Skip to main content
Oncoimmunology logoLink to Oncoimmunology
. 2014 Jun 1;3:e28694. doi: 10.4161/onci.28694

Trial Watch:

Oncolytic viruses for cancer therapy

Jonathan Pol 1,2,3,4,, Norma Bloy 1,2,3,4,, Florine Obrist 1,2,3,4, Alexander Eggermont 1, Jérôme Galon 5,6,7,8, Isabelle Cremer 5,6,7,9, Philippe Erbs 10, Jean-Marc Limacher 10, Xavier Preville 10, Laurence Zitvogel 1,11, Guido Kroemer 2,3,5,12,13,‡,*, Lorenzo Galluzzi 1,3,5,‡,*
PMCID: PMC4091053  PMID: 25097804

Abstract

Oncolytic viruses are natural or genetically modified viral species that selectively infect and kill neoplastic cells. Such an innate or exogenously conferred specificity has generated considerable interest around the possibility to employ oncolytic viruses as highly targeted agents that would mediate cancer cell-autonomous anticancer effects. Accumulating evidence, however, suggests that the therapeutic potential of oncolytic virotherapy is not a simple consequence of the cytopathic effect, but strongly relies on the induction of an endogenous immune response against transformed cells. In line with this notion, superior anticancer effects are being observed when oncolytic viruses are engineered to express (or co-administered with) immunostimulatory molecules. Although multiple studies have shown that oncolytic viruses are well tolerated by cancer patients, the full-blown therapeutic potential of oncolytic virotherapy, especially when implemented in the absence of immunostimulatory interventions, remains unclear. Here, we cover the latest advances in this active area of translational investigation, summarizing high-impact studies that have been published during the last 12 months and discussing clinical trials that have been initiated in the same period to assess the therapeutic potential of oncolytic virotherapy in oncological indications.

Keywords: adenovirus, ColoAd1, mesenchymal stem cells, MV-NIS, reolysin, talimogene laherparepvec

Introduction

The term “oncolytic virus” is commonly employed to identify a non-pathogenic viral strain that selectively infects and kills neoplastic cells while leaving their normal counterparts virtually unaffected.1 Thus, oncolytic viruses not only display a preferential tropism for transformed over non-transformed tissues, in thus far being oncotropic,2-4 but also trigger the massive demise of infected cells.5 Such a cytotoxic activity can be natural and simply reflect the so-called cytopathic effect, i.e., the lethal outcome of a replicative viral infection.6,7 Alternatively, oncolytic viruses can mediate cytotoxic effects upon the expression of (endogenous or exogenous) gene products, irrespective of their ability to drive a productive infection.1,5 Although the possibility to harness the lytic potential of viruses against cancer has been theorized as early as at the beginning of the 20th century, it is only with the advent of modern genetic engineering technologies in the late 1990s that the interest in oncolytic virotherapy has crystallized.5,8-10 Since then, dozens of viruses have been tested for their natural oncolytic activity or genetically endowed with cancer-specific cytotoxic functions or immunostimulatory properties. A precise description of these viruses goes beyond the scope of this Trial Watch. For additional details on the viral species that have been harnessed so far for oncolytic virotherapy as well as on the advantages and pitfalls associated with their use please see Refs. 1,5 and 11.

Specificity is a critical requirement for the safety of oncolytic virotherapy and multiple strategies have been developed throughout the past 2 decades to improve the oncotropism of naturally occurring viruses or endow otherwise unspecific viral strains with a highly-targeted lytic potential.1,5,12 Such a specificity can be obtained by conventional genetic engineering (1) at the transductional level, implying modifications of surface proteins that allow for the infection of cells bearing one (or a few) tumor-specific markers on their surface;13-17 (2) at the transcriptional level, based on the use of promoters that are active in neoplastic cells only to control the expression of essential viral genes;18-25 (3) at the post-transcriptional/translational level, either based on the insertion of microRNA-binding elements in non-coding regions of essential viral genes, allowing productive infections to develop only in tissues that do not express such microRNAs;26-32 or involving the cloning of genes that are absolutely required for the viral cycle downstream of internal ribosome entry sites that are inactive in selected tissues;33-35 (4) at the post-translational level, either based on “destabilization domains” that render essential viral proteins unstable in tissues that are not artificially or naturally exposed to a specific stabilizing agent,36-38 or based on viral protein precursors that can be processed only in cells expressing specific (cancer-associated) proteases;39 and (5) at a cell-wide level, harnessing the ability of attenuated viral strains to productively replicate in neoplastic cells bearing peculiar genetic or epigenetic defects, such as the hyperactivation of Harvey rat sarcoma viral oncogene homolog (HRAS) or signal transducer and activator of transcription 3 (STAT3) as well as the inactivation of p53.40-46 Among these strategies, the use of oxygen-dependent degradation domains (ODDs) stands out as a convenient approach to specifically direct the cytotoxicity of oncolytic viruses to solid tumors based on their limited oxygen supply.38

In addition, genetic engineering has been largely employed to endow oncolytic viruses with (at least hypothetically) desirable features, including (but not limited to) an increased cytotoxic potential and a superior ability to drive cell-mediated immune responses.1,5,47 Thus, the viral genome has been integrated with sequences coding for (1) enzymes that convert non-toxic pro-drugs into a lethal cytotoxic agent;48-55 (2) proteins that (at least on theoretical grounds) mediate tumor-specific lethal effects;56-58 or (3) short-hairpin RNAs targeting proteins that are necessary for the survival of neoplastic cells, such as survivin.59-61 All these approaches have been shown to improve the cytotoxic potential of oncolytic virotherapy, hence ameliorating its therapeutic profile (at least to some extent) in experimental settings.

Accumulating evidence indicates indeed that the antineoplastic effects of oncolytic virotherapy do not simply originate from cancer cell-autonomous mechanisms but involve the (re)activation of tumor-specific immune responses.62-70 Thus, the administration of oncolytic viruses to cancer patients has been associated with the insurgence of cellular as well as humoral antitumor immune responses of potential therapeutic value.71-74 Moreover, the clinical activity of oncolytic viruses seems to benefit from some extent of initial immunosuppression (which facilitates viral spread, see below) followed by the administration of immunostimulatory molecules (which exacerbate antitumor immunity).1,5 In line with this notion, oncolytic viruses have also been engineered to express (1) tumor-associated antigens (generating so-called oncolytic vaccines);75-79 (2) co-stimulatory molecules, such as CD40 ligand (CD40L) and CD80;80-84 (3) immunostimulatory cytokines, including interleukin (IL)-2,85-87 IL-12,88-95 IL-15,96-101 IL-23,102 IL-24,103-106 and granulocyte macrophage colony-stimulating factor (GM-CSF);73,89,107-113 or (4) chemokines, such as chemokine (C-C motif) ligand 7 (CCL7)114 and CCL19.115

The clinical profile of oncolytic virotherapy employed as a standalone therapeutic intervention is generally limited, for several reasons.1,5,116 These include (but may be not limited to): (1) the heterogeneous and relatively incomplete diffusion of oncolytic viruses within neoplastic lesions;117-127 (2) the equilibrium that is generally established between oncolytic viruses and continuously proliferating, non-infected cancer cells, which eventually shifts in favors of the latter owing to the insurgence of an antiviral immune response (at least in immunocompetent individuals);128-130 (3) the propensity of malignant cells to become resistant to oncolytic virotherapy,121,126,130-132 presumably reflecting their genomic instability;133,134 (4) the elevated diffusion among the population of viral species that are employed to create therapeutic strains, resulting in a significant fraction of individuals who are insensitive to some oncolytic viruses owing to neutralizing humoral immunity;135,136 (5) the elevated sensitivity of some oncolytic viruses to the complement system;137,138 (6) the sequestration of intravenously administered oncolytic viruses by the mononuclear phagocytic system (MPS) of the liver and spleen,139,140 limiting the availability of viral particles at the tumor site and (at least in some cases) causing driving serious, dose-limiting toxicities;141-143 and (7) the threats that are intrinsically associated with the use of replicating viral particles in cancer patients, who are particularly weak and often immunodepressed.144-150

Consistent efforts have been dedicated at the development of strategies that would circumvent (at least in part) these issues. For instance, several oncolytic viruses have been genetically manipulated to express endogenous (or co-administered with exogenous) inhibitors of angiogenesis.91,92,151-154 This approach may exert a dual benefit in that the inhibition of cancer-associated angiogenesis not only mediates direct antineoplastic effects,155,156 but also causes a normalization of the tumor vasculature that improves the delivery/penetration of therapeutic agents, including oncolytic viruses themselves.157-159 Along similar lines, tumor-infiltrating cells including macrophages,160,161 myeloid-derived suppressor cells,162-164 and mesenchymal stem cells (MSCs) have been harnessed as vehicles to selectively deliver oncolytic viruses to neoplastic lesions while shielding them from neutralizing antibodies and protecting them from sequestration by the MPS.163,165-170 Finally, several laboratories worldwide have demonstrated that the therapeutic profile of oncolytic virotherapy can be remarkably boosted by the co-administration of several chemotherapeutics, including (but not limited to) gemcitabine (an immunostimulatory nucleoside analog),171-173 paclitaxel (a microtubular inhibitor),174-176 temozolomide and cyclophosphamide (two alkylating agents),72,177-179 sunitinib (a relatively unspecific tyrosine kinase inhibitor),180-182 cisplatin (a non-immunogenic DNA-damaging agent),183-186 various histone deacetylase inhibitors,187 and 13-cis retinoic acid (a retinoid employed for the treatment of high risk neuroblastoma).188,189 Taken together, these findings indicate that oncolytic viruses can mediate therapeutically relevant anticancer effects in vivo. In line with this notion, clinical trials performed throughout the 2 past decades demonstrated that oncolytic virotherapy can be safely implemented in cancer patients and can exert substantial antineoplastic activity, at least in a fraction of individuals.

As it stands, no oncolytic virus is currently licensed by the US Food and Drug Administration and the European Medicines Agency for use in cancer patients (sources http://www.fda.gov/Drugs/default.htm and http://www.ema.europa.eu). Along similar lines, in spite of promising preclinical results,190-194 oncolytic virotherapy has not yet been approved as part of veterinary protocols in the US and Europe. Conversely, the Chinese State Food and Drug Administration approved a recombinant adenovirus (H101, commercialized under the name of Oncorine®) for use together with chemotherapy in refractory head and neck cancer patients as early as in November 2005.195-197

One year ago, in the May issue of OncoImmunology, we presented the scientific background to the use of oncolytic viruses in oncological indications and discussed recent clinical trials evaluating the safety and efficacy of this immunotherapeutic regimen.198 Along the lines of our monthly Trial Watch series,199-201 here we summarize the latest developments in oncolytic virotherapy.

Literature Update

Since the submission of our previous Trial Watch on this topic (March 2013),198 no less than 11 studies dealing with clinical aspects of oncolytic virotherapy have been published in peer-reviewed scientific journals (source http://www.ncbi.nlm.nih.gov/pubmed). Four of these studies tested a serotype-chimeric, infectivity-enhanced, conditionally-replicative adenovirus initially developed to specifically infect and kill ovarian cancer cells,202,203 either in its pristine configuration (Ad5/3-Delta24)204 or as a 2nd generation variant further modified to express GM-CSF (CGTG-102).72,73,205 Kim and colleagues investigated the feasibility and efficacy of intraperitoneally administered Ad5/3-Delta24 in 10 recurrent ovarian cancer patients. Nine patients completed the therapeutic protocol and only manageable Grade I/II side effects were recorded. In spite of the development of neutralizing immunity in all individuals, 6 out of 8 patients that could be evaluated for response experienced disease stabilization.204 Kanerva and coworkers tested multiple immunological and clinicopathological parameters in 115 cancer patients treated with CGTG-102, either as a single injection, either in a serial manner (3 injections over 10 wks), or in the context of a switch protocol involving the administration of viruses with modified capsid proteins to avoid neutralizing immunity. A good safety profile was recorded. Moreover, the serial regimen was associated with the induction of anticancer immune responses, with indirect indications of the mobilization of tumor-specific T cells to neoplastic lesions,164,206,207 as well as with an improved overall survival (as compared with the single injection-based protocol).73 Liikanen and collaborators assessed the clinical profile of CGTG-102 in combination with metronomic temozolomide and/or cyclophosphamide in 17 individuals affected by chemotherapy-refractory neoplasms. This chemoimmunotherapeutic regimen was well tolerated by all patients and elicited several signs of immunogenic cell death (ICD) in neoplastic cells, including the activation of autophagy and the release of high mobility group box 1 (HMGB1).208-214 This was paralleled not only by the release of pro-inflammatory cytokines but also by the activation of tumor-specific immune responses, which were observed in 10 out of 15 patients. Clinical effects were evident in 67% of patients treated with CGTG-102 plus temozolomide and/or cyclophosphamide and these individuals exhibited a trend for improved survival as compared with subjects receiving CGTG-102 only.72 Bramante et al. tested the clinical profile of CGTG-102 in 15 patients with treatment-refractory soft tissue (13/15) or primary bone (2/15) sarcoma. Of 12 patients who could be assessed for clinical outcome, 2 exhibited an objective (though minor) response, 6 stable disease, and 4 progressive disease. Median overall survival was 170 d, and 1 patient was still alive when the paper was published, approx. four years after oncolytic virotherapy.205

Different research groups investigated the clinical profile of Pexa-Vec (formerly known as JX-594), an oncolytic vaccinia virus engineered to selectively replicate in cells with alterations of the RAS pathway and to express GM-CSF.109,215-217 Kim and colleagues reported that the administration of Pexa-Vec to 3 patients with diverse tumors (in the context of a Phase I clinical trial) resulted in the insurgence of a cancer cell-specific, antibody-mediated, complement-dependent cytotoxic response, whose intensity positively correlated with overall survival.71 Breitbach and collaborators studied the effects of Pexa-Vec on the tumor vasculature in patients with treatment-refractory, histologically confirmed advanced/metastatic solid tumors (n = 18, from a Phase I clinical trial, NCT00625456) as well as in subjects with hepatic neoplasms of hepatocellular (n = 15) or colorectal (n = 1) origin (from Phase II clinical trials, NCT00554372 and NCT01171651). These authors demonstrated that Pexa-Vec infects not only neoplastic cells but also the tumor-associated (but not the normal) vasculature, resulting in its destruction and hence mediating antineoplastic effects.155

Markert and colleagues conducted a Phase I clinical trial to test the effects of the stereotactic intratumoral administration of G207, a conditionally replicating herpes simplex virus (HSV) Type 1 variant,218-220 in 9 glioblastoma patients allocated to receive a single 5 Gy dose of radiation 24 h later. Oncolytic virotherapy was well tolerated and 6 out of 9 patients exhibited stable disease or a partial response at least at one evaluation time point.221 Interestingly, interim overall survival data from a Phase III study comparing subcutaneous GM-CSF with intratumoral talimogene laherparepvec (T-vec), an oncolytic HSV variant manipulated to express GM-CSF,222,223 in over 400 patients with unresected Stage IIIB, IIIC or IV melanoma (NCT00769704) have been released by Amgen at the 2013 meeting of the American Society of Clinical Oncology (ASCO), which was held in Chicago last June. In this setting, 26% of patients receiving T-vec developed serious adverse effects, including cellulitis and pyrexia, as compared with 13% of patients receiving GM-CSF. At a predetermined interim analysis, median overall survival among T-vec- and GM-CSF-treated patients was 23.3 and 19.9 mo, respectively (HR = 0.79, 95% CI 0.61–1.02; P = 0.0746). According to Amgen representatives, such a difference, which was slightly below the threshold for statistical significance, was “pronounced in the subset patients with Stage IIIB, IIIC or IV disease (HR = 0.56, 95% CI, 0.38-0.81) or who received T-vec as first-line treatment (HR = 0.49, 95% CI, 0.33-0.74), each comprising approximately 50% of the study population” (source http://www.amgen.com/media/media_pr_detail.jsp?releaseID=1877950). Thus, although NCT00769704 met its primary endpoint of durable response rate, defined as the rate of complete or partial response lasting continuously for at least 6 mo (source http://www.amgen.com/media/media_pr_detail.jsp?releaseID=1798143),224 the actual clinical benefits of T-vec remains to be elucidated.

