Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-m8s7h Total loading time: 0 Render date: 2024-07-17T02:05:46.470Z Has data issue: false hasContentIssue false

68 - Candidate anti-herpesviral drugs; mechanisms of action and resistance

from Part VI - Antiviral therapy

Published online by Cambridge University Press:  24 December 2009

Karen K. Biron
Affiliation:
Clinical Virology, GlaxoSmithKline, Research Triangle Park, NC, 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

Research into the molecular biology of herpes replication in recent years has revealed novel targets for drug development (Fig. 68.1). The characterization and functional assay of these targets have been facilitated by advancements in gene expression, protein purification, proteonomics, bioinformatics, and efficient robotic screening technologies. The pipeline for new herpes drugs has been expanding as drug candidates have evolved more rapidly due to improvements in chemical synthesis (i.e., combinatorial and parallel synthesis methods), and with aids for drug design (X-ray crystallography, in silico computer modeling tools, as well as chemoinformatics). Many new herpes inhibitors have been reported, and most of these possess novel modes of actions. Several have entered clinical evaluation, with some later discontinued because of safety issues. This chapter will describe promising drug candidates in early development that appear to act at individual steps of the viral replication cycle, and focus on those that have the most potential for success (Table 68.1).

The chemotherapy of herpes infections was markedly advanced by the discovery of the first, highly selective antiherpetic agent, acyclovir (ACV, Zovirax®; [9-(2-hydroxyethoxymethyl)guanine]) (Elion et al., 1977). Since the introduction of this agent, there has been only incremental progress in new drug approvals for the myriad of diseases caused by this family of diverse pathogens. The drugs approved since the introduction of ACV include valacyclovir (VACV, Valtrex®, the L-valine ester prodrug of ACV, penciclovir (PCV; [9-(4-hydroxy-3-hydroxymethylbutyl-1-yl) guanine]), a related nucleoside analogue with a similar basis for drug action against HSV and VZV, and its prodrug, famciclovir (FCV, Famvir®).

Type
Chapter
Information
Human Herpesviruses
Biology, Therapy, and Immunoprophylaxis
, pp. 1219 - 1250
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

Abele, G., Cox, S., Bergman, S.et al. (1991). Antiviral activity against VZV and HSV type 1 and type 2 of the (+) and (−) enantiomers of (R,S)-9-[4-hydroxy-2-(hydroxymethyl) butyl] guanine, in comparison to other closely related acyclic nucleosides. Antivir. Chem. Chemother., 2, 163–169.CrossRefGoogle Scholar
Abenes, G., Lee, M., Haghjoo, E., Tong, T., Zhan, X., and Liu, F. (2001). Murine cytomegalovirus open reading frame M27 plays an important role in growth and virulence in mice. J. Virol., 75, 1697–1707.CrossRefGoogle ScholarPubMed
Ablashi, D. V., Chatlynne, L. G., Whitman, J. E. Jr., and Cesarman, E. (2002). Spectrum of Kaposi's sarcoma-associated herpesvirus, or human herpesvirus 8, diseases. Clin. Microbiol. Rev., 15, 439–464.CrossRefGoogle ScholarPubMed
Akanitapichat, P. and Bastow, K. F. (2002). The antiviral agent 5-chloro-1,3-dihydroxyacridone interferes with assembly and maturation of herpes simplex virus. Antiviral Res., 53, 113–126.CrossRefGoogle ScholarPubMed
Akanitapichat, P., Lowden, C. T., and Bastow, K. F. (2000).1,3-dihydroxyacridone derivatives as inhibitors of herpes virus replication. Antiviral Res., 45, 123–134.CrossRefGoogle ScholarPubMed
Aldern, K. A., Ciesla, S. L., Winegarden, K. L., and Hostetler, K. Y. (2003). Increased antiviral activity of 1-O-hexadecyloxypropyl-[2-(14)C]cidofovir in MRC-5 human lung fibroblasts is explained by unique cellular uptake and metabolism. Mol. Pharmacol., 63, 678–681.CrossRefGoogle ScholarPubMed
Arkin, M. R. and Wells, J. A. (2004). Small-molecule inhibitors of protein-protein interactions: progressing towards the dream. Nat. Rev. Drug Discov., 3, 301–317.CrossRefGoogle ScholarPubMed
Azzeh, M., Honigman, A., Taraboulos, A., Rouvinski, A., and Wolf, D. G. (2006). Structural changes in human cytomegalovirus cytoplasmic assembly sites in the absence of UL97 kinase activity. Virology, in press.CrossRefGoogle ScholarPubMed
Baba, M., Nishimura, O., Kanzaki, N.et al. (1999). A small-molecule, nonpeptide CCR5 antagonist with highly potent and selective anti-HIV-1 activity. Proc. Natl Acad Sci. USA, 96, 5698–5703.CrossRefGoogle ScholarPubMed
Bacon, T. H., Levin, M. J., Leary, J. J., Sarisky, R. T., and Sutton, D. (2003). Herpes simplex virus resistance to acyclovir and penciclovir after two decades of antiviral therapy. Clin. Microbiol. Rev., 16, 114–128.CrossRefGoogle ScholarPubMed
Baek, M. C., Krosky, P. M., He, Z., and Coen, D. M. (2002). Specific phosphorylation of exogenous protein and peptide substrates by the human cytomegalovirus UL97 protein kinase. Importance of the P+5 position. J. Biol. Chem., 277, 29593–29599.CrossRefGoogle ScholarPubMed
