Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Filovirus pathogenesis and immune evasion: insights from Ebola virus and Marburg virus

Key Points

  • The 2014–2015 Ebola virus outbreak in West Africa is the largest filovirus outbreak to date.

  • Filovirus infections in humans are acquired by direct contact with infected bodily fluid through breaks in the skin or mucosal membranes, and the virus typically targets macrophages, monocytes and dendritic cells, among other cell types.

  • Severe infection with filoviruses can lead to viral haemorrhagic fever and other coagulation abnormalities.

  • Filoviruses evade and alter the innate immune response through various mechanisms.

  • Viral protein 35 (VP35) antagonizes the innate immune system by blocking signalling through RIG-I-like receptors (RLRs) and further impairing the interferon (IFN) response.

  • VP35 can bind to double-stranded RNA (dsRNA), further suppressing the RLR-mediated IFN response, and can interact with PACT, an activator of the IFN-induced antiviral kinase protein kinase (PKR).

  • Ebola virus VP24 can bind to karyopherin-α (KPNA) proteins, preventing their interaction with tyrosine-phosphorylated signal transducer and activator of transcription 1 (STAT1), which then prevents the activation of IFN-stimulated response elements.

  • Marburg virus VP40 blocks the activation and function of Janus kinase 1 (JAK1), inhibiting the type I IFN-induced phosphorylation of STAT1 and STAT2.

  • Filoviruses adapt to their host in order to sustain viral replication and to subvert translational regulation mechanisms.

  • Marburg virus VP24 can modulate the function of kelch-like ECH-associated protein 1 (KEAP1), enabling activation of antioxidant response elements (AREs) and promoting survival of virus-infected host cells.

  • Further understanding of filoviruses is necessary to develop better therapeutics and more efficiently combat infection.

Abstract

Ebola viruses and Marburg viruses, members of the filovirus family, are zoonotic pathogens that cause severe disease in people, as highlighted by the latest Ebola virus epidemic in West Africa. Filovirus disease is characterized by uncontrolled virus replication and the activation of host responses that contribute to pathogenesis. Underlying these phenomena is the potent suppression of host innate antiviral responses, particularly the type I interferon response, by viral proteins, which allows high levels of viral replication. In this Review, we describe the mechanisms used by filoviruses to block host innate immunity and discuss the links between immune evasion and filovirus pathogenesis.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Filovirus genome structure and life cycle.
Figure 2: Subversion of IFN induction.
Figure 3: Subversion of IFN-induced signalling and of IFN-induced protein tetherin.
Figure 4: Subversion of host translation mechanisms to support virus replication.

Similar content being viewed by others

Accession codes

Accessions

Protein Data Bank

References

  1. Baize, S. et al. Emergence of Zaire Ebola virus disease in Guinea. N. Engl. J. Med. 371, 1418–1425 (2014).

    Article  CAS  PubMed  Google Scholar 

  2. Feldmann, H. & Geisbert, T. W. Ebola haemorrhagic fever. Lancet 377, 849–862 (2010).

    Article  Google Scholar 

  3. Sanchez, A., Geisbert, T. W. & Feldmann, H. in Fields Virology (eds Knipe, D. M. et al.) 1410–1448 (Lippincott Williams and Wilkins, 2007).

    Google Scholar 

  4. Gire, S. K. et al. Genomic surveillance elucidates Ebola virus origin and transmission during the 2014 outbreak. Science 345, 1369–1372 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. World Health Organization. Ebola situation report — 16 September 2015 (WHO, 2015).

  6. Kuhn, J. H. et al. Virus nomenclature below the species level: a standardized nomenclature for filovirus strains and variants rescued from cDNA. Arch. Virol. 159, 1229–1237 (2013).

    PubMed  PubMed Central  Google Scholar 

  7. Negredo, A. et al. Discovery of an ebolavirus-like filovirus in Europe. PLoS Pathog. 7, e1002304 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Miranda, M. E. & Miranda, N. L. Reston ebolavirus in humans and animals in the Philippines: a review. J. Infect. Dis. 204 (Suppl. 3), 757–760 (2011).

    Article  Google Scholar 

  9. Amman, B. R. et al. Oral shedding of Marburg virus in experimentally infected Egyptian fruit bats (Rousettus aegyptiacus). J. Wildl. Dis. 51, 113–124 (2014).

    Article  Google Scholar 

  10. Paweska, J. T. et al. Virological and serological findings in Rousettus aegyptiacus experimentally inoculated with vero cells-adapted hogan strain of Marburg virus. PLoS ONE 7, e45479 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Volchkov, V. E., Chepurnov, A. A., Volchkova, V. A., Ternovoj, V. A. & Klenk, H. D. Molecular characterization of guinea pig-adapted variants of Ebola virus. Virology 277, 147–155 (2000).

    Article  CAS  PubMed  Google Scholar 

  12. Lofts, L. L., Ibrahim, M. S., Negley, D. L., Hevey, M. C. & Schmaljohn, A. L. Genomic differences between guinea pig lethal and nonlethal Marburg virus variants. J. Infect. Dis. 196 (Suppl. 2), 305–312 (2007).

