Skip to main content
Log in

Oncolytic Viruses for Cancer Therapy

  • Leading Article
  • Published:
American Journal of Cancer

Abstract

After being recognized for their antineoplastic properties at the beginning of the last century, viruses are again being considered for use as therapeutic agents against cancer. Certain species of virus have a propensity to replicate within transformed cells, commonly rendered vulnerable because of tumor-specific defects in their defense against viral infection. Other viruses have been modified to tumor-specific growth conditions. Oncolytic viruses carry the promise to efficiently target cancer cells for destruction and spread throughout tumor tissue to reach distant neoplastic loci without causing collateral damage to healthy tissues.

In contrast to conventional cancer chemotherapy, viral antineoplastic agents require complex interactions with the host organism to reach their target and to develop oncolytic activity. Recent progress in the elucidation of the molecular mechanisms of viral pathogenesis has opened up new opportunities to manipulate virus-host interactions, generating effective antitumor strategies. On the other hand, significant obstacles towards the application of safe and efficacious viral therapies have become apparent. These frequently relate to the lack of cell culture and animal tumor models that accurately reflect the characteristics of cancerous tissues in patients.

Throughout the past century, viral therapeutics against cancer has evolved into a new class of treatment strategies characterized by unique opportunities and challenges. A growing number of oncolytic viruses have entered clinical investigation or are scheduled to do so in the near future. Great efforts are being undertaken to rekindle an old idea and, with the help of new technologies, to realize its promise of new treatment options for cancer.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Table I
Table II

Similar content being viewed by others

References

  1. Hawkins LK, Lemoine NR, Kirn D. Oncolytic biotherapy: a novel therapeutic platform. Lancet Oncol 2002; 3: 17–26

    Article  PubMed  CAS  Google Scholar 

  2. Gromeier M. Viruses as therapeutic agents against malignant disease of the central nervous system. J Nat Cancer Inst 2001; 93: 889–90

    Article  PubMed  CAS  Google Scholar 

  3. Gromeier M, Wimmer E. Viruses for the treatment of malignant glioma. Curr Opin Mol Ther 2001; 3: 503–8

    PubMed  CAS  Google Scholar 

  4. Kirn D, Martuza RL, Zwiebel J. Replication-selective virotherapy for cancer: biological principles, risk management and future directions. Nat Med 2001; 7: 781–7

    Article  PubMed  CAS  Google Scholar 

  5. Mullen JT, Tanabe KK. Viral oncolysis. Oncologist 2002; 7: 106–19

    Article  PubMed  CAS  Google Scholar 

  6. Sinkovics J, Horvath J. New developments in the virus therapy of cancer: a historical review. Intervirology 1993; 36: 193–214

    PubMed  CAS  Google Scholar 

  7. DePace NG. Rabies virus treatment of cervical cancer. Ginecologia 1912; 9: 82

    Google Scholar 

  8. Levaditi C, Nicolau S. Vaccine et neoplasmes. Ann Inst Pasteur 1923; 37: 1–3

    Google Scholar 

  9. Tomko RP, Xu R, Philipson L. HCAR and MCAR: the human and mouse cellular receptors for subgroup C adenoviruses and group B coxsackieviruses. Proc Natl Acad Sci U S A 1997; 94: 3352–6

    Article  PubMed  CAS  Google Scholar 

  10. Bergelson JM, Cunningham JA, Droguett G, et al. Isolation of a common receptor for coxsackie B viruses and adenoviruses 2 and 5. Science 1997; 275: 1320–3

    Article  PubMed  CAS  Google Scholar 

  11. Goldman MJ, Wilson JM. Expression of αvβ5 integrin is necessary for efficient adenovirus-mediated gene transfer in the human airway. J Virol 1995; 69: 5951–8

    PubMed  CAS  Google Scholar 

  12. Huang S, Endo RI, Nemerow GR. Upregulation of integrins αvβ3 and αvβ5 on human monocytes and T lymphocytes facilitates adenovirus-mediated gene delivery. J Virol 1995; 69: 2257–63