Balzarini, J. and McGuigan, C. (2002). Bicyclic pyrimidine nucleoside analogues (BCNAs) as highly selective and potent inhibitors of varicella-zoster virus replication. J. Antimicrob. Chemother., 50, 5–9.CrossRefGoogle ScholarPubMed
Balzarini, J., Sienaert, R., Liekens, S.et al. (2002). Lack of susceptibility of bicyclic nucleoside analogs, highly potent inhibitors of varicella-zoster virus, to the catabolic action of thymidine phosphorylase and dihydropyrimidine dehydrogenase. Mol. Pharmacol., 61, 1140–1145.CrossRefGoogle ScholarPubMed
Beadle, J. R., Wan, W. B., Ciesla, S. L.et al. (2006). Synthesis and antiviral evaluation of alkoxyalkyl derivatives of 9-(S)-(3-hydroxy-2-phosphonomethoxypropyl)adenine against cytomegalovirus and orthopoxviruses. J. Med. Chem., 49, 2010–2015.CrossRefGoogle ScholarPubMed
Beard, P. M., Taus, N. S., and Baines, J. D. (2002). DNA cleavage and packaging proteins encoded by genes U(L)28, U(L)15, and U(L)33 of herpes simplex virus type 1 form a complex in infected cells. J. Virol., 76, 4785–4791.CrossRefGoogle Scholar
Bernstein, D. I., Harrison, C. J., Tomai, M. A., and Miller, R. L. (2001). Daily or weekly therapy with resiquimod (R-848) reduces genital recurrences in herpes simplex virus-infected guinea pigs during and after treatment. J. Infect. Dis., 183, 844–849.CrossRefGoogle ScholarPubMed
Betz, U. A., Fischer, R., Kleymann, G., Hendrix, M., and Rubsamen-Waigmann, H. (2002). Potent in vivo antiviral activity of the herpes simplex virus primase-helicase inhibitor BAY 57-1293. Antimicrob. Agents Chemother., 46, 1766–1772.CrossRefGoogle ScholarPubMed
Bidanset, D. J., Beadle, J. R., Wan, W. B., Hostetler, K. Y., and Kern, E. R. (2004). Oral activity of ether lipid ester prodrugs of cidofovir against experimental human cytomegalovirus infection. J. Infect. Dis., 190, 499–503.CrossRefGoogle ScholarPubMed
Biron, K. K., Harvey, R. J., Chamberlain, S. C.et al. (2002). Potent and selective inhibition of human cytomegalovirus replication by 1263W94, a benzimidazole L-riboside with a unique mode of action. Antimicrob. Agents Chemother., 46, 2365–2372.CrossRefGoogle ScholarPubMed
Bogner, E., Radsak, K., and Stinski, M. F. (1998). The gene product of human cytomegalovirus open reading frame UL56 binds the pac motif and has specific nuclease activity. J. Virol., 72, 2259–2264.Google ScholarPubMed
Bogner, E. (2002). Human cytomegalovirus terminase as a target for antiviral chemotherapy. Rev. Med. Virol., 12, 115–127.CrossRefGoogle ScholarPubMed
Borthwick, A. D. (2005). Design of translactam HCMV protease inhibitors as potent antivirals. Med. Res. Rev., 25, 427–452.CrossRefGoogle ScholarPubMed
Borthwick, A. D., Weingarten, G., Haley, T. M.et al. (1998). Design and synthesis of monocyclic beta-lactams as mechanism-based inhibitors of human cytomegalovirus protease. Bioorg. Med. Chem. Lett., 8, 365–370.CrossRefGoogle ScholarPubMed
Boston Consulting Group analysis based on DiMasi, J. A. et al. (1993). As quoted by The Office of Technology Assessment in Pharmaceutical Research and Development: Cost, Risks, Rewards.
Boulware, S. L., Bronstein, J. C., Nordby, E. C., and Weber, P. C. (2001). Identification and characterization of a benzothiophene inhibitor of herpes simplex virus type 1 replication which acts at the immediate early stage of infection. Antiviral Res., 51, 111–125.CrossRefGoogle ScholarPubMed
Bournique, B., Lambert, N., Boukaiba, R., and Martinet, M. (2001). In vitro metabolism and drug interaction potential of a new highly potent anti-cytomegalovirus molecule, CMV423 (2-chloro 3-pyridine 3-yl 5,6,7,8-tetrahydroindolizine I-carboxamide). Br. J. Clin. Pharmacol., 52, 53–63.CrossRefGoogle Scholar
Braithwaite, D. K. and Ito, J. (1993). Compilation, alignment, and phylogenetic relationships of DNA polymerases. Nucl. Acids Res., 21, 787–802.CrossRefGoogle ScholarPubMed
Brideau, R. J., Knechtel, M. L., Huang, A.et al. (2002). Broad-spectrum antiviral activity of PNU-183792, a 4-oxo-dihydroquinoline, against human and animal herpesviruses. Antiviral Res., 54, 19–28.CrossRefGoogle ScholarPubMed
Brown, D. G., Visse, R., Sandhu, G.et al. (1995).Crystal structures of the thymidine kinase from herpes simplex virus type-1 in complex with deoxythymidine and ganciclovir. Nat. Struct. Biol., 2, 876–881.CrossRefGoogle Scholar
Buerger, I., Reefschlaeger, J., Bender, W.et al. (2001). A novel nonnucleoside inhibitor specifically targets cytomegalovirus DNA maturation via the UL89 and UL56 gene products. J. Virol., 75, 9077–9086.CrossRefGoogle ScholarPubMed
Buerger, I., Reefschlaeger, J., Bender, W. (2001). A novel nonnucleoside inhibitor specifically targets cytomegalovirus DNA maturation via the UL89 and UL56 gene products. J. Virol., 75, 9077–9086.CrossRefGoogle ScholarPubMed
Chee, M. S., Lawrence, G. L., and Barrell, B. G. (1989). Alpha-, beta- and gammaherpesviruses encode a putative phosphotransferase. J. Gen. Virol., 70(5), 1151–1160.CrossRefGoogle ScholarPubMed
Chou, S. (2006). UL27 and UL97 resistance mutations selected after passage of clinical CMV isolates under maribavir. 31st International Herpesvirus Workshop. 31st International Herpesvirus Workshop [Seattle, WA July 22-28; Oral Presentation 10.13].