Morris and colleagues tested the clinical profile of percutaneously administered Reolysin®, a wild-type reovirus (serotype 3 Dearing),225 in 19 patients with accessible solid tumors who failed to improve on standard therapeutic approaches. Common toxicities included Grade 1/2 local erythema and transient flu-like symptoms. Moreover, objective responses were recorded in 7/19 (37%) patients, with 1 individual exhibiting a complete response, 2 a partial response, and 4 stable disease.226

During the last 13 mo a few immunological and clinicopathological parameters have been suggested to have a prognostic or predictive value in patients treated with oncolytic virotherapy. Such parameters include polymorphisms in the gene coding for Fc fragment of IgG, low affinity IIIa, receptor (FCGR3A), perhaps because of their influence on natural killer (NK) cell antibody-dependent cellular cytotoxicity,116,227 as well as a hypointense tumor core in T2-weighted magnetic resonance imaging, perhaps indicating ongoing coagulative necrosis,228,229 Moreover, Koski and colleagues demonstrated that 18F-fluorodeoxyglucose positron emission tomography (18F-FDG PET) and computed tomography (CT) are equally reliable means to predict the long-term survival of cancer patients on oncolytic virotherapy.230 These studies may have significant implications for oncolytic viruses to become a clinical reality, as they may allow for the identification of patients who are most likely to obtain actual benefits from therapy.

Among recent (i.e., published during the last 13 mo) preclinical studies investigating the safety and efficacy of oncolytic virotherapy in experimental settings we found of particular interest the work of (1) Beug and collaborators, from the Children's Hospital of Eastern Ontario Research Institute (Ottawa, Canada), who demonstrated that the therapeutic potential of SMAC mimetics is dramatically exacerbated by oncolytic viruses, as well as by Toll-like receptor (TLR) agonists, owing to their ability to stimulate the production of interferon β (IFNβ), tumor necrosis factor α (TNFα), and tumor necrosis factor (ligand) superfamily, member 10 (TNFSF10, best known as TRAIL);231-236 (2) Zamarin and colleagues, from the Sloan Kettering Institute for Cancer Research (New York, NY, US), who proved that the intratumoral administration of oncolytic viruses can elicit tumor-specific immune reactions in distant, non-injected lesions, and that such an effect synergize with the systemic delivery of cytotoxic T lymphocyte-associated protein 4 (CTLA4)-blocking antibodies to achieve a superior antineoplastic activity;237-240 and (3) Castleton et al., from the University College London (London, United Kingdom), who provided robust evidence in support of the notion that MSCs can be efficiently employed to deliver oncolytic viruses to neoplastic lesions even in the presence of high-titer neutralizing antibodies;170 (4) Carew and coworkers, from the University of Texas Health Science Center (San Antonio, TX, US), who demonstrated that (at least some) oncolytic viruses promote the demise of transformed cells upon the establishment of endoplasmic reticulum stress, which is currently viewed as an absolute requirement for cell death to be perceived as immunogenic.208,209,241 and (5) Donnelly and colleagues, from the Leeds Institute of Cancer and Pathology (Leeds, UK), who proved that the therapeutic activity of (at least some) oncolytic viruses is significantly enhanced in animals immunized against the same viruses and pre-administered with GM-CSF, but not IL-2 or granulocyte colony-stimulating factor (G-CSF).242 This latter study has profound repercussions for the implementation of oncolytic virotherapy in patients who may be endowed with neutralizing immunity as a result of previous exposure to naturally occurring viruses.

Update on Ongoing Clinical Trials

When this Trial Watch was being prepared (March 2014), official sources listed no less than 16 clinical trials launched after March 1st, 2013 that would evaluate the efficacy and safety of oncolytic virotherapy in oncological indications (source http://www.clinicaltrials.gov).

In particular, (1) ColoAd1, a chimeric oncolytic virus developed by directed evolution,243,244 is being tested as a standalone therapeutic intervention in patients with resectable colorectal carcinoma (NCT02053220), platinum-resistant ovarian carcinoma (NCT02028117) or various neoplasms of epithelial origin (NCT02028442); (2) the safety and efficacy of MV-NIS, a strain of measles virus genetically engineered to express human solute carrier family 5, member 5 (SCL5A5, best known as sodium/iodide symporter),245-248 are being evaluated in ovarian cancer patients, receiving MV-NIS-loaded MSCs i.p. (NCT02068794), as well as in subjects with head and neck squamous cell carcinoma, who are treated with MV-NIS i.t. (NCT01846091); (3) the intratumoral or intravenous administration of VCN-01, a replication-competent adenovirus expressing human sperm adhesion molecule 1 (SPAM1, best known as PH20 hyaluronidase),249 is being assessed in individuals affected by advanced pancreatic cancer (NCT02045589) or other solid neoplasms (NCT02045602), respectively; (4) the biodistribution and shedding of intratumorally administered T-vec are being evaluated in melanoma patients (NCT02014441); (5) the safety and efficacy of Toca 511, an amphotropic replication-competent retrovirus genetically modified to express cytosine deaminase,250-252 given as a standalone therapeutic intervention are being assessed in patients undergoing surgery for recurrent brain tumors (NCT01985256); (6) a naturally occurring variant of coxsackievirus, namely, coxsackievirus A21,253-256 is being evaluated as a single agent for the systemic treatment of residual metastatic disease in subjects with NSCLC, melanoma, bladder carcinoma, and castration-resistant prostate cancer257 (NCT02043665); (7) the intravesical instillation of CG0070, a conditionally replicating oncolytic adenovirus genetically modified to express GM-CSF,108,258 is being investigated as a standalone therapeutic intervention in bladder carcinoma patients who failed Bacillus Calmette-Guérin (BCG)-based immunotherapy259,260 (NCT00109655); (8) the clinical profile of ICOVIR-5 and DNX2401, 2 oncolytic adenoviruses engineered to replicate only in cells exhibiting alterations of the retinoblastoma 1 (RB1) signaling pathway,261-266 is being assessed in subjects with advanced melanoma (NCT01864759) or other solid tumors (NCT01844661), in both scenarios as a standalone therapeutic intervention, as well as in patients with recurrent glioblastoma (in the context of temozolomide-based chemotherapy) (NCT01956734); (9) HSV-1716, a γ34.5-deficient variant of HSV,267-270 is being tested in combination with dexamethasone (a glucocorticoid) in subjects with refractory or recurrent high-grade glioma that can be removed by surgery (NCT02031965); and (10) Pexa-Vec is being tested as a single therapeutic intervention in ovarian carcinoma patients (NCT02017678) (Table 1).

Table 1. Clinical trials recently launched to evaluate the safety and efficacy of oncolytic virotherapy in cancer patients*.

Virus Indication(s) Phase Status Route Notes Ref.
CG0070 Bladder carcinoma I n.a. Intravesical As a single agent NCT00109655
ColoAd1 Colorectal carcinoma I Recruiting i.t. or i.v. As a single agent NCT02053220
Ovarian carcinoma I/II Not yet recruiting i.p. As a single agent NCT02028117
Solid tumors I/II Recruiting i.v. As a single agent NCT02028442
CVA21 Solid tumors I Not yet recruiting i.v. As a single agent NCT02043665
DNX2401 Glioblastoma I Recruiting i.t. Combined with temozolomide and/or surgery NCT01956734
HSV-1716 Glioma I Recruiting Into the tumor resection cavity Combined with dexamethasone and surgery NCT02031965
ICOVIR-5 Melanoma I Recruiting i.v. As a single agent NCT01864759
Solid tumors I/II Recruiting i.p.
(via MSCs)
As a single agent NCT01844661
MV-NIS HNSCC I Recruiting i.t. As a single agent NCT01846091
Ovarian carcinoma I/II Not yet recruiting i.p.
(via MSCs)
As a single agent NCT02068794
Pexa-Vec Ovarian carcinoma II Not recruiting i.v. As a single agent NCT02017678
T-vec Melanoma II Not yet recruiting i.t. As a single agent NCT02014441
Toca 511 Brain tumors I Recruiting i.v. Combined with 5-FC NCT01985256
VCN-01 Pancreatic cancer I Recruiting i.t. Combined with gemcitabine NCT02045589
Solid tumors I Recruiting i.v. Combined with gemcitabine NCT02045602

Abbreviations: 5-FC, 5-fluorocytosine; CVA21, coxsackievirus A21; i.a., intra arteriam; i.p., intra peritoneum; i.t., intra tumorem; i.v., intra venam; HNSCC, head and neck squamous cell carcinoma; MSC, mesenchymal stem cell; n.a., not available; T-vec, talimogene laherparepvec. *Between 2013, March 1st and the date of submission.

The following clinical studies discussed in our previous Trial Watch dealing with oncolytic virotherapy198 have changed status during the last 12 mo. NCT00602277, NCT00805376, NCT00861627, NCT00984464, NCT00998192, NCT01240538, NCT01387555, NCT01394939, NCT01469611, NCT01533194, NCT01598129, and NCT0163628 are now listed as “Active, not recruiting”; NCT01017601, NCT01274624, and NCT01438112 now appear as “Recruiting”; NCT00651157, NCT01227551, and NCT01048892 are now indicated as “Completed”; NCT01437280 has been “Terminated”; and the status of NCT00753038 and NCT01443260 is now “Unknown” (source http://www.clinicaltrials.gov). In the context of NCT00651157, a Phase II study testing Reolysin® as a standalone therapeutic intervention in Stage IV melanoma patients,225 serious adverse effects developed in a significant proportion of patients (50%) and no clinical activity was recorded (source http://clinicaltrials.gov/ct2/show/results/NCT00651157?term=NCT00651157&rank=1). Although both these trials have been completed, results are available neither for NCT01048892, a Phase I trial testing the Seneca Valley virus in combination with metronomic cyclophosphamide in patients with neuroendocrine tumors, nor for NCT01227551,271,272 a Phase II study testing coxsackievirus A21 as a standalone therapeutic intervention in patients with advanced melanoma.253-256 NCT01437280, a Phase I trial testing the safety and efficacy of a GM-CSF-encoding oncolytic adenovirus (CGTG-102) in patients with advanced tumors has been terminated prior to enrollment for undisclosed reasons.73,205

Concluding Remarks

As discussed in this Trial Watch, oncolytic virotherapy has been shown to mediate robust, therapeutically relevant antineoplastic effects in both preclinical and clinical scenarios. It is now evident that such a therapeutic activity is not a mere consequence of the cytopathic effect, but rather involves the induction of a tumor-specific immune response. Oncolytic viruses appear indeed to specifically kill transformed cells, hence releasing elevated amounts of tumor-associated antigens, and deliver to the immune system robust stimulatory signals, de facto acting as therapeutic anticancer vaccines.64 The elevated immunogenic potential of oncolytic virotherapy presumably reflects the ability of viral components to act as microbe-associated molecular patterns, hence activating multiple pattern recognition receptors,273-279 as well as to promote the emission of endogenous danger-associated molecular patterns.69,208,241,280 In line with this notion, oncolytic viruses have already been shown to improve the efficacy of multiple immunotherapeutic interventions against cancer, including peptide- as well as DNA-based vaccines.75,77,281 Conversely, a large panel of immunostimulatory agents including multiple TLR agonists259,282 and ICD inducers174,283-286 appears to boost the antineoplastic activity of oncolytic virotherapy. We believe that precisely scheduled combinatorial regimens that initially allow for the replication and dissemination of viral particles thought neoplastic lesions, and then boost the ability of oncolytic viruses to induce tumor-specific immune responses may mediate optimal antineoplastic effects. Future will tell which, if any, of the immunochemotherapeutic regimens that may be devised287 is best suited for this purpose.

Disclosure of Potential Conflicts of Interest

P.E., J.M.L., and X.P. are full-time employees of Transgene; L.Z. is part of the Board of Directors of Transgene.

Acknowledgments

Authors are supported by the Ligue contre le Cancer (équipe labelisée); Agence National de la Recherche (ANR); Association pour la recherche sur le cancer (ARC); Cancéropôle Ile-de-France; AXA Chair for Longevity Research; Institut National du Cancer (INCa); Fondation Bettencourt-Schueller; Fondation de France; Fondation pour la Recherche Médicale (FRM); the European Commission (ArtForce); the European Research Council (ERC); the LabEx Immuno-Oncology; the SIRIC Stratified Oncology Cell DNA Repair and Tumor Immune Elimination (SOCRATE); the SIRIC Cancer Research and Personalized Medicine (CARPEM); and the Paris Alliance of Cancer Research Institutes (PACRI).