    Article  CAS  Google Scholar 

  13. Bray, M., Davis, K., Geisbert, T., Schmaljohn, C. & Huggins, J. A mouse model for evaluation of prophylaxis and therapy of Ebola hemorrhagic fever. J. Infect. Dis. 179 (Suppl. 1), 248–258 (1999).

    Article  Google Scholar 

  14. Warfield, K. L. et al. Development and characterization of a mouse model for Marburg hemorrhagic fever. J. Virol. 83, 6404–6415 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Bray, M. The role of the Type I interferon response in the resistance of mice to filovirus infection. J. Gen. Virol. 82, 1365–1373 (2001).

    Article  CAS  PubMed  Google Scholar 

  16. Mehedi, M. et al. A new Ebola virus nonstructural glycoprotein expressed through RNA editing. J. Virol. 85, 5406–5414 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Sanchez, A., Trappier, S. G., Mahy, B. W., Peters, C. J. & Nichol, S. T. The virion glycoproteins of Ebola viruses are encoded in two reading frames and are expressed through transcriptional editing. Proc. Natl Acad. Sci. USA 93, 3602–3607 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Shabman, R. S. et al. Deep sequencing identifies noncanonical editing of Ebola and Marburg virus RNAs in infected cells. mBio 5, e02011–14 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Basler, C. F. & Amarasinghe, G. K. Evasion of interferon responses by Ebola and Marburg viruses. J. Interferon Cytokine Res. 29, 511–520 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Basler, C. F. et al. The Ebola virus VP35 protein functions as a type I IFN antagonist. Proc. Natl Acad. Sci. USA 97, 12289–12294 (2000). This is the first demonstration that the eVP35 protein blocks IFN responses.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Reid, S. P. et al. Ebola virus VP24 binds karyopherin α1 and blocks STAT1 nuclear accumulation. J. Virol. 80, 5156–5167 (2006). This is the first demonstration that the eVP24 protein blocks IFN-induced signalling.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Ramanan, P. et al. Structural basis for Marburg virus VP35-mediated immune evasion mechanisms. Proc. Natl Acad. Sci. USA 109, 20661–20666 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Bale, S. et al. Marburg virus VP35 can both fully coat the backbone and cap the ends of dsRNA for interferon antagonism. PLoS Pathog. 8, e1002916 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Valmas, C. et al. Marburg virus evades interferon responses by a mechanism distinct from Ebola virus. PLoS Pathog. 6, e1000721 (2010). This study demonstrates that the mVP40 protein blocks IFN-induced signalling.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Dowell, S. F. et al. Transmission of Ebola hemorrhagic fever: a study of risk factors in family members, Kikwit, Democratic Republic of the Congo, 1995. J. Infect. Dis. 179 (Suppl. 1), 87–91 (1999).

    Article  Google Scholar 

  26. Geisbert, T. W. et al. Pathogenesis of Ebola hemorrhagic fever in primate models: evidence that hemorrhage is not a direct effect of virus-induced cytolysis of endothelial cells. Am. J. Pathol. 163, 2371–2382 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Geisbert, T. W. et al. Pathogenesis of Ebola hemorrhagic fever in cynomolgus macaques: evidence that dendritic cells are early and sustained targets of infection. Am. J. Pathol. 163, 2347–2370 (2003). This study demonstrates that macrophages and dendritic cells are major targets of EBOV throughout the course of infection in non-human primates.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Zaki, S. R. et al. A novel immunohistochemical assay for the detection of Ebola virus in skin: implications for diagnosis, spread, and surveillance of Ebola hemorrhagic fever. J. Infect. Dis. 179 (Suppl. 1), 36–47 (1999).

    Article  Google Scholar 

  29. Yonezawa, A., Cavrois, M. & Greene, W. C. Studies of Ebola virus glycoprotein-mediated entry and fusion by using pseudotyped human immunodeficiency virus type 1 virions: involvement of cytoskeletal proteins and enhancement by tumor necrosis factor alpha. J. Virol. 79, 918–926 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Dube, D. et al. Cell adhesion promotes Ebola virus envelope glycoprotein-mediated binding and infection. J. Virol. 82, 7238–7242 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Martinez, O. et al. Ebola virus exploits a monocyte differentiation program to promote its entry. J. Virol. 87, 3801–3814 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Chandran, K., Sullivan, N. J., Felbor, U., Whelan, S. P. & Cunningham, J. M. Endosomal proteolysis of the Ebola virus glycoprotein is necessary for infection. Science 308, 1643–1645 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Carette, J. E. et al. Ebola virus entry requires the cholesterol transporter Niemann–Pick C1. Nature 477, 340–343 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Cote, M. et al. Small molecule inhibitors reveal Niemann–Pick C1 is essential for Ebola virus infection. Nature 477, 344–348 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Schornberg, K. L. et al. α5β1-integrin controls Ebolavirus entry by regulating endosomal cathepsins. Proc. Natl Acad. Sci. USA 106, 8003–8008 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Moller-Tank, S. & Maury, W. Phosphatidylserine receptors: enhancers of enveloped virus entry and infection. Virology 468–470, 565–580 (2014).