Chou, S., Lurain, N. S., Thompson, K. D., Miner, R. C., and Drew, W. L. (2003). Viral DNA polymerase mutations associated with drug resistance in human cytomegalovirus. J. Infect. Dis., 188, 32–39.CrossRefGoogle ScholarPubMed
Chou, S., Marousek, G. I., Senters, A. E., Davis, M. G., and Biron, K. K. (2004). Mutations in the human cytomegalovirus UL27 gene that confer resistance to maribavir. J. Virol., 78, 7124–7130.CrossRefGoogle ScholarPubMed
Chulay, J., Biron, K., Wang, L.et al. (1999). Development of novel benzimidazole riboside compounds for treatment of cytomegalovirus disease. Adv. Exp. Med Biol., 458, 129–134.CrossRefGoogle ScholarPubMed
Ciesla, S. L., Trahan, J., Wan, W. B.et al. (2003). Esterification of cidofovir with alkoxyalkanols increases oral bioavailability and diminishes drug accumulation in kidney. Antiviral Res., 59, 163–171.CrossRefGoogle ScholarPubMed
Cirone, M., Zompetta, C., Tarasi, D., Frati, L., and Faggioni, A. (1996). Infection of human T lymphoid cells by human herpesvirus 6 is blocked by two unrelated protein tyrosine kinase inhibitors, biochanin A and herbimycin. AIDS Res. Hum. Retroviruses, 12, 1629–1634.CrossRefGoogle Scholar
Crumpacker, C. S. and Schaffer, P. A. (2002). New anti-HSV therapeutics target the helicase-primase complex. Nat. Med., 8, 327–328.CrossRefGoogle ScholarPubMed
Crute, J. J. and Lehman, I. R. (1991). Herpes simplex virus-1 helicase-primase. Physical and catalytic properties. J. Biol. Chem., 266, 4484–4488.Google ScholarPubMed
Crute, J. J., Grygon, C. A., Hargrave, K. D.et al. (2002). Herpes simplex virus helicase-primase inhibitors are active in animal models of human disease. Nat. Med., 8, 386–391.CrossRefGoogle ScholarPubMed
Bolle, L., Andrei, G., Snoeck, R.et al. (2004). Potent, selective and cell-mediated inhibition of human herpesvirus 6 at an early stage of viral replication by the non-nucleoside compound CMV423. Biochem. Pharmacol., 67, 325–336.CrossRefGoogle ScholarPubMed
Clercq, E. (2002). Cidofovir in the therapy and short-term prophylaxis of poxvirus infections. Trends Pharmacol. Sci., 23, 456–458.CrossRefGoogle ScholarPubMed
Clercq, E. (2003a). Highly potent and selective inhibition of varicella-zoster virus replication by bicyclic furo[2,3-d]pyrimidine nucleoside analogues. Med. Res. Rev., 23, 253–274.CrossRefGoogle Scholar
Clercq, E. (2003b). New inhibitors of human cytomegalovirus (HCMV) on the horizon. J. Antimicrob. Chemother., 51, 1079–1083.CrossRefGoogle Scholar
Miranda, P. and Burnette, T. C. (1994). Metabolic fate and pharmacokinetics of the acyclovir prodrug valaciclovir in cynomolgus monkeys. Drug Metab. Dispos., 22, 55–59.Google ScholarPubMed
Miranda, P. and Good, S. S. (1992). Species differences in the metabolism and disposition of antiviral analogues. Antiviral Chem. Chemo., 3, 1–8.CrossRefGoogle Scholar
Diana, G. D., Oglesby, R. C., Akullian, V.et al. (1987). Structure-activity studies of 5-[[4-(4,5-dihydro-2-oxazolyl) phenoxy]alkyl]-3-methylisoxazoles: inhibitors of picornavirus uncoating. J. Med. Chem., 30, 383–388.CrossRefGoogle ScholarPubMed
Dittmer, A., Drach, J. C., Townsend, L., Fischer, A., and Bogner, E. (2005). Interaction of the putative human cytomegalovirus portal protein pUL104 with the large terminase subunit pUL56 and its inhibition by benzimidazole-D-ribonucleosides. J. Virol., 79, 14660–14667.CrossRefGoogle ScholarPubMed
Docherty, J. J., Fu, M. M., Stiffler, B. S., Limperos, R. J., Pokabla, C. M., and DeLucia, A. L. (1999). Resveratrol inhibition of herpes simplex virus replication. Antiviral Res., 43, 145–155.CrossRefGoogle ScholarPubMed
Dunkle, L. M. (1996). Lobucavir: a promising broad-spectrum antiviral agent. EleventhInternational Conference on AIDS, Vancouver, abstract Th.B.943.
Elion, G. B., Furman, P. A., Fyfe, J. A., Miranda, P., Beauchamp, L., and Schaeffer, H. J. (1977). Selectivity of action of an antiherpetic agent, 9-(2-hydroxyethoxymethyl) guanine. Proc. Natl Acad. Sci. USA, 74, 5716–5720.CrossRefGoogle ScholarPubMed
Erickson, J., Neidhart, D. J., VanDrie, J.et al. (1990). Design, activity, and 2.8 A crystal structure of a C2 symmetric inhibitor complexed to HIV-1 protease. Science, 249, 527–533.CrossRefGoogle ScholarPubMed
Evers, D. L., Komazin, G., Ptak, R. G.et al. (2004). Inhibition of human cytomegalovirus replication by benzimidazole nucleosides involves three distinct mechanisms. Antimicrob. Agents Chemother., 48, 3918–3927.CrossRefGoogle ScholarPubMed
Franklin, M. C., Wang, J., and Steitz, T. A. (2001). Structure of the replicating complex of a pol alpha family DNA polymerase. Cell, 105, 657–667.CrossRefGoogle 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, 998–1003.CrossRefGoogle ScholarPubMed
Gibson, W., Welch, A. R., and Hall, W. R. T. (1994). Assembling a herpesvirus serine maturational proteinase and a new molecular target for antivirals. Perspect Drug Discov Design, 2, 413–416.CrossRefGoogle Scholar
Good, S. S., Owens, B. S., Townsend, L. B., and Drach, J. C. (1994). The disposition in rats and monkeys of 2-bromo-5,6,-dicholoro-1-(beta-ribofuranosyl)benzimididazole (BDCRB) and its 2,5,6-trichloro congener (TCRB). Antivir. Res., 23, 103.Google Scholar
Hall, M. C. and Matson, S. W. (1999). Helicase motifs: the engine that powers DNA unwinding. Mol. Microbiol., 34, 867–877.CrossRefGoogle ScholarPubMed
Hamilton, H. W., Nishiguchi, G., Hagen, S. E.et al. (2002). Novel benzthiodiazepinones as antiherpetic agents: SAR improvement of therapeutic index by alterations of the seven-membered ring. Bioorg. Med Chem. Lett., 12, 2981–2983.CrossRefGoogle ScholarPubMed
Hasegawa, Y., Nishiyama, Y., Imaizumi, K.et al. (2000). Avoidance of bone marrow suppression using A-5021 as a nucleoside analog for retrovirus-mediated herpes simplex virus type I thymidine kinase gene therapy. Cancer Gene Ther., 7, 557–562.CrossRefGoogle ScholarPubMed
He, Z., He, Y. S., Kim, Y.et al. (1997). The human cytomegalovirus UL97 protein is a protein kinase that autophosphorylates on serines and threonines. J. Virol., 71, 405–411.Google ScholarPubMed
