    Wong HH, Kumar P, Tay FPL, Moreau D, Liu DX, Bard F 2015. Genome-wide screen reveals valosin-containing protein requirement for coronavirus exit from endosomes. J. Virol. 89:2111116–28
    [Google Scholar]
  126. 126. 
    Woo PC, Lau SK, Lam CS, Lau CC, Tsang AK et al. 2012. Discovery of seven novel mammalian and avian coronaviruses in Deltacoronavirus supports bat coronaviruses as the gene source of Alphacoronavirus and Betacoronavirus and avian coronaviruses as the gene source of Gammacoronavirus and Deltacoronavirus. J. Virol 86:3995–4008
    [Google Scholar]
  127. 127. 
    Woo PCY, Lau SKP, Chu C, Chan K, Tsoi H et al. 2005. Characterization and complete genome sequence of a novel coronavirus, coronavirus HKU1, from patients with pneumonia. J. Virol. 79:2884–95
    [Google Scholar]
  128. 128. 
    Wu C-H, Chen P-J, Yeh S-H 2014. Nucleocapsid phosphorylation and RNA helicase DDX1 recruitment enables coronavirus transition from discontinuous to continuous transcription. Cell Host Microbe 16:4462–72This paper demonstrates how a host protein regulates template read-through during HCoV genome transcription.
    [Google Scholar]
  129. 129. 
    Wu C-H, Yeh S-H, Tsay Y-G, Shieh Y-H, Kao C-L et al. 2009. Glycogen synthase kinase-3 regulates the phosphorylation of severe acute respiratory syndrome coronavirus nucleocapsid protein and viral replication. J. Biol. Chem. 284:85229–39
    [Google Scholar]
  130. 130. 
    Xu X, Liu Y, Weiss S, Arnold E, Sarafianos SG, Ding J 2003. Molecular model of SARS coronavirus polymerase: implications for biochemical functions and drug design. Nucleic Acids Res 31:247117–30
    [Google Scholar]
  131. 131. 
    Xue M, Fu F, Ma Y, Zhang X, Li L et al. 2018. The PERK arm of the unfolded protein response negatively regulates transmissible gastroenteritis virus replication by suppressing protein translation and promoting type I interferon production. J. Virol. 92:15e00431–18
    [Google Scholar]
  132. 132. 
    Yamada Y, Liu DX. 2009. Proteolytic activation of the spike protein at a novel RRRR/S motif is implicated in furin-dependent entry, syncytium formation, and infectivity of coronavirus infectious bronchitis virus in cultured cells. J. Virol. 83:178744–58
    [Google Scholar]
  133. 133. 
    Yamamoto K, Ichijo H, Korsmeyer SJ 1999. BCL-2 is phosphorylated and inactivated by an ASK1/Jun N-terminal protein kinase pathway normally activated at G2/M. Mol. Cell. Biol. 19:128469–78
    [Google Scholar]
  134. 134. 
    Yamamoto K, Yoshida H, Kokame K, Kaufman RJ, Mori K 2004. Differential contributions of ATF6 and XBP1 to the activation of endoplasmic reticulum stress-responsive cis-acting elements ERSE, UPRE and ERSE-II. J. Biochem. 136:3343–50
    [Google Scholar]
  135. 135. 
    Yang Z, Klionsky DJ. 2010. Eaten alive: a history of macroautophagy. Nat. Cell Biol. 12:9814–22
    [Google Scholar]
  136. 136. 
    Yeung M-L, Yao Y, Jia L, Chan JFW, Chan K-H et al. 2016. MERS coronavirus induces apoptosis in kidney and lung by upregulating Smad7 and FGF2. Nat. Microbiol. 1:316004
    [Google Scholar]
  137. 137. 
    Yoon S, Seger R. 2006. The extracellular signal-regulated kinase: multiple substrates regulate diverse cellular functions. Growth Factors 24:121–44
    [Google Scholar]
  138. 138. 
    Yoshida H, Matsui T, Yamamoto A, Okada T, Mori K 2001. XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell 107:7881–91
    [Google Scholar]
  139. 139. 
    Yoshida H, Okada T, Haze K, Yanagi H, Yura T et al. 2001. Endoplasmic reticulum stress-induced formation of transcription factor complex ERSF including NF-Y (CBF) and activating transcription factors 6α and 6β that activates the mammalian unfolded protein response. Mol. Cell. Biol. 21:41239–48
    [Google Scholar]
  140. 140. 
    Zarubin T, Han J. 2005. Activation and signaling of the p38 MAP kinase pathway. Cell Res 15:111–18
    [Google Scholar]
  141. 141. 
    Zhang R, Wang K, Lv W, Yu W, Xie S et al. 2014. The ORF4a protein of human coronavirus 229E functions as a viroporin that regulates viral production. Biochim. Biophys. Acta 1838:41088–95
    [Google Scholar]
  142. 142. 
    Zhang R, Wang K, Ping X, Yu W, Qian Z et al. 2015. The ns12.9 accessory protein of human coronavirus OC43 is a viroporin involved in virion morphogenesis and pathogenesis. J. Virol. 89:2211383–95
    [Google Scholar]
  143. 143. 
    Zhang X, Shi H, Chen J, Shi D, Dong H, Feng L 2015. Identification of the interaction between vimentin and nucleocapsid protein of transmissible gastroenteritis virus. Virus Res 200:56–63
    [Google Scholar]
  144. 144. 
    Zhao X, Guo F, Liu F, Cuconati A, Chang J et al. 2014. Interferon induction of IFITM proteins promotes infection by human coronavirus OC43. PNAS 111:186756–61
    [Google Scholar]
  145. 145. 
    Zhao X, Sehgal M, Hou Z, Cheng J, Shu S et al. 2018. Identification of residues controlling restriction versus enhancing activities of IFITM proteins on entry of human coronaviruses. J. Virol. 92:6e01535–17
    [Google Scholar]
  146. 146. 
    Zhao Z, Thackray LB, Miller BC, Lynn TM, Becker MM et al. 2007. Coronavirus replication does not require the autophagy gene ATG5. Autophagy 3:6581–85
    [Google Scholar]
  147. 147. 
    Zheng J, Yamada Y, Fung TS, Huang M, Chia R, Liu DX 2018. Identification of N-linked glycosylation sites in the spike protein and their functional impact on the replication and infectivity of coronavirus infectious bronchitis virus in cell culture. Virology 513:65–74
    [Google Scholar]
  148. 148. 
    Zhu L, Mou C, Yang X, Lin J, Yang Q 2016. Mitophagy in TGEV infection counteracts oxidative stress and apoptosis. Oncotarget 7:1927122–41
    [Google Scholar]
  149. 149. 
    Ziebuhr J, Snijder EJ, Gorbalenya AE 2000. Virus-encoded proteinases and proteolytic processing in the Nidovirales. J. Gen. Virol 81:4853–79
    [Google Scholar]
  150. 150. 
    Zúñiga S, Cruz JLG, Sola I, Mateos-Gómez PA, Palacio L, Enjuanes L 2010. Coronavirus nucleocapsid protein facilitates template switching and is required for efficient transcription. J. Virol. 84:42169–75
    [Google Scholar]
  151. 151. 
    Zúñiga S, Sola I, Moreno JL, Sabella P, Plana-Durán J, Enjuanes L 2007. Coronavirus nucleocapsid protein is an RNA chaperone. Virology 357:2215–27
    [Google Scholar]
/content/journals/10.1146/annurev-micro-020518-115759
Loading
/content/journals/10.1146/annurev-micro-020518-115759
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