1932

Abstract

A central dogma of molecular biology is that the sequence of a protein dictates its particular fold and the fold dictates its function. Indeed, the sequence → structure → function hypothesis has been a guiding principle by which scientists approach molecular biology. Every student knows that the genome encodes information for the progression from primary sequence to secondary, tertiary, and ultimately quaternary structure. Yet with a growing number of proteins, a fifth level has been identified: rearrangement of existing structures into distinct forms. Recent observations indicate that replication of Ebola virus depends on this fifth level. We believe other viruses with compact genomes and rapid evolution under selective pressure will be a rich source of examples of polypeptides that rearrange to gain added functions. In this review, we describe mechanisms by which viral, prokaryotic, and eukaryotic polypeptides have adopted alternate structures to control or gain function.

Associated Article

There are media items related to this article:
More than Meets the Eye: Hidden Structures in the Proteome: Video 3

Associated Article

There are media items related to this article:
More than Meets the Eye: Hidden Structures in the Proteome: Video 2

Associated Article

There are media items related to this article:
More than Meets the Eye: Hidden Structures in the Proteome: Video 1
Loading

Article metrics loading...

/content/journals/10.1146/annurev-virology-100114-054923
2016-09-29
2024-03-28
Loading full text...

Full text loading...

/deliver/fulltext/virology/3/1/annurev-virology-100114-054923.html?itemId=/content/journals/10.1146/annurev-virology-100114-054923&mimeType=html&fmt=ahah