    Article  PubMed  CAS  Google Scholar 

  37. Hunt, C. L., Kolokoltsov, A. A., Davey, R. A. & Maury, W. The Tyro3 receptor kinase Axl enhances macropinocytosis of Zaire ebolavirus. J. Virol. 85, 334–347 (2011).

    Article  CAS  PubMed  Google Scholar 

  38. Shimojima, M. et al. Tyro3 family-mediated cell entry of Ebola and Marburg viruses. J. Virol. 80, 10109–10116 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Simmons, G. et al. DC-SIGN and DC-SIGNR bind Ebola glycoproteins and enhance infection of macrophages and endothelial cells. Virology 305, 115–123 (2003).

    Article  CAS  PubMed  Google Scholar 

  40. Baize, S. et al. Inflammatory responses in Ebola virus-infected patients. Clin. Exp. Immunol. 128, 163–168 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Ebihara, H. et al. Host response dynamics following lethal infection of rhesus macaques with Zaire ebolavirus. J. Infect. Dis. 204 (Suppl. 3), 991–999 (2011).

    Article  CAS  Google Scholar 

  42. Hensley, L. E., Young, H. A., Jahrling, P. B. & Geisbert, T. W. Proinflammatory response during Ebola virus infection of primate models: possible involvement of the tumor necrosis factor receptor superfamily. Immunol. Lett. 80, 169–179 (2002).

    Article  CAS  PubMed  Google Scholar 

  43. Sanchez, A. et al. Analysis of human peripheral blood samples from fatal and nonfatal cases of Ebola (Sudan) hemorrhagic fever: cellular responses, virus load, and nitric oxide levels. J. Virol. 78, 10370–10377 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Villinger, F. et al. Markedly elevated levels of interferon (IFN)-γ, IFN-α, interleukin (IL)-2, IL-10, and tumor necrosis factor-α associated with fatal Ebola virus infection. J. Infect. Dis. 179 (Suppl. 1), 188–191 (1999).

    Article  Google Scholar 

  45. Wauquier, N., Becquart, P., Padilla, C., Baize, S. & Leroy, E. M. Human fatal Zaire Ebola virus infection is associated with an aberrant innate immunity and with massive lymphocyte apoptosis. PLoS Negl. Trop. Dis. 4, e837 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Gupta, M., Mahanty, S., Ahmed, R. & Rollin, P. E. Monocyte-derived human macrophages and peripheral blood mononuclear cells infected with ebola virus secrete MIP-1α and TNF-α and inhibit poly-IC-induced IFN-alpha in vitro. Virology 284, 20–25 (2001).

    Article  CAS  PubMed  Google Scholar 

  47. Stroher, U. et al. Infection and activation of monocytes by Marburg and Ebola viruses. J. Virol. 75, 11025–11033 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Wahl-Jensen, V. et al. Ebola virion attachment and entry into human macrophages profoundly effects early cellular gene expression. PLoS Negl. Trop. Dis. 5, e1359 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Lubaki, N. M. et al. The lack of maturation of Ebola virus-infected dendritic cells results from the cooperative effect of at least two viral domains. J. Virol. 87, 7471–7485 (2013). This study demonstrates that suppression of dendritic cell maturation by EBOVs with mutated VP35 proteins is greatly impaired.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Mahanty, S. et al. Cutting edge: impairment of dendritic cells and adaptive immunity by Ebola and Lassa viruses. J. Immunol. 170, 2797–2801 (2003). This is one of the first in vitro studies to demonstrate suppression of dendritic cell maturation by filovirus infection.

    Article  CAS  PubMed  Google Scholar 

  51. Bosio, C. M. et al. Ebola and Marburg viruses replicate in monocyte-derived dendritic cells without inducing the production of cytokines and full maturation. J. Infect. Dis. 188, 1630–1638 (2003). This is one of the first in vitro studies to demonstrate suppression of dendritic cell maturation by filovirus infection.

    Article  CAS  PubMed  Google Scholar 

  52. Yen, B., Mulder, L. C., Martinez, O. & Basler, C. F. Molecular basis for Ebola virus VP35 suppression of human dendritic cell maturation. J. Virol. 88, 12500–12510 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  53. Jin, H. et al. The VP35 protein of Ebola virus impairs dendritic cell maturation induced by virus and lipopolysaccharide. J. Gen. Virol. 91, 352–361 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Bradfute, S. B., Warfield, K. L. & Bavari, S. Functional CD8+ T cell responses in lethal Ebola virus infection. J. Immunol. 180, 4058–4066 (2008).

    Article  CAS  PubMed  Google Scholar 

  55. McElroy, A. K. et al. Human Ebola virus infection results in substantial immune activation. Proc. Natl Acad. Sci. USA 112, 4719–4724 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Geisbert, T. W. et al. Apoptosis induced in vitro and in vivo during infection by Ebola and Marburg viruses. Lab Invest. 80, 171–186 (2000).

    Article  CAS  PubMed  Google Scholar 

  57. Bradfute, S. B. et al. Lymphocyte death in a mouse model of Ebola virus infection. J. Infect. Dis. 196 (Suppl. 2), 296–304 (2007).

    Article  Google Scholar 

  58. Reed, D. S., Hensley, L. E., Geisbert, J. B., Jahrling, P. B. & Geisbert, T. W. Depletion of peripheral blood T lymphocytes and NK cells during the course of Ebola hemorrhagic fever in cynomolgus macaques. Viral Immunol. 17, 390–400 (2004).

