#PAGE_PARAMS# #ADS_HEAD_SCRIPTS# #MICRODATA#

From Scourge to Cure: Tumour-Selective Viral Pathogenesis as a New Strategy against Cancer


Tumour mutations corrupt cellular pathways, and accumulate to disrupt, dysregulate, and ultimately avoid mechanisms of cellular control. Yet the very changes that tumour cells undergo to secure their own growth success also render them susceptible to viral infection. Enhanced availability of surface receptors, disruption of antiviral sensing, elevated metabolic activity, disengagement of cell cycle controls, hyperactivation of mitogenic pathways, and apoptotic avoidance all render the malignant cell environment highly supportive to viral replication. The therapeutic use of oncolytic viruses (OVs) with a natural tropism for infecting and subsequently lysing tumour cells is a rapidly progressing area of cancer research. While many OVs exhibit an inherent degree of tropism for transformed cells, this can be further promoted through pharmacological interventions and/or the introduction of viral mutations that generate recombinant oncolytic viruses adapted to successfully replicate only in a malignant cellular environment. Such adaptations that augment OV tumour selectivity are already improving the therapeutic outlook for cancer, and there remains tremendous untapped potential for further innovation.


Vyšlo v časopise: From Scourge to Cure: Tumour-Selective Viral Pathogenesis as a New Strategy against Cancer. PLoS Pathog 10(1): e32767. doi:10.1371/journal.ppat.1003836
Kategorie: Review
prolekare.web.journal.doi_sk: https://doi.org/10.1371/journal.ppat.1003836

Souhrn

Tumour mutations corrupt cellular pathways, and accumulate to disrupt, dysregulate, and ultimately avoid mechanisms of cellular control. Yet the very changes that tumour cells undergo to secure their own growth success also render them susceptible to viral infection. Enhanced availability of surface receptors, disruption of antiviral sensing, elevated metabolic activity, disengagement of cell cycle controls, hyperactivation of mitogenic pathways, and apoptotic avoidance all render the malignant cell environment highly supportive to viral replication. The therapeutic use of oncolytic viruses (OVs) with a natural tropism for infecting and subsequently lysing tumour cells is a rapidly progressing area of cancer research. While many OVs exhibit an inherent degree of tropism for transformed cells, this can be further promoted through pharmacological interventions and/or the introduction of viral mutations that generate recombinant oncolytic viruses adapted to successfully replicate only in a malignant cellular environment. Such adaptations that augment OV tumour selectivity are already improving the therapeutic outlook for cancer, and there remains tremendous untapped potential for further innovation.


Zdroje

1. HanahanD, WeinbergRA (2000) The hallmarks of cancer. Cell 100: 57–70.

2. RussellSJ, PengKW, BellJC (2012) Oncolytic virotherapy. Nat Biotechnol 30: 658–670.

3. MendelsohnCL, WimmerE, RacanielloVR (1989) Cellular receptor for poliovirus: molecular cloning, nucleotide sequence, and expression of a new member of the immunoglobulin superfamily. Cell 56: 855–865.

4. CarlstenM, NorellH, BrycesonYT, PoschkeI, SchedvinsK, et al. (2009) Primary human tumor cells expressing CD155 impair tumor targeting by down-regulating DNAM-1 on NK cells. J Immunol 183: 4921–4930.

5. EnloeBM, JayDG (2011) Inhibition of Necl-5 (CD155/PVR) reduces glioblastoma dispersal and decreases MMP-2 expression and activity. J Neurooncol 102: 225–235.

6. MassonD, JarryA, BauryB, BlanchardieP, LaboisseC, et al. (2001) Overexpression of the CD155 gene in human colorectal carcinoma. Gut 49: 236–240.

7. NakaiR, ManiwaY, TanakaY, NishioW, YoshimuraM, et al. (2010) Overexpression of Necl-5 correlates with unfavorable prognosis in patients with lung adenocarcinoma. Cancer Sci 101: 1326–1330.

8. SuzukiK, NakamuraK, KatoK, HamadaH, TsukamotoT (2007) Exploration of target molecules for prostate cancer gene therapy. Prostate 67: 1163–1173.