Glossary

Abbreviations:

CI

confidence interval

GM-CSF

granulocyte macrophage colony-stimulating factor

HR

hazard ratio

HSV

herpes simplex virus

ICD

immunogenic cell death

IL

interleukin

MPS

mononuclear phagocytic system

MSC

mesenchymal stem cell

T-vec

talimogene laherparepvec

TLR

Toll-like receptor

Citation: Pol J, Bloy N, Obrist F, Eggermont A, Galon J, Cremer I, Erbs P, Limacher JM, Preville X, Zitvogel L, et al. Trial Watch: Oncolytic viruses for cancer therapy. OncoImmunology 2014; 3:e28694; 10.4161/onci.28694

References

  • 1.Russell SJ, Peng KW, Bell JC. Oncolytic virotherapy. Nat Biotechnol. 2012;30:658–70. doi: 10.1038/nbt.2287. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Griffith TS, Kawakita M, Tian J, Ritchey J, Tartaglia J, Sehgal I, Thompson TC, Zhao W, Ratliff TL. Inhibition of murine prostate tumor growth and activation of immunoregulatory cells with recombinant canarypox viruses. J Natl Cancer Inst. 2001;93:998–1007. doi: 10.1093/jnci/93.13.998. [DOI] [PubMed] [Google Scholar]
  • 3.Mäkelä AR, Matilainen H, White DJ, Ruoslahti E, Oker-Blom C. Enhanced baculovirus-mediated transduction of human cancer cells by tumor-homing peptides. J Virol. 2006;80:6603–11. doi: 10.1128/JVI.00528-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Singh P, Destito G, Schneemann A, Manchester M. Canine parvovirus-like particles, a novel nanomaterial for tumor targeting. J Nanobiotechnology. 2006;4:2. doi: 10.1186/1477-3155-4-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Vähä-Koskela MJ, Heikkilä JE, Hinkkanen AE. Oncolytic viruses in cancer therapy. Cancer Lett. 2007;254:178–216. doi: 10.1016/j.canlet.2007.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Galluzzi L, Brenner C, Morselli E, Touat Z, Kroemer G. Viral control of mitochondrial apoptosis. PLoS Pathog. 2008;4:e1000018. doi: 10.1371/journal.ppat.1000018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Boisgerault N, Guillerme JB, Pouliquen D, Mesel-Lemoine M, Achard C, Combredet C, Fonteneau JF, Tangy F, Grégoire M. Natural oncolytic activity of live-attenuated measles virus against human lung and colorectal adenocarcinomas. Biomed Res Int. 2013;2013:387362. doi: 10.1155/2013/387362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Stanford MM, Bell JC, Vähä-Koskela MJ. Novel oncolytic viruses: riding high on the next wave? Cytokine Growth Factor Rev. 2010;21:177–83. doi: 10.1016/j.cytogfr.2010.02.012. [DOI] [PubMed] [Google Scholar]
  • 9.Quetglas JI, John LB, Kershaw MH, Alvarez-Vallina L, Melero I, Darcy PK, Smerdou C. Virotherapy, gene transfer and immunostimulatory monoclonal antibodies. Oncoimmunology. 2012;1:1344–54. doi: 10.4161/onci.21679. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Hemminki A, Oksanen M, Merisalo-Soikkeli M. Oncolytic virotherapy trials--letter. Clin Cancer Res. 2013;19:4541–2. doi: 10.1158/1078-0432.CCR-13-1471. [DOI] [PubMed] [Google Scholar]
  • 11.Walther W, Schlag PM. Current status of gene therapy for cancer. Curr Opin Oncol. 2013;25:659–64. doi: 10.1097/CCO.0000000000000004. [DOI] [PubMed] [Google Scholar]
  • 12.Donnelly O, Harrington K, Melcher A, Pandha H. Live viruses to treat cancer. J R Soc Med. 2013;106:310–4. doi: 10.1177/0141076813494196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Ayala-Breton C, Barber GN, Russell SJ, Peng KW. Retargeting vesicular stomatitis virus using measles virus envelope glycoproteins. Hum Gene Ther. 2012;23:484–91. doi: 10.1089/hum.2011.146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Muik A, Kneiske I, Werbizki M, Wilflingseder D, Giroglou T, Ebert O, Kraft A, Dietrich U, Zimmer G, Momma S, et al. Pseudotyping vesicular stomatitis virus with lymphocytic choriomeningitis virus glycoproteins enhances infectivity for glioma cells and minimizes neurotropism. J Virol. 2011;85:5679–84. doi: 10.1128/JVI.02511-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Uchida H, Marzulli M, Nakano K, Goins WF, Chan J, Hong CS, Mazzacurati L, Yoo JY, Haseley A, Nakashima H, et al. Effective treatment of an orthotopic xenograft model of human glioblastoma using an EGFR-retargeted oncolytic herpes simplex virus. Mol Ther. 2013;21:561–9. doi: 10.1038/mt.2012.211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Bach P, Abel T, Hoffmann C, Gal Z, Braun G, Voelker I, Ball CR, Johnston IC, Lauer UM, Herold-Mende C, et al. Specific elimination of CD133+ tumor cells with targeted oncolytic measles virus. Cancer Res. 2013;73:865–74. doi: 10.1158/0008-5472.CAN-12-2221. [DOI] [PubMed] [Google Scholar]
  • 17.Nanni P, Gatta V, Menotti L, De Giovanni C, Ianzano M, Palladini A, Grosso V, Dall’ora M, Croci S, Nicoletti G, et al. Preclinical therapy of disseminated HER-2⁺ ovarian and breast carcinomas with a HER-2-retargeted oncolytic herpesvirus. PLoS Pathog. 2013;9:e1003155. doi: 10.1371/journal.ppat.1003155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Passer BJ, Cheema T, Zhou B, Wakimoto H, Zaupa C, Razmjoo M, Sarte J, Wu S, Wu CL, Noah JW, et al. Identification of the ENT1 antagonists dipyridamole and dilazep as amplifiers of oncolytic herpes simplex virus-1 replication. Cancer Res. 2010;70:3890–5. doi: 10.1158/0008-5472.CAN-10-0155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Lee CY, Bu LX, DeBenedetti A, Williams BJ, Rennie PS, Jia WW. Transcriptional and translational dual-regulated oncolytic herpes simplex virus type 1 for targeting prostate tumors. Mol Ther. 2010;18:929–35. doi: 10.1038/mt.2010.26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Cuevas Y, Hernández-Alcoceba R, Aragones J, Naranjo-Suárez S, Castellanos MC, Esteban MA, Martín-Puig S, Landazuri MO, del Peso L. Specific oncolytic effect of a new hypoxia-inducible factor-dependent replicative adenovirus on von Hippel-Lindau-defective renal cell carcinomas. Cancer Res. 2003;63:6877–84. [PubMed] [Google Scholar]
  • 21.Post DE, Van Meir EG. A novel hypoxia-inducible factor (HIF) activated oncolytic adenovirus for cancer therapy. Oncogene. 2003;22:2065–72. doi: 10.1038/sj.onc.1206464. [DOI] [PubMed] [Google Scholar]
  • 22.Lin WH, Yeh SH, Yang WJ, Yeh KH, Fujiwara T, Nii A, Chang SS, Chen PJ. Telomerase-specific oncolytic adenoviral therapy for orthotopic hepatocellular carcinoma in HBx transgenic mice. Int J Cancer. 2013;132:1451–62. doi: 10.1002/ijc.27770. [DOI] [PubMed] [Google Scholar]
  • 23.Takahashi H, Hyakusoku H, Horii C, Takahashi M, Nishimura G, Taguchi T, Kondo N, Sakakibara A, Urata Y, Sano D. Telomerase-specific oncolytic adenovirus: antitumor effects on radiation-resistant head and neck squamous cell carcinoma cells. Head Neck. 2014;36:411–8. doi: 10.1002/hed.23309. [DOI] [PubMed] [Google Scholar]
  • 24.Jin C, Yu D, Čančer M, Nilsson B, Leja J, Essand M. Tat-PTD-modified oncolytic adenovirus driven by the SCG3 promoter and ASH1 enhancer for neuroblastoma therapy. Hum Gene Ther. 2013;24:766–75. doi: 10.1089/hum.2012.132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Wang W, Ji W, Hu H, Ma J, Li X, Mei W, et al. Survivin promoter-regulated oncolytic adenovirus with Hsp70 gene exerts effective antitumor efficacy in gastric cancer immunotherapy. Oncotarget 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Sugio K, Sakurai F, Katayama K, Tashiro K, Matsui H, Kawabata K, Kawase A, Iwaki M, Hayakawa T, Fujiwara T, et al. Enhanced safety profiles of the telomerase-specific replication-competent adenovirus by incorporation of normal cell-specific microRNA-targeted sequences. Clin Cancer Res. 2011;17:2807–18. doi: 10.1158/1078-0432.CCR-10-2008. [DOI] [PubMed] [Google Scholar]
  • 27.Kelly EJ, Nace R, Barber GN, Russell SJ. Attenuation of vesicular stomatitis virus encephalitis through microRNA targeting. J Virol. 2010;84:1550–62. doi: 10.1128/JVI.01788-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Edge RE, Falls TJ, Brown CW, Lichty BD, Atkins H, Bell JC. A let-7 MicroRNA-sensitive vesicular stomatitis virus demonstrates tumor-specific replication. Mol Ther. 2008;16:1437–43. doi: 10.1038/mt.2008.130. [DOI] [PubMed] [Google Scholar]
  • 29.Kelly EJ, Hadac EM, Greiner S, Russell SJ. Engineering microRNA responsiveness to decrease virus pathogenicity. Nat Med. 2008;14:1278–83. doi: 10.1038/nm.1776. [DOI] [PubMed] [Google Scholar]
  • 30.Callegari E, Elamin BK, D’Abundo L, Falzoni S, Donvito G, Moshiri F, Milazzo M, Altavilla G, Giacomelli L, Fornari F, et al. Anti-tumor activity of a miR-199-dependent oncolytic adenovirus. PLoS One. 2013;8:e73964. doi: 10.1371/journal.pone.0073964. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Li JM, Kao KC, Li LF, Yang TM, Wu CP, Horng YM, Jia WW, Yang CT. MicroRNA-145 regulates oncolytic herpes simplex virus-1 for selective killing of human non-small cell lung cancer cells. Virol J. 2013;10:241. doi: 10.1186/1743-422X-10-241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Ylösmäki E, Lavilla-Alonso S, Jäämaa S, Vähä-Koskela M, af Hällström T, Hemminki A, Arola J, Mäkisalo H, Saksela K. MicroRNA-mediated suppression of oncolytic adenovirus replication in human liver. PLoS One. 2013;8:e54506. doi: 10.1371/journal.pone.0054506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Yang X, Chen E, Jiang H, Muszynski K, Harris RD, Giardina SL, Gromeier M, Mitra G, Soman G. Evaluation of IRES-mediated, cell-type-specific cytotoxicity of poliovirus using a colorimetric cell proliferation assay. J Virol Methods. 2009;155:44–54. doi: 10.1016/j.jviromet.2008.09.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Ammayappan A, Nace R, Peng KW, Russell SJ. Neuroattenuation of vesicular stomatitis virus through picornaviral internal ribosome entry sites. J Virol. 2013;87:3217–28. doi: 10.1128/JVI.02984-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Goetz C, Gromeier M. Preparing an oncolytic poliovirus recombinant for clinical application against glioblastoma multiforme. Cytokine Growth Factor Rev. 2010;21:197–203. doi: 10.1016/j.cytogfr.2010.02.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Banaszynski LA, Sellmyer MA, Contag CH, Wandless TJ, Thorne SH. Chemical control of protein stability and function in living mice. Nat Med. 2008;14:1123–7. doi: 10.1038/nm.1754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Glass M, Busche A, Wagner K, Messerle M, Borst EM. Conditional and reversible disruption of essential herpesvirus proteins. Nat Methods. 2009;6:577–9. doi: 10.1038/nmeth.1346. [DOI] [PubMed] [Google Scholar]
  • 38.Li J, Liu H, Li L, Wu H, Wang C, Yan Z, Wang Y, Su C, Jin H, Zhou F, et al. The combination of an oxygen-dependent degradation domain-regulated adenovirus expressing the chemokine RANTES/CCL5 and NK-92 cells exerts enhanced antitumor activity in hepatocellular carcinoma. Oncol Rep. 2013;29:895–902. doi: 10.3892/or.2012.2217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Shobana R, Samal SK, Elankumaran S. Prostate-specific antigen-retargeted recombinant newcastle disease virus for prostate cancer virotherapy. J Virol. 2013;87:3792–800. doi: 10.1128/JVI.02394-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Stojdl DF, Lichty BD, tenOever BR, Paterson JM, Power AT, Knowles S, Marius R, Reynard J, Poliquin L, Atkins H, et al. VSV strains with defects in their ability to shutdown innate immunity are potent systemic anti-cancer agents. Cancer Cell. 2003;4:263–75. doi: 10.1016/S1535-6108(03)00241-1. [DOI] [PubMed] [Google Scholar]
  • 41.Sarinella F, Calistri A, Sette P, Palù G, Parolin C. Oncolysis of pancreatic tumour cells by a gamma34.5-deleted HSV-1 does not rely upon Ras-activation, but on the PI 3-kinase pathway. Gene Ther. 2006;13:1080–7. doi: 10.1038/sj.gt.3302770. [DOI] [PubMed] [Google Scholar]
  • 42.Krishnamurthy S, Takimoto T, Scroggs RA, Portner A. Differentially regulated interferon response determines the outcome of Newcastle disease virus infection in normal and tumor cell lines. J Virol. 2006;80:5145–55. doi: 10.1128/JVI.02618-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Stojdl DF, Lichty B, Knowles S, Marius R, Atkins H, Sonenberg N, Bell JC. Exploiting tumor-specific defects in the interferon pathway with a previously unknown oncolytic virus. Nat Med. 2000;6:821–5. doi: 10.1038/77558. [DOI] [PubMed] [Google Scholar]
  • 44.Coffey MC, Strong JE, Forsyth PA, Lee PW. Reovirus therapy of tumors with activated Ras pathway. Science. 1998;282:1332–4. doi: 10.1126/science.282.5392.1332. [DOI] [PubMed] [Google Scholar]
  • 45.Cheng PH, Rao XM, McMasters KM, Zhou HS. Molecular basis for viral selective replication in cancer cells: activation of CDK2 by adenovirus-induced cyclin E. PLoS One. 2013;8:e57340. doi: 10.1371/journal.pone.0057340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Okemoto K, Wagner B, Meisen H, Haseley A, Kaur B, Chiocca EA. STAT3 activation promotes oncolytic HSV1 replication in glioma cells. PLoS One. 2013;8:e71932. doi: 10.1371/journal.pone.0071932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Cattaneo R, Miest T, Shashkova EV, Barry MA. Reprogrammed viruses as cancer therapeutics: targeted, armed and shielded. Nat Rev Microbiol. 2008;6:529–40. doi: 10.1038/nrmicro1927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Wildner O, Blaese RM, Morris JC. Therapy of colon cancer with oncolytic adenovirus is enhanced by the addition of herpes simplex virus-thymidine kinase. Cancer Res. 1999;59:410–3. [PubMed] [Google Scholar]
