Skip to main content Accessibility help
×
Hostname: page-component-7479d7b7d-k7p5g Total loading time: 0 Render date: 2024-07-14T15:14:33.021Z Has data issue: false hasContentIssue false

25 - Reactivation and lytic replication of EBV

from Part II - Basic virology and viral gene effects on host cell functions: gammaherpesviruses

Published online by Cambridge University Press:  24 December 2009

Shannon C. Kenney
Affiliation:
University of Wisconsin, Madison, WI, USA
Ann Arvin
Affiliation:
Stanford University, California
Gabriella Campadelli-Fiume
Affiliation:
Università degli Studi, Bologna, Italy
Edward Mocarski
Affiliation:
Emory University, Atlanta
Patrick S. Moore
Affiliation:
University of Pittsburgh
Bernard Roizman
Affiliation:
University of Chicago
Richard Whitley
Affiliation:
University of Alabama, Birmingham
Koichi Yamanishi
Affiliation:
University of Osaka, Japan
Get access

Summary

Viral pathogenesis

The lytic form of EBV infection is required for the production of progeny virus, and is thus essential for cell-to-cell spread of the virus, as well as transmission from host to host. Unfortunately, there is currently no cell culture system in vitro that is permissive for efficient primary lytic EBV infection. Although a recent report suggests that allowing the virus to first attach to the surface of primary B cells greatly facilitates EBV infection of epithelial cells in vitro (Shannon-Lowe et al., 2006), even this system of infection still does not result in efficient horizontal spread of virus from cell to cell. Thus, lytic EBV infection in vitro has been studied by reactivating the lytic form of infection from latently infected cell lines using a variety of inducing agents, including phorbol esters, butyrate, calcium ionophores, and B-cell receptor stimulation.

During primary infection, EBV probably initially infects oral epithelial cells in a lytic form, and then subsequently infects B-cells, where the virus usually assumes one of the latent forms of infection. In contrast to alpha and beta herpesviruses, which cause human diseases during the lytic form of viral infection but are essentially innocuous while in the latent form of infection, most illnesses attributable to EBV infection are associated with the latent forms of infection. During primary EBV infection, some individuals, particularly adolescents, develop the syndrome infectious mononucleosis (IM) approximately 1 month after being infected (Cohen, 2000; Jenson, 2000).

Type
Chapter
Information
Human Herpesviruses
Biology, Therapy, and Immunoprophylaxis
, pp. 403 - 433
Publisher: Cambridge University Press
Print publication year: 2007

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Adamson, A. L. and Kenney, S. C. (1998). Rescue of the Epstein–Barr virus BZLF1 mutant, Z(S186A), early gene activation defect by the BRLF1 gene product. Virology, 251(1), 187–197.CrossRefGoogle Scholar
Adamson, A. L. and Kenney, S. (1999). The Epstein–Barr virus BZLF1 protein interacts physically and functionally with the histone acetylase CREB-binding protein. J. Virol., 73(8), 6551–6558.Google ScholarPubMed
Adamson, A. L. and Kenney, S. (2001). Epstein–Barr virus immediate-early protein BZLF1 is SUMO-1 modified and disrupts promyelocytic leukemia bodies. J. Virol., 75(5), 2388–2399.CrossRefGoogle ScholarPubMed
Adamson, A. L., Darr, D., Holley-Guthrie, E.et al. (2000). Epstein–Barr virus immediate-early proteins BZLF1 and BRLF1 activate the ATF2 transcription factor by increasing the levels of phosphorylated p38 and c-Jun N-terminal kinases. J. Virol., 74(3), 1224–1233.CrossRefGoogle ScholarPubMed
Adler, B., Schaadt, E., Kempkes, B., Zimber-Strobl, U., Baier, B., and Bornkamm, G. W. (2002). Control of Epstein–Barr virus reactivation by activated CD40 and viral latent membrane protein 1. Proc. Natl Acad. Sci. USA, 99(1), 437–442.CrossRefGoogle ScholarPubMed
Ahsan, N., Kanda, T., Nagashima, K., and Takada, K. (2005). Epstein–Barr virus transforming protein LMP 1 plays a critical role in virus production. J. Virol., 79, 4415–4424.CrossRefGoogle Scholar
Ambinder, R. F., Robertson, K. D., and Tao, Q. (1999). DNA methylation and the Epstein–Barr virus. Semin. Cancer Biol., 9(5), 369–375.CrossRefGoogle ScholarPubMed
Andersson, J., Skoldenberg, B., Henle, W.et al. (1987). Acyclovir treatment in infectious mononucleosis: a clinical and virological study. Infection, 15 Suppl 1, S14–S20.CrossRefGoogle ScholarPubMed
Baumann, M., Mischak, H., Dammeier, S.et al. (1998). Activation of the Epstein–Barr virus transcription factor BZLF1 by 12-O-tetradecanoylphorbol-13-acetate-induced phosphorylation. J. Virol., 72(10), 8105–8114.Google ScholarPubMed
Baumann, M., Feederle, R., Kremmer, E., and Hammerschmidt, W. (1999). Cellular transcription factors recruit viral replication proteins to activate the Epstein–Barr virus origin of lytic DNA replication, oriLyt. EMBO J., 18(21), 6095–6105.CrossRefGoogle ScholarPubMed
Bell, P., Lieberman, P. M., and Maul, G. G. (2000). Lytic but not latent replication of epstein–barr virus is associated with PML and induces sequential release of nuclear domain 10 proteins. J. Virol., 74(24), 11800–11810.CrossRefGoogle Scholar
Bellows, D. S., Howell, M., Pearson, C., Hazlewood, S. A., and Hardwick, J. M. (2002). Epstein–Barr virus BALF1 is a BCL-2-like antagonist of the herpesvirus antiapoptotic BCL-2 proteins. J. Virol., 76(5), 2469–2479.CrossRefGoogle ScholarPubMed
Bernardi, R. and Pandolfi, P. P. (2003). Role of PML and the PML-nuclear body in the control of programmed cell death. Oncogene, 22(56), 9048–9057.CrossRefGoogle ScholarPubMed
Bhende, P. M., Seaman, W. T., Delecluse, H. J., and Kenney, S. C. (2004). The EBV lytic switch protein, Z, preferentially binds to and activates the methylated viral genome. Nat. Genet., 36(10), 1099–1104.CrossRefGoogle Scholar
Biggin, M., Bodescot, M., Perricaudet, M., and Farrell, P. (1987). Epstein–Barr virus gene expression in P3HR1-superinfected Raji cells. J. Virol., 61(10), 3120–3132.Google ScholarPubMed
Binne, U. K., Amon, W., and Farrell, P. J. (2002). Promoter sequences required for reactivation of Epstein-Barr virus from latency. J. Virol., 76(20), 10282–10289.CrossRefGoogle ScholarPubMed
Borras, A. M., Strominger, J. L., and Speck, S. H. (1996). Characterization of the ZI domains in the Epstein–Barr virus BZLF1 gene promoter: role in phorbol ester induction. J. Virol., 70(6), 3894–3901.Google ScholarPubMed
Borza, C. M. and Hutt-Fletcher, L. M. (2002). Alternate replication in B cells and epithelial cells switches tropism of Epstein–Barr virus. Nat. Med., 8(6), 594–599.CrossRefGoogle Scholar
Boyer, J. L., Swaminathan, S., and Silverstein, S. J. (2002). The Epstein–Barr virus SM protein is functionally similar to ICP27 from herpes simplex virus in viral infections. J. Virol., 76(18), 9420–9433.CrossRefGoogle ScholarPubMed
Boyle, S. M., Ruvolo, V., Gupta, A. K., and Swaminathan, S. (1999). Association with the cellular export receptor CRM 1 mediates function and intracellular localization of Epstein–Barr virus SM protein, a regulator of gene expression. J. Virol., 73(8), 6872–6881.Google Scholar
Bryant, H. and Farrell, P. J. (2002). Signal transduction and transcription factor modification during reactivation of Epstein–Barr virus from latency. J. Virol., 76(20), 10290–10298.CrossRefGoogle ScholarPubMed
Buisson, M., Manet, E., Trescol-Biemont, M. C., Gruffat, H., Durand, B., and Sergeant, A. (1989). The Epstein–Barr virus (EBV) early protein EB2 is a posttranscriptional activator expressed under the control of EBV transcription factors EB1 and R. J. Virol., 63(12), 5276–5284.Google ScholarPubMed
Buisson, M., Hans, F., Kusters, I., Duran, N., and Sergeant, A. (1999). The C-terminal region but not the Arg-X-Pro repeat of Epstein–Barr virus protein EB2 is required for its effect on RNA splicing and transport. J. Virol., 73(5), 4090–4100.Google Scholar
Cai, X., Schafer, A., Lu, S.et al. (2006). Epstein–Barr virus microRNAs are evolutionarily conserved and differentially expressed. PLoS Pathog. 2(3), 236–247.CrossRefGoogle ScholarPubMed
Carey, M., Kolman, J., Katz, D. A., Gradoville, L., Barberis, L., and Miller, G. (1992). Transcriptional synergy by the Epstein–Barr virus transactivator ZEBRA. J. Virol., 66(8), 4803–4813.Google ScholarPubMed
Cayrol, C. and Flemington, E. K. (1995). Identification of cellular target genes of the Epstein-Barr virus transactivator Zta: activation of transforming growth factor beta igh3 (TGF-beta igh3) and TGF-beta 1. J. Virol., 69(7), 4206–4212.Google ScholarPubMed
Cayrol, C. and Flemington, E. (1996a). G0/G1 growth arrest mediated by a region encompassing the basic leucine zipper (bZIP) domain of the Epstein–Barr virus transactivator Zta. J. Biol. Chem., 271(50), 31799–31802.CrossRefGoogle Scholar
Cayrol, C. and Flemington, E. K. (1996b). The Epstein–Barr virus bZIP transcription factor Zta causes G0/G1 cell cycle arrest through induction of cyclin-dependent kinase inhibitors. EMBO J., 15(11), 2748–2759.Google Scholar
Chang, P. J., Chang, Y. S., and Liu, S. T. (1998). Role of Rta in the translation of bicistronic BZLF1 of Epstein–Barr virus. J. Virol., 72(6), 5128–5136.Google ScholarPubMed
Chang, P. J. and Liu, S. T. (2001). Function of the intercistronic region of BRLF1-BZLF1 bicistronic mRNA in translating the zta protein of Epstein–Barr virus. J. Virol., 75(3), 1142–1151.CrossRefGoogle ScholarPubMed