9. GoetzC, DobrikovaE, ShveygertM, DobrikovM, GromeierM (2011) Oncolytic poliovirus against malignant glioma. Future Virol 6: 1045–1058.

10. GaggarA, ShayakhmetovDM, LieberA (2003) CD46 is a cellular receptor for group B adenoviruses. Nat Med 9: 1408–1412.

11. SegermanA, AtkinsonJP, MarttilaM, DennerquistV, WadellG, et al. (2003) Adenovirus type 11 uses CD46 as a cellular receptor. J Virol 77: 9183–9191.

12. AndersonBD, NakamuraT, RussellSJ, PengKW (2004) High CD46 receptor density determines preferential killing of tumor cells by oncolytic measles virus. Cancer Res 64: 4919–4926.

13. MuhlebachMD, MateoM, SinnPL, PruferS, UhligKM, et al. (2011) Adherens junction protein nectin-4 is the epithelial receptor for measles virus. Nature 480: 530–533.

14. DeryckeMS, PambuccianSE, GilksCB, KallogerSE, GhidoucheA, et al. (2010) Nectin 4 overexpression in ovarian cancer tissues and serum: potential role as a serum biomarker. Am J Clin Pathol 134: 835–845.

15. Fabre-LafayS, Garrido-UrbaniS, ReymondN, GoncalvesA, DubreuilP, et al. (2005) Nectin-4, a new serological breast cancer marker, is a substrate for tumor necrosis factor-alpha-converting enzyme (TACE)/ADAM-17. J Biol Chem 280: 19543–19550.

16. TakanoA, IshikawaN, NishinoR, MasudaK, YasuiW, et al. (2009) Identification of nectin-4 oncoprotein as a diagnostic and therapeutic target for lung cancer. Cancer Res 69: 6694–6703.

17. SkeldingKA, BarryRD, ShafrenDR (2009) Systemic targeting of metastatic human breast tumor xenografts by Coxsackievirus A21. Breast Cancer Res Treat 113: 21–30.

18. GiaginisCT, ZarrosAC, PapaefthymiouMA, PapadopouliAE, SfiniadakisIK, et al. (2008) Coxsackievirus and adenovirus receptor expression in human endometrial adenocarcinoma: possible clinical implications. World J Surg Oncol 6: 59.

19. ShafrenDR, WilliamsDT, BarryRD (1997) A decay-accelerating factor-binding strain of coxsackievirus B3 requires the coxsackievirus-adenovirus receptor protein to mediate lytic infection of rhabdomyosarcoma cells. J Virol 71: 9844–9848.

20. WangKS, KuhnRJ, StraussEG, OuS, StraussJH (1992) High-affinity laminin receptor is a receptor for Sindbis virus in mammalian cells. J Virol 66: 4992–5001.

21. BasoloF, PollinaL, PaciniF, FontaniniG, MenardS, et al. (1996) Expression of the Mr 67,000 laminin receptor is an adverse prognostic indicator in human thyroid cancer: an immunohistochemical study. Clin Cancer Res 2: 1777–1780.

22. CastronovoV, ColinC, ClaysmithAP, ChenPH, LifrangeE, et al. (1990) Immunodetection of the metastasis-associated laminin receptor in human breast cancer cells obtained by fine-needle aspiration biopsy. Am J Pathol 137: 1373–1381.

23. CioceV, CastronovoV, ShmooklerBM, GarbisaS, GrigioniWF, et al. (1991) Increased expression of the laminin receptor in human colon cancer. J Natl Cancer Inst 83: 29–36.

24. MenardS, TagliabueE, ColnaghiMI (1998) The 67 kDa laminin receptor as a prognostic factor in human cancer. Breast Cancer Res Treat 52: 137–145.

25. JamiesonKV, HubbardSR, MerueloD (2011) Structure-guided identification of a laminin binding site on the laminin receptor precursor. J Mol Biol 405: 24–32.

26. VerheijeMH, RottierPJ (2012) Retargeting of viruses to generate oncolytic agents. Adv Virol 2012: 798526.

27. NakamuraT, PengKW, HarveyM, GreinerS, LorimerIA, et al. (2005) Rescue and propagation of fully retargeted oncolytic measles viruses. Nat Biotechnol 23: 209–214.