  • 49.Tseng JC, Zanzonico PB, Levin B, Finn R, Larson SM, Meruelo D. Tumor-specific in vivo transfection with HSV-1 thymidine kinase gene using a Sindbis viral vector as a basis for prodrug ganciclovir activation and PET. J Nucl Med. 2006;47:1136–43. [PubMed] [Google Scholar]
  • 50.Foloppe J, Kintz J, Futin N, Findeli A, Cordier P, Schlesinger Y, Hoffmann C, Tosch C, Balloul JM, Erbs P. Targeted delivery of a suicide gene to human colorectal tumors by a conditionally replicating vaccinia virus. Gene Ther. 2008;15:1361–71. doi: 10.1038/gt.2008.82. [DOI] [PubMed] [Google Scholar]
  • 51.Liu Y, Deisseroth A. Oncolytic adenoviral vector carrying the cytosine deaminase gene for melanoma gene therapy. Cancer Gene Ther. 2006;13:845–55. doi: 10.1038/sj.cgt.7700962. [DOI] [PubMed] [Google Scholar]
  • 52.Leveille S, Samuel S, Goulet ML, Hiscott J. Enhancing VSV oncolytic activity with an improved cytosine deaminase suicide gene strategy. Cancer Gene Ther. 2011;18:435–43. doi: 10.1038/cgt.2011.14. [DOI] [PubMed] [Google Scholar]
  • 53.Dong X, Qu W, Ma S, Zhu Z, Zheng C, He A, Karlsson A, Xu K, Zheng X. Potent antitumoral effects of targeted promoter-driven oncolytic adenovirus armed with Dm-dNK for breast cancer in vitro and in vivo. Cancer Lett. 2013;328:95–103. doi: 10.1016/j.canlet.2012.09.003. [DOI] [PubMed] [Google Scholar]
  • 54.Hartkopf AD, Bossow S, Lampe J, Zimmermann M, Taran FA, Wallwiener D, Fehm T, Bitzer M, Lauer UM. Enhanced killing of ovarian carcinoma using oncolytic measles vaccine virus armed with a yeast cytosine deaminase and uracil phosphoribosyltransferase. Gynecol Oncol. 2013;130:362–8. doi: 10.1016/j.ygyno.2013.05.004. [DOI] [PubMed] [Google Scholar]
  • 55.Lampe J, Bossow S, Weiland T, Smirnow I, Lehmann R, Neubert W, Bitzer M, Lauer UM. An armed oncolytic measles vaccine virus eliminates human hepatoma cells independently of apoptosis. Gene Ther. 2013;20:1033–41. doi: 10.1038/gt.2013.28. [DOI] [PubMed] [Google Scholar]
  • 56.Shinoura N, Yoshida Y, Asai A, Kirino T, Hamada H. Adenovirus-mediated transfer of p53 and Fas ligand drastically enhances apoptosis in gliomas. Cancer Gene Ther. 2000;7:732–8. doi: 10.1038/sj.cgt.7700160. [DOI] [PubMed] [Google Scholar]
  • 57.Zhao L, Dong A, Gu J, Liu Z, Zhang Y, Zhang W, Wang Y, He L, Qian C, Qian Q, et al. The antitumor activity of TRAIL and IL-24 with replicating oncolytic adenovirus in colorectal cancer. Cancer Gene Ther. 2006;13:1011–22. doi: 10.1038/sj.cgt.7700969. [DOI] [PubMed] [Google Scholar]
  • 58.Zhu W, Zhang H, Shi Y, Song M, Zhu B, Wei L. Oncolytic adenovirus encoding tumor necrosis factor-related apoptosis inducing ligand (TRAIL) inhibits the growth and metastasis of triple-negative breast cancer. Cancer Biol Ther. 2013;14:1016–23. doi: 10.4161/cbt.26043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Jiang G, Li J, Zeng Z, Xian L. Lentivirus-mediated gene therapy by suppressing survivin in BALB/c nude mice bearing oral squamous cell carcinoma. Cancer Biol Ther. 2006;5:435–40. doi: 10.4161/cbt.5.4.2542. [DOI] [PubMed] [Google Scholar]
  • 60.Shen W, Tu JK, Wang XH, Fu ZX. Oncolytic adenovirus mediated Survivin RNA interference and 5-fluorouracil synergistically suppress the lymphatic metastasis of colorectal cancer. Oncol Rep. 2010;24:1285–90. doi: 10.3892/or_00000984. [DOI] [PubMed] [Google Scholar]
  • 61.Shen W, Wang CY, Wang XH, Fu ZX. Oncolytic adenovirus mediated Survivin knockdown by RNA interference suppresses human colorectal carcinoma growth in vitro and in vivo. J Exp Clin Cancer Res. 2009;28:81. doi: 10.1186/1756-9966-28-81. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Chinnery F, King CA, Elliott T, Bateman AR, James E. Viral antigen mediated NKp46 activation of NK cells results in tumor rejection via NK-DC crosstalk. Oncoimmunology. 2012;1:874–83. doi: 10.4161/onci.20636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Woller N, Kühnel F, Kubicka S. Virus infections in tumors pave the way for tumor-directed DC-vaccines. Oncoimmunology. 2012;1:208–10. doi: 10.4161/onci.1.2.18099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Bartlett DL, Liu Z, Sathaiah M, Ravindranathan R, Guo Z, He Y, Guo ZS. Oncolytic viruses as therapeutic cancer vaccines. Mol Cancer. 2013;12:103. doi: 10.1186/1476-4598-12-103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Donnelly OG, Errington-Mais F, Steele L, Hadac E, Jennings V, Scott K, Peach H, Phillips RM, Bond J, Pandha H, et al. Measles virus causes immunogenic cell death in human melanoma. Gene Ther. 2013;20:7–15. doi: 10.1038/gt.2011.205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Sampath P, Li J, Hou W, Chen H, Bartlett DL, Thorne SH. Crosstalk between immune cell and oncolytic vaccinia therapy enhances tumor trafficking and antitumor effects. Mol Ther. 2013;21:620–8. doi: 10.1038/mt.2012.257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Workenhe ST, Mossman KL. Oncolytic virotherapy and immunogenic cancer cell death: sharpening the sword for improved cancer treatment strategies. Mol Ther. 2014;22:251–6. doi: 10.1038/mt.2013.220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Workenhe ST, Simmons G, Pol JG, Lichty BD, Halford WP, Mossman KL. Immunogenic HSV-mediated oncolysis shapes the antitumor immune response and contributes to therapeutic efficacy. Mol Ther. 2014;22:123–31. doi: 10.1038/mt.2013.238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Workenhe ST, Mossman KL. Rewiring cancer cell death to enhance oncolytic viro-immunotherapy. Oncoimmunology. 2013;2:e27138. doi: 10.4161/onci.27138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Granot T, Yamanashi Y, Meruelo D. Sindbis viral vectors transiently deliver tumor-associated antigens to lymph nodes and elicit diversified antitumor CD8+ T-cell immunity. Mol Ther. 2014;22:112–22. doi: 10.1038/mt.2013.215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Kim MK, Breitbach CJ, Moon A, Heo J, Lee YK, Cho M, Lee JW, Kim SG, Kang DH, Bell JC, et al. Oncolytic and immunotherapeutic vaccinia induces antibody-mediated complement-dependent cancer cell lysis in humans. Sci Transl Med. 2013;5:185ra63. doi: 10.1126/scitranslmed.3005361. [DOI] [PubMed] [Google Scholar]
  • 72.Liikanen I, Ahtiainen L, Hirvinen ML, Bramante S, Cerullo V, Nokisalmi P, Hemminki O, Diaconu I, Pesonen S, Koski A, et al. Oncolytic adenovirus with temozolomide induces autophagy and antitumor immune responses in cancer patients. Mol Ther. 2013;21:1212–23. doi: 10.1038/mt.2013.51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Kanerva A, Nokisalmi P, Diaconu I, Koski A, Cerullo V, Liikanen I, Tähtinen S, Oksanen M, Heiskanen R, Pesonen S, et al. Antiviral and antitumor T-cell immunity in patients treated with GM-CSF-coding oncolytic adenovirus. Clin Cancer Res. 2013;19:2734–44. doi: 10.1158/1078-0432.CCR-12-2546. [DOI] [PubMed] [Google Scholar]
  • 74.Batenchuk C, Le Boeuf F, Stubbert L, Falls T, Atkins HL, Bell JC, Conrad DP. Non-replicating rhabdovirus-derived particles (NRRPs) eradicate acute leukemia by direct cytolysis and induction of antitumor immunity. Blood Cancer J. 2013;3:e123. doi: 10.1038/bcj.2013.23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Pol JG, Zhang L, Bridle BW, Stephenson KB, Rességuier J, Hanson S, Chen L, Kazdhan N, Bramson JL, Stojdl DF, et al. Maraba virus as a potent oncolytic vaccine vector. Mol Ther. 2014;22:420–9. doi: 10.1038/mt.2013.249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Bridle BW, Chen L, Lemay CG, Diallo JS, Pol J, Nguyen A, Capretta A, He R, Bramson JL, Bell JC, et al. HDAC inhibition suppresses primary immune responses, enhances secondary immune responses, and abrogates autoimmunity during tumor immunotherapy. Mol Ther. 2013;21:887–94. doi: 10.1038/mt.2012.265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Bridle BW, Clouthier D, Zhang L, Pol J, Chen L, Lichty BD, Bramson JL, Wan Y. Oncolytic vesicular stomatitis virus quantitatively and qualitatively improves primary CD8(+) T-cell responses to anticancer vaccines. Oncoimmunology. 2013;2:e26013. doi: 10.4161/onci.26013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Bridle BW, Stephenson KB, Boudreau JE, Koshy S, Kazdhan N, Pullenayegum E, Brunellière J, Bramson JL, Lichty BD, Wan Y. Potentiating cancer immunotherapy using an oncolytic virus. Mol Ther. 2010;18:1430–9. doi: 10.1038/mt.2010.98. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Bridle BW, Boudreau JE, Lichty BD, Brunellière J, Stephenson K, Koshy S, Bramson JL, Wan Y. Vesicular stomatitis virus as a novel cancer vaccine vector to prime antitumor immunity amenable to rapid boosting with adenovirus. Mol Ther. 2009;17:1814–21. doi: 10.1038/mt.2009.154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Diaconu I, Cerullo V, Hirvinen ML, Escutenaire S, Ugolini M, Pesonen SK, Bramante S, Parviainen S, Kanerva A, Loskog AS, et al. Immune response is an important aspect of the antitumor effect produced by a CD40L-encoding oncolytic adenovirus. Cancer Res. 2012;72:2327–38. doi: 10.1158/0008-5472.CAN-11-2975. [DOI] [PubMed] [Google Scholar]
  • 81.Pesonen S, Diaconu I, Kangasniemi L, Ranki T, Kanerva A, Pesonen SK, Gerdemann U, Leen AM, Kairemo K, Oksanen M, et al. Oncolytic immunotherapy of advanced solid tumors with a CD40L-expressing replicating adenovirus: assessment of safety and immunologic responses in patients. Cancer Res. 2012;72:1621–31. doi: 10.1158/0008-5472.CAN-11-3001. [DOI] [PubMed] [Google Scholar]
  • 82.Gomes EM, Rodrigues MS, Phadke AP, Butcher LD, Starling C, Chen S, Chang D, Hernandez-Alcoceba R, Newman JT, Stone MJ, et al. Antitumor activity of an oncolytic adenoviral-CD40 ligand (CD154) transgene construct in human breast cancer cells. Clin Cancer Res. 2009;15:1317–25. doi: 10.1158/1078-0432.CCR-08-1360. [DOI] [PubMed] [Google Scholar]
  • 83.Choi KJ, Kim JH, Lee YS, Kim J, Suh BS, Kim H, Cho S, Sohn JH, Kim GE, Yun CO. Concurrent delivery of GM-CSF and B7-1 using an oncolytic adenovirus elicits potent antitumor effect. Gene Ther. 2006;13:1010–20. doi: 10.1038/sj.gt.3302759. [DOI] [PubMed] [Google Scholar]
  • 84.Parviainen S, Ahonen M, Diaconu I, Hirvinen M, Karttunen Å, Vähä-Koskela M, Hemminki A, Cerullo V. CD40 ligand and tdTomato-armed vaccinia virus for induction of antitumor immune response and tumor imaging. Gene Ther. 2014;21:195–204. doi: 10.1038/gt.2013.73. [DOI] [PubMed] [Google Scholar]
  • 85.Zamarin D, Vigil A, Kelly K, García-Sastre A, Fong Y. Genetically engineered Newcastle disease virus for malignant melanoma therapy. Gene Ther. 2009;16:796–804. doi: 10.1038/gt.2009.14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Takehara Y, Satoh T, Nishizawa A, Saeki K, Nakamura M, Masuzawa M, Kaneda Y, Katayama I, Yokozeki H. Anti-tumor effects of inactivated Sendai virus particles with an IL-2 gene on angiosarcoma. Clin Immunol. 2013;149:1–10. doi: 10.1016/j.clim.2013.05.019. [DOI] [PubMed] [Google Scholar]
  • 87.Beyer M. Interleukin-2 treatment of tumor patients can expand regulatory T cells. Oncoimmunology. 2012;1:1181–2. doi: 10.4161/onci.20639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Wong RJ, Chan MK, Yu Z, Ghossein RA, Ngai I, Adusumilli PS, Stiles BM, Shah JP, Singh B, Fong Y. Angiogenesis inhibition by an oncolytic herpes virus expressing interleukin 12. Clin Cancer Res. 2004;10:4509–16. doi: 10.1158/1078-0432.CCR-04-0081. [DOI] [PubMed] [Google Scholar]