Hemmi, H., Kaisho, T., Takeuchi, O.et al. (2002). Small anti-viral compounds activate immune cells via the TLR7 MyD88-dependent signaling pathway. Nat. Immunol., 3, 196–200.CrossRefGoogle ScholarPubMed
Henry, S. P., Miner, R. C., Drew, W. L.et al. (2001). Antiviral activity and ocular kinetics of antisense oligonucleotides designed to inhibit CMV replication. Invest. Ophthalmol. Vis. Sci., 42, 2646–2651.Google ScholarPubMed
Herget, T., Freitag, M., Morbitzer, M., Kupfer, R., Stamminger, T., and Marschall, M. (2004). Novel chemical class of pUL97 protein kinase-specific inhibitors with strong anticytomegaloviral activity. Antimicrob. Agents Chemother., 48, 4154–4162.CrossRefGoogle ScholarPubMed
Holwerda, B. C. (1997). Herpesvirus proteases: targets for novel antiviral drugs. Antiviral Res., 35, 1–21.CrossRefGoogle ScholarPubMed
Hoog, S. S., Smith, W. W., Qiu, X.et al. (1997). Active site cavity of herpesvirus proteases revealed by the crystal structure of herpes simplex virus protease/inhibitor complex. Biochemistry, 56, 14023–14029.CrossRefGoogle Scholar
Hu, H. and Cohen, J. I. (2005). Varicella-zoster virus open reading frame 47 (ORF47) protein is critical for virus replication in dendritic cells and for spread to other cells. Virology, 337, 304–311.CrossRefGoogle ScholarPubMed
Huggins, J. W., Baker, R. O., Beadle, J. R., and Hostetler, K. Y. (2002). Orally active ether lipid prodrugs of cidofovir for the treatment of smallpox. Antivir. Res., 53, A66, 104.Google Scholar
Iwayama, S., Ono, N., Ohmura, Y.et al. (1998). Antiherpesvirus activities of (1′S,2′R)-9-[[1′,2′-bis(hydroxymethyl)cycloprop-1′-yl]methyl]guanine (A-5021) in cell culture. Antimicrob. Agents Chemother., 42, 1666–1670.Google Scholar
Iwayama, S., Ohmura, Y., Suzuki, K.et al. (1999). Evaluation of anti-herpesvirus activity of (1′S,2′R)-9-[[1′,2′-bis(hydroxymethyl)cycloprop-1′-yl]methyl]-guanine (A-5021) in mice. Antiviral Res., 42, 139–148.CrossRefGoogle Scholar
Jacobson, J. G., Renau, T. E., Nassiri, M. R.et al. (1999). Nonnucleoside pyrrolopyrimidines with a unique mechanism of action against human cytomegalovirus. Antimicrob. Agents Chemother., 43, 1888–1894.Google ScholarPubMed
Kato, A., Yamamoto, M., Ohno, T.et al. (2006). Herpes simplex virus 1-encoded protein kinase UL13 phosphorylates viral Us3 protein kinase and regulates nuclear localization of viral envelopment factors UL34 and UL31. J. Virol., 80, 1476–1486.CrossRefGoogle ScholarPubMed
Kawaguchi, Y. and Kato, K. (2003). Protein kinases conserved in herpesviruses potentially share a function mimicking the cellular protein kinase cdc2. Rev. Med. Virol., 13, 331–340.CrossRefGoogle Scholar
Kern, E. R. (2003). In vitro activity of potential anti-proxvirus agents. Antiviral Res., 57, 35–40.CrossRefGoogle ScholarPubMed
Kern, E. R., Collins, D. J., Wan, W. B., Beadle, J. R., Hostetler, K. Y., and Quenelle, D. C. (2004a). Oral treatment of murine cytomegalovirus infections with ether lipid esters of cidofovir. Antimicrob. Agents Chemother., 48, 3516–3522.CrossRefGoogle Scholar
Kleymann, G. (2003a). Novel agents and strategies to treat herpes simplex virus infections. Expert. Opin. Investig. Drugs, 12, 165–183.CrossRefGoogle Scholar
Kleymann, G. (2003b). Helicase-primase inhibitors. Drugs of the Future, 28, 257–265.CrossRefGoogle Scholar
Kleymann, G., Fischer, R., Betz, U. A.et al. (2002). New helicase-primase inhibitors as drug candidates for the treatment of herpes simplex disease. Nat. Med., 8, 392–398.CrossRefGoogle ScholarPubMed
Knechtel, M. L., Huang, A., Vaillancourt, V. A., and Brideau, R. J. (2002). Inhibition of clinical isolates of human cytomegalovirus and varicella zoster virus by PNU-183792, a 4-oxo-dihydroquinoline. J. Med. Virol., 68, 234–236.CrossRefGoogle ScholarPubMed
Komatsu, T., Ballestras, M. E., Barbera, A. J., Kelly-Clarke, B., and Kaye, K. M. (2004). KSHV LA NA-1 binds DNA as an oligomer and residues N-terminal to the oligomerization domain are essential for DNA replication and episome persistence. Virology, 319, 225–236.CrossRefGoogle Scholar
Koszalka, G. W., Johnson, N. W., Good, S. S. (2002). Preclinicat and toxicology studies of 1263W94, a potent and selective inhibitor of human cytomegalovirus replication. Antimicrob. Agents Chemother, 46, 2373–2380.CrossRefGoogle ScholarPubMed
Komazin, G., Ptak, R. G., Emmer, B. T., Townsend, L. B., and Drach, J. C. (2003). Resistance of human cytomegalovirus to the benzimidazole L-ribonucleoside maribavir maps to UL27. J Virol., 77, 11499–11506.CrossRefGoogle ScholarPubMed
Komazin, G., Townsend, L. B., and Drach, J. C. (2004). Role of a mutation in human cytomegalovirus gene UL104 in resistance to benzimidazole ribonucleosides. J. Virol., 78, 710–715.CrossRefGoogle ScholarPubMed
Krosky, P. M., Baek, M. C., Jahng, W. J.et al. (2003b). The human cytomegalovirus UL44 protein is a substrate for the UL97 protein kinase. J. Virol., 77, 7720–7727.CrossRefGoogle Scholar
Krosky, P. M., Baek, M. C., and Coen, D. M. (2003a). The human cytomegalovirus UL97 protein kinase, an antiviral drug target, is required at the stage of nuclear egress. J. Virol., 77, 905–914.CrossRefGoogle Scholar
Krosky, P. M., Underwood, M. R., Turk, S. R., et al. (1998). Resistance of human cytomegalovirus to benzimidazole ribonucleosides maps to two open reading frames: UL89 and UL56. J. Virol., 72, 4721–4728.Google ScholarPubMed
LaBranche, C. C., Galasso, G., Moore, J. P., Bolognesi, D. P., Hirsch, M. S., and Hammer, S. M. (2001). HIV fusion and its inhibition. Antiviral Res., 50, 95–115.CrossRefGoogle ScholarPubMed
Lalezari, J. P. (1997). New treatment options for CMV retinitis in AIDS. Adv. Nurse Pract., 5, 45–9.Google Scholar
Lalezari, J. P., Drew, W. L., Glutzer, E.et al. (1995). (S)-1[3-hydroxy-2-(phosphonylmethoxy)propyl] cytosine (cidofovir): results of a phase I/II study of a novel antiviral nucleotide analogue. J. Infect. Dis., 171, 788–796.CrossRefGoogle ScholarPubMed