28. ParaskevakouG, AllenC, NakamuraT, ZollmanP, JamesCD, et al. (2007) Epidermal growth factor receptor (EGFR)-retargeted measles virus strains effectively target EGFR- or EGFRvIII expressing gliomas. Mol Ther 15: 677–686.

29. UngerechtsG, SpringfeldC, FrenzkeME, LampeJ, JohnstonPB, et al. (2007) Lymphoma chemovirotherapy: CD20-targeted and convertase-armed measles virus can synergize with fludarabine. Cancer Res 67: 10939–10947.

30. HasegawaK, NakamuraT, HarveyM, IkedaY, ObergA, et al. (2006) The use of a tropism-modified measles virus in folate receptor-targeted virotherapy of ovarian cancer. Clin Cancer Res 12: 6170–6178.

31. CoughlanL, VallathS, SahaA, FlakM, McNeishIA, et al. (2009) In vivo retargeting of adenovirus type 5 to alphavbeta6 integrin results in reduced hepatotoxicity and improved tumor uptake following systemic delivery. J Virol 83: 6416–6428.

32. GuptaV, WangW, SosnowskiBA, HofmanFM, ChenTC (2006) Fibroblast growth factor-2-retargeted adenoviral vector for selective transduction of primary glioblastoma multiforme endothelial cells. Neurosurg Focus 20: E26.

33. HarveyTJ, BurdonD, SteeleL, IngramN, HallGD, et al. (2010) Retargeted adenoviral cancer gene therapy for tumour cells overexpressing epidermal growth factor receptor or urokinase-type plasminogen activator receptor. Gene Ther 17: 1000–1010.

34. OhIK, MokH, ParkTG (2006) Folate immobilized and PEGylated adenovirus for retargeting to tumor cells. Bioconjug Chem 17: 721–727.

35. WangW, ZhuNL, ChuaJ, SwensonS, CostaFK, et al. (2005) Retargeting of adenoviral vector using basic fibroblast growth factor ligand for malignant glioma gene therapy. J Neurosurg 103: 1058–1066.

36. GalmicheMC, RindisbacherL, WelsW, WittekR, BucheggerF (1997) Expression of a functional single chain antibody on the surface of extracellular enveloped vaccinia virus as a step towards selective tumour cell targeting. J Gen Virol 78 (Pt 11) 3019–3027.

37. PaulS, GeistM, DottK, SnaryD, Taylor-PapadimitriouJ, et al. (2007) Specific tumor cell targeting by a recombinant MVA expressing a functional single chain antibody on the surface of intracellular mature virus (IMV) particles. Viral Immunol 20: 664–671.

38. StojdlDF, LichtyBD, tenOeverBR, PatersonJM, PowerAT, et al. (2003) VSV strains with defects in their ability to shutdown innate immunity are potent systemic anti-cancer agents. Cancer Cell 4: 263–275.

39. IndraccoloS (2010) Interferon-alpha as angiogenesis inhibitor: learning from tumor models. Autoimmunity 43: 244–247.

40. KotredesKP, GameroAM (2013) Interferons as inducers of apoptosis in malignant cells. J Interferon Cytokine Res 33: 162–170.

41. SubramaniamPS, JohnsonHM (1997) A role for the cyclin-dependent kinase inhibitor p21 in the G1 cell cycle arrest mediated by the type I interferons. J Interferon Cytokine Res 17: 11–15.

42. Hervas-StubbsS, Perez-GraciaJL, RouzautA, SanmamedMF, Le BonA, et al. (2011) Direct effects of type I interferons on cells of the immune system. Clin Cancer Res 17: 2619–2627.

43. DunnGP, BruceAT, SheehanKC, ShankaranV, UppaluriR, et al. (2005) A critical function for type I interferons in cancer immunoediting. Nat Immunol 6: 722–729.

44. FiolaC, PeetersB, FournierP, ArnoldA, BucurM, et al. (2006) Tumor selective replication of Newcastle disease virus: association with defects of tumor cells in antiviral defence. Int J Cancer 119: 328–338.

45. KrishnamurthyS, TakimotoT, ScroggsRA, PortnerA (2006) Differentially regulated interferon response determines the outcome of Newcastle disease virus infection in normal and tumor cell lines. J Virol 80: 5145–5155.