Chang, Y. N., Dong, D. L., Hayward, G. S., and Hayward, S. D. (1990). The Epstein–Barr virus Zta transactivator: a member of the bZIP family with unique DNA-binding specificity and a dimerization domain that lacks the characteristic heptad leucine zipper motif. J. Virol., 64(7), 3358–3369.Google Scholar
Chatila, T., Ho, N., Liu, P.et al. (1997). The Epstein–Barr virus-induced Ca2+-/calmodulin-dependent kinase type IV/Gr promotes a Ca2(+)-dependent switch from latency to viral replication. J. Virol., 71(9), 6560–6567.Google ScholarPubMed
Chee, A. V., Lopez, P., Pandolfi, P. P., and Roizman, B. (2003). Promyelocytic leukemia protein mediates interferon-based anti-herpes simplex virus 1 effects. J. Virol., 77(12), 7101–7105.CrossRefGoogle ScholarPubMed
Chen, C. J., Deng, Z., Kim, A. Y., Blobel, G. A., and Lieberman, P. M. (2001a). Stimulation of CREB binding protein nucleosomal histone acetyltransferase activity by a class of transcriptional activators. Mol. Cell. Biol., 21(2), 476–487.CrossRefGoogle Scholar
Chen, L., Liao, G., Fujimuro, M., Semmes, O. J., and Hayward, S. D. (2001b). Properties of two EBV Mta nuclear export signal sequences. Virology, 288(1), 119–128.CrossRefGoogle Scholar
Chen, M. R., Chang, S. J., Huang, H., and Chen, J. Y. (2000). A protein kinase activity associated with Epstein–Barr virus BGLF4 phosphorylates the viral early antigen EA-D in vitro. J. Virol., 74(7), 3093–3104.CrossRefGoogle ScholarPubMed
Chevallier-Greco, A., Gruffat, H., Manet, E., Calender, A., and Sergeant, A. (1989). The Epstein–Barr virus (EBV) DR enhancer contains two functionally different domains: domain A is constitutive and cell specific, domain B is transactivated by the EBV early protein R. J. Virol., 63(2), 615–623.Google Scholar
Chi, T. and Carey, M. (1993). The ZEBRA activation domain: modular organization and mechanism of action. Mol. Cell. Biol., 13(11), 7045–7055.CrossRefGoogle ScholarPubMed
Chi, T., Lieberman, P., Ellwood, K., and Carey, M. (1995). A general mechanism for transcriptional synergy by eukaryotic activators. Nature, 377(6546), 254–257.CrossRefGoogle ScholarPubMed
Cho, M. S., Milman, G., and Hayward, S. D. (1985). A second Epstein–Barr virus early antigen gene in BamHI fragment M encodes a 48- to 50-kilodalton nuclear protein. J. Virol., 56(3), 860–866.Google ScholarPubMed
Cohen, J. I. (1999). The biology of Epstein–Barr virus: lessons learned from the virus and the host. Curr. Opin. Immunol., 11(4), 365–370.CrossRefGoogle Scholar
Cohen, J. I. (2000). Epstein–Barr virus infection. N. Engl. J. Med., 343(7), 481–492.CrossRefGoogle ScholarPubMed
Cohen, J. I. and Lekstrom, K. (1999). Epstein–Barr virus BARF1 protein is dispensable for B-cell transformation and inhibits alpha interferon secretion from mononuclear cells. J. Virol., 73(9), 7627–7632.Google ScholarPubMed
Cook, I. D., Shanahan, F., and Farrell, P. J. (1994). Epstein–Barr virus SM protein. Virology, 205(1), 217–227.CrossRefGoogle ScholarPubMed
Countryman, J. and Miller, G. (1985). Activation of expression of latent Epstein–Barr herpesvirus after gene transfer with a small cloned subfragment of heterogeneous viral DNA. Proc. Natl Acad. Sci. USA, 82(12), 4085–4089.CrossRefGoogle ScholarPubMed
Countryman, J., Jenson, H., Seibl, R., Wolf, H., and Miller, G. (1987). Polymorphic proteins encoded within BZLF1 of defective and standard Epstein–Barr viruses disrupt latency. J. Virol., 61(12), 3672–3679.Google ScholarPubMed
Cox, M. A., Leahy, J., and Hardwick, J. M. (1990). An enhancer within the divergent promoter of Epstein–Barr virus responds synergistically to the R and Z transactivators. J. Virol., 64(1), 313–321.Google Scholar
Daibata, M. and Sairenji, T. (1993). Epstein–Barr virus (EBV) replication and expressions of EA-D (BMRF1 gene product), virus-specific deoxyribonuclese, and DNA polymerase in EBV-activated Akata cells. Virology, 196(2), 900–904.CrossRefGoogle Scholar
Daibata, M., Humphreys, R. E., and Sairenji, T. (1992). Phosphorylation of the Epstein–Barr virus BZLF1 immediate-early gene product ZEBRA. Virology, 188(2), 916–920.CrossRefGoogle ScholarPubMed
Daibata, M., Speck, S. H., Mulder, C., and Sairenji, T. (1994). Regulation of the BZLF1 promoter of Epstein–Barr virus by second messengers in anti-immunoglobulintreated B cells. Virology, 198(2), 446–454.CrossRefGoogle ScholarPubMed
Darenkov, I. A., Marcarelli, M. A., Basadonna, G. P.et al. (1997). Reduced incidence of Epstein–Barr virus-associated posttransplant lymphoproliferative disorder using preemptive antiviral therapy. Transplantation, 64(6), 848–852.CrossRefGoogle ScholarPubMed
Darr, C. D., Mauser, A., and Kenney, S. (2001). Epstein–Barr virus immediate-early protein BRLF1 induces the lytic form of viral replication through a mechanims involving phosphatidylinositol-3 kinase activation. J. Virol., 75(13), 6135–6142.CrossRefGoogle Scholar
Datta, A. K. and Hood, R. E. (1981). Mechanism of inhibition of Epstein–Barr virus replication by phosphonoformic acid. Virology, 114(1), 52–59.CrossRefGoogle ScholarPubMed
Datta, A. K., Colby, B. M., Shaw, J. E., and Pagano, J. S. (1980). Acyclovir inhibition of Epstein–Barr virus replication. Proc. Natl Acad. Sci. USA, 77(9), 5163–5166.CrossRefGoogle ScholarPubMed
Davie, J. R. (2003). Inhibition of histone deacetylase activity by butyrate. J. Nutr., 133(7 Suppl), 2485S–2493S.CrossRefGoogle ScholarPubMed
Deng, Z., Chen, C. J., Zerby, D., Delecluse, H. J., and Lieberman, P. M. (2001). Identification of acidic and aromatic residues in the Zta activation domain essential for Epstein–Barr virus reactivation. J. Virol., 75(21), 10334–10347.CrossRefGoogle ScholarPubMed
Deng, Z., Chen, C. J., Chamberlin, M.et al. (2003). The CBP bromodomain and nucleosome targeting are required for Zta-directed nucleosome acetylation and transcription activation. Mol. Cell. Biol., 23(8), 2633–2644.CrossRefGoogle ScholarPubMed
Dent, P., Yacoub, A., Fisher, P. B., Hagan, M. P., and Grant, S. (2003). MAPK pathways in radiation responses. Oncogene, 22(37), 5885–5896.CrossRefGoogle ScholarPubMed
El-Guindy, A. S., Heston, L., Endo, Y., Cho, M. S., and Miller, G. (2002). Disruption of Epstein–Barr virus latency in the absence of phosphorylation of ZEBRA by protein kinase C. J. Virol., 76(22), 11199–11208.CrossRefGoogle ScholarPubMed
EI-Guindy, A. S. and Miller, G. (2004). Phosphorylation of Epstein-Barr virus ZEBRA protein at its casein kinase 2 sites mediates its ability to repress activation of a viral lytic cycle late gene by Rta. J. Virol., 78(14): 7634–7644.Google Scholar
EI-Guindy, A. S., Paek, S. Y., Countryman, J., and Miller, G. (2006). Identification of constitutive phosphorylation sites on the Epstein–Barr virus ZEBRA protein. J. Biol. Chem., 281(6), 3085–3095.CrossRefGoogle ScholarPubMed
Faggioni, A., Zompetta, C., Grimaldi, S., Barile, G., Frati, L., and Lazdins, J. (1986). Calcium modulation activates Epstein–Barr virus genome in latently infected cells. Science, 232(4757), 1554–1556.CrossRefGoogle ScholarPubMed
Fahmi, H., Cochet, C., Hmama, Z., Opolon, P., and Joab, I. (2000). Transforming growth factor beta 1 stimulates expression of the Epstein–Barr virus BZLF1 immediate-early gene product ZEBRA by an indirect mechanism which requires the MAPK kinase pathway. J. Virol., 74(13), 5810–5818.CrossRefGoogle ScholarPubMed
Falk, K. I. and Ernberg, I. (1999). Demethylation of the Epstein–Barr virus origin of lytic replication and of the immediate early gene BZLF1 is DNA replication independent. Brief report. Arch. Virol., 144(11), 2219–2227.CrossRefGoogle ScholarPubMed
Faller, D. V., Mentzer, S. J., and Perrine, S. P. (2001). Induction of the Epstein–Barr virus thymidine kinase gene with concomitant nucleoside antivirals as a therapeutic strategy for Epstein–Barr virus-associated malignancies. Curr. Opin. Oncol., 13(5), 360–367.CrossRefGoogle ScholarPubMed
Fan, M. and Chambers, T. C. (2001). Role of mitogen-activated protein kinases in the response of tumor cells to chemotherapy. Drug Resist. Updat., 4(4), 253–267.CrossRefGoogle Scholar
Farina, A., Feederle, R., Raffa, S.et al. (2005). BFRF1 of Epstein-Barr virus is essential for efficient primary viral envelopment and egress. J. Virol., 3703–3712.CrossRefGoogle ScholarPubMed
Farjot, G., Buisson, M., Dodon, Duc M., Gazzolo, L., Sergeant, A., and Mikaelian, I. (2000). Epstein–Barr virus EB2 protein exports unspliced RNA via a Crm-1-independent pathway. J. Virol., 74(13), 6068–6076.CrossRefGoogle Scholar
Farmer, D. G., McDiarmid, S. V., Winston, D.et al. (2002). Effectiveness of aggressive prophylatic and preemptive therapies targeted against cytomegaloviral and Epstein–Barr viral disease after human intestinal transplantation. Transpl. Proc., 34(3), 948–949.CrossRefGoogle ScholarPubMed
Farrell, P. J., Rowe, D. T., Rooney, C. M., and Kouzarides, T. (1989). Epstein–Barr virus BZLF1 trans-activator specifically binds to a consensus AP-1 site and is related to c-fos. EMBO. J., 8(1), 127–132.Google ScholarPubMed
Faulkner, G. C., Krajewski, A. S., and Crawford, D. H. (2000). The ins and outs of EBV infection. Trends Microbiol. 8(4), 185–189.CrossRefGoogle ScholarPubMed
Feederle, R., Kost, M., Baumann, M.et al. (2000). The Epstein–Barr virus lytic program is controlled by the co-operative functions of two transactivators. EMBO J. 19(12), 3080–3089.CrossRefGoogle ScholarPubMed
Feng, W. H., Israel, B., Raab-Traub, N., Busson, P., and Kenney, S. C. (2002a). Chemotherapy induces lytic EBV replication and confers ganciclovir susceptibility to EBV-positive epithelial cell tumors. Cancer Res., 62(6), 1920–1926.Google Scholar