Lalezari, J. P., Aberg, J. A., Wang, L. H.et al. (2002). Phase I dose escalation trial evaluating the pharmacokinetics, anti-human cytomegalovirus (HCMV) activity, and safety of 1263W94 in human immunodeficiency virus-infected men with asymptomatic HCMV shedding. Antimicrob. Agents Chemother., 46, 2969–2976.CrossRefGoogle ScholarPubMed
Lipinski, C. A., Lombardo, F., Dominy, B. W., and Feeney, P. J. (2001). Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev., 46, 3–26.CrossRefGoogle ScholarPubMed
Littler, E., Stuart, A. D., and Chee, M. S. (1992). Human cytomegalovirus UL97 open reading frame encodes a protein that phosphorylates the antiviral nucleoside analogue ganciclovir. Nature, 358, 160–162.CrossRefGoogle ScholarPubMed
Liuzzi, M., Kibler, P., Bousquet, C.et al. (2004). Isolation and characterization of herpes simplex virus type 1 resistant to aminothiazolylphenyl-based inhibitors of the viral helicase-primase. Antiviral Res., 64, 161–170.CrossRefGoogle ScholarPubMed
Loregian, A., and Coen, D. M. (2006). Selective anti-cytomegalovirus compounds discoverd by screening for inhibitors of subunit interactions of the viral polymerase, Chem. Biol., 13, 191–200.CrossRefGoogle Scholar
Lorenzi, P. L., Landowski, C. P., Brancale, A.et al. (2006). N-methylpurine DNA glycosylase and 8-oxoguanine dna glycosylase metabolize the antiviral nucleoside 2-bromo-5,6-dichloro-1-(beta-D-ribofuranosyl)benzimidazole. Drug Metab. Dispos., 34, 1070–1077.Google ScholarPubMed
Lorenzi, P. L., Landowski, C. P., Song, X.et al. (2005). Amino acid ester prodrugs of 2-bromo-5,6-dichloro-1-(beta-D-ribofuranosyl)benzimidazole enhance metabolic stability in vitro and in vivo. J. Pharmacol. Exp. Ther., 314, 883–890.CrossRefGoogle ScholarPubMed
Lowden, C. T. and Bastow, K. F. (2003). Cell culture replication of herpes simplex virus and, or human cytomegalovirus is inhibited by 3,7-dialkoxylated, 1-hydroxyacridone derivatives. Antiviral Res., 59, 143–154.CrossRefGoogle ScholarPubMed
Lowe, D. M., Alderton, W. K., , Ellis M. R.et al. (1995). Mode of action of (R)-9-[4-hydroxy-2-(hydroxymethyl)butyl]guanine against herpesviruses. Antimicrob. Agents Chemother., 39, 1802–1808.CrossRefGoogle ScholarPubMed
Lu, H. and Thomas, S. (2004). Maribavir (ViroPharma). Curr. Opin. Investig. Drugs, 5, 898–906.Google Scholar
Lurain, N. S., Weinberg, A., Crumpacker, C. S., and , Chou S. (2001). Sequencing of cytomegalovirus UL97 gene for genotypic antiviral resistance testing. Antimicrob. Agents Chemother., 45, 2775–2780.CrossRefGoogle ScholarPubMed
Ma, J. D., Nafziger, A. N., Villano, S. A., and , J. S. Jr. (2006). Maribavir pharmacokinetics and the effects of multiple-dose maribavir on cytochrome P450 (CYP) 1A2, CYP 2C9, CYP 2C19, CYP 2D6, CYP 3A, N-acetyltransferase-2, and xanthine oxidase activities in healthy adults. Antimicrob. Agents Chemother., 50, 1130–1135.CrossRefGoogle ScholarPubMed
Marschall, M., Stein-Gerlach, M., Freitage, M., Kupfer, R., Den, BM., and Stamminger, T. (2001). Inhibitors of human cytomegalovirus replication drastically reduce the activity of the viral protein kinase pUL97. J. Gen. Virol., 82, 1439–1450.CrossRefGoogle ScholarPubMed
Marschall, M., Freitag, M., Suchy, P.et al. (2003). The protein kinase pUL97 of human cytomegalovirus interacts with and phosphorylates the DNA polymerase processivity factor pUL44. Virology, 311, 60–71.CrossRefGoogle ScholarPubMed
Marschall, M., Marzi, A., aus dem, S. P.et al., (2005). Cellular p32 recruits cytomegalovirus kinase pUL97 to redistribute the nuclear lamina. J. Biol. Chem., 280, 33357–33367.CrossRefGoogle ScholarPubMed
Matthews, J. T., Terry, B. J., and Field, A. K. (1993). The structure and function of the HSV DNA replication proteins: defining novel antiviral targets. Antiviral Res., 20, 89–114.CrossRefGoogle ScholarPubMed
McGuigan, C., Jukes, A., Blewett, S.et al. (2003). Halophenyl furanopyrimidines as potent and selective anti-VZV agents. Antivir. Chem. Chemother., 14, 165–170.CrossRefGoogle ScholarPubMed
McSharry, J. J., McDonough, A., Olson, B.et al. (2001a). Susceptibilities of human cytomegalovirus clinical isolates to BAY38-4766, BAY43-9695, and ganciclovir. Antimicrob. Agents Chemother., 45, 2925–2927.CrossRefGoogle Scholar
McSharry, J. J., McDonough, A., Olson, B., Talarico, C., Davis, M., and Biron, K. K. (2001b). Inhibition of ganciclovir-susceptible and -resistant human cytomegalovirus clinical isolates by the benzimidazole L-riboside 1263W94. Clin. Diagn. Lab. Immunol., 8, 1279–1281.Google Scholar
McVoy, M. A. and Nixon, D. E. (2005). Impact of 2-bromo-5,6-dichloro-1-beta-D-ribofuranosyl benzimidazole riboside and inhibitors of DNA, RNA, and protein synthesis on human cytomegalovirus genome maturation. J. Virol., 79, 11115–11127.CrossRefGoogle ScholarPubMed
Mettenleiter, T. C. (2002). Herpesvirus assembly and egress. J Virol., 76, 1537–1547.CrossRefGoogle ScholarPubMed
Michel, D., Pavic, I., Zimmermann, A.et al. (1996). The UL97 gene product of human cytomegalovirus is an early-late protein with a nuclear localization but is not a nucleoside kinase. J. Virol., 70, 6340–6346.Google Scholar
Miller, R. L., Tomai, M. A., Harrison, C. J., and Bernstein, D. I. (2002). Immunomodulation as a treatment strategy for genital herpes: review of the evidence. Int. Immunopharmacol., 2, 443–451.CrossRefGoogle Scholar
Moffat, J. F., Zerboni, L., Sommer, M. H.et al. (1998). The ORF47 and ORF66 putative protein kinases of varicella-zoster virus determine tropism for human T cells and skin in the SCID-hu mouse. Proc. Natl Acad. Sci. USA, 95, 11969–11974.CrossRefGoogle Scholar
Nagelschmitz, J., Moeller, J. G., Stass, H. H., Wadel, C., and Kuhlmann, J. (1999). Safety, tolerability, and pharmacokinetics of single oral doses of BAY 38-4766 – a novel nonnucleosidic inhibitor of human cytomegalovirus (HCMV) replication – in healthy male subjects. In Program and Abstracts of theThirty-ninthInterscience Conference on Antimicrob Agents and Chemother, San Francisco, CA, Abstract 945, 322.