46. ShmulevitzM, PanLZ, GarantK, PanD, LeePW (2010) Oncogenic Ras promotes reovirus spread by suppressing IFN-beta production through negative regulation of RIG-I signaling. Cancer Res 70: 4912–4921.

47. StojdlDF, LichtyB, KnowlesS, MariusR, AtkinsH, et al. (2000) Exploiting tumor-specific defects in the interferon pathway with a previously unknown oncolytic virus. Nat Med 6: 821–825.

48. NaikS, NaceR, FederspielMJ, BarberGN, PengKW, et al. (2012) Curative one-shot systemic virotherapy in murine myeloma. Leukemia 26: 1870–1878.

49. NollM, BerchtoldS, LampeJ, MalekNP, BitzerM, et al. (2013) Primary resistance phenomena to oncolytic measles vaccine viruses. Int J Oncol 43: 103–112.

50. HaralambievaI, IankovI, HasegawaK, HarveyM, RussellSJ, et al. (2007) Engineering oncolytic measles virus to circumvent the intracellular innate immune response. Mol Ther 15: 588–597.

51. Le BoeufF, BatenchukC, Vaha-KoskelaM, BretonS, RoyD, et al. (2013) Model-based rational design of an oncolytic virus with improved therapeutic potential. Nat Commun 4: 1974.

52. BrunJ, McManusD, LefebvreC, HuK, FallsT, et al. (2010) Identification of genetically modified Maraba virus as an oncolytic rhabdovirus. Mol Ther 18: 1440–1449.

53. Le BoeufF, DialloJ-S, McCartJA, ThorneS, FallsT, et al. (2010) Synergistic interaction between oncolytic viruses augments tumor killing. Mol Ther 18: 888–895.

54. Alvarez-BreckenridgeCA, YuJ, PriceR, WeiM, WangY, et al. (2012) The histone deacetylase inhibitor valproic acid lessens NK cell action against oncolytic virus-infected glioblastoma cells by inhibition of STAT5/T-BET signaling and generation of gamma interferon. J Virol 86: 4566–4577.

55. BridleBW, ChenL, LemayCG, DialloJ-S, PolJ, et al. (2013) HDAC inhibition suppresses primary immune responses, enhances secondary immune responses, and abrogates autoimmunity during tumor immunotherapy. Mol Ther 21: 887–894.

56. MacTavishH, DialloJ-S, HuangB, StanfordM, Le BoeufF, et al. (2010) Enhancement of vaccinia virus based oncolysis with histone deacetylase inhibitors. PLoS ONE 5: e14462 doi:10.1371/journal.pone.0014462

57. ChangHM, PaulsonM, HolkoM, RiceCM, WilliamsBR, et al. (2004) Induction of interferon-stimulated gene expression and antiviral responses require protein deacetylase activity. Proc Natl Acad Sci U S A 101: 9578–9583.

58. NguyenTL, AbdelbaryH, ArguelloM, BreitbachC, LeveilleS, et al. (2008) Chemical targeting of the innate antiviral response by histone deacetylase inhibitors renders refractory cancers sensitive to viral oncolysis. Proc Natl Acad Sci U S A 105: 14981–14986.

59. DialloJ-S, Le BoeufF, LaiF, CoxJ, Vaha-KoskelaM, et al. (2010) A high-throughput pharmacoviral approach identifies novel oncolytic virus sensitizers. Mol Ther 18: 1123–1129.

60. Vander HeidenMG, CantleyLC, ThompsonCB (2009) Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science 324: 1029–1033.

61. AbrahamsMR, ZhangZ, ChienS, SkernsT, KotwalGJ (2005) The vaccinia virus N1L ORF may encode a multifunctional protein possibly targeting different kinases, one of which influences ATP levels in vivo. Ann N Y Acad Sci 1056: 87–99.