Feng, W. H., Westphal, E., Mauser, A.et al. (2002b). Use of adenovirus vectors expressing Epstein–Barr virus (EBV) immediate-early protein BZLF1 or BRLF1 to treat EBV-positive tumors. J. Virol., 76(21), 10951–10959.CrossRefGoogle Scholar
Feng, W. H., Hong, G., Delecluse, H. J., and Kenney, S. C. (2004). Lytic Induction therapy for Epstein–Barr virus-positive B-cell lymphomas. J. Virol., 78(4), 1893–1902..CrossRefGoogle ScholarPubMed
Fixman, E. D., Hayward, G. S., and Hayward, S. D. (1992). trans-acting requirements for replication of Epstein–Barr virus ori-Lyt. J. Virol., 66(8), 5030–5039.Google ScholarPubMed
Fixman, E. D., Hayward, G. S., and Hayward, S. D. (1995). Replication of Epstein–Barr virus oriLyt: lack of a dedicated virally encoded origin-binding protein and dependence on Zta in cotransfection assays. J. Virol., 69(5), 2998–3006.Google ScholarPubMed
Flamand, L. and Menezes, J. (1996). Cyclic AMP-responsive element-dependent activation of Epstein–Barr virus zebra promoter by human herpesvirus 6. J. Virol., 70(3), 1784–1791.Google ScholarPubMed
Fleischmann, J., Kremmer, E., Greenspan, J. S., Grasser, F. A., and Niedobitek, G. (2002). Expression of viral and human dUTPase in Epstein–Barr virus-associated diseases. J. Med. Virol., 68(4), 568–573.CrossRefGoogle ScholarPubMed
Flemington, E. and Speck, S. H. (1990a). Autoregulation of Epstein–Barr virus putative lytic switch gene BZLF1. J. Virol., 64(3), 1227–1232.Google Scholar
Flemington, E. and Speck, S. H. (1990b). Epstein–Barr virus BZLF1 trans activator induces the promoter of a cellular cognate gene, c-fos. J. Virol., 64(9), 4549–4552.Google Scholar
Flemington, E. and Speck, S. H. (1990c). Evidence for coiled-coil dimer formation by an Epstein–Barr virus transactivator that lacks a heptad repeat of leucine residues. Proc. Natl Acad. Sci. USA, 87(23), 9459–9463.CrossRefGoogle Scholar
Flemington, E. and Speck, S. H. (1990d). Identification of phorbol ester response elements in the promoter of Epstein–Barr virus putative lytic switch gene BZLF1. J. Virol., 64(3), 1217–1226.Google Scholar
Flemington, E. K., Goldfeld, A. E., and Speck, S. H. (1991). Efficient transcription of the Epstein–Barr virus immediate-early BZLF1 and BRLF1 genes requires protein synthesis. J. Virol., 65(12), 7073–7077.Google ScholarPubMed
Flemington, E. K., Borras, A. M., Lytle, J. P., and Speck, S. H. (1992). Characterization of the Epstein–Barr virus BZLF1 protein transactivation domain. J. Virol., 66(2), 922–929.Google ScholarPubMed
Flemington, E. K., Lytle, J. P., Cayrol, C., Borras, A. M., and Speck, S. H. (1994). DNA-binding-defective mutants of the Epstein–Barr virus lytic switch activator Zta transactivate with altered specificities. Mol. Cell. Biol., 14(5), 3041–3052.CrossRefGoogle ScholarPubMed
Fong, I. W., Ho, J., Toy, C., Lo, B., and Fong, M. W. (2000). Value of long-term administration of acyclovir and similar agents for protecting against AIDS-related lymphoma: case-control and historical cohort studies. Clin. Infect. Dis., 30(5), 757–761.CrossRefGoogle ScholarPubMed
Francis, A. L., Gradoville, L., and Miller, G. (1997). Alteration of a single serine in the basic domain of the Epstein–Barr virus ZEBRA protein separates its functions of transcriptional activation and disruption of latency. J. Virol., 71(4), 3054–3061.Google Scholar
Francis, A., Ragoczy, T., Gradoville, L., Heston, L., El-Guindy, A., Endo, Y., and Miller, G. (1999). Amino acid substitutions reveal distinct functions of serine 186 of the ZEBRA protein in activation of early lytic cycle genes and synergy with the Epstein–Barr virus R transactivator. J. Virol., 73(6), 4543–4551.Google ScholarPubMed
Fu, Z.Cannon, M. J. (2000). Functional analysis of the CD4(+) T-cell response to Epstein–Barr virus: T-cell-mediated activation of resting B cells and induction of viral BZLF1 expression. J. Virol., 74(14), 6675–6679.CrossRefGoogle ScholarPubMed
Fujii, K., Yokoyama, N., Kiyono, T.et al. (2000). The Epstein–Barr virus pol catalytic subunit physically interacts with the BBLF4-BSLF1-BBLF2/3 complex. J. Virol., 74(6), 2550–2557.CrossRefGoogle ScholarPubMed
Furnari, F. B., Zacny, V., Quinlivan, E. B., Kenney, S., and Pagano, J. S. (1994). RAZ, an Epstein–Barr virus transdominant repressor that modulates the viral reactivation mechanism. J. Virol., 68(3), 1827–1836.Google ScholarPubMed
Furukawa, F., Matsuzaki, K., Mori, S.et al. (2003). p38 MAPK mediates fibrogenic signal through Smad3 phosphorylation in rat myofibroblasts. Hepatology, 38(4), 879–889.CrossRefGoogle ScholarPubMed
Gan, Y. J., Razzouk, B. I., Su, T., and Sixbey, J. W. (2002). A defective, rearranged Epstein–Barr virus genome in EBER-negative and EBER-positive Hodgkin's disease. Am. J. Pathol., 160(3), 781–786.CrossRefGoogle ScholarPubMed
Gao, X., Tajima, M., and Sairenji, T. (1999). Nitric oxide down-regulates Epstein–Barr virus reactivation in epithelial cell lines. Virology, 258(2), 375–381.CrossRefGoogle ScholarPubMed
Gao, X., Ikuta, K., Tajima, M., and Sairenji, T. (2001). 12-O-tetradecanoylphorbol-13-acetate induces Epstein–Barr virus reactivation via NF-kappaB and AP-1 as regulated by protein kinase C and mitogen-activated protein kinase. Virology, 286(1), 91–99.CrossRefGoogle ScholarPubMed
Gao, Z., Krithivas, A., Finan, J. E., Semmes, O. J., Zhou, S., Wang, Y., and Hayward, S. D. (1998). The Epstein–Barr virus lytic transactivator Zta interacts with the helicase-primase replication proteins. J. Virol., 72(11), 8559–8567.Google ScholarPubMed
Gershburg, E. and Pagano, J. S. (2002). Phosphorylation of the Epstein–Barr virus (EBV) DNA polymerase processivity factor EA-D by the EBV-encoded protein kinase and effects of the L-riboside benzimidazole 1263W94. J. Virol., 76(3), 998–1003.CrossRefGoogle ScholarPubMed
Glaser, G., Vogel, M., Wolf, H., and Niller, H. H. (1998). Regulation of the Epstein-Barr viral immediate early BRLF1 promoter through a distal NF1 site. Arch. Virol., 143(10), 1967–1983.CrossRefGoogle ScholarPubMed
Goldfeld, A. E., Liu, P., Liu, S., Flemington, E. K., Strominger, J. L., and Speck, S. H. (1995). Cyclosporin A and FK506 block induction of the Epstein–Barr virus lytic cycle by anti-immunoglobulin. Virology, 209(1), 225–229.CrossRefGoogle ScholarPubMed
Gonella, R., Farina, A., Santarelli, R.et al. (2005). Characterization and intracellular localization of the Epstein-Barr virus protein BFLF 2: interactions with BFRF 1 and with the nuclear lamina. J. Virol., 79, 3713–3727.CrossRefGoogle Scholar
Gong, M. and Kieff, E. (1990). Intracellular trafficking of two major Epstein-Barr virus glycoproteins, gp350/220 and gp110. J. Virol., 64, 1507–1516.Google ScholarPubMed
Gonzalez, C. M., Wong, E. L., Bowser, B. S., Hong, G. K., Kenney, S., and Damania, B. (2006). Identification and characterization of the Orf49 protein of Kaposi's sarcoma-associated herpesvirus. J. Virol., 80(6), 3062–3070.CrossRefGoogle ScholarPubMed
Grab, L. T., Kearns, M. W., Morris, A. J., and Daniel, L. W. (2004). Differential role for phospholipase D1 and phospholipase D2 in 12-O-tetradecanoyl-13-phorbol acetate-stimulated MAPK activation, Cox-2 and IL-8 expression. Biochim. Biophys. Acta, 1636(1), 29–39.CrossRefGoogle ScholarPubMed
Gradoville, L., Grogan, E., Taylor, N., and Miller, G. (1990). Differences in the extent of activation of Epstein–Barr virus replicative gene expression among four nonproducer cell lines stably transformed by oriP/BZLF1 plasmids. Virology, 178(2), 345–354.CrossRefGoogle Scholar
Gradoville, L., Kwa, D., El-Guindy, A., and Miller, G. (2002). Protein kinase C-independent activation of the Epstein–Barr virus lytic cycle. J. Virol., 76(11), 5612–5626.CrossRefGoogle ScholarPubMed
Green, M., Reyes, J., Webber, S., and Rowe, D. (2001). The role of antiviral and immunoglobulin therapy in the prevention of Epstein–Barr virus infection and post-transplant lymphoproliferative disease following solid organ transplantation. Transpl. Infect. Dis., 3(2), 97–103.CrossRefGoogle ScholarPubMed
Greenspan, J. S., Greenspan, D., Lennette, E. T.et al. (1985). Replication of Epstein–Barr virus within the epithelial cells of oral “hairy” leukoplakia, an AIDS-associated lesion. N. Engl. J. Med., 313(25), 1564–1571.CrossRefGoogle ScholarPubMed
Grogan, E., Jenson, H., Countryman, J., Heston, L., Gradoville, L., and Miller, G. (1987). Transfection of a rearranged viral DNA fragment, WZhet, stably converts latent Epstein–Barr viral infection to productive infection in lymphoid cells. Proc. Natl Acad. Sci. USA, 84(5), 1332–1336.CrossRefGoogle ScholarPubMed
Gruffat, H. and Sergeant, A. (1994). Characterization of the DNA-binding site repertoire for the Epstein–Barr virus transcription factor R. Nucl. Acids Res., 22(7), 1172–1178.CrossRefGoogle ScholarPubMed
Gruffat, H., Manet, E., Rigolet, A., and Sergeant, A. (1990). The enhancer factor R of Epstein–Barr virus (EBV) is a sequence-specific DNA binding protein. Nucl. Acids Res., 18(23), 6835–6843.CrossRefGoogle Scholar
Gruffat, H., Duran, N., Buisson, M., Wild, F., Buckland, R., and Sergeant, A. (1992). Characterization of an R-binding site mediating the R-induced activation of the Epstein–Barr virus BMLF1 promoter. J. Virol., 66(1), 46–52.Google ScholarPubMed
Gruffat, H., Renner, O., Pich, D., and Hammerschmidt, W. (1995). Cellular proteins bind to the downstream component of the lytic origin of DNA replication of Epstein–Barr virus. J. Virol., 69(3), 1878–1886.Google ScholarPubMed
