全部 标题 作者
关键词 摘要

OALib Journal期刊
ISSN: 2333-9721
费用:99美元

查看量下载量

相关文章

更多...

Productive Entry Pathways of Human Rhinoviruses

DOI: 10.1155/2012/826301

Full-Text   Cite this paper   Add to My Lib

Abstract:

Currently, complete or partial genome sequences of more than 150 human rhinovirus (HRV) isolates are known. Twelve species A use members of the low-density lipoprotein receptor family for cell entry, whereas the remaining HRV-A and all HRV-B bind ICAM-1. HRV-Cs exploit an unknown receptor. At least all A and B type viruses depend on receptor-mediated endocytosis for infection. In HeLa cells, they are internalized mainly by a clathrin- and dynamin-dependent mechanism. Upon uptake into acidic compartments, the icosahedral HRV capsid expands by ~4% and holes open at the 2-fold axes, close to the pseudo-3-fold axes and at the base of the star-shaped dome protruding at the vertices. RNA-protein interactions are broken and new ones are established, the small internal myristoylated capsid protein VP4 is expelled, and amphipathic N-terminal sequences of VP1 become exposed. The now hydrophobic subviral particle attaches to the inner surface of endosomes and transfers its genomic (+) ssRNA into the cytosol. The RNA leaves the virus starting with the poly(A) tail at its 3′-end and passes through a membrane pore contiguous with one of the holes in the capsid wall. Alternatively, the endosome is disrupted and the RNA freely diffuses into the cytoplasm. 1. Introduction Human rhinoviruses (HRVs) are icosahedral (30?nm in diameter) and nonenveloped with a (+) ssRNA genome of ~7100 bases. Belonging to the family Picornaviridae, genus Enterovirus, they are composed of 60 copies each of four capsid proteins, VP1 to VP4. In 1987, HRVs from clinical samples were serotyped into 100 strains [1]. Recently, complete genome sequences of all known HRVs were determined. Phylogenetic analyses grouped them into 3 species; 74 HRV-A, 25 HRV-B, and 6 HRV-C [2]. Since then, many more rhinoviruses (mostly of type C) were identified in clinical specimens [3–5]. Independent from this classification, HRV-A and HRV-B are divided into two groups based upon the receptors exploited for host cell attachment; the minor receptor group, including the so far identified 12 HRV-A, bind low-density lipoprotein receptor (LDLR), very-LDLR (VLDLR), and LDLR-related protein 1 (LRP1) [6–9], while the remaining HRVA and HRV-B (constituting the majority, that is, the major group) use intercellular adhesion molecule 1 (ICAM-1) for cell entry [10]. Some major group HRVs (HRV8, 54, and 89) can also use heparan sulfate proteoglycans (HSPG) as an additional receptor [7, 11, 12] either as wild type (wt) or after adaptation to grow in cells lacking ICAM-1. This is achieved by numerous cycles alternating between

References

[1]  V. V. Hamparian, R. J. Colonno, M. K. Cooney et al., “A collaborative report: rhinoviruses-extension of the numbering system from 89 to 100,” Virology, vol. 159, no. 1, pp. 191–192, 1987.
[2]  A. C. Palmenberg, D. Spiro, R. Kuzmickas et al., “Sequencing and analyses of all known human rhinovirus genomes reveal structure and evolution,” Science, vol. 324, no. 5923, pp. 55–59, 2009.
[3]  K. E. Arden and I. M. Mackay, “Newly identified human rhinoviruses: molecular methods heat up the cold viruses,” Reviews in Medical Virology, vol. 20, no. 3, pp. 156–176, 2010.
[4]  H. Harvala, C. L. McIntyre, N. J. McLeish, J. Kondracka, et al., “High detection frequency and viral loads of human rhinovirus species A to C in fecal samples, diagnostic and clinical implications,” Journal of Medical Virology, vol. 84, pp. 536–542, 2012.
[5]  P. Simmonds, C. McIntyre, C. Savolainen-Kopra, C. Tapparel, I. M. Mackay, and T. Hovi, “Proposals for the classification of human rhinovirus species C into genotypically assigned types,” Journal of General Virology, vol. 91, no. 10, pp. 2409–2419, 2010.
[6]  F. Hofer, M. Gruenberger, H. Kowalski et al., “Members of the low density lipoprotein receptor family mediate cell entry of a minor-group common cold virus,” Proceedings of the National Academy of Sciences of the United States of America, vol. 91, no. 5, pp. 1839–1842, 1994.
[7]  M. Vlasak, M. Roivainen, M. Reithmayer et al., “The minor receptor group of human rhinovirus (HRV) includes HRV23 and HRV25, but the presence of a lysine in the VP1 HI loop is not sufficient for receptor binding,” Journal of Virology, vol. 79, no. 12, pp. 7389–7395, 2005.
[8]  T. C. Marlovits, C. Abrahamsberg, and D. Blaas, “Very-low-density lipoprotein receptor fragment shed from HeLa cells inhibits human rhinovirus infection,” Journal of Virology, vol. 72, no. 12, pp. 10246–10250, 1998.
[9]  T. C. Marlovits, T. Zechmeister, M. Gruenberger, B. Ronacher, H. Schwihla, and D. Blaas, “Recombinant soluble low density lipoprotein receptor fragment inhibits minor group rhinovirus infection in vitro,” FASEB Journal, vol. 12, no. 9, pp. 695–703, 1998.
[10]  C. R. Uncapher, C. M. DeWitt, and R. J. Colonno, “The major and minor group receptor families contain all but one human rhinovirus serotype,” Virology, vol. 180, no. 2, pp. 814–817, 1991.