    PubMed  CAS  Google Scholar 

  13. Montgomery RI, Warner MS, Lum BJ, et al. Herpes simplex virus-1 entry into cells mediated by a novel member of the TNF/NGF receptor family. Cell 1996; 87: 427–36

    Article  PubMed  CAS  Google Scholar 

  14. Cocchi F, Menotti L, Mirandola P, et al. The ectodomain of a novel member of the immunoglobulin subfamily related to the poliovirus receptor has the attributes of a bona fide receptor for herpes simplex virus types 1 and 2 in human cells. J Virol 1998; 72: 9992–10002

    PubMed  CAS  Google Scholar 

  15. Geraghty RJ, Krummenacher C, Cohen GH, et al. Entry of alphaherpesviruses mediated by poliovirus receptor-related protein 1 and poliovirus receptor. Science 1998; 280: 1618–20

    Article  PubMed  CAS  Google Scholar 

  16. Shukla D, Liu J, Blaiklock P, et al. A novel role for 3-O-sulfated heparan sulfate in herpes simplex virus 1 entry. Cell 1999; 99: 13–22

    Article  PubMed  CAS  Google Scholar 

  17. Chappell JD, Gunn VL, Wetzel JD, et al. Mutations in type 3 reovirus that determine binding to sialic acid are contained in the fibrous tail domain of viral attachment protein σ1. J Virol 1997; 71: 1834–41

    PubMed  CAS  Google Scholar 

  18. Barton ES, Forrest JC, Connolly JL, et al. Junction adhesion molecule is a receptor for reovirus. Cell 2001; 104: 441–51

    Article  PubMed  CAS  Google Scholar 

  19. Mendelsohn CL, Wimmer E, Racaniello VR. Cellular receptor for poliovirus: molecular cloning, nucleotide sequence, and expression of a new member of the immunoglobulin superfamily. Cell 1989; 56: 855–65

    Article  PubMed  CAS  Google Scholar 

  20. Shafren DR, Dorahy DJ, Greive SJ, et al. Mouse cells expressing human intercellular adhesion molecule-1 are susceptible to infection by coxsackievirus A21. J Virol 1997; 71: 785–9

    PubMed  CAS  Google Scholar 

  21. Cripe TP, Dunphy EJ, Holub AD, et al. Fiber knob modifications overcome low, heterogeneous expression of the coxsackievirus-adenovirus receptor that limits adenovirus gene transfer and oncolysis for human rhabdomyosarcoma cells. Cancer Res 2001; 61: 2953–60

    PubMed  CAS  Google Scholar 

  22. Douglas JT, Kim M, Sumerel LA, et al. Efficient oncolysis by a replicating adenovirus (ad) in vivo is critically dependent on tumor expression of primary ad receptors. Cancer Res 2001; 61: 813–7

    PubMed  CAS  Google Scholar 

  23. Miller CR, Buchsbaum DJ, Reynolds PN, et al. Differential susceptibility of primary and established human glioma cells to adenovirus infection: targeting via the epidermal growth factor receptor achieves fiber receptor-independent gene transfer. Cancer Res 1998; 58: 5738–48

    PubMed  CAS  Google Scholar 

  24. Asaoka K, Tada M, Sawamura Y, et al. Dependence of efficient adenoviral gene delivery in malignant glioma cells on the expression levels of the Coxsackievirus and adenovirus receptor. J Neurosurg 2000; 92: 1002–8

    Article  PubMed  CAS  Google Scholar 

  25. Li D, Duan L, Freimuth P, et al. Variability of adenovirus receptor density influences gene transfer efficiency and therapeutic response in head and neck cancer. Clin Cancer Res 1999; 5: 4175–81

    PubMed  CAS  Google Scholar 

  26. Li Y, Pong RC, Bergelson JM, et al. Loss of adenoviral receptor expression in human bladder cancer cells: a potential impact on the efficacy of gene therapy. Cancer Res 1999; 59: 325–30

    PubMed  CAS  Google Scholar 

  27. Wickham TJ. Targeting adenovirus. Gene Ther 2000; 7: 110–4

    Article  PubMed  CAS  Google Scholar 

  28. Krasnykh V, Dmitriev I, Navarro JG, et al. Advanced generation adenoviral vectors possess augmented gene transfer efficiency based upon coxsackie adenovirus receptor-independent cellular entry capacity. Cancer Res 2000; 60: 6784–7