Gruffat, H., Batisse, J., Pich, D.et al. (2002a). Epstein–Barr virus mRNA export factor EB2 is essential for production of infectious virus. J. Virol., 76(19), 9635–9644.CrossRefGoogle Scholar
Gruffat, H., Manet, E., and Sergeant, A. (2002b). MEF2-mediated recruitment of class II HDAC at the EBV immediate early gene BZLF1 links latency and chromatin remodeling. EMBO Rep., 3(2), 141–146.CrossRefGoogle Scholar
Gutierrez, M. I., Judde, J. G., Magrath, I. T., and Bhatia, K. G. (1996). Switching viral latency to viral lysis: a novel therapeutic approach for Epstein–Barr virus-associated neoplasia. Cancer Res., 56(5), 969–972.Google ScholarPubMed
Gutsch, D. E., Holley-Guthrie, E. A., Zhang, Q.et al. (1994). The bZIP transactivator of Epstein–Barr virus, BZLF1, functionally and physically interacts with the p65 subunit of NF-kappa B. Mol. Cell. Biol., 14(3), 1939–1948.CrossRefGoogle ScholarPubMed
Haan, K. M., Lee, S. K., and Longnecker, R. (2001). Different functional domains in the cytoplasmic tail of glycoprotein gB are involved in Epstein–Barr virus induced membrane fusion. Virology, 290, 106–114.CrossRefGoogle ScholarPubMed
Hahn, A. M., Huye, L. E., Ning, S., Webster-Cyriaque, J., and Pagano, J. S. (2005). Interferon regulatory factor 7 is negatively regulated by the Epstein–Barr virus immediate-early gene, BZLF 1. J. Virol., 79(15), 10040–10052.CrossRefGoogle ScholarPubMed
Hammerschmidt, W. and Sugden, B. (1988). Identification and characterization of oriLyt, a lytic origin of DNA replication of Epstein–Barr virus. Cell, 55(3), 427–433.CrossRefGoogle ScholarPubMed
Hardwick, J. M., Lieberman, P. M., and Hayward, S. D. (1988). A new Epstein–Barr virus transactivator, R, induces expression of a cytoplasmic early antigen. J. Virol., 62(7), 2274–2284.Google ScholarPubMed
Hardwick, J. M., Tse, L., Applegren, N., Nicholas, J., and Veliuona, M. A. (1992). The Epstein–Barr virus R transactivator (Rta) contains a complex, potent activation domain with properties different from those of VP16. J. Virol., 66(9), 5500–5508.Google ScholarPubMed
Hayakawa, J., Depatie, C., Ohmichi, M., and Mercola, D. (2003). The activation of c-Jun NH2-terminal kinase (JNK) by DNA-damaging agents serves to promote drug resistance via activating transcription factor 2 (ATF2)-dependent enhanced DNA repair. J. Biol. Chem., 278(23), 20582–20592.CrossRefGoogle ScholarPubMed
Henderson, S., Huen, D., Rowe, M., Dawson, C., Johnson, G., and Rickinson, A. (1993). Epstein–Barr virus-coded BHRF1 protein, a viral homologue of Bcl-2, protects human B cells from programmed cell death. Proc. Natl Acad. Sci. USA, 90(18), 8479–8483.CrossRefGoogle Scholar
Herrmann, K., Frangou, P., Middeldorp, J., and Niedobitek, G. (2002). Epstein–Barr virus replication in tongue epithelial cells. J. Gen. Virol., 83(12), 2995–2998.CrossRefGoogle ScholarPubMed
Herrold, R. E., Marchini, A., Frueling, S., and Longnecker, R. (1995). Glycoprotein 110, the Epstein–Barr virus homolog of herpes simplex virus glycoprotein B, is essential for Epstein–Barr virus replication in vivo. J. Virol., 70, 2049–2054.Google Scholar
Hiriart, E., Bardouillet, L., Manet, E.et al. (2003a). A region of the Epstein–Barr virus (EBV) mRNA export factor EB2 containing an arginine-rich motif mediates direct binding to RNA. J. Biol. Chem., 278(39), 37790–37798.CrossRefGoogle Scholar
Hiriart, E., Farjot, G., Gruffat, H., Nguyen, M. V., Sergeant, A., and Manet, E. (2003b). A novel nuclear export signal and a REF interaction domain both promote mRNA export by the Epstein–Barr virus EB2 protein. J. Biol. Chem., 278(1), 335–342.CrossRefGoogle Scholar
Hislop, A. D., Annels, N. E., Gudgeon, N. H., Leese, A. M., and Rickinson, A. B. (2002). Epitope-specific evolution of human CD8(+) T cell responses from primary to persistent phases of Epstein–Barr virus infection. J. Exp. Med., 195(7), 893–905.CrossRefGoogle ScholarPubMed
Holley-Guthrie, E. A., Quinlivan, E. B., Mar, E. C., and Kenney, S. (1990). The Epstein–Barr virus (EBV) BMRF1 promoter for early antigen (EA-D) is regulated by the EBV transactivators, BRLF1 and BZLF1, in a cell-specific manner. J. Virol., 64(8), 3753–3759.Google Scholar
Hong, G. K., Kumar, P., Damania, B.et al. (2005). Epstein–Barr virus lytic infection is required for efficient production of the angiogenesis factor vascular endothelial growth factor in lymphoblastoid cell lines. J. Virol., 79(22), 13984–13992.CrossRefGoogle ScholarPubMed
Hong, G. K., Gulley, M. L., Feng, W. H., Delecluse, H. J., Holley-Guthrie, E., and Kenney, S.C. (2005). Epstein-Barr virus lytic infection contributes to lymphoproliferative disease in a SCID mouse model. J. Virol., 79(22), 13993–14003.CrossRefGoogle Scholar
Hong, G. K., Delecluse, H.-J., Gruffat, H., Morrison, T. E., Feng, W., and Kenney, S. C. (2004). The BRRF1 early gene of Epstein–Barr virus encodes a transcription factor that enhances induction of lytic infection by BRLF1. J. Virol., 78(10), in press.CrossRefGoogle ScholarPubMed
Hong, Y., Holley-Guthrie, E., and Kenney, S. (1997). The bZip dimerization domain of the Epstein–Barr virus BZLF1 (Z) protein mediates lymphoid-specific negative regulation. Virology, 229(1), 36–48.CrossRefGoogle Scholar
Hsu, D. H., Malefyt, Waal R., Fiorentino, D. F.et al. (1990). Expression of interleukin-10 activity by Epstein–Barr virus protein BCRF1. Science, 250(4982), 830–832.CrossRefGoogle ScholarPubMed
Huang, J., Chen, H., Hutt-Fletcher, L., Ambinder, R. F., and Hayward, S. D. (2003). Lytic viral replication as a contributor to the detection of Epstein–Barr virus in breast cancer. J. Virol., 77(24), 13267–13274.CrossRefGoogle ScholarPubMed
Ikuta, K., Satoh, Y., Hoshikawa, Y., and Sairenji, T. (2000). Detection of Epstein–Barr virus in salivas and throat washings in healthy adults and children. Microbes Infect., 2(2), 115–120.CrossRefGoogle ScholarPubMed
Inman, G. J., Binne, U. K., Parker, G. A., Farrell, P. J., and Allday, M. J. (2001). Activators of the Epstein–Barr virus lytic program concomitantly induce apoptosis, but lytic gene expression protects from cell death. J. Virol., 75(5), 2400–2410.CrossRefGoogle ScholarPubMed
Ionescu, A. M., Schwarz, E. M., Zuscik, M. J.et al. (2003). ATF-2 cooperates with Smad3 to mediate TGF-beta effects on chondrocyte maturation. Exp. Cell. Res., 288(1), 198–207.CrossRefGoogle ScholarPubMed
Israel, B. F. and Kenney, S. C. (2003). Virally targeted therapies for EBV-associated malignancies. Oncogene, 22(33), 5122–5130.CrossRefGoogle ScholarPubMed
Israele, V., Shirley, P., and Sixbey, J. W. (1991). Excretion of the Epstein–Barr virus from the genital tract of men. J. Infect. Dis., 163(6), 1341–1343.CrossRefGoogle ScholarPubMed
Iwakiri, D. and Takada, K. (2004). Phosphatidylinositol 3-kinase is a determinant of responsiveness to B cell antigen receptor-mediated Epstein–Barr virus activation. J. Immunol., 172(3), 1561–1566.CrossRefGoogle Scholar
Jenkins, P. J., Binne, U. K., and Farrell, P. J. (2000). Histone acetylation and reactivation of Epstein–Barr virus from latency. J. Virol., 74(2), 710–720.CrossRefGoogle ScholarPubMed
Jenson, H. B. (2000). Acute complications of Epstein–Barr virus infectious mononucleosis. Curr. Opin. Pediatr., 12(3), 263–268.CrossRefGoogle ScholarPubMed
Johannsen, E., Luftig, M., Chase, M. R.et al. (2004). Proteins of purified EBV. Proc. Natl Acad. Sci. USA, 101(46), 16286–16291.CrossRefGoogle Scholar
Karimi, L., Crawford, D. H., Speck, S., and Nicholson, L. J. (1995). Identification of an epithelial cell differentiation responsive region within the BZLF1 promoter of the Epstein–Barr virus. J. Gen. Virol., 76(4), 759–765.CrossRefGoogle ScholarPubMed
Kato, K., Kawaguchi, Y., Tanaka, M.et al. (2001). Epstein–Barr virus-encoded protein kinase BGLF4 mediates hyperphosphorylation of cellular elongation factor 1delta (EF-1delta): EF-1delta is universally modified by conserved protein kinases of herpesviruses in mammalian cells. J. Gen. Virol., 82(6), 1457–1463.CrossRefGoogle ScholarPubMed
Kato, K., Yokoyama, A., Tohya, Y., Akashi, H., Nishiyama, Y., and Kawaguchi, Y. (2003). Identification of protein kinases responsible for phosphorylation of Epstein–Barr virus nuclear antigen leader protein at serine-35, which regulates its coactivator function. J. Gen. Virol., 84(12), 3381–3392.CrossRefGoogle ScholarPubMed
Kawaguchi, Y., Kato, K., Tanaka, M., Kanamori, M., Nishiyama, Y., and Yamanashi, Y. (2003). Conserved protein kinases encoded by herpesviruses and cellular protein kinase cdc2 target the same phosphorylation site in eukaryotic elongation factor 1delta. J. Virol., 77(4), 2359–2368.CrossRefGoogle ScholarPubMed
Kawanishi, M. (1995). Nitric oxide inhibits Epstein–Barr virus DNA replication and activation of latent EBV. Intervirology, 38(3–4), 206–213.CrossRefGoogle ScholarPubMed
Keating, S., Prince, S., Jones, M., and Rowe, M. (2002). The lytic cycle of Epstein–Barr virus is associated with decreased expression of cell surface major histocompatibility complex class I and class II molecules. J. Virol., 76(16), 8179–8188.CrossRefGoogle Scholar
Kenney, S., Holley-Guthrie, E., Mar, E. C., and Smith, M. (1989a). The Epstein–Barr virus BMLF1 promoter contains an enhancer element that is responsive to the BZLF1 and BRLF1 transactivators. J. Virol., 63(9), 3878–3883.Google Scholar
Kenney, S., Kamine, J., Holley-Guthrie, E., Lin, J. C., Mar, E. C., and Pagano, J. (1989b). The Epstein–Barr virus (EBV) BZLF1 immediate-early gene product differentially affects latent versus productive EBV promoters. J. Virol., 63(4), 1729–1736.Google Scholar