[11]  A. G. Khan, J. Pichler, A. Rosemann, and D. Blaas, “Human rhinovirus type 54 infection via heparan sulfate is less efficient and strictly dependent on low endosomal pH,” Journal of Virology, vol. 81, no. 9, pp. 4625–4632, 2007.
[12]  A. G. Khan, A. Pickl-Herk, L. Gajdzik, T. C. Marlovits, R. Fuchs, and D. Blaas, “Entry of a heparan sulphate-binding HRV8 variant strictly depends on dynamin but not on clathrin, caveolin, and flotillin,” Virology, vol. 412, no. 1, pp. 55–67, 2011.
[13]  M. Vlasak, I. Goesler, and D. Blaas, “Human rhinovirus type 89 variants use heparan sulfate proteoglycan for cell attachment,” Journal of Virology, vol. 79, no. 10, pp. 5963–5970, 2005.
[14]  A. Reischl, M. Reithmayer, G. Winsauer, R. Moser, I. G?sler, and D. Blaas, “Viral evolution toward change in receptor usage: adaptation of a major group human rhinovirus to grow in ICAM-1-negative cells,” Journal of Virology, vol. 75, no. 19, pp. 9312–9319, 2001.
[15]  Y. A. Bochkov, A. C. Palmenberg, W. M. Lee et al., “Molecular modeling, organ culture and reverse genetics for a newly identified human rhinovirus C,” Nature Medicine, vol. 17, no. 5, pp. 627–632, 2011.
[16]  Y. A. Bochkov and J. E. Gern, “Clinical and molecular features of human rhinovirus C,” Microbes Infect, vol. 14, pp. 485–494, 2012.
[17]  T. Heikkinen and A. J?rvinen, “The common cold,” The Lancet, vol. 361, no. 9351, pp. 51–59, 2003.
[18]  G. Waterer and R. Wunderink, “Respiratory infections: a current and future threat Series,” Respirology, vol. 14, no. 5, pp. 651–655, 2009.
[19]  A. M. Fendrick, “Viral respiratory infections due to rhinoviruses: current knowledge, new developments,” American Journal of Therapeutics, vol. 10, no. 3, pp. 193–202, 2003.
[20]  E. Farkas, “Rhinitis as a viral disease,” Ther Hung, vol. 5, pp. 3–6, 1957.
[21]  V. Peltola, M. Waris, R. ?sterback, P. Susi, T. Hyypi?, and O. Ruuskanen, “Clinical effects of rhinovirus infections,” Journal of Clinical Virology, vol. 43, no. 4, pp. 411–414, 2008.
[22]  K. B. Weiss and S. D. Sullivan, “The health economics of asthma and rhinitis. I. Assessing the economic impact,” Journal of Allergy and Clinical Immunology, vol. 107, no. 1, pp. 3–8, 2001.
[23]  N. Fuji, A. Suzuki, S. Lupisan, L. Sombrero, et al., “Detection of human rhinovirus C viral genome in blood among children with severe respiratory infections in the Philippines,” PLoS One, vol. 6, article e27247, 2011.
[24]  C. Tapparel, A. G. L'Huillier, A. L. Rougemont, M. Beghetti, C. Barazzone-Argiroffo, and L. Kaiser, “Pneumonia and pericarditis in a child with HRV-C infection: a case report,” Journal of Clinical Virology, vol. 45, no. 2, pp. 157–160, 2009.
[25]  M. K. Cooney and G. E. Kenny, “Reciprocal neutralizing cross-reaction between rhinovirus types 9 and 32,” Journal of Immunology, vol. 105, no. 2, pp. 531–533, 1970.
[26]  M. K. Cooney and G. E. Kenny, “Demonstration of dual rhinovirus infection in humans by isolation of different serotypes in human heteroploid (HeLa) and human diploid fibroblast cell cultures,” Journal of Clinical Microbiology, vol. 5, no. 2, pp. 202–207, 1977.
[27]  M. K. Cooney, G. E. Kenny, R. Tam, and J. P. Fox, “Cross relationships among 37 rhinoviruses demonstrated by virus neutralization with potent monotypic rabbit antisera,” Infection and Immunity, vol. 7, no. 3, pp. 335–340, 1973.
[28]  J. Edlmayr, K. Niespodziana, T. Popow-Kraupp et al., “Antibodies induced with recombinant VP1 from human rhinovirus exhibit cross-neutralisation,” European Respiratory Journal, vol. 37, no. 1, pp. 44–52, 2011.
[29]  U. Katpally, T. M. Fu, D. C. Freed, D. R. Casimiro, and T. J. Smith, “Antibodies to the buried N terminus of rhinovirus VP4 exhibit cross-serotypic neutralization,” Journal of Virology, vol. 83, no. 14, pp. 7040–7048, 2009.
[30]  A. M. De Palma, I. Vliegen, E. De Clercq, and J. Neyts, “Selective inhibitors of picornavirus replication,” Medicinal Research Reviews, vol. 28, no. 6, pp. 823–884, 2008.
[31]  K. Lonberg-Holm and B. D. Korant, “Early interaction of rhinoviruses with host cells,” Journal of Virology, vol. 9, no. 1, pp. 29–40, 1972.
[32]  S. Curry, M. Chow, and J. M. Hogle, “The poliovirus 135S particle is infectious,” Journal of Virology, vol. 70, no. 10, pp. 7125–7131, 1996.
[33]  J. Noble and K. Lonberg-Holm, “Interactions of components of human rhinovirus type 2 with hela cells,” Virology, vol. 51, no. 2, pp. 270–278, 1973.