62. GammonDB, GowrishankarB, DuraffourS, AndreiG, UptonC, et al. (2010) Vaccinia virus–encoded ribonucleotide reductase subunits are differentially required for replication and pathogenesis. PLoS Pathog 6: e1000984 doi:10.1371/journal.ppat.1000984

63. ElfordHL, FreeseM, PassamaniE, MorrisHP (1970) Ribonucleotide reductase and cell proliferation. I. Variations of ribonucleotide reductase activity with tumor growth rate in a series of rat hepatomas. J Biol Chem 245: 5228–5233.

64. ChungRY, SaekiY, ChioccaEA (1999) B-myb promoter retargeting of herpes simplex virus gamma34.5 gene-mediated virulence toward tumor and cycling cells. J Virol 73: 7556–7564.

65. AghiM, VistedT, DepinhoRA, ChioccaEA (2008) Oncolytic herpes virus with defective ICP6 specifically replicates in quiescent cells with homozygous genetic mutations in p16. Oncogene 27: 4249–4254.

66. KimJH, OhJY, ParkBH, LeeDE, KimJS, et al. (2006) Systemic armed oncolytic and immunologic therapy for cancer with JX-594, a targeted poxvirus expressing GM-CSF. Mol Ther 14: 361–370.

67. LiuBL, RobinsonM, HanZQ, BranstonRH, EnglishC, et al. (2003) ICP34.5 deleted herpes simplex virus with enhanced oncolytic, immune stimulating, and anti-tumour properties. Gene Ther 10: 292–303.

68. PasserBJ, CheemaT, ZhouB, WakimotoH, ZaupaC, et al. (2010) Identification of the ENT1 antagonists dipyridamole and dilazep as amplifiers of oncolytic herpes simplex virus-1 replication. Cancer Res 70: 3890–3895.

69. RuggeroD (2013) Translational control in cancer etiology. Cold Spring Harb Perspect Biol 5: a012336.

70. BjornstiMA, HoughtonPJ (2004) Lost in translation: dysregulation of cap-dependent translation and cancer. Cancer Cell 5: 519–523.

71. HollandEC, SonenbergN, PandolfiPP, ThomasG (2004) Signaling control of mRNA translation in cancer pathogenesis. Oncogene 23: 3138–3144.

72. BarberGN (2005) VSV-tumor selective replication and protein translation. Oncogene 24: 7710–7719.

73. StanfordMM, McFaddenG (2007) Myxoma virus and oncolytic virotherapy: a new biologic weapon in the war against cancer. Expert Opin Biol Ther 7: 1415–1425.

74. GarciaMA, MeursEF, EstebanM (2007) The dsRNA protein kinase PKR: virus and cell control. Biochimie 89: 799–811.

75. Kawagishi-KobayashiM, SilvermanJB, UngTL, DeverTE (1997) Regulation of the protein kinase PKR by the vaccinia virus pseudosubstrate inhibitor K3L is dependent on residues conserved between the K3L protein and the PKR substrate eIF2alpha. Mol Cell Biol 17: 4146–4158.

76. BalachandranS, BarberGN (2004) Defective translational control facilitates vesicular stomatitis virus oncolysis. Cancer Cell 5: 51–65.

77. RichterJD, SonenbergN (2005) Regulation of cap-dependent translation by eIF4E inhibitory proteins. Nature 433: 477–480.

78. ChuluunbaatarU, RollerR, FeldmanME, BrownS, ShokatKM, et al. (2010) Constitutive mTORC1 activation by a herpesvirus Akt surrogate stimulates mRNA translation and viral replication. Genes Dev 24: 2627–2639.

79. SabatiniDM (2006) mTOR and cancer: insights into a complex relationship. Nat Rev Cancer 6: 729–734.

80. LiuTC, WakimotoH, MartuzaRL, RabkinSD (2007) Herpes simplex virus Us3(-) mutant as oncolytic strategy and synergizes with phosphatidylinositol 3-kinase-Akt targeting molecular therapeutics. Clin Cancer Res 13: 5897–5902.

81. AlainT, LunX, MartineauY, SeanP, PulendranB, et al. (2010) Vesicular stomatitis virus oncolysis is potentiated by impairing mTORC1-dependent type I IFN production. Proc Natl Acad Sci U S A 107: 1576–1581.

82. ColinaR, Costa-MattioliM, DowlingRJ, JaramilloM, TaiLH, et al. (2008) Translational control of the innate immune response through IRF-7. Nature 452: 323–328.