    PubMed  CAS  Google Scholar 

  29. Spears PG. A first step toward understanding membrane fusion induced by herpes simplex virus. Mol Cell 2001; 8: 2–4

    Article  Google Scholar 

  30. Mata M, Zhang M, Hu X, et al. HveC (nectin-1) is expressed at high levels in sensory neurons, but not in motor neurons, of the rat peripheral nervous system. J Neurovirol 2001; 7: 476–80

    Article  PubMed  CAS  Google Scholar 

  31. Arita M, Koike S, Aoki J, et al. Interaction of poliovirus with its purified receptor and conformational alteration in the virion. J Virol 1998; 72: 3578–86

    PubMed  CAS  Google Scholar 

  32. Wimmer E, Harber JJ, Bibb JA, et al. Poliovirus receptors. In: Wimmer E, editor. Cellular receptors for animal viruses. Plainview (NY): Cold Spring Harbor Laboratory Press, 1994: 101–27

    Google Scholar 

  33. Gromeier M, Solecki D, Patel D, et al. Expression of the human poliovirus receptor/CD155 gene during development of the CNS: implications for the pathogenesis of poliomyelitis. Virology 2000; 273: 248–57

    Article  PubMed  CAS  Google Scholar 

  34. Bischoff JR, Kirn DH, Williams A, et al. An adenovirus mutant that replicates selectively in p53-deficient human tumor cells. Science 1996; 274: 373–6

    Article  PubMed  CAS  Google Scholar 

  35. Habib NA, Sarraf CE, Mitry RR, et al. E1B-deleted adenovirus (dl1520) gene therapy for patients with primary and secondary liver tumors. Hum Gene Ther 2000; 12: 219–26

    Article  Google Scholar 

  36. Reid TR, Galanis E, Abbruzzese J, et al. Intra-arterial administration of a replication-selective adenovirus Ci-1042 (Onyx-015) in patients with colorectal carcinoma metastatic to the liver: safety, feasibility and biological activity [abstract 549]. Proc Am Soc Clin Oncol 2001; 20: 138a

    Google Scholar 

  37. Mulvihill S, Warren R, Venook A, et al. Safety and feasibility of injection with an E1B-55kDa gene-deleted, replication-selective adenovirus (ONYX-015) into primary carcinomas of the pancreas: a phase I trial. Gene Ther 2001; 8: 308–15

    Article  PubMed  CAS  Google Scholar 

  38. Lamont JP, Nemunaitis J, Kuhn JA, et al. A prospective phase II trial of ONYX-015 adnovirus and chomterhapy in recurrent squamous cell carcinoma of the head and neck (the Baylor experience). Ann Surg Oncol 2000; 7: 588–92

    PubMed  CAS  Google Scholar 

  39. Nemunaitis J, Khuri F, Ganly I, et al. Phase II trial of intratumoral administration of ONYX-015, a replication-selective adenovirus, in patients with refractory head and neck cancer. J Clin Oncol 2001; 19: 289–98

    PubMed  CAS  Google Scholar 

  40. Rodriguez R, Schuur ER, Lim HY, et al. Prostate attenuated replication competent adenovirus (ARCA) CN706: a selective cytotoxic for prostate-specific antigen-positive prostate cancer cells. Cancer Res 1997; 57(13): 2559–63

    PubMed  CAS  Google Scholar 

  41. Yu D, Chen Y, Seng M, et al. The addition of adenovirus type 5 region E3 enables Calydon virus 787 to eliminate distant prostate tumor xenografts. Cancer Res 1999; 59: 4200–3

    PubMed  CAS  Google Scholar 

  42. Yu DC, Chen Y, Dilley J, et al. Antitumor synergy of CV787, a prostate cancer-specific adenovirus, and paclitaxel and docetaxel. Cancer Res 2001; 61: 517–25

    PubMed  CAS  Google Scholar 

  43. Kurihara T, Brough DE, Kovesdi I, et al. Selectivity of a replication-competent adenovirus for human breast carcinoma cells expressing the MUC1 antigen. J Clin Invest 2000; 106: 763–71