[34]  M. R. MacNaughton, “The structure and replication of Rhinoviruses,” Current Topics in Microbiology and Immunology, vol. 97, pp. 1–26, 1982.
[35]  W. M. Lee, S. S. Monroe, and R. R. Rueckert, “Role of maturation cleavage in infectivity of picornaviruses: activation of an infectosome,” Journal of Virology, vol. 67, no. 4, pp. 2110–2122, 1993.
[36]  A. M. Hopkins, A. W. Baird, and A. Nusrat, “ICAM-1: targeted docking for exogenous as well as endogenous ligands,” Advanced Drug Delivery Reviews, vol. 56, no. 6, pp. 763–778, 2004.
[37]  R. Rothlein, M. L. Dustin, S. D. Marlin, and T. A. Springer, “A human intercellular adhesion molecule (ICAM-1) distinct from LFA-1,” Journal of Immunology, vol. 137, no. 4, pp. 1270–1274, 1986.
[38]  L. Xing, J. M. Casasnovas, and R. H. Cheng, “Structural analysis of human rhinovirus complexed with ICAM-1 reveals the dynamics of receptor-mediated virus uncoating,” Journal of Virology, vol. 77, no. 11, pp. 6101–6107, 2003.
[39]  L. Xing, K. Tjarnlund, B. Lindqvist et al., “Distinct cellular receptor interactions in poliovirus and rhinoviruses,” EMBO Journal, vol. 19, no. 6, pp. 1207–1216, 2000.
[40]  P. R. Kolatkar, J. Bella, N. H. Olson, C. M. Bator, T. S. Baker, and M. G. Rossmann, “Structural studies of two rhinovirus serotypes complexed with fragments of their cellular receptor,” EMBO Journal, vol. 18, no. 22, pp. 6249–6259, 1999.
[41]  N. H. Olson, P. R. Kolatkar, M. A. Oliveira, R. H. Cheng, et al., “Structure of a human rhinovirus complexed with its receptor molecule,” Proceedings of the National Academy of Sciences of the United States of America, vol. 90, pp. 507–511, 1993.
[42]  G. Nurani, B. Lindqvist, and J. M. Casasnovas, “Receptor Priming of Major Group Human Rhinoviruses for Uncoating and Entry at Mild Low-pH Environments,” Journal of Virology, vol. 77, no. 22, pp. 11985–11991, 2003.
[43]  J. M. Greve, C. P. Forte, C. W. Marlor et al., “Mechanisms of receptor-mediated rhinovirus neutralization defined by two soluble forms of ICAM-1,” Journal of Virology, vol. 65, no. 11, pp. 6015–6023, 1991.
[44]  H. Hoover-Litty and J. M. Greve, “Formation of rhinovirus-soluble ICAM-1 complexes and conformational changes in the virion,” Journal of Virology, vol. 67, no. 1, pp. 390–397, 1993.
[45]  V. R. Racaniello, “Early events in poliovirus infection: virus-receptor interactions,” Proceedings of the National Academy of Sciences of the United States of America, vol. 93, no. 21, pp. 11378–11381, 1996.
[46]  V. F. Chan and F. L. Black, “Uncoating of poliovirus by isolated plasma membranes,” Journal of Virology, vol. 5, no. 3, pp. 309–312, 1970.
[47]  J. De Sena, W. R. Heifner, and N. S. Stolov, “Studies on the in vitro uncoating of poliovirus. IV. Characteristics of solubilized membrane-modifying and stabilizing factors,” Virology, vol. 128, no. 2, pp. 354–365, 1983.
[48]  J. De Sena and B. Mandel, “Studies on the in vitro uncoating of poliovirus. I. Characterization of the modifying factor and the modifying reaction,” Virology, vol. 70, no. 2, pp. 470–483, 1976.
[49]  M. Gromeier and K. Wetz, “Kinetics of poliovirus uncoating in HeLa cells in a nonacidic environment,” Journal of Virology, vol. 64, no. 8, pp. 3590–3597, 1990.
[50]  W. J. Schneider and J. Nimpf, “LDL receptor relatives at the crossroad of endocytosis and signaling,” Cellular and Molecular Life Sciences, vol. 60, no. 5, pp. 892–903, 2003.
[51]  D. W. Russell, W. J. Schneider, T. Yamamoto, K. L. Luskey, M. S. Brown, and J. L. Goldstein, “Domain map of the LDL receptor: sequence homology with the epidermal growth factor precursor,” Cell, vol. 37, no. 2, pp. 577–585, 1984.
[52]  G. Rudenko, L. Henry, K. Henderson et al., “Structure of the LDL receptor extracellular domain at endosomal pH,” Science, vol. 298, no. 5602, pp. 2353–2358, 2002.
[53]  D. K. Strickland, M. Z. Kounnas, S. E. Williams, and W. S. Argraves, “LDL receptor-related protein (LRP): a multiligand receptor,” Fibrinolysis, vol. 8, no. 1, pp. 204–215, 1994.
[54]  M. M. Hussain, D. K. Strickland, and A. Bakillah, “The mammalian low-density lipoprotein receptor family,” Annual Review of Nutrition, vol. 19, pp. 141–172, 1999.
[55]  R. Ouda, K. Onomoto, K. Takahasi et al., “Retinoic acid-inducible gene I-inducible miR-23b inhibits infections by minor group rhinoviruses through down-regulation of the very low density lipoprotein receptor,” Journal of Biological Chemistry, vol. 286, no. 29, pp. 26210–26219, 2011.