    Article  CAS  PubMed  Google Scholar 

  59. Gupta, M., Spiropoulou, C. & Rollin, P. E. Ebola virus infection of human PBMCs causes massive death of macrophages, CD4 and CD8 T cell sub-populations in vitro. Virology 364, 45–54 (2007).

    Article  CAS  PubMed  Google Scholar 

  60. Baize, S. et al. Defective humoral responses and extensive intravascular apoptosis are associated with fatal outcome in Ebola virus-infected patients. Nat. Med. 5, 423–426 (1999).

    Article  CAS  PubMed  Google Scholar 

  61. Bradfute, S. B. et al. Mechanisms and consequences of Ebolavirus-induced lymphocyte apoptosis. J. Immunol. 184, 327–335 (2010).

    Article  CAS  PubMed  Google Scholar 

  62. Rath, P. C. & Aggarwal, B. B. TNF-induced signaling in apoptosis. J. Clin. Immunol. 19, 350–364 (1999).

    Article  CAS  PubMed  Google Scholar 

  63. Simon, H. U., Haj-Yehia, A. & Levi-Schaffer, F. Role of reactive oxygen species (ROS) in apoptosis induction. Apoptosis 5, 415–418 (2000).

    Article  CAS  PubMed  Google Scholar 

  64. Snyder, C. M., Shroff, E. H., Liu, J. & Chandel, N. S. Nitric oxide induces cell death by regulating anti-apoptotic BCL-2 family members. PLoS ONE 4, e7059 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  65. Mohan, G. S., Li, W., Ye, L., Compans, R. W. & Yang, C. Antigenic subversion: a novel mechanism of host immune evasion by Ebola virus. PLoS Pathog. 8, e1003065 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Kang, E. S. et al. Hypotension during hemodialysis: role for nitric oxide. Am. J. Med. Sci. 313, 138–146 (1997).

    CAS  PubMed  Google Scholar 

  67. Geisbert, T. W. et al. Mechanisms underlying coagulation abnormalities in Ebola hemorrhagic fever: overexpression of tissue factor in primate monocytes/macrophages is a key event. J. Infect. Dis. 188, 1618–1629 (2003).

    Article  CAS  PubMed  Google Scholar 

  68. Neumann, F. J. et al. Effect of human recombinant interleukin-6 and interleukin-8 on monocyte procoagulant activity. Arterioscler. Thromb. Vasc. Biol. 17, 3399–3405 (1997).

    Article  CAS  PubMed  Google Scholar 

  69. Bray, M., Hatfill, S., Hensley, L. & Huggins, J. W. Haematological, biochemical and coagulation changes in mice, guinea-pigs and monkeys infected with a mouse-adapted variant of Ebola Zaire virus. J. Comp. Pathol. 125, 243–253 (2001).

    Article  CAS  PubMed  Google Scholar 

  70. Feldmann, H. et al. Filovirus-induced endothelial leakage triggered by infected monocytes/macrophages. J. Virol. 70, 2208–2214 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Ray, R. B. et al. Ebola virus glycoprotein-mediated anoikis of primary human cardiac microvascular endothelial cells. Virology 321, 181–188 (2004).

    Article  CAS  PubMed  Google Scholar 

  72. Wahl-Jensen, V. M. et al. Effects of Ebola virus glycoproteins on endothelial cell activation and barrier function. J. Virol. 79, 10442–10450 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Yang, Z. Y. et al. Identification of the Ebola virus glycoprotein as the main viral determinant of vascular cell cytotoxicity and injury. Nat. Med. 6, 886–889 (2000).

    Article  CAS  PubMed  Google Scholar 

  74. Schneider, W. M., Chevillotte, M. D. & Rice, C. M. Interferon-stimulated genes: a complex web of host defenses. Immunology 32, 513–545 (2014).

    Article  CAS  Google Scholar 

  75. Schoggins, J. W. Interferon-stimulated genes: roles in viral pathogenesis. Curr. Opin. Virol. 6, 40–46 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Harcourt, B. H., Sanchez, A. & Offermann, M. K. Ebola virus inhibits induction of genes by double-stranded RNA in endothelial cells. Virology 252, 179–188 (1998). This is one of the first studies to demonstrate that EBOV infection blocks cellular production of IFNs.

    Article  CAS  PubMed  Google Scholar 

  77. Harcourt, B. H., Sanchez, A. & Offermann, M. K. Ebola virus selectively inhibits responses to interferons, but not to interleukin-1β, in endothelial cells. J. Virol. 73, 3491–3496 (1999). This is one of the first studies to demonstrate that EBOV infection blocks cellular responses to IFNs.

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Kash, J. C. et al. Global suppression of the host antiviral response by Ebola- and Marburgviruses: increased antagonism of the type I interferon response is associated with enhanced virulence. J. Virol. 80, 3009–3020 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Spiropoulou, C. F. et al. RIG-I activation inhibits Ebolavirus replication. Virology 392, 11–15 (2009).