Literature Cited

  1. Elliott LH, Kiley MP, McCormick JB. 1.  1985. Descriptive analysis of Ebola virus proteins. Virology 147:169–76 [Google Scholar]
  2. Sanchez A, Kiley MP, Holloway BP, McCormick JB, Auperin DD. 2.  1993. Sequence analysis of the Ebola virus genome: organization, genetic elements, and comparison with the genome of Marburg virus. Virus Res 29:215–40 [Google Scholar]
  3. Jasenosky LD, Neumann G, Lukashevich I, Kawaoka Y. 3.  2001. Ebola virus VP40-induced particle formation and association with the lipid bilayer. J. Virol. 75:5205–14 [Google Scholar]
  4. Hoenen T, Jung S, Herwig A, Groseth A, Becker S. 4.  2010. Both matrix proteins of Ebola virus contribute to the regulation of viral genome replication and transcription. Virology 403:56–66 [Google Scholar]
  5. Valmas C, Basler CF. 5.  2011. Marburg virus VP40 antagonizes interferon signaling in a species-specific manner. J. Virol. 85:4309–17 [Google Scholar]
  6. Dessen A, Volchkov V, Dolnik O, Klenk HD, Weissenhorn W. 6.  2000. Crystal structure of the matrix protein VP40 from Ebola virus. EMBO J 19:4228–36 [Google Scholar]
  7. Gomis-Ruth FX, Dessen A, Timmins J, Bracher A, Kolesnikowa L. 7.  et al. 2003. The matrix protein VP40 from Ebola virus octamerizes into pore-like structures with specific RNA binding properties. Structure 11:423–33 [Google Scholar]
  8. Timmins J, Schoehn G, Kohlhaas C, Klenk HD, Ruigrok RW, Weissenhorn W. 8.  2003. Oligomerization and polymerization of the filovirus matrix protein VP40. Virology 312:359–68 [Google Scholar]
  9. Bornholdt ZA, Noda T, Abelson DM, Halfmann P, Wood MR. 9.  et al. 2013. Structural rearrangement of Ebola virus VP40 begets multiple functions in the virus life cycle. Cell 154:763–74 [Google Scholar]
  10. Ruigrok RW, Schoehn G, Dessen A, Forest E, Volchkov V. 10.  et al. 2000. Structural characterization and membrane binding properties of the matrix protein VP40 of Ebola virus. J. Mol. Biol. 300:103–12 [Google Scholar]
  11. Scianimanico S, Schoehn G, Timmins J, Ruigrok RH, Klenk HD, Weissenhorn W. 11.  2000. Membrane association induces a conformational change in the Ebola virus matrix protein. EMBO J 19:6732–41 [Google Scholar]
  12. Zschornig O, Arnold K, Ohki S. 12.  1993. Effect of glycosaminoglycans and PEG on fusion of Sendai virus with phosphatidylserine vesicles. Biochim. Biophys. Acta 1148:1–6 [Google Scholar]
  13. Timmins J, Scianimanico S, Schoehn G, Weissenhorn W. 13.  2001. Vesicular release of Ebola virus matrix protein VP40. Virology 283:1–6 [Google Scholar]
  14. Noda T, Sagara H, Suzuki E, Takada A, Kida H, Kawaoka Y. 14.  2002. Ebola virus VP40 drives the formation of virus-like filamentous particles along with GP. J. Virol. 76:4855–65 [Google Scholar]
  15. Hoenen T, Biedenkopf N, Zielecki F, Jung S, Groseth A. 15.  et al. 2010. Oligomerization of Ebola virus VP40 is essential for particle morphogenesis and regulation of viral transcription. J. Virol. 84:7053–63 [Google Scholar]
  16. Nanbo A, Watanabe S, Halfmann P, Kawaoka Y. 16.  2013. The spatio-temporal distribution dynamics of Ebola virus proteins and RNA in infected cells. Sci. Rep. 3:1206 [Google Scholar]
  17. Oda S, Noda T, Wijesinghe KJ, Halfmann P, Bornholdt ZA. 17.  et al. 2015. Crystal structure of Marburg virus VP40 reveals a broad, basic patch for matrix assembly and a requirement of the N-terminal domain for immunosuppression. J. Virol. 90:1839–48 [Google Scholar]
  18. Valmas C, Grosch MN, Schumann M, Olejnik J, Martinez O. 18.  et al. 2010. Marburg virus evades interferon responses by a mechanism distinct from Ebola virus. PLOS Pathog 6:e1000721 [Google Scholar]
  19. Murzin AG. 19.  2008. Metamorphic proteins. Science 320:1725–26 [Google Scholar]
  20. Jaffe EK, Lawrence SH. 20.  2012. The morpheein model of allostery: evaluating proteins as potential morpheeins. Methods Mol. Biol. 796:217–31 [Google Scholar]
  21. Jeffery CJ. 21.  2004. Molecular mechanisms for multitasking: recent crystal structures of moonlighting proteins. Curr. Opin. Struct. Biol. 14:663–68 [Google Scholar]
  22. Jaakola VP, Griffith MT, Hanson MA, Cherezov V, Chien EY. 22.  et al. 2008. The 2.6 angstrom crystal structure of a human A2A adenosine receptor bound to an antagonist. Science 322:1211–17 [Google Scholar]
  23. Hanson MA, Cherezov V, Griffith MT, Roth CB, Jaakola VP. 23.  et al. 2008. A specific cholesterol binding site is established by the 2.8 Å structure of the human β2-adrenergic receptor. Structure 16:897–905 [Google Scholar]
  24. Wynne SA, Crowther RA, Leslie AG. 24.  1999. The crystal structure of the human hepatitis B virus capsid. Mol. Cell 3:771–80 [Google Scholar]
  25. Packianathan C, Katen SP, Dann CE III, Zlotnick A. 25.  2010. Conformational changes in the hepatitis B virus core protein are consistent with a role for allostery in virus assembly. J. Virol. 84:1607–15 [Google Scholar]
  26. Zlotnick A, Tan Z, Selzer L. 26.  2013. One protein, at least three structures, and many functions. Structure 21:6–8 [Google Scholar]