[56]  J. Querol-Audí, T. Konecsni, J. Pous et al., “Minor group human rhinovirus-receptor interactions: geometry of multimodular attachment and basis of recognition,” FEBS Letters, vol. 583, no. 1, pp. 235–240, 2009.
[57]  N. Verdaguer, I. Fita, M. Reithmayer, R. Moser, and D. Blaas, “X-ray structure of a minor group human rhinovirus bound to a fragment of its cellular receptor protein,” Nature Structural and Molecular Biology, vol. 11, no. 5, pp. 429–434, 2004.
[58]  E. A. Hewat, E. Neumann, J. F. Conway et al., “The cellular receptor to human rhianovirus 2 binds around the 5-fold axis and not in the canyon: a structural view,” EMBO Journal, vol. 19, no. 23, pp. 6317–6325, 2000.
[59]  E. Neumann, R. Moser, L. Snyers, D. Blaas, and E. A. Hewat, “A cellular receptor of human rhinovirus type 2, the very-low-density lipoprotein receptor, binds to two neighboring proteins of the viral capsid,” Journal of Virology, vol. 77, no. 15, pp. 8504–8511, 2003.
[60]  T. Konecsni, U. Berka, A. Pickl-Herk et al., “Low pH-triggered beta-propeller switch of the low-density lipoprotein receptor assists rhinovirus infection,” Journal of Virology, vol. 83, no. 21, pp. 10922–10930, 2009.
[61]  A. Nicodemou, M. Petsch, T. Konecsni et al., “Rhinovirus-stabilizing activity of artificial VLDL-receptor variants defines a new mechanism for virus neutralization by soluble receptors,” FEBS Letters, vol. 579, no. 25, pp. 5507–5511, 2005.
[62]  B. Winther, “Rhinovirus infections in the upper airway,” Proceedings of the American Thoracic Society, vol. 8, no. 1, pp. 79–89, 2011.
[63]  A. G. Mosser, R. Brockman-Schneider, S. Amineva et al., “Similar frequency of rhinovirus-infectible cells in upper and lower airway epithelium,” Journal of Infectious Diseases, vol. 185, no. 6, pp. 734–743, 2002.
[64]  N. G. Papadopoulos, P. J. Bates, P. G. Bardin et al., “Rhinoviruses infect the lower airways,” Journal of Infectious Diseases, vol. 181, no. 6, pp. 1875–1884, 2000.
[65]  B. Winther, J. M. Greve, J. M. Gwaltney et al., “Surface expression of intercellular adhesion molecule 1 on epithelial cells in the human adenoid,” Journal of Infectious Diseases, vol. 176, no. 2, pp. 523–525, 1997.
[66]  B. Winther, E. Arruda, T. J. Witek et al., “Expression of ICAM-1 in nasal epithelium and levels of soluble ICAM-1 in nasal lavage fluid during human experimental rhinovirus infection,” Archives of Otolaryngology, vol. 128, no. 2, pp. 131–136, 2002.
[67]  N. Lopez-Souza, G. Dolganov, R. Dubin et al., “Resistance of differentiated human airway epithelium to infection by rhinovirus,” American Journal of Physiology, vol. 286, no. 2, pp. L373–L381, 2004.
[68]  B. Jakiela, R. Brockman-Schneider, S. Amineva, W. M. Lee, and J. E. Gern, “Basal cells of differentiated bronchial epithelium are more susceptible to rhinovirus infection,” American Journal of Respiratory Cell and Molecular Biology, vol. 38, no. 5, pp. 517–523, 2008.
[69]  N. B. Lomax and F. H. Yin, “Evidence for the role of the P2 protein of human rhinovirus in its host range change,” Journal of Virology, vol. 63, no. 5, pp. 2396–2399, 1989.
[70]  N. Brandl, A. Zemann, I. Kaupe et al., “Signal transduction and metabolism in chondrocytes is modulated by lactoferrin,” Osteoarthritis and Cartilage, vol. 18, no. 1, pp. 117–125, 2010.
[71]  B. Sherry, A. G. Mosser, R. J. Colonno, and R. R. Rueckert, “Use of monoclonal antibodies to identify four neutralization immunogens on a common cold picornavirus, human rhinovirus 14,” Journal of Virology, vol. 57, no. 1, pp. 246–257, 1986.
[72]  B. Sherry and R. Rueckert, “Evidence for at least two dominant neutralization antigens on human rhinovirus 14,” Journal of Virology, vol. 53, no. 1, pp. 137–143, 1985.
[73]  T. Skern, C. Neubauer, and L. Frasel, “A neutralizing epitope on human rhinovirus type 2 includes amino acid residues between 153 and 164 of virus capsid protein VP2,” Journal of General Virology, vol. 68, no. 2, pp. 315–323, 1987.
[74]  E. A. Hewat and D. Blaas, “Structure of a neutralizing antibody bound bivalently to human rhinovirus 2,” EMBO Journal, vol. 15, no. 7, pp. 1515–1523, 1996.
[75]  M. J. Francis, C. M. Fry, and D. J. Rowlands, “Immune response to uncoupled peptides of foot and mouth disease virus,” Immunology, vol. 61, no. 1, pp. 1–6, 1987.
[76]  E. A. Hewat, T. C. Marlovits, and D. Blaas, “Structure of a neutralizing antibody bound monovalently to human rhinovirus 2,” Journal of Virology, vol. 72, no. 5, pp. 4396–4402, 1998.
[77]  A. Pickl-Herk, D. Luque, D. Garriga, B. L. Trus, et al., “Role of the viral RNA genome in assembly and uncoating of human rhinovirus 2,” submitted.