    Article  PubMed  CAS  Google Scholar 

  44. Hallenbeck PL, Chang YN, Hay C, et al. A novel tumor-specific replication-restricted adenoviral vector for gene therapy of hepatocellular carcinoma. Hum Gene Ther 1999; 10: 1721–33

    Article  PubMed  CAS  Google Scholar 

  45. Martuza RL, Malick A, Markert JM, et al. Experimental therapy of human glioma by means of a genetically engineered virus mutant. Science 1991; 252: 854–6

    Article  PubMed  CAS  Google Scholar 

  46. Mineta T, Rabkin SD, Martuza RL. Treatment of malignant gliomas using a ganciclovir-hypersensitive, ribonucleotide reductase-deficient herpes simplex viral mutant. Cancer Res 1994; 54: 3963–6

    PubMed  CAS  Google Scholar 

  47. Andreansky SS, He B, Gillespie GY, et al. The application of genetically engineered herpes simplex viruses to the treatment of experimental brain tumors. Proc Natl Acad Sci U S A 1996; 93: 11313–8

    Article  PubMed  CAS  Google Scholar 

  48. Coukos G, Makrigiannakis A, Kang EH, et al. Qncolytic herpes simplex virus-1 lacking ICP34.5 induces p53-independent death and is efficacious against chemotherapy-resistant ovarian cancer. Clin Cancer Res 2000; 6: 3342–53

    PubMed  CAS  Google Scholar 

  49. Mineta T, Rabkin SD, Yazaki T, et al. Attenuated multi-mutated herpes simplex virus-1 for the treatment of malignant gliomas. Nat Med 1995; 1: 938–43

    Article  PubMed  CAS  Google Scholar 

  50. Markert JM, Medlock MD, Rabkin SD, et al. Conditionally replicating herpes simplex virus mutant, G207 for the treatment of malignant glioma: results of a phase I trial. Gene Ther 2000; 7: 867–74

    Article  PubMed  CAS  Google Scholar 

  51. Kooby DA, Carew IF, Halterman MW, et al. Oncolytic viral therapy for human colorectal cancer and liver metastases using a multi-mutated herpes simplex virus type-1 (G207). FASEB J 1999; 13: 1325–34

    PubMed  CAS  Google Scholar 

  52. Toda M, Rabkin SD, Martuza RL. Treatment of human breast cancer in a brain metastatic model by G207, a replication-competent multimutated herpes simplex virus 1. Hum Gene Ther 1998; 9: 2177–85

    Article  PubMed  CAS  Google Scholar 

  53. Chahlavi A, Todo T, Martuza RL, et al. Replication-competent herpes simplex virus vector G207 and cisplatin combination therapy for head and neck squamous cell carcinoma. Neoplasia 1999; 1: 162–9

    Article  PubMed  CAS  Google Scholar 

  54. Walker JR, McGeagh KG, Sundaresan P, et al. Local and systemic therapy of human prostate adenocarcinoma with the conditionally replicating herpes simplex virus vector G207. Hum Gene Ther 1999; 10: 2237–43

    Article  PubMed  CAS  Google Scholar 

  55. Meignier B, Martin B, Whitley RJ, et al. In vivo behavior of genetically engineered herpes simplex viruses R7017 and R7020: II. Studies in immunocompetent and immunosuppressed owl monkeys (Aotus trivirgatus). I Infect Dis 1990; 162: 313–21

    Article  PubMed  CAS  Google Scholar 

  56. Meignier B, Longnecker R, Roizman B. In vivo behavior of genetically engineered herpes simplex viruses R7017 and R7020: construction and evaluation in rodents. I Infect Dis 1988; 158: 602–14

    Article  PubMed  CAS  Google Scholar 

  57. McAuliffe PF, Jarnagin WR, Johnson P, et al. Effective treatment of pancreatic tumors with two multimutated herpes simplex oncolytic viruses. I Gastrointest Surg 2000; 4: 580–8