83. Costa-MattioliM, SonenbergN (2008) RAPping production of type I interferon in pDCs through mTOR. Nat Immunol 9: 1097–1099.

84. OliereS, ArguelloM, MespledeT, TumilasciV, NakhaeiP, et al. (2008) Vesicular stomatitis virus oncolysis of T lymphocytes requires cell cycle entry and translation initiation. J Virol 82: 5735–5749.

85. CorreaRJ, KomarM, TongJG, SivapragasamM, RahmanMM, et al. (2012) Myxoma virus-mediated oncolysis of ascites-derived human ovarian cancer cells and spheroids is impacted by differential AKT activity. Gynecol Oncol 125: 441–450.

86. LunX, AlainT, ZempFJ, ZhouH, RahmanMM, et al. (2010) Myxoma virus virotherapy for glioma in immunocompetent animal models: optimizing administration routes and synergy with rapamycin. Cancer Res 70: 598–608.

87. StanfordMM, ShabanM, BarrettJW, WerdenSJ, GilbertPA, et al. (2008) Myxoma virus oncolysis of primary and metastatic B16F10 mouse tumors in vivo. Mol Ther 16: 52–59.

88. FuX, TaoL, RiveraA, ZhangX (2011) Rapamycin enhances the activity of oncolytic herpes simplex virus against tumor cells that are resistant to virus replication. Int J Cancer 129: 1503–1510.

89. HomicskoK, LukashevA, IggoRD (2005) RAD001 (everolimus) improves the efficacy of replicating adenoviruses that target colon cancer. Cancer Res 65: 6882–6890.

90. LunXQ, JangJH, TangN, DengH, HeadR, et al. (2009) Efficacy of systemically administered oncolytic vaccinia virotherapy for malignant gliomas is enhanced by combination therapy with rapamycin or cyclophosphamide. Clin Cancer Res 15: 2777–2788.

91. BrassN, HeckelD, SahinU, PfreundschuhM, SybrechtGW, et al. (1997) Translation initiation factor eIF-4gamma is encoded by an amplified gene and induces an immune response in squamous cell lung carcinoma. Hum Mol Genet 6: 33–39.

92. BraunsteinS, KarpishevaK, PolaC, GoldbergJ, HochmanT, et al. (2007) A hypoxia-controlled cap-dependent to cap-independent translation switch in breast cancer. Mol Cell 28: 501–512.

93. GoetzC, GromeierM (2010) Preparing an oncolytic poliovirus recombinant for clinical application against glioblastoma multiforme. Cytokine Growth Factor Rev 21: 197–203.

94. FitzgeraldKD, SemlerBL (2011) Re-localization of cellular protein SRp20 during poliovirus infection: bridging a viral IRES to the host cell translation apparatus. PLoS Pathog 7: e1002127 doi:10.1371/journal.ppat.1002127

95. KastanMB, BartekJ (2004) Cell-cycle checkpoints and cancer. Nature 432: 316–323.

96. Ben-IsraelH, KleinbergerT (2002) Adenovirus and cell cycle control. Front Biosci 7: d1369–1395.

97. CrescenziM, SodduS, TatoF (1995) Mitotic cycle reactivation in terminally differentiated cells by adenovirus infection. J Cell Physiol 162: 26–35.

98. PoggioliGJ, DeBiasiRL, BickelR, JotteR, SpaldingA, et al. (2002) Reovirus-induced alterations in gene expression related to cell cycle regulation. J Virol 76: 2585–2594.

99. YooNK, PyoCW, KimY, AhnBY, ChoiSY (2008) Vaccinia virus-mediated cell cycle alteration involves inactivation of tumour suppressors associated with Brf1 and TBP. Cell Microbiol 10: 583–592.

100. BandaraLR, La ThangueNB (1991) Adenovirus E1a prevents the retinoblastoma gene product from complexing with a cellular transcription factor. Nature 351: 494–497.

101. ZamanianM, La ThangueNB (1992) Adenovirus E1a prevents the retinoblastoma gene product from repressing the activity of a cellular transcription factor. EMBO J 11: 2603–2610.

102. FueyoJ, Gomez-ManzanoC, AlemanyR, LeePS, McDonnellTJ, et al. (2000) A mutant oncolytic adenovirus targeting the Rb pathway produces anti-glioma effect in vivo. Oncogene 19: 2–12.

103. KimM, WilliamsonCT, PrudhommeJ, BebbDG, RiabowolK, et al. (2010) The viral tropism of two distinct oncolytic viruses, reovirus and myxoma virus, is modulated by cellular tumor suppressor gene status. Oncogene 29: 3990–3996.

104. MartinS, HarrisDT, ShislerJ (2012) The C11R gene, which encodes the vaccinia virus growth factor, is partially responsible for MVA-induced NF-kappaB and ERK2 activation. J Virol 86: 9629–9639.

105. TaoY, SongX, DengX, XieD, LeeLM, et al. (2005) Nuclear accumulation of epidermal growth factor receptor and acceleration of G1/S stage by Epstein-Barr-encoded oncoprotein latent membrane protein 1. Exp Cell Res 303: 240–251.

106. TwardzikDR, BrownJP, RanchalisJE, TodaroGJ, MossB (1985) Vaccinia virus-infected cells release a novel polypeptide functionally related to transforming and epidermal growth factors. Proc Natl Acad Sci U S A 82: 5300–5304.

107. HolbroT, CivenniG, HynesNE (2003) The ErbB receptors and their role in cancer progression. Exp Cell Res 284: 99–110.

108. EsfandiareiM, LuoH, YanagawaB, SuarezA, DabiriD, et al. (2004) Protein kinase B/Akt regulates coxsackievirus B3 replication through a mechanism which is not caspase dependent. J Virol 78: 4289–4298.

109. HsuMJ, WuCY, ChiangHH, LaiYL, HungSL (2010) PI3K/Akt signaling mediated apoptosis blockage and viral gene expression in oral epithelial cells during herpes simplex virus infection. Virus Res 153: 36–43.

110. SoaresJA, LeiteFG, AndradeLG, TorresAA, De SousaLP, et al. (2009) Activation of the PI3K/Akt pathway early during vaccinia and cowpox virus infections is required for both host survival and viral replication. J Virol 83: 6883–6899.

111. WerdenSJ, LanchburyJ, ShattuckD, NeffC, DuffordM, et al. (2009) The myxoma virus m-t5 ankyrin repeat host range protein is a novel adaptor that coordinately links the cellular signaling pathways mediated by Akt and Skp1 in virus-infected cells. J Virol 83: 12068–12083.

112. SypulaJ, WangF, MaY, BellJ, McFaddenG (2004) Myxoma virus tropism in human tumor cells. Gene Ther Mol Biol 8: 103–114.

113. FuX, TaoL, CaiR, PriggeJ, ZhangX (2006) A mutant type 2 herpes simplex virus deleted for the protein kinase domain of the ICP10 gene is a potent oncolytic virus. Mol Ther 13: 882–890.

114. MarcatoP, ShmulevitzM, PanD, StoltzD, LeePW (2007) Ras transformation mediates reovirus oncolysis by enhancing virus uncoating, particle infectivity, and apoptosis-dependent release. Mol Ther 15: 1522–1530.

115. StrongJE, CoffeyMC, TangD, SabininP, LeePW (1998) The molecular basis of viral oncolysis: usurpation of the Ras signaling pathway by reovirus. EMBO J 17: 3351–3362.

116. EndoT, TodaM, WatanabeM, IizukaY, KubotaT, et al. (2002) In situ cancer vaccination with a replication-conditional HSV for the treatment of liver metastasis of colon cancer. Cancer Gene Ther 9: 142–148.

117. MelcherA, ParatoK, RooneyCM, BellJC (2009) Thunder and lightning: immunotherapy and oncolytic viruses collide. Mol Ther 19: 1008–1016.

118. PrestwichRJ, IlettEJ, ErringtonF, DiazRM, SteeleLP, et al. (2009) Immune-mediated antitumor activity of reovirus is required for therapy and is independent of direct viral oncolysis and replication. Clin Cancer Res 15: 4374–4381.

119. VigilA, MartinezO, ChuaMA, Garcia-SastreA (2008) Recombinant Newcastle disease virus as a vaccine vector for cancer therapy. Mol Ther 16: 1883–1890.