[78]  J. M. Casasnovas and T. A. Springer, “Pathway of rhinovirus disruption by soluble intercellular adhesion molecule 1 (ICAM-1): an intermediate in which ICAM-1 is bound and RNA is released,” Journal of Virology, vol. 68, no. 9, pp. 5882–5889, 1994.
[79]  S. Martin, J. M. Casasnovas, D. E. Staunton, and T. A. Springer, “Efficient neutralization and disruption of rhinovirus by chimeric ICAM- 1/immunoglobulin molecules,” Journal of Virology, vol. 67, no. 6, pp. 3561–3568, 1993.
[80]  D. Garriga, A. Pickl-Herk, D. Luque, J. Wruss, et al., “Insights into minor group rhinovirus uncoating: the X-ray structure of the HRV2 empty capsid,” PLoS Pathog, vol. 8, article e1002473, 2012.
[81]  E. A. Hewat and D. Blaas, “Cryoelectron microscopy analysis of the structural changes associated with human rhinovirus type 14 uncoating,” Journal of Virology, vol. 78, no. 6, pp. 2935–2942, 2004.
[82]  D. M. Belnap, D. J. Filman, B. L. Trus et al., “Molecular tectonic model of virus structural transitions: the putative cell entry states of poliovirus,” Journal of Virology, vol. 74, no. 3, pp. 1342–1354, 2000.
[83]  D. Bubeck, D. J. Filman, N. Cheng, A. C. Steven, J. M. Hogle, and D. M. Belnap, “The structure of the poliovirus 135S cell entry intermediate at 10-angstrom resolution reveals the location of an externalized polypeptide that binds to membranes,” Journal of Virology, vol. 79, no. 12, pp. 7745–7755, 2005.
[84]  J. Lin, N. Cheng, M. Chow, D. J. Filman, et al., “An externalized polypeptide partitions between two distinct sites on genome-released poliovirus particles,” Journal of Virology, vol. 85, pp. 9974–9983, 2011.
[85]  N. Verdaguer, D. Blaas, and I. Fita, “Structure of human rhinovirus serotype 2 (HRV2),” Journal of Molecular Biology, vol. 300, no. 5, pp. 1179–1194, 2000.
[86]  K. Lonberg-Holm and N. M. Whiteley, “Physical and metabolic requirements for early interaction of poliovirus and human rhinovirus with HeLa cells,” Journal of Virology, vol. 19, no. 3, pp. 857–870, 1976.
[87]  E. Metschnikoff, “Lecture on phagocytosis and immunity,” British Medical Journal, vol. 1, pp. 213–217, 1891.
[88]  J. Mercer, M. Schelhaas, and A. Helenius, “Virus entry by endocytosis,” Annual Review of Biochemistry, vol. 79, pp. 803–833, 2010.
[89]  J. L. Goldstein, R. G. W. Anderson, and M. S. Brown, “Coated pits, coated vesicles, and receptor-mediated endocytosis,” Nature, vol. 279, no. 5715, pp. 679–685, 1979.
[90]  H. T. McMahon and E. Boucrot, “Molecular mechanism and physiological functions of clathrin-mediated endocytosis,” Nature Reviews Molecular Cell Biology, vol. 12, pp. 517–533, 2011.
[91]  S. D. Conner and S. L. Schmid, “Regulated portals of entry into the cell,” Nature, vol. 422, pp. 37–44, 2003.
[92]  S. H. Hansen, K. Sandvig, and B. Van Deurs, “Molecules internalized by clathrin-independent endocytosis are delivered to endosomes containing transferrin receptors,” Journal of Cell Biology, vol. 123, no. 1, pp. 89–97, 1993.
[93]  L. Pelkmans, T. Bürli, M. Zerial, and A. Helenius, “Caveolin-stabilized membrane domains as multifunctional transport and sorting devices in endocytic membrane traffic,” Cell, vol. 118, no. 6, pp. 767–780, 2004.
[94]  B. D. Grant and J. G. Donaldson, “Pathways and mechanisms of endocytic recycling,” Nature Reviews Molecular Cell Biology, vol. 10, no. 9, pp. 597–608, 2009.
[95]  F. R. Maxfield and T. E. McGraw, “Endocytic recycling,” Nature Reviews Molecular Cell Biology, vol. 5, no. 2, pp. 121–132, 2004.
[96]  P. L. Tuma and A. L. Hubbard, “Transcytosis: crossing cellular barriers,” Physiological Reviews, vol. 83, no. 3, pp. 871–932, 2003.
[97]  T. Nishi and M. Forgac, “The vacuolar (H+)-ATPases—nature's most versatile proton pumps,” Nature Reviews Molecular Cell Biology, vol. 3, no. 2, pp. 94–103, 2002.
[98]  V. Marshansky and M. Futai, “The V-type H+-ATPase in vesicular trafficking: targeting, regulation and function,” Current Opinion in Cell Biology, vol. 20, no. 4, pp. 415–426, 2008.
[99]  C. C. Scott and J. Gruenberg, “Ion flux and the function of endosomes and lysosomes: PH is just the start: the flux of ions across endosomal membranes influences endosome function not only through regulation of the luminal pH,” BioEssays, vol. 33, no. 2, pp. 103–110, 2011.
[100]  I. Mellman, R. Fuchs, and A. Helenius, “Acidification of the endocytic and exocytic pathways,” Annual Review of Biochemistry, vol. 55, pp. 663–700, 1986.
[101]  J. Huotari and A. Helenius, “Endosome maturation,” The EMBO Journal, vol. 30, pp. 3481–3500, 2011.