    Article  CAS  Google Scholar 

  58. Cozzi PI, Malhotra S, McAuliffe P, et al. Intravesical oncolytic viral therapy using attenuated, replication-competent herpes simplex viruses G207 and Nvl020 is effective in the treatment of bladder cancer in an orthotopic syngeneic model. FASEB J 2001; 15: 1306–8

    PubMed  CAS  Google Scholar 

  59. Wong RJ, Kim SH, Joe JK, et al. Effective treatment of head and neck squamous cell carcinoma by an oncolytic herpes simplex virus. J Am Coll Surg 2001; 193: 12–21

    Article  PubMed  CAS  Google Scholar 

  60. Miyatake S, Iyer A, Martuza RL, et al. Transcriptional targeting of herpes simplex virus for cell-specific replication. J Virol 1997; 71: 5124–32

    PubMed  CAS  Google Scholar 

  61. Miyatake S, Tani S, Feigenbaum F, et al. Hepatoma-specific anti-tumor activity of an albumin enhancer/promoter regulated herpes simplex virus in vivo. Gene Ther 1999; 6: 564–72

    Article  PubMed  CAS  Google Scholar 

  62. Mastrangelo MJ, Maguire Jr HC, Eisenlohr LC, et al. Intratumoral recombinant GM-CSF-encoding virus as gene therapy in patients with cutaneous melanoma. Cancer Gene Ther 1999; 6: 409–22

    Article  PubMed  CAS  Google Scholar 

  63. McCart JA, Ward JM, Lee J, et al. Systemic cancer therapy with a tumor-selective vaccinia virus mutant lacking thymidine kinase and vaccinia growth factor genes. Cancer Res 2001; 61: 8751–7

    PubMed  CAS  Google Scholar 

  64. Timiryasova TM, Li j, Chen B, et al. Antitumor effect of vaccinia virus in glioma model. Oncol Res 1999; 11: 133–44

    PubMed  CAS  Google Scholar 

  65. Coffey MC, Strong JE, Forsyth PA, et al. Reovirus therapy of tumors with activated Ras pathway. Science 1998; 282: 1332–4

    Article  PubMed  CAS  Google Scholar 

  66. Stojdl DF, Lichty B, Knowles S, et al. Exploiting tumor-specific defects in the interferon pathway with a previously unknown oncolytic virus. Nat Med 2000; 6: 821–5

    Article  PubMed  CAS  Google Scholar 

  67. Gromeier M, Lachmann S, Rosenfeld MR, et al. Nonpathogenic intergeneric poliovirus recombinants for the treatment of glioma. Proc Natl Acad Sci U S A 2000; 97: 6803–8

    Article  PubMed  CAS  Google Scholar 

  68. Shafren ER, Au GG, Barry G. Therapeutic application of coxsackievirus A15 and A21 as novel oncolytic anti-tumour agents for control of malignant human melanoma [abstract]. Europic 2002; 19: K8

    Google Scholar 

  69. Cassel WA, Garrett RE. Newcastle disease virus as an anti-neoplastic agent. Cancer 1965; 18: 863–8

    Article  PubMed  CAS  Google Scholar 

  70. Csatary LK, Bakacs T. Use of Newcastle disease virus vaccine (MTH-68/H) in a patient with high-grade glioblastoma. JAMA 1999; 281: 1588–9

    Article  PubMed  CAS  Google Scholar 

  71. Pecora AL, Rizvi N, Cohen GI, et al. Phase I trial of intravenous administration of PV701, an oncolytic virus, in patients with advanced solid cancers. J Clin Oncol 2002; 20: 2251–66

    Article  PubMed  CAS  Google Scholar 

  72. Solecki DJ, Gromeier M, Mueller S, et al. Expression of the human poliovirus receptor/CD155 gene is activated by sonic hedgehog. J Biol Chem 2002; 277(28): 25691–102

    Article  CAS  Google Scholar 

  73. Kinzler KW, Bigner SH, Bigner DD, et al. Identification of an amplified, highly expressed gene in a human glioma. Science 1987; 236: 70–3

    Article  PubMed  CAS  Google Scholar 

  74. Goodrich LV, Milenkovic L, Higgins KM, et al. Altered neural cell fates and medulloblastoma in mouse patched mutants. Science 1997; 277: 1109–13