    Article  CAS  PubMed  Google Scholar 

  80. Hartman, A. L. et al. Inhibition of IRF-3 activation by VP35 is critical for the high level of virulence of Ebola virus. J. Virol. 82, 2699–2704 (2008). This study demonstrates that disruption of VP35 IRF3-antagonist function attenuates EBOV replication in a mouse model.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Prins, K. C. et al. Mutations abrogating VP35 interaction with double-stranded RNA render Ebola virus avirulent in guinea pigs. J. Virol. 84, 3004–3015 (2010). This study demonstrates that disruption of VP35 IFN-antagonist function renders EBOV non-lethal in a guinea pig model.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Mahanty, S. et al. Protection from lethal infection is determined by innate immune responses in a mouse model of Ebola virus infection. Virology 312, 415–424 (2003).

    Article  CAS  PubMed  Google Scholar 

  83. Huang, I. C. et al. Distinct patterns of IFITM-mediated restriction of filoviruses, SARS coronavirus, and influenza A virus. PLoS Pathog. 7, e1001258 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Okumura, A., Pitha, P. M. & Harty, R. N. ISG15 inhibits Ebola VP40 VLP budding in an L-domain-dependent manner by blocking Nedd4 ligase activity. Proc. Natl Acad. Sci. USA 105, 3974–3979 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Bray, M., Raymond, J. L., Geisbert, T. & Baker, R. O. 3-Deazaneplanocin A induces massively increased interferon-α production in Ebola virus-infected mice. Antiviral Res. 55, 151–159 (2002).

    Article  CAS  PubMed  Google Scholar 

  86. Jahrling, P. B. et al. Evaluation of immune globulin and recombinant interferon-α2b for treatment of experimental Ebola virus infections. J. Infect. Dis. 179 (Suppl. 1), 224–234 (1999).

    Article  Google Scholar 

  87. Basler, C. F. et al. The Ebola virus VP35 protein inhibits activation of interferon regulatory factor 3. J. Virol. 77, 7945–7956 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Prins, K. C., Cardenas, W. B. & Basler, C. F. Ebola virus protein VP35 impairs the function of interferon regulatory factor-activating kinases IKKε and TBK-1. J. Virol. 83, 3069–3077 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Chang, T. H. et al. Ebola Zaire virus blocks type I interferon production by exploiting the host SUMO modification machinery. PLoS Pathog. 5, e1000493 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  90. Kubota, T. et al. Virus infection triggers SUMOylation of IRF3 and IRF7, leading to the negative regulation of type I interferon gene expression. J. Biol. Chem. 283, 25660–25670 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Leung, D. W. et al. Structure of the Ebola VP35 interferon inhibitory domain. Proc. Natl Acad. Sci. USA 106, 411–416 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Cardenas, W. B. et al. Ebola virus VP35 protein binds double-stranded RNA and inhibits alpha/beta interferon production induced by RIG-I signaling. J. Virol. 80, 5168–5178 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Hartman, A. L., Towner, J. S. & Nichol, S. T. A C-terminal basic amino acid motif of Zaire ebolavirus VP35 is essential for type I interferon antagonism and displays high identity with the RNA-binding domain of another interferon antagonist, the NS1 protein of influenza A virus. Virology 328, 177–184 (2004).

    Article  CAS  PubMed  Google Scholar 

  94. Leung, D. W. et al. Structural basis for dsRNA recognition and interferon antagonism by Ebola VP35. Nat. Struct. Mol. Biol. 17, 165–172 (2010). This study describes the structural basis of VP35 IFN-antagonist and dsRNA-binding activities.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Bale, S. et al. Ebolavirus VP35 coats the backbone of double-stranded RNA for interferon antagonism. J. Virol. 87, 10385–10388 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Luthra, P. et al. Mutual antagonism between the Ebola virus VP35 protein and the RIG-I activator PACT determines infection outcome. Cell Host Microbe 14, 74–84 (2013). This study demonstrates that interaction of eVP35 with the cellular protein PACT contributes to its suppression of IFN production.

    Article  CAS  PubMed  Google Scholar 

  97. Patel, R. C. & Sen, G. C. PACT, a protein activator of the interferon-induced protein kinase, PKR. EMBO J. 17, 4379–4390 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Kok, K. H. et al. The double-stranded RNA-binding protein PACT functions as a cellular activator of RIG-I to facilitate innate antiviral response. Cell Host Microbe 9, 299–309 (2011).

    Article  CAS  PubMed  Google Scholar 

  99. Mellman, I. & Steinman, R. M. Dendritic cells: specialized and regulated antigen processing machines. Cell 106, 255–258 (2001).

    Article  CAS  PubMed  Google Scholar 

  100. Leung, L. W. et al. Ebolavirus VP35 suppresses IFN production from conventional but not plasmacytoid dendritic cells. Immunol. Cell Biol. 89, 792–802 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Reid, S. P., Cardenas, W. B. & Basler, C. F. Homo-oligomerization facilitates the interferon-antagonist activity of the Ebolavirus VP35 protein. Virology 341, 179–189 (2005).