[102]  R. Fuchs, A. Ellinger, M. Pavelka, I. Mellman, and H. Klapper, “Rat liver endocytic coated vesicles do not exhibit ATP-dependent acidification in vitro,” Proceedings of the National Academy of Sciences of the United States of America, vol. 91, no. 11, pp. 4811–4815, 1994.
[103]  D. J. Yamashiro and F. R. Maxfield, “Acidification of endocytic compartments and the intracellular pathways of ligands and receptors,” Journal of Cellular Biochemistry, vol. 26, no. 4, pp. 231–246, 1984.
[104]  M. J. Clague, “Molecular aspects of the endocytic pathway,” Biochemical Journal, vol. 336, no. 2, pp. 271–282, 1998.
[105]  G. Baravalle, D. Schober, M. Huber, N. Bayer, R. F. Murphy, and R. Fuchs, “Transferrin recycling and dextran transport to lysosomes is differentially affected by bafilomycin, nocodazole, and low temperature,” Cell and Tissue Research, vol. 320, no. 1, pp. 99–113, 2005.
[106]  S. Mukherjee, R. N. Ghosh, and F. R. Maxfield, “Endocytosis,” Physiological Reviews, vol. 77, no. 3, pp. 759–803, 1997.
[107]  N. D. Sonawane and A. S. Verkman, “Determinants of [Cl-] in recycling and late endosomes and Golgi complex measured using fluorescent ligands,” Journal of Cell Biology, vol. 160, no. 7, pp. 1129–1138, 2003.
[108]  N. Bayer, D. Schober, E. Prchla, R. F. Murphy, D. Blaas, and R. Fuchs, “Effect of bafilomycin A1 and nocodazole on endocytic transport in HeLa cells: implications for viral uncoating and infection,” Journal of Virology, vol. 72, no. 12, pp. 9645–9655, 1998.
[109]  R. N. Ghosh, D. L. Gelman, and F. R. Maxfield, “Quantification of low density lipoprotein and transferrin endocytic sorting in HEp2 cells using confocal microscopy,” Journal of Cell Science, vol. 107, no. 8, pp. 2177–2189, 1994.
[110]  D. R. Sheff, E. A. Daro, M. Hull, and I. Mellman, “The receptors recycling pathway contains two distinct populations of early endosomes with different sorting functions,” Journal of Cell Biology, vol. 145, no. 1, pp. 123–139, 1999.
[111]  W. A. Dunn, A. L. Hubbard, and N. N. Aronson, “Low temperature selectively inhibits fusion between pinocytic vesicles and lysosomes during heterophagy of 125I-asialofetuin by the perfused rat liver,” Journal of Biological Chemistry, vol. 255, no. 12, pp. 5971–5978, 1980.
[112]  R. Fuchs and D. Blaas, “Uncoating of human rhinoviruses,” Reviews in Medical Virology, vol. 20, no. 5, pp. 281–297, 2010.
[113]  K. Hall, M. E. B. Zajdel, and G. E. Blair, “Unity and diversity in the human adenoviruses: exploiting alternative entry pathways for gene therapy,” Biochemical Journal, vol. 431, no. 3, pp. 321–336, 2010.
[114]  E. de Vries, D. M. Tscherne, M. J. Wienholts et al., “Dissection of the influenza a virus endocytic routes reveals macropinocytosis as an alternative entry pathway,” PLoS Pathogens, vol. 7, no. 3, article e1001329, 2011.
[115]  J. Mercer and A. Helenius, “Gulping rather than sipping: macropinocytosis as a way of virus entry,” Current Opinion in Microbiology, vol. 15, pp. 490–499, 2012.
[116]  A. Jurgeit, R. McDowell, S. Moese, E. Meldrum, et al., “Inhibition of endosomal acidification by a proton carrying small compound has broad anti-picornavirus effects,” EUROPIC, vol. 2012, p. 112, 2012.
[117]  T. Maurin, D. Fenard, G. Lambeau, and A. Doglio, “An Envelope-determined Endocytic Route of Viral Entry Allows HIV-1 to Escape from Secreted Phospholipase A2 Entry Blockade,” Journal of Molecular Biology, vol. 367, no. 3, pp. 702–714, 2007.
[118]  C. M. Finnegan and R. Blumenthal, “Fenretinide inhibits HIV infection by promoting viral endocytosis,” Antiviral Research, vol. 69, no. 2, pp. 116–123, 2006.
[119]  D. Marchant, A. Sall, X. Si et al., “ERK MAP kinase-activated Arf6 trafficking directs coxsackievirus type B3 into an unproductive compartment during virus host-cell entry,” Journal of General Virology, vol. 90, no. 4, pp. 854–862, 2009.
[120]  M. Brabec, D. Schober, E. Wagner et al., “Opening of size-selective pores in endosomes during human rhinovirus serotype 2 in vivo uncoating monitored by single-organelle flow analysis,” Journal of Virology, vol. 79, no. 2, pp. 1008–1016, 2005.
[121]  M. Brabec-Zaruba, B. Pfanzagl, D. Blaas, and R. Fuchs, “Site of human rhinovirus RNA uncoating revealed by fluorescent in situ hybridization,” Journal of Virology, vol. 83, no. 8, pp. 3770–3777, 2009.
[122]  M. Huber, M. Brabec, N. Bayer, D. Blaas, and R. Fuchs, “Elevated endosomal pH in HeLa cells overexpressing mutant dynamin can affect infection by pH-sensitive viruses,” Traffic, vol. 2, no. 10, pp. 727–736, 2001.