  • 89.Zhang SN, Choi IK, Huang JH, Yoo JY, Choi KJ, Yun CO. Optimizing DC vaccination by combination with oncolytic adenovirus coexpressing IL-12 and GM-CSF. Mol Ther. 2011;19:1558–68. doi: 10.1038/mt.2011.29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Parker JN, Gillespie GY, Love CE, Randall S, Whitley RJ, Markert JM. Engineered herpes simplex virus expressing IL-12 in the treatment of experimental murine brain tumors. Proc Natl Acad Sci U S A. 2000;97:2208–13. doi: 10.1073/pnas.040557897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Passer BJ, Cheema T, Wu S, Wu CL, Rabkin SD, Martuza RL. Combination of vinblastine and oncolytic herpes simplex virus vector expressing IL-12 therapy increases antitumor and antiangiogenic effects in prostate cancer models. Cancer Gene Ther. 2013;20:17–24. doi: 10.1038/cgt.2012.75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Zhang W, Fulci G, Wakimoto H, Cheema TA, Buhrman JS, Jeyaretna DS, Stemmer Rachamimov AO, Rabkin SD, Martuza RL. Combination of oncolytic herpes simplex viruses armed with angiostatin and IL-12 enhances antitumor efficacy in human glioblastoma models. Neoplasia. 2013;15:591–9. doi: 10.1593/neo.13158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Freytag SO, Barton KN, Zhang Y. Efficacy of oncolytic adenovirus expressing suicide genes and interleukin-12 in preclinical model of prostate cancer. Gene Ther. 2013;20:1131–9. doi: 10.1038/gt.2013.40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Roth JC, Cassady KA, Cody JJ, Parker JN, Price KH, Coleman JM, et al. Evaluation of the Safety and Biodistribution of M032, an Attenuated HSV-1 Virus Expressing hIL-12, After Intracerebral Administration to Aotus Non-Human Primates. Hum Gene Ther Clin Dev 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Malvicini M, Alaniz L, Bayo J, Garcia M, Piccioni F, Fiore E, Atorrasagasti C, Aquino JB, Matar P, Mazzolini G. Single low-dose cyclophosphamide combined with interleukin-12 gene therapy is superior to a metronomic schedule in inducing immunity against colorectal carcinoma in mice. Oncoimmunology. 2012;1:1038–47. doi: 10.4161/onci.20684. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Wong HC, Jeng EK, Rhode PR. The IL-15-based superagonist ALT-803 promotes the antigen-independent conversion of memory CD8(+) T cells into innate-like effector cells with antitumor activity. Oncoimmunology. 2013;2:e26442. doi: 10.4161/onci.26442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Stephenson KB, Barra NG, Davies E, Ashkar AA, Lichty BD. Expressing human interleukin-15 from oncolytic vesicular stomatitis virus improves survival in a murine metastatic colon adenocarcinoma model through the enhancement of anti-tumor immunity. Cancer Gene Ther. 2012;19:238–46. doi: 10.1038/cgt.2011.81. [DOI] [PubMed] [Google Scholar]
  • 98.Liu J, Wennier S, Reinhard M, Roy E, MacNeill A, McFadden G. Myxoma virus expressing interleukin-15 fails to cause lethal myxomatosis in European rabbits. J Virol. 2009;83:5933–8. doi: 10.1128/JVI.00204-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.van Rikxoort M, Michaelis M, Wolschek M, Muster T, Egorov A, Seipelt J, Doerr HW, Cinatl J., Jr. Oncolytic effects of a novel influenza A virus expressing interleukin-15 from the NS reading frame. PLoS One. 2012;7:e36506. doi: 10.1371/journal.pone.0036506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Gaston DC, Odom CI, Li L, Markert JM, Roth JC, Cassady KA, Whitley RJ, Parker JN. Production of bioactive soluble interleukin-15 in complex with interleukin-15 receptor alpha from a conditionally-replicating oncolytic HSV-1. PLoS One. 2013;8:e81768. doi: 10.1371/journal.pone.0081768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Vincent M, Quéméner A, Jacques Y. Antitumor activity of an immunocytokine composed of an anti-GD2 antibody and the IL-15 superagonist RLI. Oncoimmunology. 2013;2:e26441. doi: 10.4161/onci.26441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Choi IK, Li Y, Oh E, Kim J, Yun CO. Oncolytic adenovirus expressing IL-23 and p35 elicits IFN-γ- and TNF-α-co-producing T cell-mediated antitumor immunity. PLoS One. 2013;8:e67512. doi: 10.1371/journal.pone.0067512. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Jiang G, Jiang AJ, Cheng Q, Tian H, Li LT, Zheng JN. A dual-regulated oncolytic adenovirus expressing interleukin-24 sensitizes melanoma cells to temozolomide via the induction of apoptosis. Tumour Biol. 2013;34:1263–71. doi: 10.1007/s13277-013-0701-7. [DOI] [PubMed] [Google Scholar]
  • 104.Lou W, Chen Q, Ma L, Liu J, Yang Z, Shen J, Cui Y, Bian XW, Qian C. Oncolytic adenovirus co-expressing miRNA-34a and IL-24 induces superior antitumor activity in experimental tumor model. J Mol Med (Berl) 2013;91:715–25. doi: 10.1007/s00109-012-0985-x. [DOI] [PubMed] [Google Scholar]
  • 105.He B, Huang X, Liu X, Xu B. Cancer targeting gene-viro-therapy for pancreatic cancer using oncolytic adenovirus ZD55-IL-24 in immune-competent mice. Mol Biol Rep. 2013;40:5397–405. doi: 10.1007/s11033-013-2638-8. [DOI] [PubMed] [Google Scholar]
  • 106.Fang L, Cheng Q, Bai J, Qi YD, Liu JJ, Li LT, Zheng JN. An oncolytic adenovirus expressing interleukin-24 enhances antitumor activities in combination with paclitaxel in breast cancer cells. Mol Med Rep. 2013;8:1416–24. doi: 10.3892/mmr.2013.1680. [DOI] [PubMed] [Google Scholar]
  • 107.Pesonen S, Diaconu I, Cerullo V, Escutenaire S, Raki M, Kangasniemi L, Nokisalmi P, Dotti G, Guse K, Laasonen L, et al. Integrin targeted oncolytic adenoviruses Ad5-D24-RGD and Ad5-RGD-D24-GMCSF for treatment of patients with advanced chemotherapy refractory solid tumors. Int J Cancer. 2012;130:1937–47. doi: 10.1002/ijc.26216. [DOI] [PubMed] [Google Scholar]
  • 108.Burke JM, Lamm DL, Meng MV, Nemunaitis JJ, Stephenson JJ, Arseneau JC, Aimi J, Lerner S, Yeung AW, Kazarian T, et al. A first in human phase 1 study of CG0070, a GM-CSF expressing oncolytic adenovirus, for the treatment of nonmuscle invasive bladder cancer. J Urol. 2012;188:2391–7. doi: 10.1016/j.juro.2012.07.097. [DOI] [PubMed] [Google Scholar]
  • 109.Heo J, Reid T, Ruo L, Breitbach CJ, Rose S, Bloomston M, Cho M, Lim HY, Chung HC, Kim CW, et al. Randomized dose-finding clinical trial of oncolytic immunotherapeutic vaccinia JX-594 in liver cancer. Nat Med. 2013;19:329–36. doi: 10.1038/nm.3089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Thorne SH. The role of GM-CSF in enhancing immunotherapy of cancer. Immunotherapy. 2013;5:817–9. doi: 10.2217/imt.13.65. [DOI] [PubMed] [Google Scholar]
  • 111.Grossardt C, Engeland CE, Bossow S, Halama N, Zaoui K, Leber MF, Springfeld C, Jaeger D, von Kalle C, Ungerechts G. Granulocyte-macrophage colony-stimulating factor-armed oncolytic measles virus is an effective therapeutic cancer vaccine. Hum Gene Ther. 2013;24:644–54. doi: 10.1089/hum.2012.205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Liu H, Yuan SJ, Chen YT, Xie YB, Cui L, Yang WZ, Yang DX, Tian YT. Preclinical evaluation of herpes simplex virus armed with granulocyte-macrophage colony-stimulating factor in pancreatic carcinoma. World J Gastroenterol. 2013;19:5138–43. doi: 10.3748/wjg.v19.i31.5138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Egilmez NK, Harden JL, Rowswell-Turner RB. Chemoimmunotherapy as long-term maintenance therapy for cancer. Oncoimmunology. 2012;1:563–5. doi: 10.4161/onci.19369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Dempe S, Lavie M, Struyf S, Bhat R, Verbeke H, Paschek S, Berghmans N, Geibig R, Rommelaere J, Van Damme J, et al. Antitumoral activity of parvovirus-mediated IL-2 and MCP-3/CCL7 delivery into human pancreatic cancer: implication of leucocyte recruitment. Cancer Immunol Immunother. 2012;61:2113–23. doi: 10.1007/s00262-012-1279-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Li J, O’Malley M, Sampath P, Kalinski P, Bartlett DL, Thorne SH. Expression of CCL19 from oncolytic vaccinia enhances immunotherapeutic potential while maintaining oncolytic activity. Neoplasia. 2012;14:1115–21. doi: 10.1593/neo.121272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Alvarez-Breckenridge CA, Yu J, Caligiuri MA, Chiocca EA. Uncovering a novel mechanism whereby NK cells interfere with glioblastoma virotherapy. Oncoimmunology. 2013;2:e23658. doi: 10.4161/onci.23658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Colombo F, Barzon L, Franchin E, Pacenti M, Pinna V, Danieli D, Zanusso M, Palù G. Combined HSV-TK/IL-2 gene therapy in patients with recurrent glioblastoma multiforme: biological and clinical results. Cancer Gene Ther. 2005;12:835–48. doi: 10.1038/sj.cgt.7700851. [DOI] [PubMed] [Google Scholar]
  • 118.Lun X, Alain T, Zemp FJ, Zhou H, Rahman MM, Hamilton MG, McFadden G, Bell J, Senger DL, Forsyth PA. Myxoma virus virotherapy for glioma in immunocompetent animal models: optimizing administration routes and synergy with rapamycin. Cancer Res. 2010;70:598–608. doi: 10.1158/0008-5472.CAN-09-1510. [DOI] [PubMed] [Google Scholar]
  • 119.Ram Z, Culver KW, Oshiro EM, Viola JJ, DeVroom HL, Otto E, Long Z, Chiang Y, McGarrity GJ, Muul LM, et al. Therapy of malignant brain tumors by intratumoral implantation of retroviral vector-producing cells. Nat Med. 1997;3:1354–61. doi: 10.1038/nm1297-1354. [DOI] [PubMed] [Google Scholar]
  • 120.Smyth JW, Fleeton MN, Sheahan BJ, Atkins GJ. Treatment of rapidly growing K-BALB and CT26 mouse tumours using Semliki Forest virus and its derived vector. Gene Ther. 2005;12:147–59. doi: 10.1038/sj.gt.3302390. [DOI] [PubMed] [Google Scholar]
  • 121.Vähä-Koskela MJ, Kallio JP, Jansson LC, Heikkilä JE, Zakhartchenko VA, Kallajoki MA, Kähäri VM, Hinkkanen AE. Oncolytic capacity of attenuated replicative semliki forest virus in human melanoma xenografts in severe combined immunodeficient mice. Cancer Res. 2006;66:7185–94. doi: 10.1158/0008-5472.CAN-05-2214. [DOI] [PubMed] [Google Scholar]
  • 122.Cordaro TA, de Visser KE, Tirion FH, Graus YM, Haanen JB, Kioussis D, Kruisbeek AM. Tumor size at the time of adoptive transfer determines whether tumor rejection occurs. Eur J Immunol. 2000;30:1297–307. doi: 10.1002/(SICI)1521-4141(200005)30:5<1297::AID-IMMU1297>3.0.CO;2-C. [DOI] [PubMed] [Google Scholar]
  • 123.Li ZY, Ni S, Yang X, Kiviat N, Lieber A. Xenograft models for liver metastasis: Relationship between tumor morphology and adenovirus vector transduction. Mol Ther. 2004;9:650–7. doi: 10.1016/j.ymthe.2004.01.021. [DOI] [PubMed] [Google Scholar]
  • 124.Shayakhmetov DM, Li ZY, Ni S, Lieber A. Targeting of adenovirus vectors to tumor cells does not enable efficient transduction of breast cancer metastases. Cancer Res. 2002;62:1063–8. [PubMed] [Google Scholar]
  • 125.Stohrer M, Boucher Y, Stangassinger M, Jain RK. Oncotic pressure in solid tumors is elevated. Cancer Res. 2000;60:4251–5. [PubMed] [Google Scholar]
  • 126.Bilbao R, Bustos M, Alzuguren P, Pajares MJ, Drozdzik M, Qian C, Prieto J. A blood-tumor barrier limits gene transfer to experimental liver cancer: the effect of vasoactive compounds. Gene Ther. 2000;7:1824–32. doi: 10.1038/sj.gt.3301312. [DOI] [PubMed] [Google Scholar]
  • 127.Cairns R, Papandreou I, Denko N. Overcoming physiologic barriers to cancer treatment by molecularly targeting the tumor microenvironment. Mol Cancer Res. 2006;4:61–70. doi: 10.1158/1541-7786.MCR-06-0002. [DOI] [PubMed] [Google Scholar]
  • 128.Grote D, Russell SJ, Cornu TI, Cattaneo R, Vile R, Poland GA, Fielding AK. Live attenuated measles virus induces regression of human lymphoma xenografts in immunodeficient mice. Blood. 2001;97:3746–54. doi: 10.1182/blood.V97.12.3746. [DOI] [PubMed] [Google Scholar]
  • 129.Peng KW, TenEyck CJ, Galanis E, Kalli KR, Hartmann LC, Russell SJ. Intraperitoneal therapy of ovarian cancer using an engineered measles virus. Cancer Res. 2002;62:4656–62. [PubMed] [Google Scholar]
  • 130.Kanerva A, Zinn KR, Peng KW, Ranki T, Kangasniemi L, Chaudhuri TR, Desmond RA, Wang M, Takayama K, Hakkarainen T, et al. Noninvasive dual modality in vivo monitoring of the persistence and potency of a tumor targeted conditionally replicating adenovirus. Gene Ther. 2005;12:87–94. doi: 10.1038/sj.gt.3302387. [DOI] [PubMed] [Google Scholar]
  • 131.Alain T, Kim M, Johnston RN, Urbanski S, Kossakowska AE, Forsyth PA, Lee PW. The oncolytic effect in vivo of reovirus on tumour cells that have survived reovirus cell killing in vitro. Br J Cancer. 2006;95:1020–7. doi: 10.1038/sj.bjc.6603363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Kim YT, Ganly I, Brown R, Stuart D. Acquired resistance to cytolysis of the E1B-attenuated adenovirus, dl1520, in ovarian tumour cell lines. Cancer Gene Ther. 2003;10:589–90. doi: 10.1038/sj.cgt.7700607. [DOI] [PubMed] [Google Scholar]
  • 133.Vitale I, Galluzzi L, Castedo M, Kroemer G. Mitotic catastrophe: a mechanism for avoiding genomic instability. Nat Rev Mol Cell Biol. 2011;12:385–92. doi: 10.1038/nrm3115. [DOI] [PubMed] [Google Scholar]
  • 134.Vitale I, Galluzzi L, Senovilla L, Criollo A, Jemaà M, Castedo M, Kroemer G. Illicit survival of cancer cells during polyploidization and depolyploidization. Cell Death Differ. 2011;18:1403–13. doi: 10.1038/cdd.2010.145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Fisher KD, Stallwood Y, Green NK, Ulbrich K, Mautner V, Seymour LW. Polymer-coated adenovirus permits efficient retargeting and evades neutralising antibodies. Gene Ther. 2001;8:341–8. doi: 10.1038/sj.gt.3301389. [DOI] [PubMed] [Google Scholar]
  • 136.Massari I, Donnini A, Argentati K, Straino S, Mangoni A, Gaetano C, Viticchi C, Capogrossi M, Provinciali M. Age-dependent effects of repeated immunization with a first generation adenovirus vector on the immune response and transgene expression in young and old rats. Exp Gerontol. 2002;37:823–31. doi: 10.1016/S0531-5565(02)00011-6. [DOI] [PubMed] [Google Scholar]
  • 137.Ikeda K, Wakimoto H, Ichikawa T, Jhung S, Hochberg FH, Louis DN, Chiocca EA. Complement depletion facilitates the infection of multiple brain tumors by an intravascular, replication-conditional herpes simplex virus mutant. J Virol. 2000;74:4765–75. doi: 10.1128/JVI.74.10.4765-4775.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Pensiero MN, Wysocki CA, Nader K, Kikuchi GE. Development of amphotropic murine retrovirus vectors resistant to inactivation by human serum. Hum Gene Ther. 1996;7:1095–101. doi: 10.1089/hum.1996.7.9-1095. [DOI] [PubMed] [Google Scholar]
