Skip to main content Accessibility help
×
Hostname: page-component-5c6d5d7d68-thh2z Total loading time: 0 Render date: 2024-08-18T11:24:18.972Z Has data issue: false hasContentIssue false

33 - HSV-1 and 2: molecular basis of HSV latency and reactivation

from Part III - Pathogenesis, clinical disease, host response, and epidemiology: HSV-1 and HSV-2

Published online by Cambridge University Press:  24 December 2009

Chris M. Preston
Affiliation:
Medical Research Council Virology Unit, Scotland, UK
Stacey Efstathiou
Affiliation:
Division of Virology, Department of Pathology, University of Cambridge, UK
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

Introduction

Primary infection with HSV-1 or HSV-2 results in productive replication of the virus at the site of infection, following the pattern of gene expression described elsewhere in this volume. During this initial phase, virus enters sensory neurons via their termini and retrograde transport takes the genome to the neuronal nuclei in the sensory ganglia that innervate the infected dermatome. At early times after infection, virus replication occurs in ganglionic neurons but within a few days no virus can be detected. The genome, however, persists in neurons in a latent state from which it reactivates periodically to resume replication and produce infectious virus. This reactivation event may be “spontaneous” but is generally thought to be provoked by stress stimuli that act on the neuron, or at a peripheral site innervated by the infected ganglion, or systemically. Three phases of latency are recognized. Establishment occurs during the period following primary infection, and although virus replication can be detected in a proportion of neurons during this phase, the initiation and normal progression of productive infection and cell death is arrested in those neurons destined to become latently infected. Unravelling the way in which the seemingly inexorable progression of the gene expression program is blocked constitutes a major challenge for the molecular virologist. The maintenance phase of latency is characterized by the lifelong retention of the HSV genome in a silent state, characterized by repression of all viral lytic genes. One region, encoding the latency-associated transcripts (LATs), remains active during latency.

Type
Chapter
Information
Human Herpesviruses
Biology, Therapy, and Immunoprophylaxis
, pp. 602 - 615
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

Ahmed, M. and Fraser, N. W. (2001). Herpes simplex virus type 1 2-kilobase latency-associated transcript intron associates with ribosomal proteins and splicing factors. J. Virol., 75, 12070–12080.CrossRefGoogle ScholarPubMed
Ahmed, M., Lock, M., Miller, C. G., and Fraser, N. W. (2002). Regions of the herpes simplex virus type 1 latency-associated transcript that protect cells from apoptosis in vitro and protect neuronal cells in vivo. J. Virol., 76, 717–729.CrossRefGoogle ScholarPubMed
Akhova, O., Bainbridge, M. and Misra, V. (2005). The neuronal host cell factor-binding protein Zhangfei inhibits herpes simplex virus replication. J. Virol. 79, 14708–14718.CrossRefGoogle ScholarPubMed
Amelio, A. L., Giordani, N. V., Kubat, N. J., O'Neil, J. E. and Bloom, D. C. (2006). Deacetylation of the herpes simplex virus type 1 latency-associated transcript (LAT) enhancer and a decrease in LAT abundance precede an increase in ICP0 transcriptional permissiveness at early times postexplant. J. Virol. 80, 2063–2068.CrossRefGoogle Scholar
Arthur, J., Efstathiou, S., and Simmons, A. (1993). Intranuclear foci containing low abundance herpes simplex virus latency-associated transcripts visualized by non-isotopic in situ hybridization. J. Gen. Virol., 74, 1363–1370.CrossRefGoogle ScholarPubMed
Arthur, J. L., Scarpini, C. G., Connor, V., Lachmann, R. H., Tolkovsky, A. M., and Efstathiou, S. (2001). Herpes simplex virus type 1 promoter activity during latency establishment, maintenance, and reactivation in primary dorsal root neurons in vitro. J. Virol., 75, 3885–3895.CrossRefGoogle ScholarPubMed
Berthomme, H., Thomas, J., Texier, P., Epstein, A., and Feldman, L. T. (2001). Enhancer and long-term expression functions of herpes simplex virus type 1 latency-associated promoter are both located in the same region. J. Virol., 75, 4386–4393.CrossRefGoogle ScholarPubMed
Bloom, D. C., Hill, J. M., Devi-Rao, G. B., Wagner, E. K., Feldman, L. T., and Stevens, J. G. (1996). A 348-base-pair region in the latency-associated transcript facilitates herpes simplex virus type 1 reactivation. J. Virol., 70, 2449–2459.Google ScholarPubMed
Bloom, D. C., Stevens, J. G., Hill, J. M., and Tran, R. K. (1997). Mutagenesis of a cAMP response element within the latency-associated transcript promoter of HSV-1 reduces adrenergic reactivation. Virology, 236, 202–207.CrossRefGoogle ScholarPubMed
Boutell, C., Sadis, S., and Everett, R. D. (2002). Herpes simplex virus type 1 immediate-early protein ICP0 and its isolated RING finger domain act as ubiquitin E3 ligases in vitro. J. Virol., 76, 841–850.CrossRefGoogle ScholarPubMed
Burton, E. A., Hong, C. S., and Glorioso, J. C. (2003). The stable 2.0-kilobase intron of the herpes simplex virus type 1 latency-associated transcript does not function as an antisense repressor of ICP0 in nonneuronal cells. J. Virol., 77, 3516–3530.CrossRefGoogle Scholar
Cantin, E. M., Hinton, D. R., Chen, J. D. and Openshaw, H. (1995). Gamma interferon expression during acute and latent nervous system infection by herpes simplex virus type 1. J. Virol. 69, 4898–4905.Google ScholarPubMed
Chen, X., Schmidt, M. C., Goins, W. F., and Glorioso, J. C. (1995). Two herpes simplex virus type 1 latency-active promoters differ in their contributions to latency-associated transcript expression during lytic and latent infections. J. Virol., 69, 7899–7908.Google ScholarPubMed
Chen, X.-P., Li, J., Mata, M., Goss, J., Wolfe, D., Glorioso, J. C. and Fink, D. J. (2000). Herpes simplex virus type 1 ICP0 protein does not accumulate in the nucleus of primary neurons in culture. J. Virol., 74, 10132–10141.CrossRefGoogle Scholar
Chen, X.-P., Mata, M., Kelley, M., Glorioso, J. C., and Fink, D. J. (2002). The relationship of herpes simplex virus latency associated transcript expression to genome copy number: a quantitative study using laser capture microdissection. Journal of Neurovirology, 8, 204–210.CrossRefGoogle ScholarPubMed
Cleary, M. A., Stern, S., Tanaka, M., and Herr, W. (1993). Differential positive control by Oct-1 and Oct-2 – activation of a transcriptionally silent motif through Oct-1 and VP16 recruitment. Genes Dev., 7, 72–83.CrossRefGoogle Scholar
Cui, C., Griffiths, A., Li, G. L., Silva, L. M., Kramer, M. F., Gaasterland, T., Wang, X. J. and Coen, D. M. (2006). Prediction and identification of herpes simplex virus 1-encoded microRNAs. J. Virol. 80, 5499–5508.CrossRefGoogle ScholarPubMed
Colgin, M. A., Smith, R. L., and Wilcox, C. L. (2001). Inducible cyclic AMP early repressor produces reactivation of latent herpes simplex virus type 1 in neurons in vitro. J. Virol., 75, 2912–2920.CrossRefGoogle ScholarPubMed
Davido, D. J. and Leib, D. A. (1996). Role of cis-acting sequences of the ICP0 promoter of herpes simplex virus type 1 in viral pathogenesis, latency and reactivation. J. Gen. Virol., 77, 1853–1863.CrossRefGoogle Scholar
Davido, D. J., Leib, D. A., and Schaffer, P. A. (2002). The cyclin-dependent kinase inhibitor roscovitine inhibits the transactivating activity and alters the posttranslational modification of herpes simplex virus type 1 ICP0. J. Virol., 76, 1077–1088.CrossRefGoogle ScholarPubMed
Deshmane, S. L. and Fraser, N. W. (1989). During latency, herpes simplex virus type 1 DNA is associated with nucleosomes in a chromatin structure. J. Virol., 63, 943–947.Google Scholar
Devireddy, L. R. and Jones, C. J. (2000). Olf-1, a neuron-specific transcription factor, can activate the herpes simplex virus type 1-infected cell protein 0 promoter. J. Biol. Chem., 275, 77–81.CrossRefGoogle ScholarPubMed
Dobson, A., Sederati, F., Devi-Rao, G., Flanagan, W. M., Farrell, M., Stevens, J., Wagner, E., and Feldman, L. T. (1989). Identification of the latency-associated transcript promoter by expression of rabbit beta-globin mRNA in sensory nerve ganglia latently infected with a recombinant herpes simplex virus. J. Virol., 63: 3844–3851.Google ScholarPubMed
Drolet, B. S., Perng, G. C., Cohen, J.et al., (1998). The region of herpes simplex virus type 1 LAT gene involved in spontaneous reactivation does not encode a functional protein. Virology, 242, 221–232.CrossRefGoogle Scholar
Efstathiou, S., Minson, A. C., Field, H. J., Anderson, J. R. and Wildy, P. (1986). Detection of herpes simplex virus-specific DNA sequences in latently infected mice and in humans. J. Virol. 57, 446–455.Google ScholarPubMed
Everett, R. D. (2000). ICP0, a regulator of herpes simplex virus during lytic and latent infection. Bioessays, 22, 761–770.3.0.CO;2-A>CrossRefGoogle ScholarPubMed
Feldman, L. T., Ellison, A. R., Voytek, C. C., Yang, L., Krause, P., and Margolis, T. (2002). Spontaneous molecular reactivation of herpes simplex virus type 1 latency in mice. Proc. Natl Acad. Sci. USA, 99, 978–983.CrossRefGoogle ScholarPubMed
Frazier, D. P., Cox, D., Godshalk, E. M., and Schaffer, P. A. (1996). Identification of cis-acting sequences in the promoter of the herpes simplex virus type 1 latency-associated transcripts required for activation by nerve growth factor and sodium butyrate in PC12 cells. J Virol., 70, 7433–7444.Google ScholarPubMed
Freiman, R. N. and Herr, W. (1997). Viral mimicry: common mode of association with HCF by VP16 and the cellular protein LZIP. Genes Dev., 11, 3122–3127.CrossRefGoogle ScholarPubMed
Garber, D. A., Schaffer, P. A., and Knipe, D. M. (1997). A LAT-associated function reduces productive-cycle gene expression during acute infection of murine sensory neurons with herpes simplex virus type 1. J. Virol., 71, 5885–5893.Google ScholarPubMed
Gu, H., Liang, Y., Mandel, G. and Roizman, B. (2005). Components of the REST/CoREST/histone deacetylase repressor complex arre disrupted, modified, and translocated in HSV-1-infected cells. Proc Natl Acad Sci U S A, 102, 7571–7576.CrossRefGoogle Scholar
Gupta, A., Gartner, J. J., Sethupathy, P., Hatzigeorgiou, A. G. and Fraser, N. W. (2006). Anti-apoptotic function of a microRNA encoded by the HSV-1 latency-associated transcript. Nature 442, 82–85.CrossRefGoogle ScholarPubMed
Hagmann, M., Georgiev, O., Schaffner, W., and Douville, P. (1995). Transcription factors interacting with herpes simplex virus α gene promoters in sensory neurons. Nucl. Acids Res., 23, 4978–4985.CrossRefGoogle ScholarPubMed
Halford, W. P. and Schaffer, P. A. (2001). ICP0 is required for efficient reactivation of herpes simplex virus type 1 from neuronal latency. J. Virol., 75, 3240–3249.CrossRefGoogle ScholarPubMed
Halford, W. P., Gebhardt, B. M., and Carr, D. J. (1996). Mechanisms of herpes simplex virus type 1 reactivation. J. Virol., 70, 5051–5060.Google ScholarPubMed
Herrera, F. J. and Triezenberg, S. J. (2004). VP16-dependent association of chromatin-modifying coactivators and underrepresentation of histones at immediate-early gene promoters during herpes simplex virus infection. J. Virol. 78, 9689–9696.CrossRefGoogle ScholarPubMed
Inman, M., Perng, G. -C., Henderson, G., Ghiasi, H., Nesburn, A. B., Wechsler, S. L., and Jones, C. (2001). Region of herpes simplex virus type 1 latency-associated transcript sufficient for wild-type spontaneous reactivation promotes cell survival in tissue culture. J. Virol., 75, 3636–3646.CrossRefGoogle ScholarPubMed
Jackson, S. A. and DeLuca, N. A. (2003). Relationship of herpes simplex virus genome configuration to productive and persistent infections. Proc. Natl Acad. Sci. USA, 100, 7871–7876.CrossRefGoogle ScholarPubMed
Jamieson, D. R. S., Robinson, L. H., Daksis, J. I., Nicholl, M. J., and Preston, C. M. (1995). Quiescent viral genomes in human fibroblasts after infection with herpes simplex virus Vmw65 mutants. J. Gen. Virol., 76, 1417–1431.CrossRefGoogle ScholarPubMed
Jin, L., Perng, G. C., Mott, K. R.et al., (2005). A herpes simplex virus type 1 mutant expressing a baculovirus inhibitor of apoptosis gene in place of latency-associated transcript has a wild-type reactivation phenotype in the mouse. J. Virol. 79, 12286–12295.CrossRefGoogle Scholar
Kent, J. R., Zeng, P.-Y., Atanasiu, D., Gardner, J., Fraser, N. W. and Berger, S. L. (2004). During lytic infection herpes simplex virus type 1 is associated with histones bearing modifications that correlate with active transcription. J. Virol. 78, 10178–10186.CrossRefGoogle ScholarPubMed
Khanna, K. M., Bonneau, R. H., Kinchington, P. R., and Hendricks, R. L. (2003). Herpes simplex virus -specific memory CD8+ T cells are selectively activated and retained in latently infected sensory ganglia. Immunity, 18, 593–603.CrossRefGoogle ScholarPubMed
Kramer, M. F., Chen, S. -H., Knipe, D. M., and Coen, D. M. (1998). Accumulation of viral transcripts and DNA during establishment of latency by herpes simplex virus. J. Virol., 72, 1177–1185.Google ScholarPubMed
Kramer, M. F., Cook, W. J., Roth, F. P.et al., (2003). Latent herpes simplex virus infection of sensory neurons alters neuronal gene expression. J. Virol., 77, 9533–9541.CrossRefGoogle ScholarPubMed
Kristie, T. M. and Roizman, B. (1988). Differentiation and DNA contact points of the host proteins binding at the cis site for virion-mediated induction of herpes simplex virus 1 α genes. J. Virol., 62, 1145–1157.Google Scholar
Kristie, T. M., Vogel, J. L., and Sears, A. E. (1999). Nuclear localization of the C1 factor (host cell factor) in sensory neurons correlates with reactivation of herpes simplex virus from latency. Proc. Natl Acad. Sci. USA, 96, 1229–1233.CrossRefGoogle ScholarPubMed
Kubat, N. J., Amelio, A. L., Giordani, N. V. and Bloom, D. C. (2004a). The herpes simplex virus type 1 latency-associated transcript (LAT) enhancer/rcr is hyperacetylated during latency independently of LAT transcription. J. Virol. 78, 12508–12518.CrossRefGoogle Scholar
Kubat, N. J., Tran, R. K., McAnany, P. K. and Bloom, D. C. (2004b). Specific histone tail modification and not DNA methylation is a determinant of herpes simplex virus type 1 latent gene expression. J. Virol. 78, 1139–1149.CrossRefGoogle Scholar
Lachmann, R. H. and Efstathiou, S. (1997). Utilization of the herpes simplex virus type 1 latency-associated regulatory region to drive stable reporter gene expression in the nervous system. J. Virol., 71, 3197–3207.Google Scholar
Latchman, D. S. (1999). POU family transcription factors in the nervous system. J. Cell. Physiol., 80, 1271–1282.Google Scholar
Lociano, R. L. and Wilson, A. C. (2002). An activation domain in the C-terminal subunit of HCF-1 is important for transactivation by VP16 and LZIP. Proc. Natl Acad. Sci. USA, 99, 13403–13408.CrossRefGoogle Scholar
Lu, R. and Misra, V. (2000). Potential role for luman, the cellular homologue of herpes simplex virus VP16 (alpha gene trans-inducing factor), in herpesvirus latency. J. Virol., 74, 934–943.CrossRefGoogle Scholar
Lomonte, P., Thomas, J., Texier, P., Caron, C., Khochbin, S. and Epstein, A. L. (2004). Functional interaction between class II histone deacetylases and ICP0 of herpes simplex type 1. J. Virol. 78, 6744–6757.CrossRefGoogle Scholar
Mador, N., Goldenberg, D., Cohen, O., Panet, A., and Steiner I., (1998). Herpes simplex virus type 1 latency-associated transcripts suppress viral replication and reduce immediate-early gene mRNA levels in a neuronal cell line. J. Virol., 72, 5067–5075.Google Scholar
Margolis, T., Sederati, F., Dobson, A. T., Feldman, L. T., and Stevens, J. G. (1992). Pathways of viral gene expression during acute neuronal infection with HSV-1. Virology, 189, 150–160.CrossRefGoogle ScholarPubMed
Marquart, M. E., Zheng, X., Tran, R. K., Thompson, H. W., Bloom, D. C., and Hill, J. M. (2001). A cAMP response element within the latency-associated transcript promoter of HSV-1 facilitates induced ocular reactivation in a mouse hyperthermia model. Virology, 284, 62–69.CrossRefGoogle Scholar
Marshall, K. R., Lachmann, R. H., Efstathiou, S., Rinaldi, A., and Preston, C. M. (2000). Long-term transgene expression in mice infected with a herpes simplex virus type 1 mutant severely impaired for immediate-early gene expression. J. Virol., 74, 956–964.CrossRefGoogle ScholarPubMed
Nicosia, M., Deshmane, S. L., Zabolotny, J. M., Valyi-Nagy, T., and Fraser, N. W. (1993). Herpes simplex virus type 1 latency-associated transcript (LAT) promoter deletion mutants can express a 2-kilobase transcript mapping to the LAT region. J. Virol., 67, 7276–7283.Google ScholarPubMed
O'Reilly, D., Hanscombe, O., and O'Hare, P. (1997). A single serine residue at position 375 of VP16 is critical for complex assembly with Oct-1 and HCF and is a target of phosphorylation by casein kinase II. EMBO J., 16, 2420–2430.CrossRefGoogle ScholarPubMed
O'Rourke, D. and O'Hare, P. (1993). Mutually exclusive binding of two cellular factors within a critical promoter region of the gene for the IE110k protein of herpes simplex virus. J. Virol., 67, 7201–7214.Google ScholarPubMed
Perng, G., Ghiasi, H., Slanina, S., Nesburn, A., and Wechsler, S. (1996). The spontaneous reactivation function of the herpes simplex virus type 1 LAT gene resides completely within the first 1.5 kilobases of the 8.3- kilobase primary transcript. J. Virol., 70, 976–984.Google ScholarPubMed
Perng, G., Jones, C., Ciacci-Zanella, J.et al., (2000). Virus-induced neuronal apoptosis blocked by the herpes simplex virus latency-associated transcript. Science, 287, 1500–1503.CrossRefGoogle ScholarPubMed
Pfeffer, S., Sewer, A., Lagos-Quintana, M., Sheridan, R., Sander, C., Grasser, F. A., Dyk, L. F., Ho, C. K., Shuman, S., Chien, M., Fusso, J. J., Ju, J., Randall, G., Lindenbach, B. D., Rice, C. M., Simon, V., Ho, D. D., Zavolan, M. and Tuschl, T. (2005). Identification of microRNAs of the herpesvirus family. Nature Methods 2, 269–276.CrossRefGoogle ScholarPubMed
Preston, C. M. and Nicholl, M. J. (1997). Repression of gene expression upon infection of cells with herpes simplex virus type 1 mutants impaired for immediate-early protein synthesis. J. Virol., 71, 7807–7813.Google ScholarPubMed
Raggo, C., Rapin, N., Stirling, J.et al., (2002). Luman, the cellular counterpart of herpes simplex virus VP16, is processed by regulated intramembrane proteolysis. Mol. Cell. Biol., 22, 5639–5649.CrossRefGoogle ScholarPubMed
Ramakrishnan, R., Poliani, P. L., Levine, M., Glorioso, J. C., and Fink, D. J. (1996). Detection of herpes simplex virus type 1 latency-associated transcript expression in trigeminal ganglia by in situ reverse transcriptase PCR. J. Virol., 70, 6519–6523.Google ScholarPubMed
Rock, D. L. and Fraser, N. W. (1983). Detection of HSV-1 genome in central nervous system of latently infected mice. Nature, 302, 523–525.CrossRefGoogle ScholarPubMed
Samaniego, L. A., Neiderhiser, L., and DeLuca, N. A. (1998). Persistence and expression of the herpes simplex virus genome in the absence of immediate-early proteins. J. Virol., 72, 3307–3320.Google ScholarPubMed
Sawtell, N. M. (1997). Comprehensive quantification of herpes simplex virus latency at the single-cell level. J. Virol., 71, 5423–5431.Google ScholarPubMed
Sawtell, N. M. (1998). The probability of in vivo reactivation of herpes simplex virus type 1 increases with the number of latently infected neurons in the ganglia. J. Virol., 72, 6888–6892.Google ScholarPubMed
Sawtell, N. M. and Thompson, R. L. (1992a). Herpes simplex virus type 1 latency-associated transcription unit promotes anatomical site-dependent establishment and reactivation from latency. J. Virol., 66, 2157–2169.Google Scholar
Sawtell, N. M. and Thompson, R. L. (1992b). Rapid in vivo reactivation of herpes simplex virus in latently infected murine ganglionic neurons after transient hyperthermia. J. Virol., 66, 2150–2156.Google Scholar
Simmons, A., Slobedman, B., Speck, P., Arthur, J., and Efstathiou, S. (1992). Two patterns of persistence of herpes simplex virus DNA sequences in the nervous system of latently infected mice. J. Gen. Virol., 73: 1287–1291.CrossRefGoogle ScholarPubMed
Smith, R. L., Escudero, J. M., and Wilcox, C. L. (1992). Activation of second messenger pathways activates latent herpes simplex virus in neuronal cultures. Virology, 188, 311–318.CrossRefGoogle Scholar
Stevens, J., Wagner, E., Devi-Rao, G. B., Cook, M. L., and Feldman, L. T. (1987). RNA complementary to a herpesvirus alpha gene mRNA is predominant in latently infected neurons. Science, 235, 1056–1059.CrossRefGoogle Scholar
Stevens, J. G. and Cook, M. L. (1971). Latent herpes simplex virus in spinal ganglia of mice. Science, 173, 843–845.CrossRefGoogle Scholar
Thomas, S. K., Lilley, C. E., Latchman, D. S., and Coffin, R. S. (2002). A protein encoded by the herpes simplex virus (HSV) type 1 2-kilobase latency-associated transcript is phosphorylated, localized to the nucleus, and overcomes the repression of expression from exogenous promoters when inserted into the quiescent HSV genome. J. Virol., 76, 4056–4067.CrossRefGoogle ScholarPubMed
Thompson, R. L. and Sawtell, N. M. (1997). The herpes simplex virus type 1 latency-associated transcript gene regulates the establishment of latency. J. Virol., 71, 5432–5440.Google ScholarPubMed
Thompson, R. L. and Sawtell, N. M. (2000). Replication of herpes simplex virus type 1 within trigeminal ganglia is required for high frequency but not high viral genome copy number latency. J. Virol., 74, 965–974.CrossRefGoogle Scholar
Thompson, R. L. and Sawtell, N. M. (2001). Herpes simplex virus type 1 latency-associated transcript gene promotes neuronal survival. J. Virol., 75, 6660–6675.CrossRefGoogle ScholarPubMed
Tsavachidou, D., Podrzucki, W., Seykora, J., and Berger, S. L. (2001). Gene array analysis reveals changes in peripheral nervous system gene expression following stimuli that result in reactivation of latent herpes simplex virus type 1: induction of transcription factor Bcl-3. J. Virol., 75, 9909–9917.CrossRefGoogle ScholarPubMed
Sant, C., Hagglund, R., Lopez, P., and Roizman, B. (2001). The infected cell protein 0 of herpes simplex virus 1 dynamically interacts with proteasomes, binds and activates the cdc34 E2 ubiquitin-conjugating enzyme, and possesses in vitro E3 ubiquitin ligase activity. Proc. Natl Acad. Sci. USA, 98, 8815–8820.CrossRefGoogle Scholar
Vogel, J. L. and Kristie, T. M. (2000). The novel coactivator C1 (HCF) coordinates multiprotein enhancer formation and mediates transcription activation by GABP. EMBO J., 19, 683–690.CrossRefGoogle ScholarPubMed
Wald, A., Corey, L., Cone, R., Hobson, A., Davis, G., and Zeh, J. (1997). Frequent genital herpes simplex virus 2 shedding in immunocompetent women. Effect of acyclovir treatment. J. Clin. Investig., 99, 1092–1097.CrossRefGoogle ScholarPubMed
Wang, Q. Y., Zhou, C. H., Johnson, K. E., Colgrove, R. C., Coen, D. M. and Knipe, D. M. (2005). Herpesviral latency-associated transcript gene promotes assembly of heterochromatin on viral lytic-gene promoters in latent infection. Proc. Natl. Acad. Sci. USA 102, 16055–16059.CrossRefGoogle ScholarPubMed
Wilcox, C. L. and Johnson, E. (1988). Characterization of nerve growth factor-dependent herpes simplex virus latency in neurons in vitro. J. Virol., 62, 393–399.Google ScholarPubMed
Wilson, A. C., LaMarco, K., Peterson, M. G., and Herr, W. (1993). The VP16 accessory protein HCF-1 is a family of polypeptides processed from a large precursor protein. Cell, 74, 115–125.CrossRefGoogle Scholar
Yoshikawa, T., Hill, J. M., Stanberry, L. R., Bourne, N., Kurawadwala, J. F., and Krause, P. R. (1996). The characteristic site-specific reactivation phenotypes of HSV-1 and HSV-2 depend upon the latency-associated transcript region. J. Exp. Med., 184, 659–664.CrossRefGoogle ScholarPubMed
Zabolotny, J. M., Krummenacher, C., and Fraser, N. W. (1997). The herpes simplex virus type 1 2.0-kilobase latency-associated transcript is a stable intron which branches at a guanosine. J. Virol., 71, 4199–4208.Google Scholar

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
×