[123]  A. G. Khan, A. Pickl-Herk, L. Gajdzik, T. C. Marlovits, R. Fuchs, and D. Blaas, “Human rhinovirus 14 enters rhabdomyosarcoma cells expressing ICAM-1 by a clathrin-, caveolin-, and flotillin-independent pathway,” Journal of Virology, vol. 84, no. 8, pp. 3984–3992, 2010.
[124]  D. Schober, P. Kronenberger, E. Prchla, D. Blaas, and R. Fuchs, “Major and minor receptor group human rhinoviruses penetrate from endosomes by different mechanisms,” Journal of Virology, vol. 72, no. 2, pp. 1354–1364, 1998.
[125]  L. Snyers, H. Zwickl, and D. Blaas, “Human rhinovirus type 2 is internalized by clathrin-mediated endocytosis,” Journal of Virology, vol. 77, pp. 5360–5369, 2003.
[126]  G. Abraham and R. J. Colonno, “Many rhinovirus serotypes share the same cellular receptor,” Journal of Virology, vol. 51, no. 2, pp. 340–345, 1984.
[127]  E. J. Stott and R. A. Killington, “Rhinoviruses,” Annual Review of Microbiology, vol. 26, pp. 503–524, 1972.
[128]  L. DeTulleo and T. Kirchhausen, “The clathrin endocytic pathway in viral infection,” EMBO Journal, vol. 17, no. 16, pp. 4585–4593, 1998.
[129]  H. P. Grunert, K. U. Wolf, K. D. Langner, D. Sawitzky, K. O. Habermehl, and H. Zeichhardt, “Internalization of human rhinovirus 14 into HeLa and ICAM-1-transfected BHK cells,” Medical Microbiology and Immunology, vol. 186, no. 1, pp. 1–9, 1997.
[130]  L. Gajdzig, J. Edlmayr, R. Valenta, D. Blaas, et al., “Entry of HRV89 involves dynamin and clathrin,” EUROPIC, vol. 2012, p. 23, 2012.
[131]  D. E. Staunton, A. Gaur, P. Y. Chan, and T. A. Springer, “Internalization of a major group human rhinovirus does not require cytoplasmic or transmembrane domains of ICAM-1,” Journal of Immunology, vol. 148, no. 10, pp. 3271–3274, 1992.
[132]  N. Bayer, E. Prchla, M. Schwab, D. Blaas, and R. Fuchs, “Human rhinovirus HRV14 uncoats from early endosomes in the presence of bafilomycin,” FEBS Letters, vol. 463, no. 1-2, pp. 175–178, 1999.
[133]  J. M. Casasnovas and T. A. Springer, “Kinetics and thermodynamics of virus binding to receptor. Studies with rhinovirus, intercellular adhesion molecule-1 (ICAM-1), and surface plasmon resonance,” Journal of Biological Chemistry, vol. 270, no. 22, pp. 13216–13224, 1995.
[134]  M. J. Clague, S. Urbe, F. Aniento, and J. Gruenberg, “Vacuolar ATPase activity is required for endosomal carrier vesicle formation,” Journal of Biological Chemistry, vol. 269, no. 1, pp. 21–24, 1994.
[135]  J. F. Presley, S. Mayor, T. E. McGraw, K. W. Dunn, and F. R. Maxfield, “Bafilomycin A1 treatment retards transferrin receptor recycling more than bulk membrane recycling,” Journal of Biological Chemistry, vol. 272, no. 21, pp. 13929–13936, 1997.
[136]  N. Bayer, D. Schober, M. Hüttinger, D. Blaas, and R. Fuchs, “Inhibition of Clathrin-dependent Endocytosis Has Multiple Effects on Human Rhinovirus Serotype 2 Cell Entry,” Journal of Biological Chemistry, vol. 276, no. 6, pp. 3952–3962, 2001.
[137]  M. Brabec, G. Baravalle, D. Blaas, and R. Fuchs, “Conformational changes, plasma membrane penetration, and infection by human rhinovirus type 2: role of receptors and low pH,” Journal of Virology, vol. 77, no. 9, pp. 5370–5377, 2003.
[138]  E. Prchla, E. Kuechler, D. Blaas, and R. Fuchs, “Uncoating of human rhinovirus serotype 2 from late endosomes,” Journal of Virology, vol. 68, no. 6, pp. 3713–3723, 1994.
[139]  G. Baravalle, M. Brabec, L. Snyers, D. Blaas, and R. Fuchs, “Human rhinovirus type 2-antibody complexes enter and infect cells via Fc-γ receptor IIB1,” Journal of Virology, vol. 78, no. 6, pp. 2729–2737, 2004.
[140]  G. Bilek, N. M. Matscheko, A. Pickl-Herk et al., “Liposomal nanocontainers as models for viral infection: monitoring viral genomic RNA transfer through lipid membranes,” Journal of Virology, vol. 85, no. 16, pp. 8368–8375, 2011.
[141]  S. P. Amineva, A. G. Aminev, J. E. Gern, and A. C. Palmenberg, “Comparison of rhinovirus A infection in human primary epithelial and HeLa cells,” Journal of General Virology, vol. 92, pp. 2549–2557, 2011.
[142]  D. A. Knight and S. T. Holgate, “The airway epithelium: structural and functional properties in health and disease,” Respirology, vol. 8, no. 4, pp. 432–446, 2003.
[143]  Y. J. Jang, S. H. Lee, H. J. Kwon, Y. S. Chung, and B. J. Lee, “Development of rhinovirus study model using organ culture of turbinate mucosa,” Journal of Virological Methods, vol. 125, no. 1, pp. 41–47, 2005.