    Article  CAS  PubMed  Google Scholar 

  102. Prins, K. C. et al. Basic residues within the Ebolavirus VP35 protein are required for its viral polymerase cofactor function. J. Virol. 84, 10581–10591 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Kimberlin, C. R. et al. Ebolavirus VP35 uses a bimodal strategy to bind dsRNA for innate immune suppression. Proc. Natl Acad. Sci. USA 107, 314–319 (2009). This study describes the structural basis of VP35 IFN-antagonist and dsRNA-binding activities.

    Article  PubMed  PubMed Central  Google Scholar 

  104. Leung, D. W. et al. Structural and functional characterization of Reston Ebola VP35 interferon inhibitory domain. J. Mol. Biol. 399, 347–357 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Platanias, L. C. Mechanisms of type-I- and type-II-interferon-mediated signalling. Nat. Rev. Immunol. 5, 375–386 (2005).

    Article  CAS  PubMed  Google Scholar 

  106. McBride, K. M. & Reich, N. C. The ins and outs of STAT1 nuclear transport. Sci. STKE 2003, Re13 (2003).

    PubMed  Google Scholar 

  107. Reid, S. P., Valmas, C., Martinez, O., Sanchez, F. M. & Basler, C. F. Ebola virus VP24 proteins inhibit the interaction of NPI-1 subfamily karyopherin α proteins with activated STAT1. J. Virol. 81, 13469–13477 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Zhang, A. P. et al. The Ebola virus interferon antagonist VP24 directly binds STAT1 and has a novel, pyramidal fold. PLoS Pathog. 8, e1002550 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Edwards, M. R. et al. The Marburg virus VP24 protein interacts with Keap1 to activate the cytoprotective antioxidant response pathway. Cell Rep. 6, 1017–1025 (2014). This study explains the molecular basis of the interaction between mVP24 and the cellular protein KEAP1, which activates the antioxidant response.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Xu, W. et al. Ebola virus VP24 targets a unique NLS binding site on karyopherin alpha 5 to selectively compete with nuclear import of phosphorylated STAT1. Cell Host Microbe 16, 187–200 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Shabman, R. S., Gulcicek, E. E., Stone, K. L. & Basler, C. F. The Ebola virus VP24 protein prevents hnRNP C1/C2 binding to karyopherin α1 and partially alters its nuclear import. J. Infect. Dis. 204 (Suppl. 3), 904–910 (2011).

    Article  CAS  Google Scholar 

  112. Rodig, S. J. et al. Disruption of the Jak1 gene demonstrates obligatory and nonredundant roles of the Jaks in cytokine-induced biologic responses. Cell 93, 373–383 (1998).

    Article  CAS  PubMed  Google Scholar 

  113. Feng, Z., Cerveny, M., Yan, Z. & He, B. The VP35 protein of Ebola virus inhibits the antiviral effect mediated by double-stranded RNA-dependent protein kinase PKR. J. Virol. 81, 182–192 (2007).

    Article  CAS  PubMed  Google Scholar 

  114. Schümann, M., Gantke, T. & Mühlberger, E. Ebola virus VP35 antagonizes PKR activity through its C-terminal interferon inhibitory domain. J. Virol. 83, 8993–8997 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  115. Peters, G. A., Hartmann, R., Qin, J. & Sen, G. C. Modular structure of PACT: distinct domains for binding and activating PKR. Mol. Cell. Biol. 21, 1908–1920 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Strong, J. E. et al. Stimulation of Ebola virus production from persistent infection through activation of the Ras/MAPK pathway. Proc. Natl Acad. Sci. USA 105, 17982–17987 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Neil, S. J., Zang, T. & Bieniasz, P. D. Tetherin inhibits retrovirus release and is antagonized by HIV-1 Vpu. Nature 451, 425–430 (2008).

    Article  CAS  PubMed  Google Scholar 

  118. Van Damme, N. et al. The interferon-induced protein BST-2 restricts HIV-1 release and is downregulated from the cell surface by the viral Vpu protein. Cell Host Microbe 3, 245–252 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Jouvenet, N. et al. Broad-spectrum inhibition of retroviral and filoviral particle release by tetherin. J. Virol. 83, 1837–1844 (2009).

    Article  CAS  PubMed  Google Scholar 

  120. Neil, S. J., Sandrin, V., Sundquist, W. I. & Bieniasz, P. D. An interferon-α-induced tethering mechanism inhibits HIV-1 and Ebola virus particle release but is counteracted by the HIV-1 Vpu protein. Cell Host Microbe 2, 193–203 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Radoshitzky, S. R. et al. Infectious Lassa virus, but not filoviruses, is restricted by BST-2/tetherin. J. Virol. 84, 10569–10580 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Kaletsky, R. L., Francica, J. R., Agrawal-Gamse, C. & Bates, P. Tetherin-mediated restriction of filovirus budding is antagonized by the Ebola glycoprotein. Proc. Natl Acad. Sci. USA 106, 2886–2891 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Kuhl, A. et al. The Ebola virus glycoprotein and HIV-1 Vpu employ different strategies to counteract the antiviral factor tetherin. J. Infect. Dis. 204 (Suppl. 3), 850–860 (2011).