Kenney, S., Kamine, J., Holley-Guthrie, E.et al. (1989c). The Epstein–Barr virus immediate-early gene product, BMLF1, acts in trans by a posttranscriptional mechanism which is reporter gene dependent. J. Virol., 63(9), 3870–3877.Google Scholar
Kiehl, A., and Dorsky, D. I. (1991). Cooperation of EBV DNA polymerase and EAD(BMRF1) in vitro and colocalization in nuclei of infected cells. Virology, 184(1), 330–340.CrossRefGoogle Scholar
Kiehl, A., and Dorsky, D. I. (1995). Bipartite DNA-binding region of the Epstein–Barr virus BMRF1 product essential for DNA polymerase accessory function. J. Virol., 69(3), 1669–1677.Google ScholarPubMed
Kouzarides, T., Packham, G., Cook, A., and Farrell, P. J. (1991). The BZLF1 protein of EBV has a coiled coil dimerisation domain without a heptad leucine repeat but with homology to the C/EBP leucine zipper. Oncogene, 6(2), 195–204.Google Scholar
Krappmann, D., Patke, A., Heissmeyer, V., and Scheidereit, C. (2001). B-cell receptor- and phorbol ester-induced NF-kappaB and c-Jun N-terminal kinase activation in B cells requires novel protein kinase C's. Mol. Cell. Biol., 21(19), 6640–6650.CrossRefGoogle Scholar
Kraus, R. J., Mirocha, S. J., Stephany, H. M., Puchalski, J. R., and Mertz, J. E. (2001). Identification of a novel element involved in regulation of the lytic switch BZLF1 gene promoter of Epstein–Barr virus. J. Virol., 75(2), 867–877.CrossRefGoogle ScholarPubMed
Kraus, R. J., Perrigoue, J. G., and Mertz, J. E. (2003). ZEB negatively regulates the lytic-switch BZLF1 gene promoter of Epstein–Barr virus. J. Virol., 77(1), 199–207.CrossRefGoogle ScholarPubMed
Kudoh, A., Fujita, M., Kiyono, T.et al. (2003). Reactivation of lytic replication from B cells latently infected with Epstein–Barr virus occurs with high S-phase cyclin-dependent kinase activity while inhibiting cellular DNA replication. J. Virol., 77(2), 851–861.CrossRefGoogle Scholar
Kudoh, A., Daikoku, T., Sugaya, Y.et al. (2004). Inhibition of S-phase cyclin-dependent kinase activity blocks expression of Epstein–Barr virus immediate-early and early genes, preventing viral lytic replication. J. Virol., 78(1), 104–115.CrossRefGoogle ScholarPubMed
Laichalk, L. L. and Thorley-Lawson, D. A. (2005). Terminal differentiation into plasma cells initiates the replicative cycle of Epstein–Barr virus in vivo. J. Virol., 79(2), 1296–307.CrossRefGoogle ScholarPubMed
Lake, C. M. and Hutt-Fletcher, L. M. (2000). Epstein–Barr virus that lacks glycoprotein gN is impaired in assembly and infection. J. Virol., 74, 11162–11172.CrossRefGoogle ScholarPubMed
Lake, C. M. and Hutt-Fletcher, L. M. (2004). The Epstein–Barr virus BFRF 1 and BFLF 2 proteins interact and coexpression alters their cellular localization. Virology, 320, 99–106.CrossRefGoogle Scholar
Lau, R., Middeldorp, J., and Farrell, P. J. (1993). Epstein–Barr virus gene expression in oral hairy leukoplakia. Virology, 195(2), 463–474.CrossRefGoogle ScholarPubMed
Roux, F., Sergeant, A., and Corbo, L. (1996). Epstein–Barr virus (EBV) EB1/Zta protein provided in trans and competent for the activation of productive cycle genes does not activate the BZLF1 gene in the EBV genome. J. Gen. Virol., 77, 501–509.CrossRefGoogle Scholar
Li, M., Linseman, D. A., Allen, M. P.et al. (2001). Myocyte enhancer factor 2A and 2D undergo phosphorylation and caspase-mediated degradation during apoptosis of rat cerebellar granule neurons. J. Neurosci., 21(17), 6544–6552.CrossRefGoogle ScholarPubMed
Li, Q. X., Young, L. S., Niedobitek, G.et al. (1992). Epstein–Barr virus infection and replication in a human epithelial cell system. Nature, 356(6367), 347–350.CrossRefGoogle Scholar
Li, Y., Webster-Cyriaque, J., Tomlinson, C. C., Yohe, M., and Kenney, S. (2004). Fatty acid synthase expression is induced by the Epstein–Barr virus immediate-early protein BRLF1 and required for lytic viral gene expression. J. Virol., 78(8), in press.Google ScholarPubMed
Liang, C. L., Chen, J. L., Hsu, Y. P., Ou, J. T., and Chang, Y. S. (2002). Epstein–Barr virus BZLF1 gene is activated by transforming growth factor-beta through cooperativity of Smads and c-Jun/c-Fos proteins. J. Biol. Chem., 277(26), 23345–23357.CrossRefGoogle ScholarPubMed
Liao, G., Wu, F. Y., and Hayward, S. D. (2001). Interaction with the Epstein–Barr virus helicase targets Zta to DNA replication compartments. J. Virol., 75(18), 8792–8802.CrossRefGoogle ScholarPubMed
Lieberman, P. M. and Berk, A. J. (1990). In vitro transcriptional activation, dimerization, and DNA-binding specificity of the Epstein–Barr virus Zta protein. J. Virol., 64(6), 2560–2568.Google ScholarPubMed
Lieberman, P. M. and Berk, A. J. (1991). The Zta trans-activator protein stabilizes TFIID association with promoter DNA by direct protein-protein interaction. Genes Dev., 5(12B), 2441–2454.CrossRefGoogle ScholarPubMed
Lieberman, P. M. and Berk, A. J. (1994). A mechanism for TAFs in transcriptional activation: activation domain enhancement of TFIID-TFIIA – promoter DNA complex formation. Genes Dev., 8(9), 995–1006.CrossRefGoogle ScholarPubMed
Lieberman, P. M., Hardwick, J. M., and Hayward, S. D. (1989). Responsiveness of the Epstein–Barr virus NotI repeat promoter to the Z transactivator is mediated in a cell-type-specific manner by two independent signal regions. J. Virol., 63(7), 3040–3050.Google Scholar
Lieberman, P. M., Hardwick, J. M., Sample, J., Hayward, G. S., and Hayward, S. D. (1990). The zta transactivator involved in induction of lytic cycle gene expression in Epstein–Barr virus-infected lymphocytes binds to both AP-1 and ZRE sites in target promoter and enhancer regions. J. Virol., 64(3), 1143–1155.Google ScholarPubMed
Lieberman, P. M., Ozer, J., and Gursel, D. B. (1997). Requirement for transcription factor IIA (TFIIA)-TFIID recruitment by an activator depends on promoter structure and template competition. Mol. Cell. Biol., 17(11), 6624–6632.CrossRefGoogle ScholarPubMed
Lin, J. C. and Machida, H. (1988). Comparison of two bromovinyl nucleoside analogs, 1-beta-D-arabinofuranosyl-E-5-(2-bromovinyl)uracil and E-5-(2-bromovinyl)-2-deoxyuridine, with acyclovir in inhibition of Epstein–Barr virus replication. Antimicrob. Agents Chemother., 32(7), 1068–1072.CrossRefGoogle ScholarPubMed
Lin, J. C., Nelson, D. J., Lambe, C. U., and Choi, E. I. (1986). Metabolic activation of 9([2-hydroxy-1-(hydroxymethyl) ethoxy]methy1)guanine in human lymphoblastoid cell lines infected with Epstein–Barr virus. J. Virol., 60(2), 569–573.Google Scholar
Lin, J. C., DeClercq, E., and Pagano, J. S. (1987). Novel acyclic adenosine analogs inhibit Epstein–Barr virus replication. Antimicrob. Agents Chemother., 31(9), 1431–1433.CrossRefGoogle ScholarPubMed
Ling, P. D., Lednicky, J. A., Keitel, W. A.et al. (2003). The dynamics of herpesvirus and polyomavirus reactivation and shedding in healthy adults: a 14-month longitudinal study. J. Infect. Dis., 187(10), 1571–1580.CrossRefGoogle ScholarPubMed
Littler, E., and Arrand, J. R. (1988). Characterization of the Epstein–Barr virus-encoded thymidine kinase expressed in heterologous eucaryotic and procaryotic systems. J. Virol., 62(10), 3892–3895.Google ScholarPubMed
Littler, E., Zeuthen, J., McBride, A. A.et al. (1986). Identification of an Epstein–Barr virus-coded thymidine kinase. EMBO J., 5(8), 1959–1966.Google ScholarPubMed
Liu, C., Sista, N. D., and Pagano, J. S. (1996). Activation of the Epstein–Barr virus DNA polymerase promoter by the BRLF1 immediate-early protein is mediated through USF and E2F. J. Virol., 70(4), 2545–2555.Google ScholarPubMed
Liu, P. and Speck, S. H. (2003). Synergistic autoactivation of the Epstein–Barr virus immediate-early BRLF1 promoter by Rta and Zta. Virology, 310(2), 199–206.CrossRefGoogle ScholarPubMed
Liu, P., Liu, S., and Speck, S. H. (1998). Identification of a negative cis element within the ZII domain of the Epstein–Barr virus lytic switch BZLF1 gene promoter. J. Virol., 72(10), 8230–8239.Google ScholarPubMed
Liu, S., Borras, A. M., Liu, P., Suske, G., and Speck, S. H. (1997a). Binding of the ubiquitous cellular transcription factors Sp1 and Sp3 to the ZI domains in the Epstein–Barr virus lytic switch BZLF1 gene promoter. Virology, 228(1), 11–18.CrossRefGoogle Scholar
Liu, S., Liu, P., Borras, A., Chatila, T., and Speck, S. H. (1997b). Cyclosporin A-sensitive induction of the Epstein–Barr virus lytic switch is mediated via a novel pathway involving a MEF2 family member. EMBO J., 16(1), 143–153.CrossRefGoogle Scholar
Lu, J., Chen, S. Y., Chua, H. H.et al. (2000). Upregulation of tyrosine kinase TKT by the Epstein–Barr virus transactivator Zta. J. Virol., 74(16), 7391–7399.CrossRefGoogle ScholarPubMed
Lu, J., Chua, H. H., Chen, S. Y., Chen, J. Y., and Tsai, C. H. (2003). Regulation of matrix metalloproteinase-1 by Epstein–Barr virus proteins. Cancer Res., 63(1), 256–262.Google ScholarPubMed
Lucht, E., Biberfeld, P., and Linde, A. (1995). Epstein–Barr virus (EBV) DNA in saliva and EBV serology of HIV-1-infected persons with and without hairy leukoplakia. J. Infect., 31(3), 189–194.CrossRefGoogle ScholarPubMed
MacCallum, P., Karimi, L., and Nicholson, L. J. (1999). Definition of the transcription factors which bind the differentiation responsive element of the Epstein–Barr virus BZLF1 Z promoter in human epithelial cells. J. Gen. Virol., 80(6), 1501–1512.CrossRefGoogle Scholar