  27. DiMattia MA, Watts NR, Stahl SJ, Grimes JM, Steven AC. 27.  et al. 2013. Antigenic switching of hepatitis B virus by alternative dimerization of the capsid protein. Structure 21:133–42 [Google Scholar]
  28. Elgouhari HM, Abu-Rajab Tamimi TI, Carey WD. 28.  2008. Hepatitis B virus infection: understanding its epidemiology, course, and diagnosis. Clevel. Clin. J. Med. 75:881–89 [Google Scholar]
  29. Lopez-Pelegrin M, Ksiazek M, Karim AY, Guevara T, Arolas JL. 29.  et al. 2015. A novel mechanism of latency in matrix metalloproteinases. J. Biol. Chem. 290:4728–40 [Google Scholar]
  30. Chang YG, Cohen SE, Phong C, Myers WK, Kim YI. 30.  et al. 2015. A protein fold switch joins the circadian oscillator to clock output in cyanobacteria. Science 349:324–28 [Google Scholar]
  31. Prusiner SB. 31.  1982. Novel proteinaceous infectious particles cause scrapie. Science 216:136–44 [Google Scholar]
  32. Holmes BB, Diamond MI. 32.  2012. Cellular mechanisms of protein aggregate propagation. Curr. Opin. Neurol. 25:721–26 [Google Scholar]
  33. Prusiner SB. 33.  2012. A unifying role for prions in neurodegenerative diseases. Science 336:1511–13 [Google Scholar]
  34. Jucker M, Walker LC. 34.  2013. Self-propagation of pathogenic protein aggregates in neurodegenerative diseases. Nature 501:45–51 [Google Scholar]
  35. Suzuki G, Shimazu N, Tanaka M. 35.  2012. A yeast prion, Mod5, promotes acquired drug resistance and cell survival under environmental stress. Science 336:355–59 [Google Scholar]
  36. Kuloglu ES, McCaslin DR, Markley JL, Volkman BF. 36.  2002. Structural rearrangement of human lymphotactin, a C chemokine, under physiological solution conditions. J. Biol. Chem. 277:17863–70 [Google Scholar]
  37. Kuloglu ES, McCaslin DR, Kitabwalla M, Pauza CD, Markley JL, Volkman BF. 37.  2001. Monomeric solution structure of the prototypical ‘C’ chemokine lymphotactin. Biochemistry 40:12486–96 [Google Scholar]
  38. Kelner GS, Kennedy J, Bacon KB, Kleyensteuber S, Largaespada DA. 38.  et al. 1994. Lymphotactin: a cytokine that represents a new class of chemokine. Science 266:1395–99 [Google Scholar]
  39. Tuinstra RL, Peterson FC, Elgin ES, Pelzek AJ, Volkman BF. 39.  2007. An engineered second disulfide bond restricts lymphotactin/XCL1 to a chemokine-like conformation with XCR1 agonist activity. Biochemistry 46:2564–73 [Google Scholar]
  40. Tuinstra RL, Peterson FC, Kutlesa S, Elgin ES, Kron MA, Volkman BF. 40.  2008. Interconversion between two unrelated protein folds in the lymphotactin native state. PNAS 105:5057–62 [Google Scholar]
  41. Tyler RC, Murray NJ, Peterson FC, Volkman BF. 41.  2011. Native-state interconversion of a metamorphic protein requires global unfolding. Biochemistry 50:7077–79 [Google Scholar]
  42. Bornholdt ZA, Prasad BV. 42.  2008. X-ray structure of NS1 from a highly pathogenic H5N1 influenza virus. Nature 456:985–88 [Google Scholar]
  43. Carrillo B, Choi JM, Bornholdt ZA, Sankaran B, Rice AP, Prasad BV. 43.  2014. The influenza A virus protein NS1 displays structural polymorphism. J. Virol. 88:4113–22 [Google Scholar]
  44. Bornholdt ZA, Prasad BV. 44.  2006. X-ray structure of influenza virus NS1 effector domain. Nat. Struct. Mol. Biol. 13:559–60 [Google Scholar]
  45. Hale BG, Barclay WS, Randall RE, Russell RJ. 45.  2008. Structure of an avian influenza A virus NS1 protein effector domain. Virology 378:1–5 [Google Scholar]
  46. Kerry PS, Ayllon J, Taylor MA, Hass C, Lewis A. 46.  et al. 2011. A transient homotypic interaction model for the influenza A virus NS1 protein effector domain. PLOS ONE 6:e17946 [Google Scholar]
  47. Kerry PS, Long E, Taylor MA, Russell RJ. 47.  2011. Conservation of a crystallographic interface suggests a role for β-sheet augmentation in influenza virus NS1 multifunctionality. Acta Crystallogr. F 67:858–61 [Google Scholar]
  48. Xia S, Monzingo AF, Robertus JD. 48.  2009. Structure of NS1A effector domain from the influenza A/Udorn/72 virus. Acta Crystallogr. D 65:11–17 [Google Scholar]
  49. Aramini JM, Hamilton K, Ma LC, Swapna GV, Leonard PG. 49.  et al. 2014. 19F NMR reveals multiple conformations at the dimer interface of the nonstructural protein 1 effector domain from influenza A virus. Structure 22:515–25 [Google Scholar]
  50. Aramini JM, Ma LC, Zhou L, Schauder CM, Hamilton K. 50.  et al. 2011. Dimer interface of the effector domain of non-structural protein 1 from influenza A virus: an interface with multiple functions. J. Biol. Chem. 286:26050–60 [Google Scholar]
  51. Xia S, Robertus JD. 51.  2010. X-ray structures of NS1 effector domain mutants. Arch. Biochem. Biophys. 494:198–204 [Google Scholar]
  52. Hale BG. 52.  2014. Conformational plasticity of the influenza A virus NS1 protein. J. Gen. Virol. 95:2099–105 [Google Scholar]
  53. Das K, Ma LC, Xiao R, Radvansky B, Aramini J. 53.  et al. 2008. Structural basis for suppression of a host antiviral response by influenza A virus. PNAS 105:13093–98 [Google Scholar]
  54. Noah DL, Twu KY, Krug RM. 54.  2003. Cellular antiviral responses against influenza A virus are countered at the posttranscriptional level by the viral NS1A protein via its binding to a cellular protein required for the 3′ end processing of cellular pre-mRNAS. Virology 307:386–95 [Google Scholar]
  55. Hale BG, Batty IH, Downes CP, Randall RE. 55.  2008. Binding of influenza A virus NS1 protein to the inter-SH2 domain of p85 suggests a novel mechanism for phosphoinositide 3-kinase activation. J. Biol. Chem. 283:1372–80 [Google Scholar]
  56. Neumann P, Lieber D, Meyer S, Dautel P, Kerth A. 56.  et al. 2009. Crystal structure of the Borna disease virus matrix protein (BDV-M) reveals ssRNA binding properties. PNAS 106:3710–15 [Google Scholar]
  57. Battisti AJ, Meng G, Winkler DC, McGinnes LW, Plevka P. 57.  et al. 2012. Structure and assembly of a paramyxovirus matrix protein. PNAS 109:13996–4000 [Google Scholar]
  58. Leyrat C, Renner M, Harlos K, Huiskonen JT, Grimes JM. 58.  2014. Structure and self-assembly of the calcium binding matrix protein of human metapneumovirus. Structure 22:136–48 [Google Scholar]
  59. Gaudier M, Gaudin Y, Knossow M. 59.  2002. Crystal structure of vesicular stomatitis virus matrix protein. EMBO J 21:2886–92 [Google Scholar]
  60. Graham SC, Assenberg R, Delmas O, Verma A, Gholami A. 60.  et al. 2008. Rhabdovirus matrix protein structures reveal a novel mode of self-association. PLOS Pathog 4:e1000251 [Google Scholar]
  61. Money VA, McPhee HK, Mosely JA, Sanderson JM, Yeo RP. 61.  2009. Surface features of a Mononegavirales matrix protein indicate sites of membrane interaction. PNAS 106:4441–46 [Google Scholar]
  62. Forster A, Maertens GN, Farrell PJ, Bajorek M. 62.  2015. Dimerization of matrix protein is required for budding of respiratory syncytial virus. J. Virol. 89:4624–35 [Google Scholar]
  63. Balch WE, Dunphy WG, Braell WA, Rothman JE. 63.  1984. Reconstitution of the transport of protein between successive compartments of the Golgi measured by the coupled incorporation of N-acetylglucosamine. Cell 39:405–16 [Google Scholar]
  64. Nishimura N, Balch WE. 64.  1997. A di-acidic signal required for selective export from the endoplasmic reticulum. Science 277:556–58 [Google Scholar]
  65. Lane DP, Crawford LV. 65.  1979. T antigen is bound to a host protein in SV40-transformed cells. Nature 278:261–63 [Google Scholar]
  66. Melero JA, Stitt DT, Mangel WF, Carroll RB. 66.  1979. Identification of new polypeptide species (48–55K) immunoprecipitable by antiserum to purified large T antigen and present in SV40-infected and -transformed cells. Virology 93:466–80 [Google Scholar]
  67. Linzer DI, Levine AJ. 67.  1979. Characterization of a 54K dalton cellular SV40 tumor antigen present in SV40-transformed cells and uninfected embryonal carcinoma cells. Cell 17:43–52 [Google Scholar]
  68. Kress M, May E, Cassingena R, May P. 68.  1979. Simian virus 40-transformed cells express new species of proteins precipitable by anti-simian virus 40 tumor serum. J. Virol. 31:472–83 [Google Scholar]
  69. Chang C, Martin RG, Livingston DM, Luborsky SW, Hu CP, Mora PT. 69.  1979. Relationship between T-antigen and tumor-specific transplantation antigen in simian virus 40-transformed cells. J. Virol. 29:69–75 [Google Scholar]
  70. Zinkernagel RM, Doherty PC. 70.  1973. Cytotoxic thymus-derived lymphocytes in cerebrospinal fluid of mice with lymphocytic choriomeningitis. J. Exp. Med. 138:1266–69 [Google Scholar]
  71. Oldstone MB, Dixon FJ, Mitchell GF, McDevitt HO. 71.  1973. Histocompatibility-linked genetic control of disease susceptibility: murine lymphocytic choriomeningitis virus infection. J. Exp. Med. 137:1201–12 [Google Scholar]
  72. Doherty PC, Zinkernagel RM. 72.  1975. H-2 compatibility is required for T-cell-mediated lysis of target cells infected with lymphocytic choriomeningitis virus. J. Exp. Med. 141:502–7 [Google Scholar]
  73. Zinkernagel RM, Dunlop MB, Blanden RV, Doherty PC, Shreffler DC. 73.  1976. H-2 compatibility requirement for virus-specific T-cell-mediated cytolysis: evaluation of the role of H-2I region and non-H-2 genes in regulating immune response. J. Exp. Med. 144:519–32 [Google Scholar]
  74. Rice G, Stedman K, Snyder J, Wiedenheft B, Willits D. 74.  et al. 2001. Viruses from extreme thermal environments. PNAS 98:13341–45 [Google Scholar]
  75. DiMaio F, Yu X, Rensen E, Krupovic M, Prangishvili D, Egelman EH. 75.  2015. A virus that infects a hyperthermophile encapsidates A-form DNA. Science 348:914–17 [Google Scholar]
  76. Speir JA, Munshi S, Wang G, Baker TS, Johnson JE. 76.  1995. Structures of the native and swollen forms of cowpea chlorotic mottle virus determined by X-ray crystallography and cryo-electron microscopy. Structure 3:63–78 [Google Scholar]
  77. Singh P, Gonzalez MJ, Manchester M. 77.  2006. Viruses and their use in nanotechnology. Drug Dev. Res. 67:23–41 [Google Scholar]
  78. Dessen A, Forest E, Volchkov V, Dolnik O, Klenk HD, Weissenhorn W. 78.  2000. Crystallization and preliminary X-ray analysis of the matrix protein from Ebola virus. Acta Crystallogr. D 56:758–60 [Google Scholar]
/content/journals/10.1146/annurev-virology-100114-054923
Loading
/content/journals/10.1146/annurev-virology-100114-054923
Loading

Data & Media loading...

Supplemental Material

    Morphs modeling how the VP40 dimer might transition to the matrix-assembling linear hexamer (this video) and the RNA-binding octameric ring (). In each of these videos, N-terminal domains are blue and C-terminal domains are orange. In this video, construction of the hexamer form the dimer, it is modeled that three VP40 dimers gather at the membrane surface. It is hypothesized that an electrostatic interaction with lipid head groups triggers conformational rearrangement and assembly of the hexamer. Note that in the crystal structure of this hexamer (PDB: 4LDD), the central VP40 is upside down. In , the crystal structure of the octameric ring (PDB: 1H2D; Gomis-Ruth et al. 2003, 11:423–33) reveals a 3-nt RNA bound to each N-terminal domain. RNA is illustrated here as the trigger that begins the conformational rearrangement. However, it is not yet known what the trigger of this rearrangement is. We thank Dr. Zachary Bornholdt for the use of these models and videos, originally published as supplemental figures in Bornholdt et al. 2013, 154:763–74.

    Morphs modeling how the VP40 dimer might transition to the matrix-assembling linear hexamer () and the RNA-binding octameric ring (this video). In each of these videos, N-terminal domains are blue and C-terminal domains are orange. In , construction of the hexamer form the dimer, it is modeled that three VP40 dimers gather at the membrane surface. It is hypothesized that an electrostatic interaction with lipid head groups triggers conformational rearrangement and assembly of the hexamer. Note that in the crystal structure of this hexamer (PDB: 4LDD), the central VP40 is upside down. In this video, the crystal structure of the octameric ring (PDB: 1H2D; Gomis-Ruth et al. 2003, 11:423–33) reveals a 3-nt RNA bound to each N-terminal domain. RNA is illustrated here as the trigger that begins the conformational rearrangement. However, it is not yet known what the trigger of this rearrangement is. We thank Dr. Zachary Bornholdt for the use of these models and videos, originally published as supplemental figures in Bornholdt et al. 2013, 154:763–74.

    A modeled morph between the two folds of lymphotactin. Residues 61–70 have been modeled into the final structure to facilitate the final morph and allow the complete α-helix to be included. The beginning structure is the originally disordered Ltn10 monomeric form that functions as a chemokine (PDB: 1J8I). The final structure is one protomer of the biologically active dimer, Ltn40, which functions in chemotaxis (PDB: 1J8I).

  • 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