  • 139.Bernt KM, Ni S, Gaggar A, Li ZY, Shayakhmetov DM, Lieber A. The effect of sequestration by nontarget tissues on anti-tumor efficacy of systemically applied, conditionally replicating adenovirus vectors. Mol Ther. 2003;8:746–55. doi: 10.1016/j.ymthe.2003.07.006. [DOI] [PubMed] [Google Scholar]
  • 140.Underhill DM, Ozinsky A. Phagocytosis of microbes: complexity in action. Annu Rev Immunol. 2002;20:825–52. doi: 10.1146/annurev.immunol.20.103001.114744. [DOI] [PubMed] [Google Scholar]
  • 141.Doronin K, Shashkova EV, May SM, Hofherr SE, Barry MA. Chemical modification with high molecular weight polyethylene glycol reduces transduction of hepatocytes and increases efficacy of intravenously delivered oncolytic adenovirus. Hum Gene Ther. 2009;20:975–88. doi: 10.1089/hum.2009.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Ferguson MS, Lemoine NR, Wang Y. Systemic delivery of oncolytic viruses: hopes and hurdles. Adv Virol. 2012;2012:805629. doi: 10.1155/2012/805629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Brenner C, Galluzzi L, Kepp O, Kroemer G. Decoding cell death signals in liver inflammation. J Hepatol. 2013;59:583–94. doi: 10.1016/j.jhep.2013.03.033. [DOI] [PubMed] [Google Scholar]
  • 144.Chisholm J, Howe K, Taj M, Zambon M. Influenza immunisation in children with solid tumours. Eur J Cancer. 2005;41:2280–7. doi: 10.1016/j.ejca.2005.07.006. [DOI] [PubMed] [Google Scholar]
  • 145.Cooksley CD, Avritscher EB, Bekele BN, Rolston KV, Geraci JM, Elting LS. Epidemiology and outcomes of serious influenza-related infections in the cancer population. Cancer. 2005;104:618–28. doi: 10.1002/cncr.21203. [DOI] [PubMed] [Google Scholar]
  • 146.Hicks KL, Chemaly RF, Kontoyiannis DP. Common community respiratory viruses in patients with cancer: more than just “common colds”. Cancer. 2003;97:2576–87. doi: 10.1002/cncr.11353. [DOI] [PubMed] [Google Scholar]
  • 147.Nemunaitis J, Khuri F, Ganly I, Arseneau J, Posner M, Vokes E, Kuhn J, McCarty T, Landers S, Blackburn A, 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. doi: 10.1200/JCO.2001.19.2.289. [DOI] [PubMed] [Google Scholar]
  • 148.Geevarghese SK, Geller DA, de Haan HA, Hörer M, Knoll AE, Mescheder A, Nemunaitis J, Reid TR, Sze DY, Tanabe KK, et al. Phase I/II study of oncolytic herpes simplex virus NV1020 in patients with extensively pretreated refractory colorectal cancer metastatic to the liver. Hum Gene Ther. 2010;21:1119–28. doi: 10.1089/hum.2010.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Hacein-Bey-Abina S, von Kalle C, Schmidt M, Le Deist F, Wulffraat N, McIntyre E, Radford I, Villeval JL, Fraser CC, Cavazzana-Calvo M, et al. A serious adverse event after successful gene therapy for X-linked severe combined immunodeficiency. N Engl J Med. 2003;348:255–6. doi: 10.1056/NEJM200301163480314. [DOI] [PubMed] [Google Scholar]
  • 150.Hacein-Bey-Abina S, Von Kalle C, Schmidt M, McCormack MP, Wulffraat N, Leboulch P, Lim A, Osborne CS, Pawliuk R, Morillon E, et al. LMO2-associated clonal T cell proliferation in two patients after gene therapy for SCID-X1. Science. 2003;302:415–9. doi: 10.1126/science.1088547. [DOI] [PubMed] [Google Scholar]
  • 151.Thaci B, Ulasov IV, Ahmed AU, Ferguson SD, Han Y, Lesniak MS. Anti-angiogenic therapy increases intratumoral adenovirus distribution by inducing collagen degradation. Gene Ther. 2013;20:318–27. doi: 10.1038/gt.2012.42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Tsuji T, Nakamori M, Iwahashi M, Nakamura M, Ojima T, Iida T, Katsuda M, Hayata K, Ino Y, Todo T, et al. An armed oncolytic herpes simplex virus expressing thrombospondin-1 has an enhanced in vivo antitumor effect against human gastric cancer. Int J Cancer. 2013;132:485–94. doi: 10.1002/ijc.27681. [DOI] [PubMed] [Google Scholar]
  • 153.Wang H, Wei F, Li H, Ji X, Li S, Chen X. Combination of oncolytic adenovirus and endostatin inhibits human retinoblastoma in an in vivo mouse model. Int J Mol Med. 2013;31:377–85. doi: 10.3892/ijmm.2012.1197. [DOI] [PubMed] [Google Scholar]
  • 154.Li LX, Zhang YL, Zhou L, Ke ML, Chen JM, Fu X, Ye CL, Wu JX, Liu RY, Huang W. Antitumor efficacy of a recombinant adenovirus encoding endostatin combined with an E1B55KD-deficient adenovirus in gastric cancer cells. J Transl Med. 2013;11:257. doi: 10.1186/1479-5876-11-257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Breitbach CJ, Arulanandam R, De Silva N, Thorne SH, Patt R, Daneshmand M, Moon A, Ilkow C, Burke J, Hwang TH, et al. Oncolytic vaccinia virus disrupts tumor-associated vasculature in humans. Cancer Res. 2013;73:1265–75. doi: 10.1158/0008-5472.CAN-12-2687. [DOI] [PubMed] [Google Scholar]
  • 156.Gil M, Seshadri M, Komorowski MP, Abrams SI, Kozbor D. Targeting CXCL12/CXCR4 signaling with oncolytic virotherapy disrupts tumor vasculature and inhibits breast cancer metastases. Proc Natl Acad Sci U S A. 2013;110:E1291–300. doi: 10.1073/pnas.1220580110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Jain RK. Normalization of tumor vasculature: an emerging concept in antiangiogenic therapy. Science. 2005;307:58–62. doi: 10.1126/science.1104819. [DOI] [PubMed] [Google Scholar]
  • 158.Goel S, Duda DG, Xu L, Munn LL, Boucher Y, Fukumura D, Jain RK. Normalization of the vasculature for treatment of cancer and other diseases. Physiol Rev. 2011;91:1071–121. doi: 10.1152/physrev.00038.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Huang Y, Goel S, Duda DG, Fukumura D, Jain RK. Vascular normalization as an emerging strategy to enhance cancer immunotherapy. Cancer Res. 2013;73:2943–8. doi: 10.1158/0008-5472.CAN-12-4354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Long KB, Beatty GL. Harnessing the antitumor potential of macrophages for cancer immunotherapy. Oncoimmunology. 2013;2:e26860. doi: 10.4161/onci.26860. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Devaud C, John LB, Westwood JA, Darcy PK, Kershaw MH. Immune modulation of the tumor microenvironment for enhancing cancer immunotherapy. Oncoimmunology. 2013;2:e25961. doi: 10.4161/onci.25961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Chandra D, Gravekamp C. Myeloid-derived suppressor cells: Cellular missiles to target tumors. Oncoimmunology. 2013;2:e26967. doi: 10.4161/onci.26967. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Pan PY, Chen HM, Chen SH. Myeloid-derived suppressor cells as a Trojan horse: A cellular vehicle for the delivery of oncolytic viruses. Oncoimmunology. 2013;2:e25083. doi: 10.4161/onci.25083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Senovilla L, Vacchelli E, Galon J, Adjemian S, Eggermont A, Fridman WH, Sautès-Fridman C, Ma Y, Tartour E, Zitvogel L, et al. Trial watch: Prognostic and predictive value of the immune infiltrate in cancer. Oncoimmunology. 2012;1:1323–43. doi: 10.4161/onci.22009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Muthana M, Rodrigues S, Chen YY, Welford A, Hughes R, Tazzyman S, Essand M, Morrow F, Lewis CE. Macrophage delivery of an oncolytic virus abolishes tumor regrowth and metastasis after chemotherapy or irradiation. Cancer Res. 2013;73:490–5. doi: 10.1158/0008-5472.CAN-12-3056. [DOI] [PubMed] [Google Scholar]
  • 166.Adair RA, Scott KJ, Fraser S, Errington-Mais F, Pandha H, Coffey M, Selby P, Cook GP, Vile R, Harrington KJ, et al. Cytotoxic and immune-mediated killing of human colorectal cancer by reovirus-loaded blood and liver mononuclear cells. Int J Cancer. 2013;132:2327–38. doi: 10.1002/ijc.27918. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Eisenstein S, Coakley BA, Briley-Saebo K, Ma G, Chen HM, Meseck M, Ward S, Divino C, Woo S, Chen SH, et al. Myeloid-derived suppressor cells as a vehicle for tumor-specific oncolytic viral therapy. Cancer Res. 2013;73:5003–15. doi: 10.1158/0008-5472.CAN-12-1597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Mader EK, Butler G, Dowdy SC, Mariani A, Knutson KL, Federspiel MJ, Russell SJ, Galanis E, Dietz AB, Peng KW. Optimizing patient derived mesenchymal stem cells as virus carriers for a phase I clinical trial in ovarian cancer. J Transl Med. 2013;11:20. doi: 10.1186/1479-5876-11-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Ong HT, Federspiel MJ, Guo CM, Ooi LL, Russell SJ, Peng KW, Hui KM. Systemically delivered measles virus-infected mesenchymal stem cells can evade host immunity to inhibit liver cancer growth. J Hepatol. 2013;59:999–1006. doi: 10.1016/j.jhep.2013.07.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Castleton A, Dey A, Beaton B, Patel B, Aucher A, Davis DM, Fielding AK. Human mesenchymal stromal cells deliver systemic oncolytic measles virus to treat acute lymphoblastic leukemia in the presence of humoral immunity. Blood. 2014;123:1327–35. doi: 10.1182/blood-2013-09-528851. [DOI] [PubMed] [Google Scholar]
  • 171.Esaki S, Goshima F, Kimura H, Murakami S, Nishiyama Y. Enhanced antitumoral activity of oncolytic herpes simplex virus with gemcitabine using colorectal tumor models. Int J Cancer. 2013;132:1592–601. doi: 10.1002/ijc.27823. [DOI] [PubMed] [Google Scholar]
  • 172.Amit M, Gil Z. Macrophages increase the resistance of pancreatic adenocarcinoma cells to gemcitabine by upregulating cytidine deaminase. Oncoimmunology. 2013;2:e27231. doi: 10.4161/onci.27231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Galluzzi L, Senovilla L, Zitvogel L, Kroemer G. The secret ally: immunostimulation by anticancer drugs. Nat Rev Drug Discov. 2012;11:215–33. doi: 10.1038/nrd3626. [DOI] [PubMed] [Google Scholar]
  • 174.Roulstone V, Twigger K, Zaidi S, Pencavel T, Kyula JN, White C, McLaughlin M, Seth R, Karapanagiotou EM, Mansfield D, et al. Synergistic cytotoxicity of oncolytic reovirus in combination with cisplatin-paclitaxel doublet chemotherapy. Gene Ther. 2013;20:521–8. doi: 10.1038/gt.2012.68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Zeng WG, Li JJ, Hu P, Lei L, Wang JN, Liu RB. An oncolytic herpes simplex virus vector, G47Δ, synergizes with paclitaxel in the treatment of breast cancer. Oncol Rep. 2013;29:2355–61. doi: 10.3892/or.2013.2359. [DOI] [PubMed] [Google Scholar]
  • 176.Senovilla L, Vitale I, Martins I, Tailler M, Pailleret C, Michaud M, Galluzzi L, Adjemian S, Kepp O, Niso-Santano M, et al. An immunosurveillance mechanism controls cancer cell ploidy. Science. 2012;337:1678–84. doi: 10.1126/science.1224922. [DOI] [PubMed] [Google Scholar]
  • 177.Peng KW, Myers R, Greenslade A, Mader E, Greiner S, Federspiel MJ, Dispenzieri A, Russell SJ. Using clinically approved cyclophosphamide regimens to control the humoral immune response to oncolytic viruses. Gene Ther. 2013;20:255–61. doi: 10.1038/gt.2012.31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Vacchelli E, Galluzzi L, Fridman WH, Galon J, Sautès-Fridman C, Tartour E, Kroemer G. Trial watch: Chemotherapy with immunogenic cell death inducers. Oncoimmunology. 2012;1:179–88. doi: 10.4161/onci.1.2.19026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Walter S, Weinschenk T, Reinhardt C, Singh-Jasuja H. Single-dose cyclophosphamide synergizes with immune responses to the renal cell cancer vaccine IMA901. Oncoimmunology. 2013;2:e22246. doi: 10.4161/onci.22246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Jha BK, Dong B, Nguyen CT, Polyakova I, Silverman RH. Suppression of antiviral innate immunity by sunitinib enhances oncolytic virotherapy. Mol Ther. 2013;21:1749–57. doi: 10.1038/mt.2013.112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Zitvogel L, Galluzzi L, Smyth MJ, Kroemer G. Mechanism of action of conventional and targeted anticancer therapies: reinstating immunosurveillance. Immunity. 2013;39:74–88. doi: 10.1016/j.immuni.2013.06.014. [DOI] [PubMed] [Google Scholar]
  • 182.Boehrer S, Adès L, Braun T, Galluzzi L, Grosjean J, Fabre C, Le Roux G, Gardin C, Martin A, de Botton S, et al. Erlotinib exhibits antineoplastic off-target effects in AML and MDS: a preclinical study. Blood. 2008;111:2170–80. doi: 10.1182/blood-2007-07-100362. [DOI] [PubMed] [Google Scholar]
  • 183.Liu RY, Peng JL, Li YQ, Huang BJ, Lin HX, Zhou L, Luo HL, Huang W. Tumor-specific cytolysis caused by an E1B55K-attenuated adenovirus in nasopharyngeal carcinoma is augmented by cisplatin. Anat Rec (Hoboken) 2013;296:1833–41. doi: 10.1002/ar.22813. [DOI] [PubMed] [Google Scholar]
  • 184.Galluzzi L, Senovilla L, Vitale I, Michels J, Martins I, Kepp O, Castedo M, Kroemer G. Molecular mechanisms of cisplatin resistance. Oncogene. 2012;31:1869–83. doi: 10.1038/onc.2011.384. [DOI] [PubMed] [Google Scholar]
  • 185.Galluzzi L, Morselli E, Vitale I, Kepp O, Senovilla L, Criollo A, Servant N, Paccard C, Hupé P, Robert T, et al. miR-181a and miR-630 regulate cisplatin-induced cancer cell death. Cancer Res. 2010;70:1793–803. doi: 10.1158/0008-5472.CAN-09-3112. [DOI] [PubMed] [Google Scholar]
  • 186.Galluzzi L, Vitale I, Senovilla L, Olaussen KA, Pinna G, Eisenberg T, Goubar A, Martins I, Michels J, Kratassiouk G, et al. Prognostic impact of vitamin B6 metabolism in lung cancer. Cell Rep. 2012;2:257–69. doi: 10.1016/j.celrep.2012.06.017. [DOI] [PubMed] [Google Scholar]
  • 187.Cody JJ, Markert JM, Hurst DR. Histone deacetylase inhibitors improve the replication of oncolytic herpes simplex virus in breast cancer cells. PLoS One. 2014;9:e92919. doi: 10.1371/journal.pone.0092919. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Nomura M, Shimbo T, Miyamoto Y, Fukuzawa M, Kaneda Y. 13-Cis retinoic acid can enhance the antitumor activity of non-replicating Sendai virus particle against neuroblastoma. Cancer Sci. 2013;104:238–44. doi: 10.1111/cas.12063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Carlson LM, De Geer A, Sveinbjørnsson B, Orrego A, Martinsson T, Kogner P, Levitskaya J. The microenvironment of human neuroblastoma supports the activation of tumor-associated T lymphocytes. Oncoimmunology. 2013;2:e23618. doi: 10.4161/onci.23618. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Gentschev I, Patil SS, Adelfinger M, Weibel S, Geissinger U, Frentzen A, Chen NG, Yu YA, Zhang Q, Ogilvie G, et al. Characterization and evaluation of a new oncolytic vaccinia virus strain LIVP6.1.1 for canine cancer therapy. Bioengineered. 2013;4:84–9. doi: 10.4161/bioe.22462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Patil SS, Gentschev I, Adelfinger M, Donat U, Hess M, Weibel S, Nolte I, Frentzen A, Szalay AA. Virotherapy of canine tumors with oncolytic vaccinia virus GLV-1h109 expressing an anti-VEGF single-chain antibody. PLoS One. 2012;7:e47472. doi: 10.1371/journal.pone.0047472. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Gentschev I, Adelfinger M, Josupeit R, Rudolph S, Ehrig K, Donat U, Weibel S, Chen NG, Yu YA, Zhang Q, et al. Preclinical evaluation of oncolytic vaccinia virus for therapy of canine soft tissue sarcoma. PLoS One. 2012;7:e37239. doi: 10.1371/journal.pone.0037239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Patil SS, Gentschev I, Nolte I, Ogilvie G, Szalay AA. Oncolytic virotherapy in veterinary medicine: current status and future prospects for canine patients. J Transl Med. 2012;10:3. doi: 10.1186/1479-5876-10-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Hwang CC, Umeki S, Kubo M, Hayashi T, Shimoda H, Mochizuki M, Maeda K, Baba K, Hiraoka H, Coffey M, et al. Oncolytic reovirus in canine mast cell tumor. PLoS One. 2013;8:e73555. doi: 10.1371/journal.pone.0073555. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Cerullo V, Koski A, Vähä-Koskela M, Hemminki A. Chapter eight--Oncolytic adenoviruses for cancer immunotherapy: data from mice, hamsters, and humans. Adv Cancer Res. 2012;115:265–318. doi: 10.1016/B978-0-12-398342-8.00008-2. [DOI] [PubMed] [Google Scholar]
  • 196.Liang M. Clinical development of oncolytic viruses in China. Curr Pharm Biotechnol. 2012;13:1852–7. doi: 10.2174/138920112800958760. [DOI] [PubMed] [Google Scholar]
  • 197.Räty JK, Pikkarainen JT, Wirth T, Ylä-Herttuala S. Gene therapy: the first approved gene-based medicines, molecular mechanisms and clinical indications. Curr Mol Pharmacol. 2008;1:13–23. doi: 10.2174/1874467210801010013. [DOI] [PubMed] [Google Scholar]
  • 198.Vacchelli E, Eggermont A, Sautès-Fridman C, Galon J, Zitvogel L, Kroemer G, Galluzzi L. Trial watch: Oncolytic viruses for cancer therapy. Oncoimmunology. 2013;2:e24612. doi: 10.4161/onci.24612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Aranda F, Vacchelli E, Obrist F, Eggermont A, Fridman WH, Galon J, et al. Trial Watch: Adoptive cell transfer for anticancer immunotherapy. OncoImmunology. 2014;3:e28344. doi: 10.4161/onci.28344. Forthcoming. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Vacchelli E, Vitale I, Tartour E, Eggermont A, Sautès-Fridman C, Galon J, Zitvogel L, Kroemer G, Galluzzi L. Trial Watch: Anticancer radioimmunotherapy. Oncoimmunology. 2013;2:e25595. doi: 10.4161/onci.25595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Semeraro M, Vacchelli E, Eggermont A, Galon J, Zitvogel L, Kroemer G, Galluzzi L. Trial Watch: Lenalidomide-based immunochemotherapy. Oncoimmunology. 2013;2:e26494. doi: 10.4161/onci.26494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Raki M, Kanerva A, Ristimaki A, Desmond RA, Chen DT, Ranki T, Sarkioja M, Kangasniemi L, Hemminki A. Combination of gemcitabine and Ad5/3-Delta24, a tropism modified conditionally replicating adenovirus, for the treatment of ovarian cancer. Gene Ther. 2005;12:1198–205. doi: 10.1038/sj.gt.3302517. [DOI] [PubMed] [Google Scholar]
  • 203.Kanerva A, Zinn KR, Chaudhuri TR, Lam JT, Suzuki K, Uil TG, Hakkarainen T, Bauerschmitz GJ, Wang M, Liu B, et al. Enhanced therapeutic efficacy for ovarian cancer with a serotype 3 receptor-targeted oncolytic adenovirus. Mol Ther. 2003;8:449–58. doi: 10.1016/S1525-0016(03)00200-4. [DOI] [PubMed] [Google Scholar]
  • 204.Kim KH, Dmitriev IP, Saddekni S, Kashentseva EA, Harris RD, Aurigemma R, Bae S, Singh KP, Siegal GP, Curiel DT, et al. A phase I clinical trial of Ad5/3-Δ24, a novel serotype-chimeric, infectivity-enhanced, conditionally-replicative adenovirus (CRAd), in patients with recurrent ovarian cancer. Gynecol Oncol. 2013;130:518–24. doi: 10.1016/j.ygyno.2013.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Bramante S, Koski A, Kipar A, Diaconu I, Liikanen I, Hemminki O, Vassilev L, Parviainen S, Cerullo V, Pesonen SK, et al. Serotype chimeric oncolytic adenovirus coding for GM-CSF for treatment of sarcoma in rodents and humans. Int J Cancer. 2013 doi: 10.1002/ijc.28696. Forthcoming. [DOI] [PubMed] [Google Scholar]
  • 206.Fridman WH, Galon J, Pagès F, Tartour E, Sautès-Fridman C, Kroemer G. Prognostic and predictive impact of intra- and peritumoral immune infiltrates. Cancer Res. 2011;71:5601–5. doi: 10.1158/0008-5472.CAN-11-1316. [DOI] [PubMed] [Google Scholar]
  • 207.Zitvogel L, Kepp O, Kroemer G. Immune parameters affecting the efficacy of chemotherapeutic regimens. Nat Rev Clin Oncol. 2011;8:151–60. doi: 10.1038/nrclinonc.2010.223. [DOI] [PubMed] [Google Scholar]
  • 208.Kroemer G, Galluzzi L, Kepp O, Zitvogel L. Immunogenic cell death in cancer therapy. Annu Rev Immunol. 2013;31:51–72. doi: 10.1146/annurev-immunol-032712-100008. [DOI] [PubMed] [Google Scholar]
  • 209.Kepp O, Galluzzi L, Martins I, Schlemmer F, Adjemian S, Michaud M, Sukkurwala AQ, Menger L, Zitvogel L, Kroemer G. Molecular determinants of immunogenic cell death elicited by anticancer chemotherapy. Cancer Metastasis Rev. 2011;30:61–9. doi: 10.1007/s10555-011-9273-4. [DOI] [PubMed] [Google Scholar]
  • 210.Honeychurch J, Dive C, Illidge TM. Synchronous apoptosis in established tumors leads to the induction of adaptive immunity. Oncoimmunology. 2013;2:e24501. doi: 10.4161/onci.24501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Senovilla L, Vitale I, Martins I, Kepp O, Galluzzi L, Zitvogel L, Castedo M, Kroemer G. An anticancer therapy-elicited immunosurveillance system that eliminates tetraploid cells. Oncoimmunology. 2013;2:e22409. doi: 10.4161/onci.22409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Galluzzi L, Kepp O, Kroemer G. Immunogenic cell death in radiation therapy. Oncoimmunology. 2013;2:e26536. doi: 10.4161/onci.26536. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Ma Y, Adjemian S, Yang H, Catani JP, Hannani D, Martins I, Michaud M, Kepp O, Sukkurwala AQ, Vacchelli E, et al. ATP-dependent recruitment, survival and differentiation of dendritic cell precursors in the tumor bed after anticancer chemotherapy. Oncoimmunology. 2013;2:e24568. doi: 10.4161/onci.24568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Ma Y, Adjemian S, Mattarollo SR, Yamazaki T, Aymeric L, Yang H, Portela Catani JP, Hannani D, Duret H, Steegh K, et al. Anticancer chemotherapy-induced intratumoral recruitment and differentiation of antigen-presenting cells. Immunity. 2013;38:729–41. doi: 10.1016/j.immuni.2013.03.003. [DOI] [PubMed] [Google Scholar]
  • 215.Park BH, Hwang T, Liu TC, Sze DY, Kim JS, Kwon HC, Oh SY, Han SY, Yoon JH, Hong SH, et al. Use of a targeted oncolytic poxvirus, JX-594, in patients with refractory primary or metastatic liver cancer: a phase I trial. Lancet Oncol. 2008;9:533–42. doi: 10.1016/S1470-2045(08)70107-4. [DOI] [PubMed] [Google Scholar]
  • 216.Breitbach CJ, Burke J, Jonker D, Stephenson J, Haas AR, Chow LQ, Nieva J, Hwang TH, Moon A, Patt R, et al. Intravenous delivery of a multi-mechanistic cancer-targeted oncolytic poxvirus in humans. Nature. 2011;477:99–102. doi: 10.1038/nature10358. [DOI] [PubMed] [Google Scholar]
  • 217.Hwang TH, Moon A, Burke J, Ribas A, Stephenson J, Breitbach CJ, Daneshmand M, De Silva N, Parato K, Diallo JS, et al. A mechanistic proof-of-concept clinical trial with JX-594, a targeted multi-mechanistic oncolytic poxvirus, in patients with metastatic melanoma. Mol Ther. 2011;19:1913–22. doi: 10.1038/mt.2011.132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Markert JM, Medlock MD, Rabkin SD, Gillespie GY, Todo T, Hunter WD, Palmer CA, Feigenbaum F, Tornatore C, Tufaro F, 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. doi: 10.1038/sj.gt.3301205. [DOI] [PubMed] [Google Scholar]
  • 219.Todo T, Martuza RL, Rabkin SD, Johnson PA. Oncolytic herpes simplex virus vector with enhanced MHC class I presentation and tumor cell killing. Proc Natl Acad Sci U S A. 2001;98:6396–401. doi: 10.1073/pnas.101136398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Markert JM, Liechty PG, Wang W, Gaston S, Braz E, Karrasch M, Nabors LB, Markiewicz M, Lakeman AD, Palmer CA, et al. Phase Ib trial of mutant herpes simplex virus G207 inoculated pre-and post-tumor resection for recurrent GBM. Mol Ther. 2009;17:199–207. doi: 10.1038/mt.2008.228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Markert JM, Razdan SN, Kuo HC, Cantor A, Knoll A, Karrasch M, Nabors LB, Markiewicz M, Agee BS, Coleman JM, et al. A Phase 1 Trial of Oncolytic HSV-1, G207, Given in Combination With Radiation for Recurrent GBM Demonstrates Safety and Radiographic Responses. Mol Ther. 2014 doi: 10.1038/mt.2014.22. Forthcoming. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Liu BL, Robinson M, Han ZQ, Branston RH, English C, Reay P, McGrath Y, Thomas SK, Thornton M, Bullock P, et al. ICP34.5 deleted herpes simplex virus with enhanced oncolytic, immune stimulating, and anti-tumour properties. Gene Ther. 2003;10:292–303. doi: 10.1038/sj.gt.3301885. [DOI] [PubMed] [Google Scholar]
  • 223.Galluzzi L, Lugli E. Cancer immunotherapy turns viral. Oncoimmunology. 2013;2:e24802. doi: 10.4161/onci.24802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Pol JG, Marguerie M, Arulanandam R, Bell JC, Lichty BD. Panorama from the oncolytic virotherapy summit. Mol Ther. 2013;21:1814–8. doi: 10.1038/mt.2013.207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Kelly K, Nawrocki S, Mita A, Coffey M, Giles FJ, Mita M. Reovirus-based therapy for cancer. Expert Opin Biol Ther. 2009;9:817–30. doi: 10.1517/14712590903002039. [DOI] [PubMed] [Google Scholar]
  • 226.Morris DG, Feng X, DiFrancesco LM, Fonseca K, Forsyth PA, Paterson AH, Coffey MC, Thompson B. REO-001: A phase I trial of percutaneous intralesional administration of reovirus type 3 dearing (Reolysin®) in patients with advanced solid tumors. Invest New Drugs. 2013;31:696–706. doi: 10.1007/s10637-012-9865-z. [DOI] [PubMed] [Google Scholar]
  • 227.Hirvinen M, Heiskanen R, Oksanen M, Pesonen S, Liikanen I, Joensuu T, Kanerva A, Cerullo V, Hemminki A. Fc-gamma receptor polymorphisms as predictive and prognostic factors in patients receiving oncolytic adenovirus treatment. J Transl Med. 2013;11:193. doi: 10.1186/1479-5876-11-193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Hemminki O, Immonen R, Närväinen J, Kipar A, Paasonen J, Jokivarsi KT, Yli-Ollila H, Soininen P, Partanen K, Joensuu T, et al. In vivo magnetic resonance imaging and spectroscopy identifies oncolytic adenovirus responders. Int J Cancer. 2014;134:2878–90. doi: 10.1002/ijc.28615. [DOI] [PubMed] [Google Scholar]
  • 229.Haabeth OA, Bogen B, Corthay A. A model for cancer-suppressive inflammation. Oncoimmunology. 2012;1:1146–55. doi: 10.4161/onci.21542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Koski A, Ahtinen H, Liljenback H, Roivainen A, Koskela A, Oksanen M, Partanen K, Laasonen L, Kairemo K, Joensuu T, et al. [(18)F]-fluorodeoxyglucose positron emission tomography and computed tomography in response evaluation of oncolytic adenovirus treatments of patients with advanced cancer. Hum Gene Ther. 2013;24:1029–41. doi: 10.1089/hum.2013.123. [DOI] [PubMed] [Google Scholar]