Mahot, S., Sergeant, A., Drouet, E., and Gruffat, H. (2003). A novel function for the Epstein–Barr virus transcription factor EB1/Zta: induction of transcription of the hIL-10 gene. J. Gen. Virol., 84(4), 965–974.CrossRefGoogle ScholarPubMed
Malouf, M. A., Chhajed, P. N., Hopkins, P., Plit, M., Turner, J., and Glanville, A. R. (2002). Anti-viral prophylaxis reduces the incidence of lymphoproliferative disease in lung transplant recipients. J. Heart Lung Transpl., 21(5), 547–554.CrossRefGoogle ScholarPubMed
Manet, E., Gruffat, H., Trescol-Biemont, M. C.et al. (1989). Epstein–Barr virus bicistronic mRNAs generated by facultative splicing code for two transcriptional trans-activators. EMBO J., 8(6), 1819–1826.Google ScholarPubMed
Manet, E., Rigolet, A., Gruffat, H., Giot, J. F., and Sergeant, A. (1991). Domains of the Epstein–Barr virus (EBV) transcription factor R required for dimerization, DNA binding and activation. Nucl. Acids Res., 19(10), 2661–2667.CrossRefGoogle ScholarPubMed
Manet, E., Allera, C., Gruffat, H., Mikaelian, I., Rigolet, A., and Sergeant, A. (1993). The acidic activation domain of the Epstein–Barr virus transcription factor R interacts in vitro with both TBP and TFIIB and is cell-specifically potentiated by a prolinerich region. Gene Expr., 3(1), 49–59.Google Scholar
Mansouri, A., Ridgway, L. D., Korapati, A. L.et al. (2003). Sustained activation of JNK/p38 MAPK pathways in response to cisplatin leads to Fas ligand induction and cell death in ovarian carcinoma cells. J. Biol. Chem., 278(21), 19245–19256.CrossRefGoogle ScholarPubMed
Marschall, M., Stein-Gerlach, M., Freitag, M., Kupfer, R, Bogaard, M., and Stamminger, T. (2002). Direct targeting of human cytomegalovirus protein kinase pUL97 by kinase inhibitors is a novel principle for antiviral therapy. J. Gen. Virol., 83(5), 1013–1023.CrossRefGoogle ScholarPubMed
Marshall, W. L., Yim, C., Gustafson, E.et al. (1999). Epstein–Barr virus encodes a novel homolog of the bcl-2 oncogene that inhibits apoptosis and associates with Bax and Bak. J. Virol., 73(6), 5181–5185.Google ScholarPubMed
Matthews, R. P., Guthrie, C. R., Wailes, L. M., Zhao, X., Means, A. R., and McKnight, G. S. (1994). Calcium/calmodulin-dependent protein kinase types II and IV differentially regulate CREB-dependent gene expression. Mol. Cell. Biol., 14(9), 6107–6116.CrossRefGoogle ScholarPubMed
Mauser, A., Holley-Guthrie, E., Simpson, D., Kaufmann, W., and Kenney, S. (2002a). The Epstein–Barr virus immediate-early protein BZLF1 induces both a G(2) and a mitotic block. J. Virol., 76(19), 10030–10037.CrossRefGoogle Scholar
Mauser, A., Holley-Guthrie, E., Zanation, A.et al. (2002b). The Epstein–Barr virus immediate-early protein BZLF1 induces expression of E2F-1 and other proteins involved in cell cycle progression in primary keratinocytes and gastric carcinoma cells. J. Virol., 76(24), 12543–12552.CrossRefGoogle Scholar
Mauser, A., Saito, S., Appella, E., Anderson, C. W., Seaman, W. T., and Kenney, S. (2002c). The Epstein–Barr virus immediate-early protein BZLF1 regulates p53 function through multiple mechanisms. J. Virol., 76(24), 12503–12512.CrossRefGoogle Scholar
McDiarmid, S. V., Jordan, S., Kim, G. S.et al. (1998). Prevention and preemptive therapy of postransplant lymphoproliferative disease in pediatric liver recipients. Transplantation, 66(12), 1604–1611.CrossRefGoogle ScholarPubMed
McKinsey, T. A., Zhang, C. L., Lu, J., and Olson, E. N. (2000). Signal-dependent nuclear export of a histone deacetylase regulates muscle differentiation. Nature, 408(6808), 106–111.Google ScholarPubMed
Meerbach, A., Holy, A., Wutzler, P., Clercq, E., and Neyts, J. (1998). Inhibitory effects of novel nucleoside and nucleotide analogues on Epstein–Barr virus replication. Antivir. Chem. Chemother. 9(3), 275–282.CrossRefGoogle ScholarPubMed
Mellinghoff, I., Daibata, M., Humphreys, R. E., Mulder, C., Takada, K., and Sairenji, T. (1991). Early events in Epstein–Barr virus genome expression after activation: regulation by second messengers of B cell activation. Virology, 185(2), 922–928.CrossRefGoogle ScholarPubMed
Mentzer, S. J., Fingeroth, J., Reilly, J. J., Perrine, S. P., and Faller, D. V. (1998). Arginine butyrate-induced susceptibility to ganciclovir in an Epstein–Barr-virus-associated lymphoma. Blood Cells Mol. Dis., 24(2), 114–119.CrossRefGoogle Scholar
Mentzer, S. J., Perrine, S. P., and Faller, D. V. (2001). Epstein–Barr virus post-transplant lymphoproliferative disease and virus-specific therapy: pharmacological reactivation of viral target genes with arginine butyrate. Transpl. Infect. Dis., 3(3), 177–185.CrossRefGoogle Scholar
Mettenleiter, T. C. (2002). Herpesvirus assembly and egress. J. Virol., 76(4), 1537–1547.CrossRefGoogle ScholarPubMed
Mikaelian, I., Manet, E., and Sergeant, A. (1993). The bZIP motif of the Epstein–Barr virus (EBV) transcription factor EB1 mediates a direct interaction with TBP. C R Acad. Sci. III, 316(12), 1424–1432.Google ScholarPubMed
Miller, C. L., Lee, J. H., Kieff, E., Burkhardt, A. L., Bolen, J. B., and Longnecker, R. (1994a). Epstein–Barr virus protein LMP2A regulates reactivation from latency by negatively regulating tyrosine kinases involved in sIg-mediated signal transduction. Infect. Agents. Dis., 3(2–3), 128–136.Google Scholar
Miller, C. L., Lee, J. H., Kieff, E., and Longnecker, R. (1994b). An integral membrane protein (LMP2) blocks reactivation of Epstein–Barr virus from latency following surface immunoglobulin crosslinking. Proc. Natl Acad. Sci. USA, 91(2), 772–776.CrossRefGoogle Scholar
Miller, C. L., Burkhardt, A. L., Lee, J. H.et al. (1995). Integral membrane protein 2 of Epstein–Barr virus regulates reactivation from latency through dominant negative effects on protein-tyrosine kinases. Immunity, 2(2), 155–166.CrossRefGoogle ScholarPubMed
Miyazaki, I., Cheung, R. K., and Dosch, H. M. (1993). Viral interleukin 10 is critical for the induction of B cell growth transformation by Epstein–Barr virus. J. Exp. Med., 178(2), 439–447.CrossRefGoogle ScholarPubMed
Montalvo, E. A., Shi, Y., Shenk, T. E., and Levine, A. J. (1991). Negative regulation of the BZLF1 promoter of Epstein–Barr virus. J. Virol., 65(7), 3647–3655.Google ScholarPubMed
Montalvo, E. A., Cottam, M., Hill, S., and Wang, Y. J. (1995). YY1 binds to and regulates cis-acting negative elements in the Epstein–Barr virus BZLF1 promoter. J. Virol., 69(7), 4158–4165.Google ScholarPubMed
Moore, S. M., Cannon, J. S., Tanhehco, Y. C., Hamzeh, F. M., and Ambinder, R. F. (2001). Induction of Epstein–Barr virus kinases to sensitize tumor cells to nucleoside analogues. Antimicrob. Agents Chemother., 45(7), 2082–2091.CrossRefGoogle ScholarPubMed
Morrison, T. E. and Kenney, S. C. (2004). BZLF1, an Epstein–Barr virus immediate-early protein, induces p65 nuclear translocation while inhibiting p65 transcriptional function. Virology, 328(2), 219–232.CrossRefGoogle ScholarPubMed
Morrison, T. E., Mauser, A., Wong, A., Ting, J. P., and Kenney, S. C. (2001). Inhibition of IFN-gamma signaling by an Epstein–Barr virus immediate-early protein. Immunity 15(5), 787–799.CrossRefGoogle ScholarPubMed
Morrison, T. E., Mauser, A., Klingelhutz, A., and Kenney, S. C. (2004). Epstein–Barr virus immediate-early protein BZLF1 inhibits tumor necrosis factor alpha-induced signaling and apoptosis by downregulating tumor necrosis factor receptor 1. J. Virol., 78(1), 544–549.CrossRefGoogle ScholarPubMed
Niedobitek, G., Agathanggelou, A., Herbst, H., Whitehead, L., Wright, D. H., and Young, L. S. (1997). Epstein–Barr virus (EBV) infection in infectious mononucleosis: virus latency, replication and phenotype of EBV-infected cells. J. Pathol., 182(2), 151–159.3.0.CO;2-3>CrossRefGoogle ScholarPubMed
Niedobitek, G., Agathanggelou, A., Steven, N., and Young, L. S. (2000). Epstein–Barr virus (EBV) in infectious mononucleosis: detection of the virus in tonsillar B lymphocytes but not in desquamated oropharyngeal epithelial cells. Mol. Pathol., 53(1), 37–42.CrossRefGoogle ScholarPubMed
Nonkwelo, C. B. and Long, W. K. (1993). Regulation of Epstein–Barr virus BamHI-H divergent promoter by DNA methylation. Virology, 197(1), 205–215.CrossRefGoogle ScholarPubMed
Oertel, S. H., Ruhnke, M. S., Anagnostopoulos, I.et al. (1999). Treatment of Epstein–Barr virus-induced posttransplantation lymphoproliferative disorder with foscarnet alone in an adult after simultaneous heart and renal transplantation. Transplantation, 67(5), 765–767.CrossRefGoogle Scholar
Packham, G., Economou, A., Rooney, C. M., Rowe, D. T., and Farrell, P. J. (1990). Structure and function of the Epstein–Barr virus BZLF1 protein. J. Virol., 64(5), 2110–2116.Google ScholarPubMed
Paulson, E. J. and Speck, S. H. (1999). Differential methylation of Epstein–Barr virus latency promoters facilitates viral persistence in healthy seropositive individuals. J. Virol., 73(12), 9959–9968.Google ScholarPubMed
Paulson, E. J., Fingeroth, J. D., Yates, J. L., and Speck, S. H. (2002). Methylation of the EBV genome and establishment of restricted latency in low-passage EBV-infected 293 epithelial cells. Virology, 299(1), 109–121.CrossRefGoogle ScholarPubMed
Pegtel, D. M., Middeldorp, J., and Thorley-Lawson, D. A. (2004). Epstein–Barr virus infection in ex vivo tonsil epithelial cell cultures of asymptomatic carriers. J. Virol., 78(22), 12613–12624.CrossRefGoogle ScholarPubMed
Petosa, C., Morand, P., Baudin, F., Moulin, M., Artero, J. B., and Muller, C. W. (2006). Structural basis of lytic cycle activation by the Epstein–Barr virus ZEBRA protein. Mol. Cell., 21(4), 565–572.CrossRefGoogle ScholarPubMed