Naesens, L., Stephens, C. E., Andrei, G.et al. (2006). Antiviral properties of new arylsulfone derivatives with activity against human betaherpesviruses. Antiviral Res., in press.
Newcomb, W. W. and Brown, J. C. (2002). Inhibition of herpes simplex virus replication by WAY-150138: assembly of capsids depleted of the portal and terminase proteins involved in DNA encapsidation. J Virol., 76, 10084–10088.CrossRefGoogle ScholarPubMed
Newcomb, W. W., Juhas, R. M., Thomsen, D. R.et al. (2001). The UL6 gene product forms the portal for entry of DNA into the herpes simplex virus capsid. J. Virol., 75, 10923–10932.CrossRefGoogle ScholarPubMed
Newcomb, W. W., Juhas, R. M., Thomsen, D. R.et al. (2001). The UL6 gene product forms the portal for entry of DNA into the herpes simplex virus capsid. J. Virol., 75, 10923–10932.CrossRefGoogle ScholarPubMed
Newcomb, W. W., Thomsen, D. R., Homa, F. L., and Brown, J. C. (2003). Assembly of the herpes simplex virus capsid: identification of soluble scaffold-portal complexes and their role in formation of portal-containing capsids. J. Virol., 77, 9862–9871.CrossRefGoogle ScholarPubMed
Neyts, J. and Clercq, E. (2001). The anti-herpesvirus activity of (1′S,2′R)-9-[[1′,2′-bis(hydroxymethyl)-cycloprop-1′-yl]methyl]guanine is markedly potentiated by the immunosuppressive agent mycophenolate mofetil. Antiviral Res., 49, 121–127.CrossRefGoogle Scholar
Neyts, J., Naesens, L., Ying, C., Bolle, L., and Clercq, E. (2001). Anti-herpesvirus activity of (1′S,2′R)-9-[[1′,2′-bis(hydroxymethyl)-cycloprop-1′-yl]methyl] x guanine (A-5021) in vitro and in vivo. Antiviral Res., 49, 115–120.CrossRefGoogle Scholar
Neyts, J., Leyssen, P., Verbeken, E., and Clercq, E. (2004). Efficacy of cidofovir in a murine model of disseminated progressive vaccinia. Antimicrob. Agents Chemother., 48, 2267–2273.CrossRefGoogle Scholar
Ng, T. I., Shi, Y., Huffaker, H. J.et al. (2001). Selection and characterization of varicella-zoster virus variants resistant to (R)-9-[4-hydroxy-2-(hydroxymethyl)butyl]guanine. Antimicrob. Agents Chemother., 45, 1629–1636.CrossRefGoogle ScholarPubMed
Oien, N. L., Brideau, R. J., Hopkins, T. A.et al. (2002). Broad-spectrum antiherpes activities of 4-hydroxyquinoline carboxamides, a novel class of herpesvirus polymerase inhibitors. Antimicrob. Agents Chemother., 46, 724–730.CrossRefGoogle ScholarPubMed
Okano, M. (2003). The evolving therapeutic approaches for Epstein-Barr virus infection in immunocompetent and immunocompromised individuals. Curr. Drug Targets. Immune. Endocr. Metabol. Disord., 3, 137–142.CrossRefGoogle ScholarPubMed
Ono, N., Iwayama, S., Suzuki, K.et al. (1998). Mode of action of (1′S,2′R)-9-[[1′,2′-bis(hydroxymethyl) cycloprop-1′-yl]methyl]guanine (A-5021) against herpes simplex virus type 1 and type 2 and varicella-zoster virus. Antimicrob. Agents Chemother., 42, 2095–2102.Google Scholar
Painter, G. R. and Hostetler, K. Y. (2004). Design and development of oral drugs for the prophylaxis and treatment of smallpox infection. Trends Biotechnol., 22, 423–427.CrossRefGoogle ScholarPubMed
Prichard, M. N., Britt, W. J., Daily, S. L., Hartline, C. B., and Kern, E. R. (2005). Human cytomegalovirus UL97 Kinase is required for the normal intranuclear distribution of pp65 and virion morphogenesis. J. Virol., 79, 15494–15502.CrossRefGoogle ScholarPubMed
Prichard, M. N., Gao, N., Jairath, S.et al. (1999). A recombinant human cytomegalovirus with a large deletion in UL97 has a severe replication deficiency. J. Virol., 73, 5663–5670.Google Scholar
Przech, A. J., Yu, D., and Weller, S. K. (2003). Point mutations in exon I of the herpes simplex virus putative terminase subunit, UL15, indicate that the most conserved residues are essential for cleavage and packaging. J. Virol., 77, 9613–9621.CrossRefGoogle ScholarPubMed
Qiu, X. Y. and Abdelmeguid, S. S. (1999). Human herpes proteases In Dunn, B. M., ed. Proteases of Infectious Agents. San Diego: Academic Press, 93–115.Google Scholar
Razonable, R. R., Brown, R. A., Humar, A., Covington, E., Alecock, E., and Paya, C. V. (2005). Herpesvirus infections in solid organ transplant patients at high risk of primary cytomegalovirus disease. J. Infect. Dis., 192, 1331–1339.CrossRefGoogle ScholarPubMed
Rechtsteiner, G., Warger, T., Osterloh, P., Schild, H., and Radsak, M. P. (2005). Cutting edge: priming of CTL by transcutaneous peptide immunization with imiquimod. J. Immunol., 174, 2476–2480.CrossRefGoogle ScholarPubMed
Reefschlaeger, J., Bender, W., Hallenberger, S.et al. (2001). Novel non-nucleoside inhibitors of cytomegaloviruses (BAY 38-4766): in vitro and in vivo antiviral activity and mechanism of action. J. Antimicrob. Chemother., 48, 757–767.CrossRefGoogle ScholarPubMed
Roizman, B., Whitley, R. J., and Lopez, C., eds. (1993). The Human Herpesvireses. New York, NY: Raven Press.Google Scholar
Safrin, S., Cherrington, J., and Jaffe, H. S. (1997). Clinical uses of cidofovir. Rev. Med. Virol., 7, 145–156.3.0.CO;2-0>CrossRefGoogle ScholarPubMed
Sakuma, T., Saijo, M., Suzutani, T.et al. (1991). Antiviral activity of oxetanocins against varicella-zoster virus. Antimicrob. Agents Chemother., 35, 1512–1514.CrossRefGoogle ScholarPubMed
Schang, L. M. (2002). Cyclin-dependent kinases as cellular targets for antiviral drugs. J. Antimicrob. Chemother., 50, 779–792.CrossRefGoogle ScholarPubMed
Schang, L. M., Bantly, A., Knockaert, M.et al. (2002). Pharmacological cyclin-dependent kinase inhibitors inhibit replication of wild-type and drug-resistant strains of herpes simplex virus and human immunodeficiency virus type 1 by targeting cellular, not viral, proteins. J. Virol., 76, 7874–7882.CrossRefGoogle Scholar
Scheffczik, H., Savva, C. G., Holzenburg, A., Kolesnikova, L., and Bogner, E. (2002). The terminase subunits pUL56 and pUL89 of human cytomegalovirus are DNA-metabolizing proteins with toroidal structure. Nucl. Acids Res., 30, 1695–1703.CrossRefGoogle ScholarPubMed
Scolnick, E. M., Richards, F. M., Eisenberg, D. S., and Kim, P. S., eds. (2001). Drug Discovery and Design (Advances in Protein Chemistry)51, Academic Press.Google Scholar
Sekiyama, T., Hatsuya, S., Tanaka, Y.et al. (1998). Synthesis and antiviral activity of novel acyclic nucleosides: discovery of a cyclopropyl nucleoside with potent inhibitory activity against herpesviruses. J. Med. Chem., 41, 1284–1298.CrossRefGoogle ScholarPubMed
Shin, Y. K., Cai, G. Y., Weinberg, A., Leary, J. J., and Levin, M. J. (2001). Frequency of acyclovir-resistant herpes simplex virus in clinical specimens and laboratory isolates. J. Clin. Microbiol., 39, 913–917.CrossRefGoogle ScholarPubMed
Shugar, D. (1999). Viral and host-cell protein kinases: enticing antiviral targets and relevance of nucleoside, and viral thymidine, kinases. Pharmacol. Ther., 82, 315–335.CrossRefGoogle ScholarPubMed
Sienaert, R., Andrei, G., Snoeck, R., De Clercq, E., McGuigan, C., and Balzarini, J. (2004). Inactivity of the bicyclic pyrimidine nucleoside analogues against simian varicella virus (SVV) does not correlate with their substrate activity for SVV-encoded thymidine kinase. Biochem. Biophys. Res. Commun., 315, 877–883.CrossRefGoogle Scholar
Sienaert, R., Naesens, L., Brancale, A., De Clercq, E., McGuigan, C., and Balzarini, J. (2002). Specific recognition of the bicyclic pyrimidine nucleoside analogs, a new class of highly potent and selective inhibitors of varicella-zoster virus (VZV), by the VZV-encoded thymidine kinase. Mol. Pharmacol., 61, 249–254.CrossRefGoogle Scholar
Simpson-Holley, M., Baines, J., Roller, R., and Knipe, D. M. (2004). Herpes simplex virus 1 U(L)31 and U(L)34 gene products promote the late maturation of viral replication compartments to the nuclear periphery. J. Virol., 78, 5591–5600.CrossRefGoogle ScholarPubMed
Slater, M. J., Cockerill, S., Baxter, R.et al. (1999). Indolocarbazoles: potent, selective inhibitors of human cytomegalovirus replication. Bioorg. Med. Chem., 7, 1067–1074.CrossRefGoogle ScholarPubMed
Smith, R. F. and Smith, T. F. (1989). Identification of new protein kinase-related genes in three herpesviruses, herpes simplex virus, varicella-zoster virus, and Epstein–Barr virus. J. Virol., 63, 450–455.Google ScholarPubMed
Snoeck, R., Andrei, G., Bodaghi, B.et al. (2002). 2-Chloro-3-pyridin-3-yl-5,6,7,8-tetrahydroindolizine-1-carboxamide (CMV423), a new lead compound for the treatment of human cytomegalovirus infections. Antiviral Res., 55, 413–424.CrossRefGoogle Scholar
Soike, K. F., Bohm, R., Huang, J. L., and, Oberg, B. (1993). Efficacy of (-)-9-[4-hydroxy-2-(hydroxymethyl)butyl]guanine in African green monkeys infected with simian varicella virus. Antimicrob. Agents Chemother., 37, 1370–1372.CrossRefGoogle ScholarPubMed
Spector, F. C., Liang, L., Giordano, H., Sivaraja, M., and Peterson, M. G. (1998). Inhibition of herpes simplex virus replication by a 2-amino thiazole via interactions with the helicase component of the UL5-UL8-UL52 complex. J. Virol., 72, 6979–6987.Google ScholarPubMed
Spence, R. A., Kati, W. M., Anderson, K. S., and Johnson, K. A. (1995). Mechanism of inhibition of HIV-1 reverse transcriptase by nonnucleoside inhibitors. Science, 267, 988–993.CrossRefGoogle ScholarPubMed
Stamminger, T., Gstaiger, M., Weinzierl, K., Lorz, K., Winkler, M., and Schaffner, W. (2002). Open reading frame UL26 of human cytomegalovirus encodes a novel tegument protein that contains a strong transcriptional activation domain. J. Virol., 76, 4836–4847.CrossRefGoogle ScholarPubMed
Sullivan, V., Talarico, C. L., Stanat, S. C., Davis, M., Coen, D. M., and Biron, K. K. (1992). A protein kinase homologue controls phosphorylation of ganciclovir in human cytomegalovirus-infected cells. Nature, 359, 85.CrossRefGoogle ScholarPubMed
Supuran, C. T., Casini, A., and Scozzafava, A. (2003). Protease inhibitors of the sulfonamide type: anticancer, antiinflammatory, and antiviral agents. Med. Res. Rev., 23, 535–558.CrossRefGoogle ScholarPubMed
Talarico, C. L., Burnette, T. C., Miller, W. H.et al. (1999). Acyclovir is phosphorylated by the human cytomegalovirus UL97 protein. Antimicrob. Agents Chemother., 43, 1941–1946.Google ScholarPubMed
Tenney, D. J., Yamanaka, G., Voss, S. M.et al. (1997). Lobucavir is phosphorylated in human cytomegalovirus-infected and -uninfected cells and inhibits the viral DNA polymerase. Antimicrob. Agents Chemother., 41, 2680–2685.Google ScholarPubMed
Thoma, C., Borst, E., Messerle, M., Rieger, M., Hwang, J. S. and Bogner, E. (2006). Identification of the interaction domain of the small terminase subunit pUL89 with the large subunit pUL56 of human cytomegalovirus. Biochemistry, 45, 8855–8863.CrossRefGoogle ScholarPubMed
Thomsen, D. R., Oien, N. L., Hopkins, T. A.et al. (2003). Amino acid changes within conserved region III of the herpes simplex virus and human cytomegalovirus DNA polymerases confer resistance to 4-oxo-dihydroquinolines, a novel class of herpesvirus antiviral agents. J. Virol., 77, 1868–1876.CrossRefGoogle ScholarPubMed
Thomsen, L. L., Topley, P., Daly, M. G., Brett, S. J., and Tite, J. P. (2004). Imiquimod and resiquimod in a mouse model: adjuvants for DNA vaccination by particle-mediated immunotherapeutic delivery. Vaccine, 22, 1799–1809.CrossRefGoogle Scholar
Townsend, L. B. and Revankar, G. R. (1970). Benzimidazole nucleosides, nucleotides, and related derivatives. Chem. Rev., 70, 389–438.CrossRefGoogle ScholarPubMed
Townsend, L. B., Devivar, R. V., Turk, S. R., Nassiri, M. R., and Drach, J. C. (1995). Design, synthesis, and antiviral activity of certain 2,5,6-trihalo-l-(beta-D-ribofuranosyl)benzimidazoles. J. Med. Chem., 38, 4098–4105.CrossRefGoogle ScholarPubMed
Townsend, L. B., Gudmundsson, K. S., Daluge, S. M.et al. (1999). Studies designed to increase the stability and antiviral activity (HCMV) of the active benzimidazole nucleoside, TCRB. Nucleosides Nucleotides, 18, 509–519.CrossRefGoogle ScholarPubMed
Tsai, C. J., Lin, S. L., Wolfson, H. J., and Nussinov, R. (1997). Studies of protein-protein interfaces: a statistical analysis of the hydrophobic effect. Protein Sci., 6, 53–64.CrossRefGoogle ScholarPubMed
Underwood, M. R., Harvey, R. J., Stanat, S. C.et al. (1998). Inhibition of human cytomegalovirus DNA maturation by a benzimidazole ribonucleoside is mediated through the UL89 gene product. J. Virol., 72, 717–725.Google ScholarPubMed
Vaillancourt, V. A., Cudahy, M. M., Staley, S. A.et al. (2000). Naphthalene carboxamides as inhibitors of human cytomegalovirus DNA polymerase. Bioorg. Med. Chem. Lett., 10, 2079–2081.CrossRefGoogle ScholarPubMed
Van Zeijl, M., Fairhurst, J., Baum, E. Z., Sun, L., and Jones, T. R. (1997). The human cytomegalovirus UL97 protein is phosphory lated and a component of virions. Virology, 231, 72–80.CrossRefGoogle Scholar
Zeijl, M., Fairhurst, J., Jones, T. R.et al. (2000). Novel class of thiourea compounds that inhibit herpes simplex virus type 1 DNA cleavage and encapsidation: resistance maps to the UL6 gene. J. Virol., 74, 9054–9061.CrossRefGoogle ScholarPubMed
Visalli, R. J. and Zeijl, M. (2003). DNA encapsidation as a target for antiherpesvirus drug therapy. Antiviral Res., 59, 73–87.CrossRefGoogle Scholar
Visalli, R. J., Fairhurst, J., Srinivas, S.et al. (2003). Identification of small molecule compounds that selectively inhibit varicella-zoster virus replication. J. Virol., 77, 2349–2358.CrossRefGoogle ScholarPubMed
Wang, L. H., Peck, R. W., Yin, Y., Allanson, J., Wiggs, R., and Wire, M. W. (2003). Phase I safety and pharmacokinetic trials of 1263W94, a novel oral anti-human cytomegalovirus agent, in healthy and human immunodeficiency virus-infected subject. Antimicrob. Agents Chemother., 47, 1334–1342.CrossRefGoogle Scholar
Wan, W. B., Beadle, J. R., Hartline, C.et al. (2005). Comparison of the antiviral activities of alkoxyalkyl and alkyl esters of cidofovir against human and murine cytomegalovirus replication in vitro. Antimicrob. Agents. Chemother., 49, 656–662.CrossRefGoogle ScholarPubMed
Wathen, M. W. (2002). Non-nucleoside inhibitors of herpesviruses. Rev. Med. Virol., 12, 167–178.CrossRefGoogle Scholar
Waxman, L. and Darke, P. L. (2000). The herpesvirus proteases as targets for antiviral chemotherapy. Antivir. Chem. Chemother., 11, 1–22.CrossRefGoogle ScholarPubMed
Wang, Y., Abel, K., Lantz, K., Krieg, A. M., McChesney, M. B., and Miller, C. J. (2005). The Toll-like receptor 7 (TLR7) agonist, imiquimod, and the TLR9 agonist, CpG ODN, induce antiviral cytokines and chemokines but do not prevent vaginal transmission of simian immunodeficiency virus when applied intravaginally to rhesus macaques. J. Virol., 79, 14355–14370.CrossRefGoogle Scholar
Weber, O., Bender, W., Eckenberg, P.et al. (2001). Inhibition of murine cytomegalovirus and human cytomegalovirus by a novel non-nucleosidic compound in vivo. Antiviral Res., 49, 179–189.CrossRefGoogle ScholarPubMed
White, C. A., Stow, N. D., Patel, A. H., Hughes, M., and Preston, V. G. (2003). Herpes simplex virus type 1 portal protein UL6 interacts with the putative terminase subunits UL15 and UL28. J Virol., 77, 6351–6358.CrossRefGoogle ScholarPubMed
Williams, S. L., Hartline, C. B., Kushner, N. L.et al. (2003). In vitro activities of benzimidazole D- and L-ribonucleosides against herpesviruses. Antimicrob. Agents Chemother., 47, 2186–2192.CrossRefGoogle Scholar
Williams-Aziz, S. L., Hartline, C. B., Harden, E. A.et al. (2005). Comparative activities of lipid esters of cidofovir and cyclic cidofovir against replication of herpesviruses in vitro. Antimicrob. Agents Chemother., 49, 3724–3733.CrossRefGoogle ScholarPubMed
Wolf, D. G., Courcelle, C. T., Prichard, M. N., and Mocarski, E. S. (2001). Distinct and separate roles for herpesvirus-conserved UL97 kinase in cytomegalovirus DNA synthesis and encapsidation. Proc. Natl Acad. Sci.USA, 98, 1895–1900.CrossRefGoogle ScholarPubMed
Wolf, D. G., Courcelle, C. T., Prichard, M. N., and Mocarski, E. S. (2001). Distinct and separate roles for herpesvirus-conserved UL97 kinase in cytomegalovirus DNA synthesis and encapsidation. Proc. Natl Acad. Sci. U.S.A, 98, 1895–1900.CrossRefGoogle ScholarPubMed
Zacny, V. L., Gershburg, E., Davis, M. G., Biron, K. K., and Pagano, J. S. (1999). Inhibition of Epstein-Barr virus replication by a benzimidazole L-riboside: novel antiviral mechanism of 5,6-dichloro-2-(isopropylamino)-1-beta-L-ribofuranosyl-1H-benzimidazole. J Virol., 73, 7271–7277.Google ScholarPubMed
Zimmermann, A., Wilts, H., Lenhardt, M., Hahn, M., and Mertens, T. (2000). Fifteenth International Conference on Antiviral Res. Indolocarbazoles exhibit strong antiviral activity against human cytomegalovirus and are potent inhibitors of the pUL97 protein kinase. Antiviral Res., 48, 49–60.CrossRefGoogle 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
×