    Article  CAS  Google Scholar 

  124. Lopez, L. A. et al. Anti-tetherin activities of HIV-1 Vpu and Ebola virus glycoprotein do not involve removal of tetherin from lipid rafts. J. Virol. 86, 5467–5480 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Lopez, L. A. et al. Ebola virus glycoprotein counteracts BST-2/Tetherin restriction in a sequence-independent manner that does not require tetherin surface removal. J. Virol. 84, 7243–7255 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Wek, R. C., Jiang, H. Y. & Anthony, T. G. Coping with stress: eIF2 kinases and translational control. Biochem. Soc. Trans. 34, 7–11 (2006).

    Article  CAS  PubMed  Google Scholar 

  127. Shabman, R. S. et al. An upstream open reading frame modulates Ebola virus polymerase translation and virus replication. PLoS Pathog. 9, e1003147 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Pichlmair, A. et al. Viral immune modulators perturb the human molecular network by common and unique strategies. Nature 487, 486–490 (2012).

    Article  CAS  PubMed  Google Scholar 

  129. Page, A. et al. Marburgvirus hijacks Nrf2-dependent pathway by targeting Nrf2-negative regulator Keap1. Cell Rep. 6, 1026–1036 (2014). This study describes the interaction of mVP24 with the cellular protein KEAP1 and demonstrates a role for the host antioxidant response in MARV pathogenesis in a mouse model.

    Article  CAS  PubMed  Google Scholar 

  130. Copple, I. M. The keap1–nrf2 cell defense pathway — a promising therapeutic target? Adv. Pharmacol. 63, 43–79 (2012).

    Article  CAS  PubMed  Google Scholar 

  131. Zhang, A. P., Bornholdt, Z. A., Abelson, D. M. & Saphire, E. O. Crystal structure of Marburg virus VP24. J. Virol. 88, 5859–5863 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  132. Mateo, M. et al. VP24 is a molecular determinant of Ebola virus virulence in guinea pigs. J. Infect. Dis. 204 (Suppl. 3), 1011–1020 (2011).

    Article  CAS  Google Scholar 

  133. Ebihara, H. et al. Molecular determinants of Ebola virus virulence in mice. PLoS Pathog. 2, e73 (2006).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  134. Warfield, K. L. et al. Development of a model for Marburgvirus based on severe-combined immunodeficiency mice. Virol. J. 4, 108 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  135. Lofts, L. L., Wells, J. B., Bavari, S. & Warfield, K. L. Key genomic changes necessary for an in vivo lethal mouse Marburgvirus variant selection process. J. Virol. 85, 3905–3917 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Valmas, C. & Basler, C. F. Marburg virus VP40 antagonizes interferon signaling in a species-specific manner. J. Virol. 85, 4309–4317 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Feagins, A. R. & Basler, C. F. Amino acid residue at position 79 of Marburg virus VP40 confers interferon antagonism in mouse cells. J. Infect. Dis. 212, 219–225 (2015).

    Article  CAS  Google Scholar 

  138. Feagins, A. R. & Basler, C. F. The VP40 protein of Marburg virus exhibits impaired budding and increased sensitivity to human tetherin following mouse adaptation. J. Virol. 88, 14440–14450 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  139. Feldmann, H., Slenczka, W. & Klenk, H. D. in Imported Virus Infections (eds Schwarz, T. F. & Siedl, G.) 77–100 (Springer, 1996).

    Book  Google Scholar 

  140. Slenczka, W. & Klenk, H. D. Forty years of Marburg virus. J. Infect. Dis. 196 (Suppl. 2), 131–135 (2007).

    Article  Google Scholar 

  141. Johnson, K. M., Lange, J. V., Webb, P. A. & Murphy, F. A. Isolation and partial characterisation of a new virus causing acute haemorrhagic fever in Zaire. Lancet 1, 569–571 (1977).

    Article  CAS  PubMed  Google Scholar 

  142. World Health Organization. Ebola haemorrhagic fever in Sudan, 1976. Bull. World Health Organ. 56, 247–270 (1978).

  143. Towner, J. S. et al. Marburgvirus genomics and association with a large hemorrhagic fever outbreak in Angola. J. Virol. 80, 6497–6516 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Bausch, D. G. et al. Marburg hemorrhagic fever associated with multiple genetic lineages of virus. New Engl. J. Med. 355, 909–919 (2006).

    Article  CAS  PubMed  Google Scholar 

  145. Jahrling, P. B. et al. Preliminary report: isolation of Ebola virus from monkeys imported to USA. Lancet 335, 502–505 (1990).

    Article  CAS  PubMed  Google Scholar 

  146. Formenty, P. et al. Human infection due to Ebola virus, subtype Cote d'Ivoire: clinical and biologic presentation. J. Infect. Dis. 179 (Suppl. 1), 48–53 (1999).

    Article  Google Scholar 

  147. Conrad, J. L. et al. Epidemiologic investigation of Marburg virus disease, Southern Africa, 1975. Am. J. Trop. Med. Hyg. 27, 1210–1215 (1978).