    Article  PubMed  CAS  Google Scholar 

  75. Merrill M, Bernhard G, Sampson J, et al. Molecular targeting of malignant glioma with oncolytic poliovirus recombinants. In press

  76. Masson D, Jarry A, Baury B, et al. Overexpression of the CD155 gene in human colorectal carcinoma. Gut 2001; 49: 236–40

    Article  PubMed  CAS  Google Scholar 

  77. Muller U, Steinhoff U, Reis LFL, et al. Functional roles of the type I and II interferons in antiviral defense. Science 1994; 264: 1918–21

    Article  PubMed  CAS  Google Scholar 

  78. Dreiding P, Staeheli P, Haller O. Interferon-induced protein Mx accumulates in nuclei of mouse cells expressing resistance to influenza viruses. Virology 1985; 140: 192–6

    Article  PubMed  CAS  Google Scholar 

  79. Staeheli P, Haller O, Boll W, et al. Mx protein: constitutive expression in 3T3 cells transformed with cloned Mx cDNA confers selective resistance to influenza virus. Cell 1986; 44: 147–58

    Article  PubMed  CAS  Google Scholar 

  80. Chebath J, Benech P, Revel M, et al. Constitutive expression of (2′–5′) oligo A synthetase confers resistance to picornavirus infection. Nature 1987; 330: 587–8

    Article  PubMed  CAS  Google Scholar 

  81. Zaragoza C, Ocampo C, Saura M, et al. The role of inducible nitric oxide synthase in the host response to Coxsackievirus myocarditis. Proc Natl Acad Sei U S A 1998; 95: 2469–74

    Article  CAS  Google Scholar 

  82. Der SD, Lau AS. Involvement of the double-stranded-RNA-dependent kinase PKR in interferon expression and interferon-mediated antiviral activity. Proc Natl Acad Sci U S A 1995; 92: 8841–5

    Article  PubMed  CAS  Google Scholar 

  83. Colamonici OR, Domanski P, Platanias LC, et al. Correlation between interferon (IFN) alpha resistance and deletion of the IFN alpha/beta genes in acute leukemia cell lines suggests selection against the IFN system. Blood 1992; 80: 744–9

    PubMed  CAS  Google Scholar 

  84. Wong LH, Krauer KG, Hatzinisiriou I, et al. Interferon-resistant human melanoma cells are deficient in ISGF3 components, STAT1, STAT2, and p48-ISGF3gamma. J Biol Chem 1997; 272: 28779–85

    Article  PubMed  CAS  Google Scholar 

  85. Abril E, Mendez RE, Garcia A, et al. Characterization of a gastric tumor cell line defective in MHC class I inducibility by both alpha- and gamma-interferon. Tissue Antigens 1996; 47: 391–8

    Article  PubMed  CAS  Google Scholar 

  86. Samuel CE. Reoviruses and the interferon system. Curr Top Microbiol Immunol 1998; 233: 125–45

    Article  PubMed  CAS  Google Scholar 

  87. Duncan MR, Stanish SM, Cox DC. Differential sensitivity of normal and transformed human cells to reovirus infection. J Virol 1978; 28: 444–9

    PubMed  CAS  Google Scholar 

  88. Hashiro G, Loh PC, Yau JT. The preferential cytotoxicity of reovirus for certain transformed cell lines. Arch Virol 1977; 54: 307–15

    Article  PubMed  CAS  Google Scholar 

  89. Strong JE, Coffey MC, Tang D, et al. The molecular basis of viral oncolysis: usurpation of the Ras signaling pathway by reovirus. EMBO J 1998; 17: 3351–62

    Article  PubMed  CAS  Google Scholar 

  90. Belkowski LS, Sen GC. Inhibition of vesicular stomatitis viral mRNA synthesis by interferons. J Virol 1987; 61: 653–60

    PubMed  CAS  Google Scholar 

  91. Stojdl DF, Abraham N, Knowles S, et al. The murine double-stranded RNA-dependent protein kinase PKR is required for resistance to vesicular stomatitis virus. J Virol 2000; 74: 9580–5