[144]  T. Suzuki, M. Yamaya, M. Kamanaka et al., “Type 2 rhinovirus infection of cultured human tracheal epithelial cells: role of LDL receptor,” American Journal of Physiology, vol. 280, no. 3, pp. L409–L420, 2001.
[145]  H. Wan, H. L. Winton, C. Soeller et al., “Tight junction properties of the immortalized human bronchial epithelial cell lines Calu-3 anmd 16HBE14o-,” European Respiratory Journal, vol. 15, no. 6, pp. 1058–1068, 2000.
[146]  X. Wang, C. Lau, S. Wiehler et al., “Syk is downstream of intercellular adhesion molecule-1 and mediates human rhinovirus activation of p38 MAPK in airway epithelial cells,” Journal of Immunology, vol. 177, no. 10, pp. 6859–6870, 2006.
[147]  C. Lau, P. Castellanos, D. Ranev, X. Wang, and C. W. Chow, “HRV signaling in airway epithelial cells is regulated by ITAM-mediated recruitment and activation of Syk,” Protein and Peptide Letters, vol. 18, no. 5, pp. 518–529, 2011.
[148]  C. Lau, X. Wang, L. Song et al., “Syk associates with clathrin and mediates phosphatidylinositol 3-kinase activation during human rhinovirus internalization,” Journal of Immunology, vol. 180, no. 2, pp. 870–880, 2008.
[149]  D. C. Newcomb, U. Sajjan, S. Nanua et al., “Phosphatidylinositol 3-kinase is required for rhinovirus-induced airway epithelial cell interleukin-8 expression,” Journal of Biological Chemistry, vol. 280, no. 44, pp. 36952–36961, 2005.
[150]  J. K. Bentley, D. C. Newcomb, A. M. Goldsmith, Y. Jia, U. S. Sajjan, and M. B. Hershenson, “Rhinovirus activates interleukin-8 expression via a Src/p110β phosphatidylinositol 3-kinase/Akt pathway in human airway epithelial cells,” Journal of Virology, vol. 81, no. 3, pp. 1186–1194, 2007.
[151]  S. Dreschers, P. Franz, C. A. Dumitru, B. Wilker, K. Jahnke, and E. Gulbins, “Infections with human rhinovirus induce the formation of distinct functional membrane domains,” Cellular Physiology and Biochemistry, vol. 20, no. 1-4, pp. 241–254, 2007.
[152]  H. Grassmé, A. Riehle, B. Wilker, and E. Gulbins, “Rhinoviruses infect human epithelial cells via ceramide-enriched membrane platforms,” Journal of Biological Chemistry, vol. 280, no. 28, pp. 26256–26262, 2005.
[153]  C. A. Dumitru, S. Dreschers, and E. Gulbins, “Rhinoviral infections activate p38MAP-Kinases via membrane rafts and RhoA,” Cellular Physiology and Biochemistry, vol. 17, no. 3-4, pp. 159–166, 2006.
[154]  M. Terajima, M. Yamaya, K. Sekizawa et al., “Rhinovirus infection of primary cultures of human tracheal epithelium: role of ICAM-1 and IL-1β,” American Journal of Physiology, vol. 273, no. 4, pp. L749–L759, 1997.
[155]  D. Schober, E. Prchla, D. Blaas, and R. Fuchs, “Role of viral receptor and low pH on in vivo uncoating of human rhinovirus serotype 14,” Berichte Pathologie, vol. 121, p. 237, 1995.
[156]  T. Skern, H. Torgersen, H. Auer, E. Kuechler, and D. Blaas, “Human rhinovirus mutants resistant to low pH,” Virology, vol. 183, no. 2, pp. 757–763, 1991.
[157]  M. Brabec, D. Blaas, and R. Fuchs, “Wortmannin delays transfer of human rhinovirus serotype 2 to late endocytic compartments,” Biochemical and Biophysical Research Communications, vol. 348, no. 2, pp. 741–749, 2006.
[158]  U. Berka, A. Khan, D. Blaas, and R. Fuchs, “Human rhinovirus type 2 uncoating at the plasma membrane is not affected by a pH gradient but is affected by the membrane potential,” Journal of Virology, vol. 83, no. 8, pp. 3778–3787, 2009.
[159]  B. Brandenburg, L. Y. Lee, M. Lakadamyali, M. J. Rust, X. Zhuang, and J. M. Hogle, “Imaging poliovirus entry in live cells,” PLoS Biology, vol. 5, no. 7, article e183, 2007.
[160]  S. Harutyunyan, M. Kumar, A. Sedivy, X. Subirats, et al., “RNA release from the minor group rhinovirus HRV2 starts from its 3'-end,” EUROPIC, vol. 2012, p. 9, 2012.
[161]  H. C. Levy, M. Bostina, D. J. Filman, and J. M. Hogle, “Catching a virus in the act of RNA release: a novel poliovirus uncoating intermediate characterized by cryo-electron microscopy,” Journal of Virology, vol. 84, no. 9, pp. 4426–4441, 2010.
[162]  M. Bostina, H. Levy, D. J. Filman, and J. M. Hogle, “Poliovirus RNA is released from the capsid near a twofold symmetry axis,” Journal of Virology, vol. 85, no. 2, pp. 776–783, 2011.
[163]  X. Peng, E. Y. Chan, Y. Li, D. L. Diamond, M. J. Korth, and M. G. Katze, “Virus-host interactions: from systems biology to translational research,” Current Opinion in Microbiology, vol. 12, no. 4, pp. 432–438, 2009.
[164]  L. Pelkmans, “Systems biology of virus infection in mammalian cells,” Current Opinion in Microbiology, vol. 12, no. 4, pp. 429–431, 2009.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133