    Article  CAS  PubMed  Google Scholar 

  148. Swanepoel, R. et al. Studies of reservoir hosts for Marburg virus. Emerg. Infect. Dis. 13, 1847–1851 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Towner, J. S. et al. Isolation of genetically diverse Marburg viruses from Egyptian fruit bats. PLoS Pathog. 5, e1000536 (2009). This study describes the first isolation of infectious MARV from bats caught in the wild.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  150. Timen, A. et al. Response to imported case of Marburg hemorrhagic fever, the Netherlands. Emerg. Infect. Dis. 15, 1171–1175 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  151. Towner, J. S. et al. Marburg virus infection detected in a common African bat. PLoS ONE 2, e764 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  152. Amman, B. R. et al. Seasonal pulses of Marburg virus circulation in juvenile Rousettus aegyptiacus bats coincide with periods of increased risk of human infection. PLoS Pathog. 8, e1002877 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  153. Aleksandrowicz, P. et al. Ebola virus enters host cells by macropinocytosis and clathrin-mediated endocytosis. J. Infect. Dis. 204 (Suppl. 3), 957–967 (2011).

    Article  CAS  Google Scholar 

  154. Saeed, M. F., Kolokoltsov, A. A., Albrecht, T. & Davey, R. A. Cellular entry of Ebola virus involves uptake by a macropinocytosis-like mechanism and subsequent trafficking through early and late endosomes. PLoS Pathog. 6, e1001110 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  155. Nanbo, A. et al. Ebolavirus is internalized into host cells via macropinocytosis in a viral glycoprotein-dependent manner. PLoS Pathog. 6, e1001121 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  156. Sakurai, Y. et al. Two-pore channels control Ebola virus host cell entry and are drug targets for disease treatment. Science 347, 995–998 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Muhlberger, E. Filovirus replication and transcription. Future Virol. 2, 205–215 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  158. Jasenosky, L. D., Neumann, G., Lukashevich, I. & Kawaoka, Y. Ebola virus VP40-induced particle formation and association with the lipid bilayer. J. Virol. 75, 5205–5214 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Noda, T. et al. Ebola virus VP40 drives the formation of virus-like filamentous particles along with GP. J. Virol. 76, 4855–4865 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Harty, R. N., Brown, M. E., Wang, G., Huibregtse, J. & Hayes, F. P. A PPxY motif within the VP40 protein of Ebola virus interacts physically and functionally with a ubiquitin ligase: implications for filovirus budding. Proc. Natl Acad. Sci. USA 97, 13871–13876 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Hartlieb, B. & Weissenhorn, W. Filovirus assembly and budding. Virology 344, 64–70 (2006).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors thank C. Schwall for critical reading of this manuscript. The authors were supported by US National Institutes of Health (NIH) grants U19 AI109945 to C.F.B. (Basler- PI), I.M. and G.K.A.; R01 AI059536 and U19 AI109664 to C.F.B.; and U19 AI070489 (Holtzman- PI) and R01 AI081914 to G.K.A. The authors are also supported by the US Department of Defense and the Defense Threat Reduction Agency grants HDTRA1-12-1-0051 and HDTRA1-14-1-0013 to C.F.B. and G.K.A. The content of this Review does not necessarily reflect the position or the policy of the federal government, and no official endorsement should be inferred.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christopher F. Basler.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

PowerPoint slides

Glossary

Negative-strand RNA viruses

Viruses possessing single-stranded RNA genomes that are negative sense (that is, complementary to the viral mRNAs).

Haemorrhagic fevers

Severe illnesses typically associated with bleeding that are caused by various viral infections, including filovirus infections.

Index case

The first identified case of a particular disease or epidemic, or the case from which the disease initially progressed.

Type I interferon

A group that comprises interferon-β and multiple types of interferon-α.

Macropinocytosis

An actin-dependent form of endocytosis that results in the formation of large fluid-filled vacuoles.

Cathepsins

Endosomal cysteine proteases that cleave the filovirus glycoprotein during viral entry.

Phosphatidylserine

A phospholipid found on the surface of apoptotic bodies and on viral membranes that can promote intracellular uptake through interaction with specific cell surface receptors.

Lymphopenia

Low levels of lymphocytes in the blood.

Isotype class switching

A recombination process that alters the isotype of an antibody (for example, from immunoglobulin M to immunoglobulin G).

Necrosis

A non-apoptotic form of cell death.

Fibrin

A non-globular protein involved in blood clotting.

Extrinsic pathway of coagulation

The blood coagulation pathway that is initiated at the site of injury in response to tissue factor.

Thrombocytopenia

Low levels of platelets in the blood.

Pro-thrombin time

A blood test that measures clotting time.

Sumoylation

A post-translational modification that influences protein stability and other cellular processes.

Monopartite

Involving one part. In the case of nuclear localization signals, the recognized sequence is composed of a single stretch of amino acids.

Bipartite

Involving two parts. In the case of nuclear localization signals, the recognized sequence is composed of two separate amino acid sequences connected by a linker region.

293 cells

Human embryonic kidney-derived cell line.

Lipid rafts

Microdomains within a lipid membrane. It is postulated that lipid rafts support the concentration of viral proteins for more efficient viral assembly and budding.

Kelch

A repeated motif within proteins that predicts a conserved βpropeller structure.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Messaoudi, I., Amarasinghe, G. & Basler, C. Filovirus pathogenesis and immune evasion: insights from Ebola virus and Marburg virus. Nat Rev Microbiol 13, 663–676 (2015). https://doi.org/10.1038/nrmicro3524

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrmicro3524

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing