Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-wq2xx Total loading time: 0 Render date: 2024-04-23T06:43:07.015Z Has data issue: false hasContentIssue false

8 - Early events pre-initiation of alphaherpes viral gene expression

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

Published online by Cambridge University Press:  24 December 2009

Thomas M. Kristie
Affiliation:
National Institute of Allergy and Infectious Diseases, National Institutes of Health, Bethesda, MD, 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

The regulated transcription of the HSV IE (immediate–early, α) genes has been a model system for elucidating principles and mechanisms of combinatorial-differential regulation, basic RNAPII -directed transcription, and multiprotein assembly specificities. The regulation exemplifies viral mechanisms dedicated to the recruitment of cellular components into complex viral–host interactions that illustrate general parameters of protein–protein, DNA –protein, RNA transcription, and protein complex assembly. Continued studies hold promise of advancing the understanding of the complexities of biochemical interactions in gene expression as well as complex cellular response pathways. The regulation of the IE genes within specific contexts may also lead to the understanding of signals and pathways which modulate viral infection and determine the extent of lytic-latent infection. While HSV has been extensively studied and will represent the focus of this review, the regulatory domain of the VZV IE gene (IE62) contains similar elements and is regulated by similar mechanisms.

The HSV IE regulatory domains: multiple sites for differential regulation

The regulatory domains of the HSV IE genes have been the focus of numerous studies that have defined the sequence elements and their contributions to the basal and induced levels of transcription. These IE domains typically consist of a reiterated inducible enhancer core element (consensus: TAATGARAT) that is flanked by binding sites for members of the ets and kruppel transcription family (Fig. 8.1, left) (Roizman and Sears, 1996; Vogel and Kristie, 2001).

Type
Chapter
Information
Human Herpesviruses
Biology, Therapy, and Immunoprophylaxis
, pp. 112 - 127
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

Adams, J., Kelso, R., and Cooley, L. (2000). The kelch repeat superfamily of proteins: propellers of cell function. Trends Cell Biol., 10, 17–24.CrossRefGoogle ScholarPubMed
Ajuh, P. M., Browne, G. J., Hawkes, N. A., Cohen, P. T., Roberts, S. G., and Lamond, A. I. (2000). Association of a protein phosphatase 1 activity with the human factor C1 (HCF) complex. Nucl. Acids Res., 28, 678–686.CrossRefGoogle ScholarPubMed
Ajuh, P., Chusainow, J., Ryder, U., and Lamond, A. I. (2002). A novel function for human factor C1 (HCF-1), a host protein required for herpes simplex virus infection, in pre-mRNA splicing. EMBO J., 21, 6590–6602.CrossRefGoogle Scholar
Aurrekoetxea-Hernandez, K., and Buetti, E. (2000). Synergistic action of GA-binding protein and glucocorticoid receptor in transcription from the mouse mammary tumor virus promoter. J. Virol., 74, 4988–4998.CrossRefGoogle Scholar
Babb, R., Cleary, M. A., and Herr, W. (1997). OCA-B is a functional analog of VP16 but targets a separate surface of the Oct-1 POU domain. Mol. Cell Biol., 17, 7295–7305.CrossRefGoogle ScholarPubMed
Babb, R., Huang, C. C., Aufiero, D. J., and Herr, W. (2001). DNA recognition by the herpes simplex virus transactivator VP16: a novel DNA-binding structure. Mol. Cell Biol., 21, 4700–4712.CrossRefGoogle ScholarPubMed
Bannert, N., Avots, A., Baier, M., Serfling, E., and Kurth, R. (1999). GA-binding protein factors, in concert with the coactivator CREB binding protein/p300, control the induction of the interleukin 16 promoter in T lymphocytes. Proc. Natl Acad. Sci. USA, 96, 1541–1546.CrossRefGoogle ScholarPubMed
Black, A. R., Black, J. D., and Azizkhan-Clifford, J. (2001). Sp1 and kruppel-like factor family of transcription factors in cell growth regulation and cancer. J. Cell Physiol., 188, 143–160.CrossRefGoogle ScholarPubMed
Boulon, S., Dantonel, J. C., Binet, V.et al. (2002). Oct-1 potentiates CREB-driven cyclin D1 promoter activation via a phospho-CREB- and CREB binding protein-independent mechanism. Mol. Cell Biol., 22, 7769–7779.CrossRefGoogle Scholar
Brownlees, J., Gough, G., Thomas, S.et al. (1999). Distinct responses of the herpes simplex virus and varicella zoster virus immediate early promoters to the cellular transcription factors Brn-3a and Brn-3b. Int. J. Biochem. Cell Biol., 31, 451–461.CrossRefGoogle ScholarPubMed
Chasman, D., Cepek, K., Sharp, P. A., and Pabo, C. O. (1999). Crystal structure of an OCA-B peptide bound to an Oct-1 POU domain/octamer DNA complex: specific recognition of a protein–DNA interface. Genes Dev., 13, 2650–2657.CrossRefGoogle Scholar
Clerc, R. G., Corcoran, L. M., LeBowitz, J. H., Baltimore, D., and Sharp, P. A. (1988). The B-cell-specific Oct-2 protein contains POU box- and homeo box-type domains. Genes Dev., 2, 1570–1581.CrossRefGoogle ScholarPubMed
Coenjaerts, F. E., Oosterhout, J. A., and Vliet, P. C. (1994). The Oct-1 POU domain stimulates adenovirus DNA replication by a direct interaction between the viral precursor terminal protein-DNA polymerase complex and the POU homeodomain. EMBO J., 13, 5401–5409.Google ScholarPubMed
Cosma, M. P. (2002). Ordered recruitment: gene-specific mechanism of transcription activation. Mol. Cell, 10, 227–236.CrossRefGoogle ScholarPubMed
Jong, R. N., Mysiak, M. E., Meijer, L. A., Linden, M., and Vliet, P. C. (2002). Recruitment of the priming protein pTP and DNA binding occur by overlapping Oct-1 POU homeodomain surfaces. EMBO J., 21, 725–735.CrossRefGoogle ScholarPubMed
Brousse, F. C., Birkenmeier, E. H., King, D. S., Rowe, L. B., and McKnight, S. L. (1994). Molecular and genetic characterization of GABP beta. Genes Dev., 8, 1853–1865.CrossRefGoogle ScholarPubMed
Flory, E., Hoffmeyer, A., Smola, U., Rapp, U. R., and Bruder, J. T. (1996). Raf-1 kinase targets GA-binding protein in transcriptional regulation of the human immunodeficiency virus type 1 promoter. J. Virol., 70, 2260–2268.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
Fromm, L. and Burden, S. J. (1998). Synapse-specific and neuregulin-induced transcription require an ets site that binds GABPalpha/GABPbeta. Genes Dev., 12, 3074–3083.CrossRefGoogle ScholarPubMed
Gold, M. O., Tassan, J. P., Nigg, E. A., Rice, A. P., and Herrmann, C. H. (1996). Viral transactivators E1A and VP16 interact with a large complex that is associated with CTD kinase activity and contains CDK8. Nucl. Acids Res., 24, 3771–3777.CrossRefGoogle ScholarPubMed
Gonzalez, M. I. and Robins, D. M. (2001). Oct-1 preferentially interacts with androgen receptor in a DNA-dependent manner that facilitates recruitment of SRC-1. J. Biol. Chem., 276, 6420–6428.CrossRefGoogle Scholar
Goto, H., Motomura, S., Wilson, A. C.et al. (1997). A single-point mutation in HCF causes temperature-sensitive cell-cycle arrest and disrupts VP16 function. Genes Dev., 11, 726–737.CrossRefGoogle ScholarPubMed
Grenfell, S. J., Latchman, D. S., and Thomas, N. S. (1996). Oct-1 [corrected] and Oct-2 DNA-binding site specificity is regulated in vitro by different kinases. Biochem J., 315(3), 889–893.CrossRefGoogle ScholarPubMed
Gugneja, S., Virbasius, J. V., and Scarpulla, R. C. (1995). Four structurally distinct, non-DNA-binding subunits of human nuclear respiratory factor 2 share a conserved transcriptional activation domain. Mol. Cell Biol., 15, 102–111.CrossRefGoogle Scholar
Gugneja, S., Virbasius, C. M., and Scarpulla, R. C. (1996). Nuclear respiratory factors 1 and 2 utilize similar glutamine-containing clusters of hydrophobic residues to activate transcription. Mol. Cell Biol., 16, 5708–5716.CrossRefGoogle ScholarPubMed
Gunther, M., Laithier, M., and Brison, O. (2000). A set of proteins interacting with transcription factor Sp1 identified in a two-hybrid screening. Mol. Cell Biochem., 210, 131–142.CrossRefGoogle Scholar
Hall, D. B. and Struhl, K. (2002). The VP16 activation domain interacts with multiple transcriptional components as determined by protein-protein cross-linking in vivo. J. Biol. Chem., 277, 46043–46050.CrossRefGoogle ScholarPubMed
Herrera, F. J. and Triezenberg, S. J. (2004). VP16-dependent association of chromatin-modifying coactivators and underrpresentation of histones at immediate-early gene promoters during herpes simplex virus infection. J. Virol., 78(18), 9689–9696.CrossRefGoogle ScholarPubMed
Herrmann, C. H., Gold, M. O., and Rice, A. P. (1996). Viral transactivators specifically target distinct cellular protein kinases that phosphorylate the RNA polymerase II C-terminal domain. Nucl. Acids Res., 24, 501–508.CrossRefGoogle ScholarPubMed
Hoffmeyer, A., Avots, A., Flory, E., Weber, C. K., Serfling, E., and Rapp, U. R. (1998). The GABP-responsive element of the interleukin-2 enhancer is regulated by JNK/SAPK-activating pathways in T lymphocytes. J. Biol. Chem., 273, 10112–10119.CrossRefGoogle ScholarPubMed
Hovde, S., Hinkley, C. S., Strong, K.et al. (2002). Activator recruitment by the general transcription machinery: X-ray structural analysis of the Oct-1 POU domain/human U1 octamer/SNAP190 peptide ternary complex. Genes Dev., 16, 2772–2777.CrossRefGoogle ScholarPubMed
Hughes, T. A., Boissiere, S., and O'Hare, P. (1999). Analysis of functional domains of the host cell factor involved in VP16 complex formation. J. Biol. Chem., 274, 16437–16443.CrossRefGoogle ScholarPubMed
Inamoto, S., Segil, N., Pan, Z. Q., Kimura, M., and Roeder, R. G. (1997). The cyclin-dependent kinase-activating kinase (CAK) assembly factor, MAT1, targets and enhances CAK activity on the POU domains of octamer transcription factors. J. Biol. Chem., 272, 29852–29858.CrossRefGoogle ScholarPubMed
Jin, S., Fan, F., Fan, W.et al. (2001). Transcription factors Oct-1 and NF-YA regulate the p53-independent induction of the GADD45 following DNA damage. Oncogene, 20, 2683–2690.CrossRefGoogle ScholarPubMed
Johnson, K. M., Mahajan, S. S., and Wilson, A. C. (1999). Herpes simplex virus transactivator VP16 discriminates between HCF-1 and a novel family member, HCF-2. J. Virol., 73, 3930–3940.Google Scholar
Jones, K. A. and Tjian, R. (1985). Sp1 binds to promoter sequences and activates herpes simplex virus ‘immediate-early’ gene transcription in vitro. Nature, 317, 179–182.CrossRefGoogle ScholarPubMed
Julien, E. and Herr, W. (2003). Proteolytic processing is necessary to separate and ensure proper cell growth and cytokinesis functions of HCF-1. Embo J., 22, 2360–2369.CrossRefGoogle ScholarPubMed
Kaczynski, J., Cook, T., and Urrutia, R. (2003). Sp1- and Kruppel-like transcription factors. Genome Biol., 4, 206.CrossRefGoogle ScholarPubMed
Khurana, B. and Kristie, T. M. (2004). A protein sequestering system reveals control of cellular programs by the transcriptional coactivator HCF-1. J. Biol. Chem., 279(32), 33673–33683.CrossRefGoogle ScholarPubMed
Kim, D. B. and DeLuca, N. A. (2002). Phosphorylation of transcription factor Sp1 during herpes simplex virus type 1 infection. J. Virol., 76, 6473–6479.CrossRefGoogle ScholarPubMed
Kinchington, P. R., Hougland, J. K., Arvin, A. M., Ruyechan, W. T., and Hay, J. (1992). The varicella-zoster virus immediate–early protein IE62 is a major component of virus particles. J. Virol., 66, 359–366.Google ScholarPubMed
Klemm, J. D., Rould, M. A., Aurora, R., Herr, W., and Pabo, C. O. (1994). Crystal structure of the Oct-1 POU domain bound to an octamer site: DNA recognition with tethered DNA-binding modules. Cell, 77, 21–32.CrossRefGoogle Scholar
Knez, J., Piluso, D.et al. (2006). Host cell factor-1 and E2F4 intract via multiple determinants in each protein. Mol. Cell Biochem. Apr 22; [Epub ahead of print].CrossRefGoogle Scholar
Kristie, T. M. and Sharp, P. A. (1990). Interactions of the Oct-1 POU subdomains with specific DNA sequences and with the HSV alpha-trans-activator protein. Genes Dev., 4, 2383–2396.CrossRefGoogle ScholarPubMed
Kristie, T. M. and Sharp, P. A. (1993). Purification of the cellular C1 factor required for the stable recognition of the Oct-1 homeodomain by the herpes simplex virus alpha-trans-induction factor (VP16). J. Biol. Chem., 268, 6525–6534.Google Scholar
Kristie, T. M., LeBowitz, J. H., and Sharp, P. A. (1989). The octamer-binding proteins form multi-protein–DNA complexes with the HSV alpha TIF regulatory protein. Embo J., 8, 4229–4238.Google ScholarPubMed
Kristie, T. M., Pomerantz, J. L., Twomey, T. C., Parent, S. A., and Sharp, P. A. (1995). The cellular C1 factor of the herpes simplex virus enhancer complex is a family of polypeptides. J. Biol. Chem., 270, 4387–4394.CrossRefGoogle ScholarPubMed
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
Krumm, A., Hickey, L. B., and Groudine, M. (1995). Promoter-proximal pausing of RNA polymerase II defines a general rate-limiting step after transcription initiation. Genes Dev., 9, 559–572.CrossRefGoogle ScholarPubMed
Boissiere, S., Walker, S., and O'Hare, P. (1997). Concerted activity of host cell factor subregions in promoting stable VP16 complex assembly and preventing interference by the acidic activation domain. Mol. Cell Biol., 17, 7108–7118.CrossRefGoogle Scholar
Boissiere, S., Hughes, T., and O'Hare, P. (1999). HCF-dependent nuclear import of VP16. EMBO J., 18, 480–489.CrossRefGoogle ScholarPubMed
Lai, J. S. and Herr, W. (1997). Interdigitated residues within a small region of VP16 interact with Oct-1, HCF, and DNA. Mol. Cell Biol., 17, 3937–3946.CrossRefGoogle ScholarPubMed
Lai, J. S., Cleary, M. A., and Herr, W. (1992). A single amino acid exchange transfers VP16-induced positive control from the Oct-1 to the Oct-2 homeo domain. Genes Dev., 6, 2058–2065.CrossRefGoogle ScholarPubMed
LaMarco, K., Thompson, C. C., Byers, B. P., Walton, E. M., and McKnight, S. L. (1991). Identification of Ets- and notch-related subunits in GA binding protein. Science, 253, 789–792.CrossRefGoogle ScholarPubMed
Latchman, D. S. (1996). The Oct-2 transcription factor. Int. J. Biochem. Cell Biol., 28, 1081–1083.CrossRefGoogle ScholarPubMed
Latchman, D. S. (1999). POU family transcription factors in the nervous system. J. Cell Physiol., 179, 126–133.3.0.CO;2-M>CrossRefGoogle Scholar
Lillycrop, K. A., Dawson, S. J., Estridge, J. K., Gerster, T., Matthias, P., and Latchman, D. S. (1994). Repression of a herpes simplex virus immediate–early promoter by the Oct-2 transcription factor is dependent on an inhibitory region at the N terminus of the protein. Mol. Cell Biol., 14, 7633–7642.CrossRefGoogle Scholar
Lin, J., Puigserver, P., Donovan, J., Tarr, P., and Spiegelman, B. M. (2002). Peroxisome proliferator-activated receptor gamma coactivator 1beta (PGC-1beta), a novel PGC-1-related transcription coactivator associated with host cell factor. J. Biol. Chem., 277, 1645–1648.CrossRefGoogle Scholar
Liu, Y., Gong, W., Huang, C. C., Herr, W., and Cheng, X. (1999). Crystal structure of the conserved core of the herpes simplex virus transcriptional regulatory protein VP16. Genes Dev., 13, 1692–1703.CrossRefGoogle ScholarPubMed
Lu, R. and Misra, V. (2000a). 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
Lu, R. and Misra, V. (2000b). Zhangfei: a second cellular protein interacts with herpes simplex virus accessory factor HCF in a manner similar to Luman and VP16. Nucl. Acids Res., 28, 2446–2454.CrossRefGoogle Scholar
Lu, R., Yang, P., O'Hare, P., and Misra, V. (1997). Luman, a new member of the CREB/ATF family, binds to herpes simplex virus VP16-associated host cellular factor. Mol. Cell Biol., 17, 5117–5126.CrossRefGoogle ScholarPubMed
Lu, R., Yang, P., Padmakumar, S., and Misra, V. (1998). The herpesvirus transactivator VP16 mimics a human basic domain leucine zipper protein, luman, in its interaction with HCF. J. Virol., 72, 6291–6297.Google ScholarPubMed
Luciano, 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 ScholarPubMed
Luciano, R. L. and Wilson, A. C. (2003). HCF-1 function as a coactivator for the zinc finger protein Krox20. J. Biol. Chem., 278(51), 51116–51124.CrossRefGoogle ScholarPubMed
Mahajan, S. S. and Wilson, A. C. (2000). Mutations in host cell factor 1 separate its role in cell proliferation from recruitment of VP16 and LZIP. Mol. Cell Biol., 20, 919–928.CrossRefGoogle ScholarPubMed
Mahajan, S. S.Little, M. M., Vazquez, R., and Wilson, A. C. (2002). Interaction of HCF-1 with a cellular nuclear export factor. J. Biol. Chem., 277, 44292–44299.CrossRefGoogle ScholarPubMed
McKee, T. A. and Preston, C. M. (1991). Identification of two protein binding sites within the varicella-zoster virus major immediate early gene promoter. Virus Res., 20, 59–69.CrossRefGoogle ScholarPubMed
Memedula, S. and Belmont, A. S. (2003). Sequential recruitment of HAT and SWI/SNF components to condensed chromatin by VP16. Curr. Biol., 13, 241–246.CrossRefGoogle ScholarPubMed
Mittal, V., Ma, B., and Hernandez, N. (1999). SNAP(c): a core promoter factor with a built-in DNA-binding damper that is deactivated by the Oct-1 POU domain. Genes Dev., 13, 1807–1821.CrossRefGoogle ScholarPubMed
Moriuchi, H., Moriuchi, M., and Cohen, J. I. (1995). Proteins and cis-acting elements associated with transactivation of the varicella-zoster virus (VZV) immediate–early gene 62 promoter by VZV open reading frame 10 protein. J. Virol., 69, 4693–4701.Google ScholarPubMed
Narayanan, A., Nogueira, M. L., and Kristie, T. M. (2005). Combinatorial transcription of herpes simplex virus and varicella zoster virus immediate early genes is strictly determined by the cellular co-activator HCF-1. J. Biol. Chem., 280, 1369–1375.CrossRefGoogle Scholar
Nishikawa, J., Kokubo, T., Horikoshi, M., Roeder, R. G., and Nakatani, Y. (1997). Drosophila TAF(ii)230 and the transcriptional activator VP16 bind competitively to the TATA box-binding domain of the TATA box-binding protein. Proc. Natl Acad. Sci. USA, 94, 85–90.CrossRefGoogle ScholarPubMed
Nogueira, M. L., Wang, V. E. H., Tantin, D., Sharp, P. A., and Kristie, T. M. (2004). Herpes simplex virus infections are arrested in Oct-1 deficient cells. Proc. Natl Acad. Sci. USA, 101, 1473–1478.CrossRefGoogle 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 Scholar
Ouyang, L., Jacob, K. K., and Stanley, F. M. (1996). GABP mediates insulin-increased prolactin gene transcription. J. Biol. Chem., 271, 10425–10428.CrossRefGoogle ScholarPubMed
Perera, L. P., Mosca, J. D., Sadeghi-Zadeh, M., Ruyechan, W. T., and Hay, J. (1992). The varicella-zoster virus immediate early protein, IE62, can positively regulate its cognate promoter. Virology, 191, 346–354.CrossRefGoogle ScholarPubMed
Phillips, K. and Luisi, B. (2000). The virtuoso of versatility: POU proteins that flex to fit. J. Mol. Biol., 302, 1023–1039.CrossRefGoogle Scholar
Piluso, D., Bilan, P., and Capone, J. P. (2002). Host cell factor-1 interacts with and antagonizes transactivation by the cell cycle regulatory factor Miz-1. J. Biol. Chem., 277, 46799–46808.CrossRefGoogle ScholarPubMed
Pomerantz, J. L., Kristie, T. M., and Sharp, P. A. (1992). Recognition of the surface of a homeo domain protein. Genes Dev., 6, 2047–2057.CrossRefGoogle ScholarPubMed
Puigserver, P. and Spiegelman, B. M. (2003). Peroxisome proliferator-activated receptor-gamma coactivator 1 alpha (PGC-1 alpha): transcriptional coactivator and metabolic regulator. Endocr. Rev., 24, 78–90.CrossRefGoogle ScholarPubMed
Regier, J. L., Shen, F., and Triezenberg, S. J. (1993). Pattern of aromatic and hydrophobic amino acids critical for one of two subdomains of the VP16 transcriptional activator. Proc. Natl Acad. Sci. USA, 90, 883–887.CrossRefGoogle ScholarPubMed
Roizman, B. and Sears, A. E. (1996). Herpes simplex viruses and their replication. In Fundamental Virology, ed. Fields, B. N., Knipe, D. M., and Howley, P. M., Philadelphia: Lippincott-Raven Publishers: 1043–1107.Google Scholar
Ryu, H., Lee, J., Zaman, K.et al. (2003). Sp1 and Sp3 are oxidative stress-inducible, antideath transcription factors in cortical neurons. J. Neurosci., 23, 3597–3606.CrossRefGoogle ScholarPubMed
Sawada, J., Simizu, N., Suzuki, F.et al. (1999). Synergistic transcriptional activation by hGABP and select members of the activation transcription factor/cAMP response element-binding protein family. J. Biol. Chem., 274, 35475–35482.CrossRefGoogle ScholarPubMed
Scarpulla, R. C. (2002). Nuclear activators and coactivators in mammalian mitochondrial biogenesis. Biochim. Biophys. Acta., 1576, 1–14.CrossRefGoogle ScholarPubMed
Scarr, R. B. and Sharp, P. A. (2002). PDCD2 is a negative regulator of HCF-1 (C1). Oncogene, 21, 5245–5254.CrossRefGoogle Scholar
Scarr, R. B., Smith, M. R., Beddall, M., and Sharp, P. A. (2000). A novel 50-kilodalton fragment of host cell factor 1 (C1) in G(0) cells. Mol. Cell Biol., 20, 3568–3575.CrossRefGoogle Scholar
Shen, F., Triezenberg, S. J., Hensley, P., Porter, D., and Knutson, J. R. (1996). Transcriptional activation domain of the herpesvirus protein VP16 becomes conformationally constrained upon interaction with basal transcription factors. J. Biol. Chem., 271, 4827–4837.Google ScholarPubMed
Simmen, K. A., Newell, A., Robinson, M.et al. (1997). Protein interactions in the herpes simplex virus type 1 VP16-induced complex: VP16 peptide inhibition and mutational analysis of host cell factor requirements. J. Virol., 71, 3886–3894.Google ScholarPubMed
Sturm, R. A., Das, G., and Herr, W. (1988). The ubiquitous octamer-binding protein Oct-1 contains a POU domain with a homeo box subdomain. Genes Dev., 2, 1582–1599.CrossRefGoogle ScholarPubMed
Suske, G. (1999). The Sp-family of transcription factors. Gene, 238, 291–300.CrossRefGoogle ScholarPubMed
Tanaka, M., Lai, J. S., and Herr, W. (1992). Promoter-selective activation domains in Oct-1 and Oct-2 direct differential activation of an snRNA and mRNA promoter. Cell, 68, 755–767.CrossRefGoogle ScholarPubMed
Thompson, C. C., Brown, T. A., and McKnight, S. L. (1991). Convergence of Ets- and notch-related structural motifs in a heteromeric DNA binding complex. Science, 253, 762–768.CrossRefGoogle Scholar
Tiedt, R., Bartholdy, B. A., Matthias, G., Newell, J. W., and Matthias, P. (2001). The RING finger protein Siah-1 regulates the level of the transcriptional coactivator OBF-1. EMBO J., 20, 4143–4152.CrossRefGoogle ScholarPubMed
Triezenberg, S. J., LaMarco, K. L., and McKnight, S. L. (1988). Evidence of DNA: protein interactions that mediate HSV-1 immediate early gene activation by VP16. Genes Dev., 2, 730–742.CrossRefGoogle ScholarPubMed
Vignali, M., Steger, D. J., Neely, K. E., and Workman, J. L. (2000). Distribution of acetylated histones resulting from Gal4-VP16 recruitment of SAGA and NuA4 complexes. EMBO J., 19, 2629–2640.CrossRefGoogle ScholarPubMed
Vogel, J. L. and Kristie, T. M. (2000a). Autocatalytic proteolysis of the transcription factor-coactivator C1 (HCF): a potential role for proteolytic regulation of coactivator function. Proc. Natl Acad. Sci. USA, 97, 9425–9430.CrossRefGoogle Scholar
Vogel, J. L. and Kristie, T. M. (2000b). The novel coactivator C1 (HCF) coordinates multiprotein enhancer formation and mediates transcription activation by GABP. EMBO J., 19, 683–690.CrossRefGoogle Scholar
Vogel, J. L. and Kristie, T. M. (2001). The C1 factor (HCF). In The Encyclopedia of Molecular Medicine, ed. Creighton, T. E., New York: John Wiley and Sons, Inc.: 732–735.Google Scholar
Vogel, J. L. and Kristie, T. M. (2006). Site-specific proteolysis of the transcriptional coactivator HCF-1 can regulate its interaction with protein cofactors. Proc. Natl Acad. Sci. USA, 103(18), 6817–6822.CrossRefGoogle ScholarPubMed
Walker, S., Greaves, R., and O'Hare, P. (1993). Transcriptional activation by the acidic domain of Vmw65 requires the integrity of the domain and involves additional determinants distinct from those necessary for TFIIB binding. Mol. Cell Biol., 13, 5233–5244.CrossRefGoogle ScholarPubMed
Wilson, A. C., LaMarco, K., Peterson, M. G., and Herr, W. (1993). The VP16 accessory protein HCF is a family of polypeptides processed from a large precursor protein. Cell, 74, 115–125.CrossRefGoogle ScholarPubMed
Wilson, A. C., Freiman, R. N., Goto, H., Nishimoto, T., and Herr, W. (1997). VP16 targets an amino-terminal domain of HCF involved in cell cycle progression. Mol. Cell Biol., 17, 6139–6146.CrossRefGoogle ScholarPubMed
Wilson, A. C., Boutros, M., Johnson, K. M., and Herr, W. (2000). HCF-1 amino- and carboxy-terminal subunit association through two separate sets of interaction modules: involvement of fibronectin type 3 repeats. Mol. Cell Biol., 20, 6721–6730.CrossRefGoogle ScholarPubMed
Wu, T. J., Monokian, G., Mark, D. F., and Wobbe, C. R. (1994). Transcriptional activation by herpes simplex virus type 1 VP16 in vitro and its inhibition by oligopeptides. Mol. Cell Biol., 14, 3484–3493.CrossRefGoogle ScholarPubMed
Wysocka, J. and Herr, W. (2003). The herpes simplex virus VP16-induced complex: the makings of a regulatory switch. Trends Biochem. Sci., 28, 294–304.CrossRefGoogle ScholarPubMed
Wysocka, J., Reilly, P. T., and Herr, W. (2001). Loss of HCF-1-chromatin association precedes temperature-induced growth arrest of tsBN67 cells. Mol. Cell Biol., 21, 3820–3829.CrossRefGoogle ScholarPubMed
Wysocka, J., Myers, M. P., Laherty, C. D., Eisenman, R. N., and Herr, W. (2003). Human Sin3 deacetylase and trithorax-related Set1/Ash2 histone H3-K4 methyltransferase are tethered together selectively by the cell-proliferation factor HCF-1. Genes Dev., 17, 896–911.CrossRefGoogle ScholarPubMed
Xiao, H., Pearson, A., Coulombe, B.et al. (1994). Binding of basal transcription factor TFIIH to the acidic activation domains of VP16 and p53. Mol. Cell Biol., 14, 7013–7024.CrossRefGoogle ScholarPubMed
Yan, G. Z. and Ziff, E. B. (1997). Nerve growth factor induces transcription of the p21 WAF1/CIP1 and cyclin D1 genes in PC12 cells by activating the Sp1 transcription factor. J. Neurosci., 17, 6122–6132.CrossRefGoogle ScholarPubMed
Yudkovsky, N., Ranish, J. A., and Hahn, S. (2000). A transcription reinitiation intermediate that is stabilized by activator. Nature, 408, 225–229.CrossRefGoogle ScholarPubMed
Zhao, H., Jin, S., Fan, F., Fan, W., Tong, T., and Zhan, Q. (2000). Activation of the transcription factor Oct-1 in response to DNA damage. Cancer Res., 60, 6276–6280.Google ScholarPubMed
Zheng, L., Roeder, R. G., and Luo, Y. (2003). S phase activation of the histone H2B promoter by OCA-S, a coactivator complex that contains GAPDH as a key component. Cell, 114, 255–266.CrossRefGoogle ScholarPubMed
Zwilling, S., Annweiler, A., and Wirth, T. (1994). The POU domains of the Oct1 and Oct2 transcription factors mediate specific interaction with TBP. Nucl. Acids Res., 22, 1655–1662.CrossRefGoogle ScholarPubMed
Zwilling, S., Konig, H., and Wirth, T. (1995). High mobility group protein 2 functionally interacts with the POU domains of octamer transcription factors. Embo J., 14, 1198–1208.Google 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
×