Pfeffer, S., Zavolan, M., Grasser, F. A.et al. (2004). Identification of virus-encoded microRNAs. Science, 304(5671), 734–736.CrossRefGoogle ScholarPubMed
Poppers, J., Mulvey, M., Perez, C., Khoo, D., and Mohr, I. (2003). Identification of a lytic-cycle Epstein–Barr virus gene product that can regulate PKR activation. J. Virol., 77(1), 228–236.CrossRefGoogle ScholarPubMed
Portes-Sentis, S., Sergeant, A., and Gruffat, H. (1997). A particular DNA structure is required for the function of a cis-acting component of the Epstein–Barr virus OriLyt origin of replication. Nucl. Acids Res., 25(7), 1347–1354.CrossRefGoogle ScholarPubMed
Quinlivan, E. B., Holley-Guthrie, E. A., Norris, M., Gutsch, D., Bachenheimer, S. L., and Kenney, S. C. (1993). Direct BRLF1 binding is required for cooperative BZLF1/BRLF1 activation of the Epstein–Barr virus early promoter, BMRF1. Nucl. Acids Res., 21(14), 1999–2007.CrossRefGoogle ScholarPubMed
Ragoczy, T. and Miller, G. (1999). Role of the Epstein–Barr virus RTA protein in activation of distinct classes of viral lytic cycle genes. J. Virol., 73(12), 9858–9866.Google ScholarPubMed
Ragoczy, T. and Miller, G. (2001). Autostimulation of the Epstein–Barr virus BRLF1 promoter is mediated through consensus Sp1 and Sp3 binding sites. J. Virol., 75(11), 5240–5251.CrossRefGoogle ScholarPubMed
Ragoczy, T., Heston, L., and Miller, G. (1998). The Epstein–Barr virus Rta protein activates lytic cycle genes and can disrupt latency in B lymphocytes. J. Virol., 72(10), 7978–7984.Google ScholarPubMed
Resnick, L., Herbst, J. S., and Raab-Traub, N. (1990). Oral hairy leukoplakia. J. Am. Acad. Dermatol., 22(6 Pt 2), 1278–1282.CrossRefGoogle ScholarPubMed
Rodriguez, A., Armstrong, M., Dwyer, D., and Flemington, E. (1999). Genetic dissection of cell growth arrest functions mediated by the Epstein–Barr virus lytic gene product, Zta. J. Virol., 73(11), 9029–9038.Google ScholarPubMed
Rodriguez, A., Jung, E. J., and Flemington, E. K. (2001a). Cell cycle analysis of Epstein–Barr virus-infected cells following treatment with lytic cycle-inducing agents. J. Virol., 75(10), 4482–4489.CrossRefGoogle Scholar
Rodriguez, A., Jung, E. J., Yin, Q., Cayrol, C., and Flemington, E. K. (2001b). Role of cmyc regulation in Zta-mediated induction of the cyclin-dependent kinase inhibitors p21 and p27 and cell growth arrest. Virology, 284(2), 159–169.CrossRefGoogle Scholar
Rooney, C., Taylor, N., Countryman, J., Jenson, H., Kolman, J., and Miller, G. (1988). Genome rearrangements activate the Epstein–Barr virus gene whose product disrupts latency. Proc. Natl Acad. Sci. USA, 85(24), 9801–9805.CrossRefGoogle ScholarPubMed
Rooney, C. M., Rowe, D. T., Ragot, T., and Farrell, P. J. (1989). The spliced BZLF1 gene of Epstein–Barr virus (EBV) transactivates an early EBV promoter and induces the virus productive cycle. J. Virol., 63(7), 3109–3116.Google ScholarPubMed
Rousset, F., Garcia, E., Defrance, T.et al. (1992). Interleukin 10 is a potent growth and differentiation factor for activated human B lymphocytes. Proc. Natl Acad. Sci. USA, 89(5), 1890–1893.CrossRefGoogle ScholarPubMed
Roychowdhury, S., Peng, R., Baiocchi, R. A.et al. (2003). Experimental treatment of Epstein–Barr virus-associated primary central nervous system lymphoma. Cancer Res., 63(5), 965–971.Google ScholarPubMed
Ruvolo, V., Wang, E., Boyle, S., and Swaminathan, S. (1998). The Epstein–Barr virus nuclear protein SM is both a post-transcriptional inhibitor and activator of gene expression. Proc. Natl Acad. Sci. USA, 95(15), 8852–8857.CrossRefGoogle ScholarPubMed
Ruvolo, V., Gupta, A. K., and Swaminathan, S. (2001). Epstein–Barr virus SM protein interacts with mRNA in vivo and mediates a gene-specific increase in cytoplasmic mRNA. J. Virol., 75(13), 6033–6041.CrossRefGoogle ScholarPubMed
Ruvolo, V., Navarro, L., Sample, C. E., David, M., Sung, S., and Swaminathan, S. (2003). The Epstein–Barr virus SM protein induces STAT1 and interferon-stimulated gene expression. J. Virol., 77(6), 3690–3701.CrossRefGoogle ScholarPubMed
Ruvolo, V., Sun, L., Howard, K.et al. (2004). Functional analysis of Epstein–Barr virus SM protein: identification of amino acids essential for structure, transactivation, splicing inhibition, and virion production. J. Virol., 78(1), 340–352.CrossRefGoogle ScholarPubMed
Salek-Ardakani, S., Arrand, J. R., and Mackett, M. (2002). Epstein–Barr virus encoded interleukin-10 inhibits HLA-class I, ICAM-1, and B7 expression on human monocytes: implications for immune evasion by EBV. Virology, 304(2), 342–351.CrossRefGoogle ScholarPubMed
Salomoni, P. and Pandolfi, P. P. (2002). The role of PML in tumor suppression. Cell, 108(2), 165–170.CrossRefGoogle ScholarPubMed
Sarisky, R. T., Gao, Z., Lieberman, P. M., Fixman, E. D., Hayward, G. S., and Hayward, S. D. (1996). A replication function associated with the activation domain of the Epstein–Barr virus Zta transactivator. J. Virol., 70(12), 8340–8347.Google ScholarPubMed
Satoh, T., Hoshikawa, Y., Satoh, Y., Kurata, T., and Sairenji, T. (1999). The interaction of mitogen-activated protein kinases to Epstein–Barr virus activation in Akata cells. Virus Genes, 18(1), 57–64.CrossRefGoogle ScholarPubMed
Schepers, A., Pich, D., and Hammerschmidt, W. (1993a). A transcription factor with homology to the AP-1 family links RNA transcription and DNA replication in the lytic cycle of Epstein–Barr virus. EMBO J., 12(10), 3921–3929.Google Scholar
Schepers, A., Pich, D., Mankertz, J., and Hammerschmidt, W. (1993b). Cis-acting elements in the lytic origin of DNA replication of Epstein–Barr virus. J. Virol., 67(7), 4237–4245.Google Scholar
Schepers, A., Pich, D., and Hammerschmidt, W. (1996). Activation of oriLyt, the lytic origin of DNA replication of Epstein–Barr virus, by BZLF1. Virology, 220(2), 367–376.CrossRefGoogle ScholarPubMed
Segouffin, C., Gruffat, H., and Sergeant, A. (1996). Repression by RAZ of Epstein–Barr virus bZIP transcription factor EB1 is dimerization independent. J. Gen. Virol., 77(7), 1529–1536.CrossRefGoogle ScholarPubMed
Segouffin-Cariou, C., Farjot, G., Sergeant, A., and Gruffat, H. (2000). Characterization of the Epstein–Barr virus BRRF1 gene, located between early genes BZLF1 and BRLF1. J. Gen. Virol., 81(7), 1791–1799.CrossRefGoogle ScholarPubMed
Semmes, O. J., Chen, L., Sarisky, R. T., Gao, Z., Zhong, L., and Hayward, S. D. (1998). Mta has properties of an RNA export protein and increases cytoplasmic accumulation of Epstein–Barr virus replication gene mRNA. J. Virol., 72(12), 9526–9534.Google ScholarPubMed
Serio, T. R., Cahill, N., Prout, M. E., and Miller, G. (1998). A functionally distinct TATA box required for late progression through the Epstein–Barr virus life cycle. J. Virol., 72(10), 8338–8343.Google ScholarPubMed
Serio, T. R., Kolman, J. L., and Miller, G. (1997). Late gene expression from the Epstein–Barr virus BcLF1 and BFRF3 promoters does not require DNA replication in cis. J. Virol., 71(11), 8726–8734.Google Scholar
Shannon-Lowe, C. D., Neuhierl, B., Baldwin, G., Rickinson, A. B., and Delecluse, H. J. (2006). Resting B cells as a transfer vehicle for Epstein–Barr virus infection of epithelial cells. Proc. Natl Acad. Sci. USA, 103(18), 7065–7070.CrossRefGoogle Scholar
Shen, B. and Goodman, H. M. (2004). Uridine addition after microRNA-directed cleavage. Science, 306(5698), 997.CrossRefGoogle ScholarPubMed
Shimizu, N. and Takada, K. (1993). Analysis of the BZLF1 promoter of Epstein–Barr virus: identification of an anti-immunoglobulin response sequence. J. Virol., 67(6), 3240–3245.Google ScholarPubMed
Shu, C. H., Chang, Y. S., Liang, C. L., Liu, S. T., Lin, C. Z., and Chang, P. (1992). Distribution of type A and type B EBV in normal individuals and patients with head and neck carcinomas in Taiwan. J. Virol. Methods, 38(1), 123–130.CrossRefGoogle ScholarPubMed
Sinclair, A. J., Brimmell, M., Shanahan, F., and Farrell, P. J. (1991). Pathways of activation of the Epstein–Barr virus productive cycle. J. Virol., 65(5), 2237–2244.Google ScholarPubMed
Sista, N. D., Pagano, J. S., Liao, W., and Kenney, S. (1993). Retinoic acid is a negative regulator of the Epstein–Barr virus protein (BZLF1) that mediates disruption of latent infection. Proc. Natl Acad. Sci. USA, 90(9), 3894–3898.CrossRefGoogle ScholarPubMed
Sixbey, J. W., Lemon, S. M., and Pagano, J. S. (1986). A second site for Epstein–Barr virus shedding: the uterine cervix. Lancet, 2(8516), 1122–1124.CrossRefGoogle ScholarPubMed
Speck, S. H., Chatila, T., and Flemington, E. (1997). Reactivation of Epstein–Barr virus: regulation and function of the BZLF1 gene. Trends Microbiol., 5(10), 399–405.CrossRefGoogle ScholarPubMed
Steven, N. M., Annels, N. E., Kumar, A., Leese, A. M., Kurilla, M. G., and Rickinson, A. B. (1997). Immediate early and early lytic cycle proteins are frequent targets of the Epstein–Barr virus-induced cytotoxic T cell response. J. Exp. Med., 185(9), 1605–1617.CrossRefGoogle ScholarPubMed
Strockbine, L. D., Cohen, J. I., Farrah, T.et al. (1998). The Epstein–Barr virus BARF1 gene encodes a novel, soluble colony-stimulating factor-1 receptor. J. Virol., 72(5), 4015–4021.Google ScholarPubMed
Stuart, A. D., Stewart, J. P., Arrand, J. R., and Mackett, M. (1995). The Epstein–Barr virus encoded cytokine viral interleukin-10 enhances transformation of human B lymphocytes. Oncogene, 11(9), 1711–1719.Google ScholarPubMed
Swaminathan, S., Hesselton, R., Sullivan, J., and Kieff, E. (1993). Epstein–Barr virus recombinants with specifically mutated BCRF1 genes. J. Virol., 67(12), 7406–7413.Google ScholarPubMed
Swenson, J. J., Mauser, A. E., Kaufmann, W. K., and Kenney, S. C. (1999). The Epstein–Barr virus protein BRLF1 activates S phase entry through E2F1 induction. J. Virol., 73(8), 6540–6550.Google ScholarPubMed
Swenson, J. J., Holley-Guthrie, E., and Kenney, S. C. (2001). Epstein–Barr virus immediate-early protein BRLF1 interacts with CBP, promoting enhanced BRLF1 transactivation. J. Virol., 75(13), 6228–6234.CrossRefGoogle ScholarPubMed
Szyf, M., Eliasson, L., Mann, V., Klein, G., and Razin, A. (1985). Cellular and viral DNA hypomethylation associated with induction of Epstein–Barr virus lytic cycle. Proc. Natl Acad. Sci. USA, 82(23), 8090–8094.CrossRefGoogle ScholarPubMed
Takada, K. and Ono, Y. (1989). Synchronous and sequential activation of latently infected Epstein–Barr virus genomes. J. Virol., 63(1), 445–449.Google ScholarPubMed
Takada, K., Shimizu, N., Sakuma, S., and Ono, Y. (1986). Trans activation of the latent Epstein–Barr virus (EBV) genome after transfection of the EBV DNA fragment. J. Virol., 57(3), 1016–1022.Google ScholarPubMed
Takagi, S., Takada, K., and Sairenji, T. (1991). Formation of intranuclear replication compartments of Epstein–Barr virus with redistribution of BZLF1 and BMRF1 gene products. Virology, 185(1), 309–315.CrossRefGoogle ScholarPubMed
Tao, Q., Srivastava, G., Chan, A. C., Chung, L. P., Loke, S. L., and Ho, F. C. (1995). Evidence for lytic infection by Epstein–Barr virus in mucosal lymphocytes instead of nasopharyngeal epithelial cells in normal individuals. J. Med. Virol., 45(1), 71–77.CrossRefGoogle ScholarPubMed
Tarodi, B., Subramanian, T., and Chinnadurai, G. (1994). Epstein–Barr virus BHRF1 protein protects against cell death induced by DNA-damaging agents and heterologous viral infection. Virology, 201(2), 404–407.CrossRefGoogle ScholarPubMed
Taylor, N., Countryman, J., Rooney, C., Katz, D., and Miller, G. (1989). Expression of the BZLF1 latency-disrupting gene differs in standard and defective Epstein–Barr viruses. J. Virol., 63(4), 1721–1728.Google ScholarPubMed
Thomas, C., Danke, A. sreiter, Wolf, H., and Schwarzmann, F. (2003). The BZLF1 promoter of Epstein–Barr virus is controlled by E box-/HI-motif-binding factors during virus latency. J. Gen. Virol., 84(4), 959–964.CrossRefGoogle Scholar
Torre, D. and Tambini, R. (1999). Acyclovir for treatment of infectious mononucleosis: a meta-analysis. Scand. J. Infect. Dis., 31(6), 543–547.Google ScholarPubMed
Tovey, M. G., Lenoir, G., and Begon-Lours, J. (1978). Activation of latent Epstein–Barr virus by antibody to human IgM. Nature, 276(5685), 270–272.CrossRefGoogle ScholarPubMed
Tsurumi, T., Daikoku, T., Kurachi, R., and Nishiyama, Y. (1993). Functional interaction between Epstein–Barr virus DNA polymerase catalytic subunit and its accessory subunit in vitro. J. Virol., 67(12), 7648–7653.Google ScholarPubMed
Tsurumi, T., Daikoku, T., and Nishiyama, Y. (1994). Further characterization of the interaction between the Epstein–Barr virus DNA polymerase catalytic subunit and its accessory subunit with regard to the 3′-to-5′ exonucleolytic activity and stability of initiation complex at primer terminus. J. Virol., 68(5), 3354–3363.Google Scholar
Urier, G., Buisson, M., Chambard, P., and Sergeant, A. (1989). The Epstein–Barr virus early protein EBI activates transcription from different responsive elements including AP-1 binding sites. EMBO J., 8(5), 1447–1453.Google Scholar
Walling, D. M., Edmiston, S. N., Sixbey, J. W., Abdel-Hamid, M., Resnick, L., and Raab-Traub, N. (1992). Coinfection with multiple strains of the Epstein–Barr virus in human immunodeficiency virus-associated hairy leukoplakia. Proc. Natl Acad. Sci. USA, 89(14), 6560–6564.CrossRefGoogle ScholarPubMed
Walling, D. M., Flaitz, C. M., Nichols, C. M., Hudnall, S. D., and Adler-Storthz, K. (2001). Persistent productive Epstein–Barr virus replication in normal epithelial cells in vivo. J. Infect. Dis., 184(12), 1499–1507.CrossRefGoogle ScholarPubMed
Walling, D. M., Brown, A. L., Etienne, W., Keitel, W. A., and Ling, P. D. (2003a). Multiple Epstein–Barr virus infections in healthy individuals. J. Virol., 77(11), 6546–6550.CrossRefGoogle Scholar
Walling, D. M., Flaitz, C. M., and Nichols, C. M. (2003b). Epstein–Barr virus replication in oral hairy leukoplakia: response, persistence, and resistance to treatment with valacyclovir. J. Infect. Dis., 188(6), 883–890.CrossRefGoogle Scholar
Wang, Y. C., Huang, J. M., and Montalvo, E. A. (1997). Characterization of proteins binding to the ZII element in the Epstein–Barr virus BZLF1 promoter: transactivation by ATF1. Virology, 227(2), 323–330.CrossRefGoogle ScholarPubMed
Westphal, E. M., Mauser, A., Swenson, J., Davis, M. G., Talarico, C. L., and Kenney, S. C. (1999). Induction of lytic Epstein–Barr virus (EBV) infection in EBV-associated malignancies using adenovirus vectors in vitro and in vivo. Cancer Res., 59(7), 1485–1491.Google ScholarPubMed
Westphal, E. M., Blackstock, W., Feng, W., Israel, B., and Kenney, S. C. (2000). Activation of lytic Epstein–Barr virus (EBV) infection by radiation and sodium butyrate in vitro and in vivo: a potential method for treating EBV-positive malignancies. Cancer Res., 60(20), 5781–5788.Google ScholarPubMed
Wu, F. Y., Chen, H., Wang, S. E.et al. (2003). CCAAT/enhancer binding protein alpha interacts with ZTA and mediates ZTA-induced p21(CIP-1) accumulation and G(1) cell cycle arrest during the Epstein–Barr virus lytic cycle. J. Virol., 77(2), 1481–1500.CrossRefGoogle Scholar
Yoshizaki, T., Sato, H., Murono, S., Pagano, J. S., and Furukawa, M. (1999). Matrix metalloproteinase 9 is induced by the Epstein–Barr virus BZLF1 transactivator. Clin. Exp. Metastasis, 17(5), 431–436.CrossRefGoogle ScholarPubMed
Young, L. S., Lau, R., Rowe, M.et al. (1991). Differentiation-associated expression of the Epstein–Barr virus BZLF1 transactivator protein in oral hairy leukoplakia. J. Virol., 65(6), 2868–2874.Google ScholarPubMed
Yuan, J., Cahir-McFarland, E., Zhao, B., and Kieff, E. (2006). Virus and cell RNAs expressed during Epstein–Barr virus replicationJ. Virol., 80(5), 2548–2565.CrossRefGoogle ScholarPubMed
Zacny, V. L., Wilson, J., and Pagano, J. S. (1998). The Epstein–Barr virus immediate-early gene product, BRLF1, interacts with the retinoblastoma protein during the viral lytic cycle. J. Virol., 72(10), 8043–8051.Google ScholarPubMed
Zalani, S., Holley-Guthrie, E. A., Gutsch, D. E., and Kenney, S. C. (1992). The Epstein–Barr virus immediate-early promoter BRLF1 can be activated by the cellular Sp1 transcription factor. J. Virol., 66(12), 7282–7292.Google ScholarPubMed
Zalani, S., Holley-Guthrie, E., and Kenney, S. (1995). The Zif268 cellular transcription factor activates expression of the Epstein–Barr virus immediate-early BRLF1 promoter. J. Virol., 69(6), 3816–3823.Google ScholarPubMed
Zalani, S., Holley-Guthrie, E., and Kenney, S. (1996). Epstein–Barr viral latency is disrupted by the immediate-early BRLF1 protein through a cell-specific mechanism. Proc. Natl Acad. Sci. USA, 93(17), 9194–9199.CrossRefGoogle ScholarPubMed
Zalani, S., Coppage, A., Holley-Guthrie, E., and Kenney, S. (1997). The cellular YY1 transcription factor binds a cis-acting, negatively regulating element in the Epstein–Barr virus BRLF1 promoter. J. Virol., 71(4), 3268–3274.Google ScholarPubMed
Zerby, D., Chen, C. J., Poon, E., Lee, D., Shiekhattar, R., and Lieberman, P. M. (1999). The amino-terminal C/H1 domain of CREB binding protein mediates zta transcriptional activation of latent Epstein–Barr virus. Mol. Cell. Biol., 19(3), 1617–1626.CrossRefGoogle ScholarPubMed
Zetterberg, H., Jansson, A., Rymo, L.et al. (2002). The Epstein–Barr virus ZEBRA protein activates transcription from the early lytic F promoter by binding to a promoter-proximal AP-1-like site. J. Gen. Virol., 83(8), 2007–2014.CrossRefGoogle ScholarPubMed
Zhang, Q., Gutsch, D., and Kenney, S. (1994). Functional and physical interaction between p53 and BZLF1: implications for Epstein–Barr virus latency. Mol. Cell. Biol., 14(3), 1929–1938.CrossRefGoogle ScholarPubMed
Zhang, Q., Hong, Y., Dorsky, D.et al. (1996). Functional and physical interactions between the Epstein–Barr virus (EBV) proteins BZLF 1 and BMRF1: Effects on EBV transcription and lytic replication. J. Virol., 70(8), 5131–5142.Google Scholar
Zhang, Q., Holley-Guthrie, E., Ge, J. Q., Dorsky, D., and Kenney, S. (1997). The Epstein–Barr virus (EBV) DNA polymerase accessory protein, BMRF1, activates the essential downstream component of the EBV oriLyt. Virology, 230(1), 22–34.CrossRefGoogle ScholarPubMed
Zhang, Q., Holley-Guthrie, E., Dorsky, D., and Kenney, S. (1999a). Identification of transactivator and nuclear localization domains in the Epstein–Barr virus DNA polymerase accessory protein, BMRF1. J. Gen. Virol., 80(1), 69–74.CrossRefGoogle Scholar
Zhang, Q., Wang, Y. C., and Montalvo, E. A. (1999b). Smubp-2 represses the Epstein–Barr virus lytic switch promoter. Virology, 255(1), 160–170.CrossRefGoogle Scholar
Zimmermann, J. and Hammerschmidt, W. (1995). Structure and role of the terminal repeats of Epstein–Barr virus in processing and packaging of virion DNA. J. Virol., 69(5), 3147–3155.Google ScholarPubMed
Hausen, zur H., Neill, O' F. J., Freese, U. K., and Hecker, E. (1978). Persisting oncogenic herpesvirus induced by the tumour promotor TPA. Nature, 272(5651), 373–375.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×