    Article  PubMed  CAS  Google Scholar 

  92. Balachandran S, Roberts PC, Brown LE, et al. Essential role for the dsRNA-dependent protein kinase PKR in innate immunity to viral infection. Immunity 2000; 13: 129–41

    Article  PubMed  CAS  Google Scholar 

  93. Balachandran S, Porosnicu M, Barber GN. Oncolytic activity of vesicular stomatitis virus is effective against tumors exhibiting aberrant p53, Ras, or myc function and involves the induction of apoptosis. J Virol 2001; 75: 3474–9

    Article  PubMed  CAS  Google Scholar 

  94. Levaditi C, Haber P. Affinité du virus de la peste aviaire pour les cellules neoplastiques. Compt Rend Soc Biol 1936; 202: 2018–20

    Google Scholar 

  95. Roberts MS, Buasen PT, Incao BA, et al. PV701, a naturally attenuated strain of Newcastle disease virus, has a broad spectrum of oncolytic activity against human tumor xenografts [abstract 2441]. Proc Am Assoc Cancer Res 2001; 42: 454

    Google Scholar 

  96. Lorence RM, Roberts MS, Groene WS, et al. Regression of human tumor xenografts following intravenous treatment using PV701, a naturally attenuated oncolytic strain of Newcastle disease virus [abstract 2442]. Proc Am Assoc Cancer Res 2001; 42: 454

    Google Scholar 

  97. Tenser RB, Miller RL, Rapp F. Trigeminal ganglion infection by thymidine kinase-negative mutants of herpes simplex virus. Science 1979; 205: 915–7

    Article  PubMed  CAS  Google Scholar 

  98. Coen DM, Kosz-Vnenchak M, Jacobson JG, et al. Thymidine kinase-negative herpes simplex virus mutants establish latency in mouse trigeminal ganglia but do not reactivate. Proc Natl Acad Sci U S A 1989; 86: 4736–40

    Article  PubMed  CAS  Google Scholar 

  99. Goldstein DJ, Weller SK. Herpes simplex virus type 1-induced ribonucleotide reductase activity is dispensable for virus growth and DNA synthesis: isolation and characterization of an ICP6 lacZ insertion mutant. J Virol 1988; 62: 196–205

    PubMed  CAS  Google Scholar 

  100. Whitley RJ, Kern ER, Chatterjee S, et al. Replication, establishment of latency, and induced reactivation of herpes simplex virus gamma 1 34.5 deletion mutants in rodent models. J Clin Invest 1993; 91: 2837–43

    Article  PubMed  CAS  Google Scholar 

  101. Chou J, Kern ER, Whitley RJ, et al. Mapping of herpes simplex virus-1 neurovirulence to gamma 134.5, a gene nonessential for growth in culture. Science 1990; 250: 1262–6

    Article  PubMed  CAS  Google Scholar 

  102. Chou J, Roizman B. The gamma 1(34.5) gene of herpes simplex virus 1 precludes neuroblastoma cells from triggering total shutoff of protein synthesis characteristic of programed cell death in neuronal cells. Proc Natl Acad Sci U S A 1992; 89: 3266–70

    Article  PubMed  CAS  Google Scholar 

  103. Chou J, Chen JJ, Gross M, et al. Association of a M(r) 90,000 phosphoprotein with protein kinase PKR in cells exhibiting enhanced phosphorylation of translation initiation factor eIF-2 alpha and premature shutoff of protein synthesis after infection with gamma 134.5- mutants of herpes simplex virus 1. Proc Natl Acad Sci U S A 1995; 92: 10516–20

    Article  PubMed  CAS  Google Scholar 

  104. Gale Jr M, Katze MG. Molecular mechanisms of interferon resistance mediated by viral-directed inhibition of PKR, the interferon-induced protein kinase. Pharmacol Ther 1998; 78: 29–46

    Article  PubMed  CAS  Google Scholar 

  105. Buller RM, Smith GL, Cremer K, et al. Decreased virulence of recombinant vaccinia virus expression vectors is associated with a thymidine kinase-negative phenotype. Nature 1985; 317: 813–5