120. GuoZS, NaikA, O'MalleyME, PopovicP, DemarcoR, et al. (2005) The enhanced tumor selectivity of an oncolytic vaccinia lacking the host range and antiapoptosis genes SPI-1 and SPI-2. Cancer Res 65: 9991–9998.

121. VijayalingamS, SubramanianT, RyerseJ, VarvaresM, ChinnaduraiG (2009) Down-regulation of multiple cell survival proteins in head and neck cancer cells by an apoptogenic mutant of adenovirus type 5. Virology 392: 62–72.

122. JinJ, LiuH, YangC, LiG, LiuX, et al. (2009) Effective gene-viral therapy of leukemia by a new fiber chimeric oncolytic adenovirus expressing TRAIL: in vitro and in vivo evaluation. Mol Cancer Ther 8: 1387–1397.

123. SathaiahM, ThirunavukkarasuP, O'MalleyME, KavanaghMA, RavindranathanR, et al. (2012) Oncolytic poxvirus armed with Fas ligand leads to induction of cellular Fas receptor and selective viral replication in FasR-negative cancer. Cancer Gene Ther 19: 192–201.

124. ChalikondaS, KivlenMH, O'MalleyME, Eric DongXD, McCartJA, et al. (2008) Oncolytic virotherapy for ovarian carcinomatosis using a replication-selective vaccinia virus armed with a yeast cytosine deaminase gene. Cancer Gene Ther 15: 115–125.

125. ZhengFQ, XuY, YangRJ, WuB, TanXH, et al. (2009) Combination effect of oncolytic adenovirus therapy and herpes simplex virus thymidine kinase/ganciclovir in hepatic carcinoma animal models. Acta Pharmacol Sin 30: 617–627.

126. MaS, QuW, MaoL, ZhuZ, JiaL, et al. (2012) Antitumor effects of oncolytic adenovirus armed with Drosophila melanogaster deoxyribonucleoside kinase in colorectal cancer. Oncol Rep 27: 1443–1450.

127. SunamuraM, OonumaM, MotoiF, AbeH, SaitohY, et al. (2002) Gene therapy for pancreatic cancer targeting the genomic alterations of tumor suppressor genes using replication-selective oncolytic adenovirus. Hum Cell 15: 138–150.

128. SchepelmannS, OgilvieLM, HedleyD, FriedlosF, MartinJ, et al. (2007) Suicide gene therapy of human colon carcinoma xenografts using an armed oncolytic adenovirus expressing carboxypeptidase G2. Cancer Res 67: 4949–4955.

129. StubdalH, PerinN, LemmonM, HolmanP, BauzonM, et al. (2003) A prodrug strategy using ONYX-015-based replicating adenoviruses to deliver rabbit carboxylesterase to tumor cells for conversion of CPT-11 to SN-38. Cancer Res 63: 6900–6908.

130. MahoneyDJ, LefebvreC, AllanK, BrunJ, SanaeiCA, et al. (2011) Virus-tumor interactome screen reveals ER stress response can reprogram resistant cancers for oncolytic virus-triggered caspase-2 cell death. Cancer Cell 20: 443–456.

131. HanahanD, WeinbergRA (2011) Hallmarks of cancer: the next generation. Cell 144: 646–674.

Štítky
Hygiena a epidemiológia Infekčné lekárstvo Laboratórium

Článok vyšiel v časopise

PLOS Pathogens


2014 Číslo 1
Najčítanejšie tento týždeň
Najčítanejšie v tomto čísle
Kurzy

Zvýšte si kvalifikáciu online z pohodlia domova

Získaná hemofilie - Povědomí o nemoci a její diagnostika
nový kurz

Eozinofilní granulomatóza s polyangiitidou
Autori: doc. MUDr. Martina Doubková, Ph.D.

Všetky kurzy
Prihlásenie
Zabudnuté heslo

Zadajte e-mailovú adresu, s ktorou ste vytvárali účet. Budú Vám na ňu zasielané informácie k nastaveniu nového hesla.

Prihlásenie

Nemáte účet?  Registrujte sa

#ADS_BOTTOM_SCRIPTS#