    Article  PubMed  CAS  Google Scholar 

  106. Buller RM, Chakrabarti S, Cooper JA, et al. Deletion of the vaccinia virus growth factor gene reduces virus virulence. J Virol 1988; 62: 866–74

    PubMed  CAS  Google Scholar 

  107. Flint J, Shenk T. Viral transactivating proteins. Annu Rev Genet 1997; 31: 177–212

    Article  PubMed  CAS  Google Scholar 

  108. De Stanchina E, McCurrach ME, Zindy F, et al. E1A signaling to p53 involves the p19(ARF) tumour suppressor. Genes Dev 1998; 12: 2434–42

    Article  PubMed  Google Scholar 

  109. Miyashita T, Reed JC. Tumour suppressor p53 is a direct transcriptional activator of the human bax gene. Cell 1995; 80: 293–9

    Article  PubMed  CAS  Google Scholar 

  110. Yew PR, Berk AJ. Inhibition of p53 transactivation required for transformation by adenovirus early 1B protein. Nature 1992; 357: 82–5

    Article  PubMed  CAS  Google Scholar 

  111. Yew PR, Liu X, Berk AJ. Adenovirus E1B oncoprotein tethers a transcriptional repression domain to p53. Genes Dev 1994; 8: 190–202

    Article  PubMed  CAS  Google Scholar 

  112. Kao CC, Yew PR, Berk AJ. Domains required for in vitro association between the cellular p53 and the adenovirus 2 E1B 55K proteins. Virology 1990; 179: 806–14

    Article  PubMed  CAS  Google Scholar 

  113. Steegenga WT, Riteco N, Jochemsen AG, et al. The large E1B protein together with E4orf6 protein target p53 for active degradation in adenovirus infected cells. Oncogene 1998; 16: 349–57

    Article  PubMed  CAS  Google Scholar 

  114. Dix BR, Edwards S J, Braithwaite AW. Does the antitumor adenovirus ONYX-015/ dl1520 selectively target cells defective in the p53 pathway? J Virol 2001; 75: 5443–7

    Article  PubMed  CAS  Google Scholar 

  115. Goodrum FD, Ornelles DA. p53 status does not determine outcome of E1b 55-kilodalton mutant adenovirus lytic infection. J Virol 1998; 72: 9479–90

    PubMed  CAS  Google Scholar 

  116. Harada JN, Berk AJ. p53-independent and —dependent requirements for E1B-55K in adenovirus type 5 replication. J Virol 1999; 73: 5333–44

    PubMed  CAS  Google Scholar 

  117. Rothmann T, Hengstermann A, Whitaker J, et al. Replication of ONYX-015, a potential anticancer adenovirus, is independent of p53 status in tumor cells. J Virol 1998; 72: 9470–8

    PubMed  CAS  Google Scholar 

  118. Turnell AS, Grand RJ, Gallimore H. The replicative capacities of large E1B-null group A and group C adenoviruses are independent of host cell p53 status. J Virol 1999; 73: 2074–83

    PubMed  CAS  Google Scholar 

  119. Koch P, Gatfield J, Lober C, et al. Efficient replication of adenovirus despite the overexpression of active and nondegradable p53. Cancer Res 2001; 61: 5941–7

    PubMed  CAS  Google Scholar 

  120. Jackson RJ, Kaminski A. Internal initiation of translation in eukaryotes: the picornavirus paradigm and beyond. RNA 1995; 1: 985–1000

    PubMed  CAS  Google Scholar 

  121. Gromeier M, Alexander L, Wimmer E. Internal ribosomal entry site substitution eliminates neurovirulence in intergeneric poliovirus recombinants. Proc Natl Acad Sci U S A 1996; 93: 2370–5

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements

The author was supported by Public Health Service grant CA87537 and by a Career Award from Burroughs Wellcome Fund. The author has no conflicts of interest that are directly relevant to the content of this manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Matthias Gromeier.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Gromeier, M. Oncolytic Viruses for Cancer Therapy. Am J Cancer 2, 313–323 (2003). https://doi.org/10.2165/00024669-200302050-00002

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.2165/00024669-200302050-00002

Keywords

Navigation