  • 231.Beug ST, Tang VA, LaCasse EC, Cheung HH, Beauregard CE, Brun J, Nuyens JP, Earl N, St-Jean M, Holbrook J, et al. Smac mimetics and innate immune stimuli synergize to promote tumor death. Nat Biotechnol. 2014;32:182–90. doi: 10.1038/nbt.2806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Ma Y, Yamazaki T, Yang H, Kepp O, Galluzzi L, Zitvogel L, Smyth MJ, Kroemer G. Tumor necrosis factor is dispensable for the success of immunogenic anticancer chemotherapy. Oncoimmunology. 2013;2:e24786. doi: 10.4161/onci.24786. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Becker C, Bopp T, Steinbrink K. Interferon α interferes with immunological tolerance. Oncoimmunology. 2013;2:e27528. doi: 10.4161/onci.27528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Smits EL, Anguille S, Berneman ZN. Interferon α may be back on track to treat acute myeloid leukemia. Oncoimmunology. 2013;2:e23619. doi: 10.4161/onci.23619. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.James BR, Griffith TS. Activation of systemic antitumor immunity via TRAIL-induced apoptosis. Oncoimmunology. 2012;1:1178–80. doi: 10.4161/onci.20638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Munich S, Sobo-Vujanovic A, Buchser WJ, Beer-Stolz D, Vujanovic NL. Dendritic cell exosomes directly kill tumor cells and activate natural killer cells via TNF superfamily ligands. Oncoimmunology. 2012;1:1074–83. doi: 10.4161/onci.20897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Zamarin D, Holmgaard RB, Subudhi SK, Park JS, Mansour M, Palese P, Merghoub T, Wolchok JD, Allison JP. Localized oncolytic virotherapy overcomes systemic tumor resistance to immune checkpoint blockade immunotherapy. Sci Transl Med. 2014;6:226ra32. doi: 10.1126/scitranslmed.3008095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Aranda F, Vacchelli E, Eggermont A, Galon J, Fridman WH, Zitvogel L, Kroemer G, Galluzzi L. Trial Watch: Immunostimulatory monoclonal antibodies in cancer therapy. Oncoimmunology. 2014;3:e27297. doi: 10.4161/onci.27297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239.Vacchelli E, Eggermont A, Galon J, Sautès-Fridman C, Zitvogel L, Kroemer G, Galluzzi L. Trial watch: Monoclonal antibodies in cancer therapy. Oncoimmunology. 2013;2:e22789. doi: 10.4161/onci.22789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Mavilio D, Lugli E. Inhibiting the inhibitors: Checkpoints blockade in solid tumors. Oncoimmunology. 2013;2:e26535. doi: 10.4161/onci.26535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Carew JS, Espitia CM, Zhao W, Kelly KR, Coffey M, Freeman JW, Nawrocki ST. Reolysin is a novel reovirus-based agent that induces endoplasmic reticular stress-mediated apoptosis in pancreatic cancer. Cell Death Dis. 2013;4:e728. doi: 10.1038/cddis.2013.259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Donnelly O, Ilett E, Kottke T, Thompson J, Vile R, Melcher A. Cytokine-enhanced intravenous oncolytic virotherapy. Lancet. 2014;383:S42. doi: 10.1016/S0140-6736(14)60305-6. Forthcoming. [DOI] [Google Scholar]
  • 243.Bauzon M, Hermiston TW. Oncolytic viruses: the power of directed evolution. Adv Virol. 2012;2012:586389. doi: 10.1155/2012/586389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Kuhn I, Harden P, Bauzon M, Chartier C, Nye J, Thorne S, Reid T, Ni S, Lieber A, Fisher K, et al. Directed evolution generates a novel oncolytic virus for the treatment of colon cancer. PLoS One. 2008;3:e2409. doi: 10.1371/journal.pone.0002409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Li H, Peng KW, Russell SJ. Oncolytic measles virus encoding thyroidal sodium iodide symporter for squamous cell cancer of the head and neck radiovirotherapy. Hum Gene Ther. 2012;23:295–301. doi: 10.1089/hum.2011.128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Opyrchal M, Allen C, Iankov I, Aderca I, Schroeder M, Sarkaria J, Galanis E. Effective radiovirotherapy for malignant gliomas by using oncolytic measles virus strains encoding the sodium iodide symporter (MV-NIS) Hum Gene Ther. 2012;23:419–27. doi: 10.1089/hum.2011.158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Penheiter AR, Wegman TR, Classic KL, Dingli D, Bender CE, Russell SJ, Carlson SK. Sodium iodide symporter (NIS)-mediated radiovirotherapy for pancreatic cancer. AJR Am J Roentgenol. 2010;195:341–9. doi: 10.2214/AJR.09.3672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 248.Liu C, Russell SJ, Peng KW. Systemic therapy of disseminated myeloma in passively immunized mice using measles virus-infected cell carriers. Mol Ther. 2010;18:1155–64. doi: 10.1038/mt.2010.43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Guedan S, Rojas JJ, Gros A, Mercade E, Cascallo M, Alemany R. Hyaluronidase expression by an oncolytic adenovirus enhances its intratumoral spread and suppresses tumor growth. Mol Ther. 2010;18:1275–83. doi: 10.1038/mt.2010.79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Huang TT, Hlavaty J, Ostertag D, Espinoza FL, Martin B, Petznek H, Rodriguez-Aguirre M, Ibañez CE, Kasahara N, Gunzburg W, et al. Toca 511 gene transfer and 5-fluorocytosine in combination with temozolomide demonstrates synergistic therapeutic efficacy in a temozolomide-sensitive glioblastoma model. Cancer Gene Ther. 2013;20:544–51. doi: 10.1038/cgt.2013.51. [DOI] [PubMed] [Google Scholar]
  • 251.Perez OD, Logg CR, Hiraoka K, Diago O, Burnett R, Inagaki A, Jolson D, Amundson K, Buckley T, Lohse D, et al. Design and selection of Toca 511 for clinical use: modified retroviral replicating vector with improved stability and gene expression. Mol Ther. 2012;20:1689–98. doi: 10.1038/mt.2012.83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Ostertag D, Amundson KK, Lopez Espinoza F, Martin B, Buckley T, Galvão da Silva AP, Lin AH, Valenta DT, Perez OD, Ibañez CE, et al. Brain tumor eradication and prolonged survival from intratumoral conversion of 5-fluorocytosine to 5-fluorouracil using a nonlytic retroviral replicating vector. Neuro Oncol. 2012;14:145–59. doi: 10.1093/neuonc/nor199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Skelding KA, Barry RD, Shafren DR. Enhanced oncolysis mediated by Coxsackievirus A21 in combination with doxorubicin hydrochloride. Invest New Drugs. 2012;30:568–81. doi: 10.1007/s10637-010-9614-0. [DOI] [PubMed] [Google Scholar]
  • 254.Skelding KA, Barry RD, Shafren DR. Systemic targeting of metastatic human breast tumor xenografts by Coxsackievirus A21. Breast Cancer Res Treat. 2009;113:21–30. doi: 10.1007/s10549-008-9899-2. [DOI] [PubMed] [Google Scholar]
  • 255.Au GG, Lindberg AM, Barry RD, Shafren DR. Oncolysis of vascular malignant human melanoma tumors by Coxsackievirus A21. Int J Oncol. 2005;26:1471–6. doi: 10.3892/ijo.26.6.1471. [DOI] [PubMed] [Google Scholar]
  • 256.Shafren DR, Au GG, Nguyen T, Newcombe NG, Haley ES, Beagley L, Johansson ES, Hersey P, Barry RD. Systemic therapy of malignant human melanoma tumors by a common cold-producing enterovirus, coxsackievirus a21. Clin Cancer Res. 2004;10:53–60. doi: 10.1158/1078-0432.CCR-0690-3. [DOI] [PubMed] [Google Scholar]
  • 257.Galluzzi L. New immunotherapeutic paradigms for castration-resistant prostate cancer. OncoImmunology. 2013;2:e26084. doi: 10.4161/onci.26084. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Ramesh N, Ge Y, Ennist DL, Zhu M, Mina M, Ganesh S, Reddy PS, Yu DC. CG0070, a conditionally replicating granulocyte macrophage colony-stimulating factor--armed oncolytic adenovirus for the treatment of bladder cancer. Clin Cancer Res. 2006;12:305–13. doi: 10.1158/1078-0432.CCR-05-1059. [DOI] [PubMed] [Google Scholar]
  • 259.Vacchelli E, Eggermont A, Sautès-Fridman C, Galon J, Zitvogel L, Kroemer G, Galluzzi L. Trial Watch: Toll-like receptor agonists for cancer therapy. Oncoimmunology. 2013;2:e25238. doi: 10.4161/onci.25238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 260.Vacchelli E, Galluzzi L, Eggermont A, Fridman WH, Galon J, Sautès-Fridman C, Tartour E, Zitvogel L, Kroemer G. Trial watch: FDA-approved Toll-like receptor agonists for cancer therapy. Oncoimmunology. 2012;1:894–907. doi: 10.4161/onci.20931. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 261.Alonso MM, Cascallo M, Gomez-Manzano C, Jiang H, Bekele BN, Perez-Gimenez A, Lang FF, Piao Y, Alemany R, Fueyo J. ICOVIR-5 shows E2F1 addiction and potent antiglioma effect in vivo. Cancer Res. 2007;67:8255–63. doi: 10.1158/0008-5472.CAN-06-4675. [DOI] [PubMed] [Google Scholar]
  • 262.Cascallo M, Alonso MM, Rojas JJ, Perez-Gimenez A, Fueyo J, Alemany R. Systemic toxicity-efficacy profile of ICOVIR-5, a potent and selective oncolytic adenovirus based on the pRB pathway. Mol Ther. 2007;15:1607–15. doi: 10.1038/sj.mt.6300239. [DOI] [PubMed] [Google Scholar]
  • 263.Alonso MM, Gomez-Manzano C, Jiang H, Bekele NB, Piao Y, Yung WK, Alemany R, Fueyo J. Combination of the oncolytic adenovirus ICOVIR-5 with chemotherapy provides enhanced anti-glioma effect in vivo. Cancer Gene Ther. 2007;14:756–61. doi: 10.1038/sj.cgt.7701067. [DOI] [PubMed] [Google Scholar]
  • 264.Alonso MM, Jiang H, Yokoyama T, Xu J, Bekele NB, Lang FF, Kondo S, Gomez-Manzano C, Fueyo J. Delta-24-RGD in combination with RAD001 induces enhanced anti-glioma effect via autophagic cell death. Mol Ther. 2008;16:487–93. doi: 10.1038/sj.mt.6300400. [DOI] [PubMed] [Google Scholar]
  • 265.Jiang H, Gomez-Manzano C, Aoki H, Alonso MM, Kondo S, McCormick F, Xu J, Kondo Y, Bekele BN, Colman H, et al. Examination of the therapeutic potential of Delta-24-RGD in brain tumor stem cells: role of autophagic cell death. J Natl Cancer Inst. 2007;99:1410–4. doi: 10.1093/jnci/djm102. [DOI] [PubMed] [Google Scholar]
  • 266.Fueyo J, Alemany R, Gomez-Manzano C, Fuller GN, Khan A, Conrad CA, Liu TJ, Jiang H, Lemoine MG, Suzuki K, et al. Preclinical characterization of the antiglioma activity of a tropism-enhanced adenovirus targeted to the retinoblastoma pathway. J Natl Cancer Inst. 2003;95:652–60. doi: 10.1093/jnci/95.9.652. [DOI] [PubMed] [Google Scholar]
  • 267.MacKie RM, Stewart B, Brown SM. Intralesional injection of herpes simplex virus 1716 in metastatic melanoma. Lancet. 2001;357:525–6. doi: 10.1016/S0140-6736(00)04048-4. [DOI] [PubMed] [Google Scholar]
  • 268.Coukos G, Makrigiannakis A, Kang EH, Caparelli D, Benjamin I, Kaiser LR, Rubin SC, Albelda SM, Molnar-Kimber KL. Use of carrier cells to deliver a replication-selective herpes simplex virus-1 mutant for the intraperitoneal therapy of epithelial ovarian cancer. Clin Cancer Res. 1999;5:1523–37. [PubMed] [Google Scholar]
  • 269.Kucharczuk JC, Randazzo B, Chang MY, Amin KM, Elshami AA, Sterman DH, Rizk NP, Molnar-Kimber KL, Brown SM, MacLean AR, et al. Use of a “replication-restricted” herpes virus to treat experimental human malignant mesothelioma. Cancer Res. 1997;57:466–71. [PubMed] [Google Scholar]
  • 270.Randazzo BP, Kucharczuk JC, Litzky LA, Kaiser LR, Brown SM, MacLean A, Albelda SM, Fraser NW. Herpes simplex 1716--an ICP 34.5 mutant--is severely replication restricted in human skin xenografts in vivo. Virology. 1996;223:392–5. doi: 10.1006/viro.1996.0493. [DOI] [PubMed] [Google Scholar]
  • 271.Poirier JT, Dobromilskaya I, Moriarty WF, Peacock CD, Hann CL, Rudin CM. Selective tropism of Seneca Valley virus for variant subtype small cell lung cancer. J Natl Cancer Inst. 2013;105:1059–65. doi: 10.1093/jnci/djt130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Reddy PS, Burroughs KD, Hales LM, Ganesh S, Jones BH, Idamakanti N, Hay C, Li SS, Skele KL, Vasko AJ, et al. Seneca Valley virus, a systemically deliverable oncolytic picornavirus, and the treatment of neuroendocrine cancers. J Natl Cancer Inst. 2007;99:1623–33. doi: 10.1093/jnci/djm198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Sieben M, Schäfer P, Dinsart C, Galle PR, Moehler M. Activation of the human immune system via toll-like receptors by the oncolytic parvovirus H-1. Int J Cancer. 2013;132:2548–56. doi: 10.1002/ijc.27938. [DOI] [PubMed] [Google Scholar]
  • 274.Guillerme JB, Boisgerault N, Roulois D, Ménager J, Combredet C, Tangy F, Fonteneau JF, Gregoire M. Measles virus vaccine-infected tumor cells induce tumor antigen cross-presentation by human plasmacytoid dendritic cells. Clin Cancer Res. 2013;19:1147–58. doi: 10.1158/1078-0432.CCR-12-2733. [DOI] [PubMed] [Google Scholar]
  • 275.Cerullo V, Diaconu I, Romano V, Hirvinen M, Ugolini M, Escutenaire S, Holm SL, Kipar A, Kanerva A, Hemminki A. An oncolytic adenovirus enhanced for toll-like receptor 9 stimulation increases antitumor immune responses and tumor clearance. Mol Ther. 2012;20:2076–86. doi: 10.1038/mt.2012.137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276.Xia M, Gonzalez P, Li C, Meng G, Jiang A, Wang H, Gao Q, Debatin KM, Beltinger C, Wei J. Mitophagy Enhances Oncolytic Measles Virus Replication by Mitigating DDX58/RIG-I-Like Receptor Signaling. J Virol. 2014;88:5152–64. doi: 10.1128/JVI.03851-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Kaneda Y. The RIG-I/MAVS signaling pathway in cancer cell-selective apoptosis. Oncoimmunology. 2013;2:e23566. doi: 10.4161/onci.23566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Berchtold S, Lampe J, Weiland T, Smirnow I, Schleicher S, Handgretinger R, Kopp HG, Reiser J, Stubenrauch F, Mayer N, et al. Innate immune defense defines susceptibility of sarcoma cells to measles vaccine virus-based oncolysis. J Virol. 2013;87:3484–501. doi: 10.1128/JVI.02106-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 279.Matsushima-Miyagi T, Hatano K, Nomura M, Li-Wen L, Nishikawa T, Saga K, Shimbo T, Kaneda Y. TRAIL and Noxa are selectively upregulated in prostate cancer cells downstream of the RIG-I/MAVS signaling pathway by nonreplicating Sendai virus particles. Clin Cancer Res. 2012;18:6271–83. doi: 10.1158/1078-0432.CCR-12-1595. [DOI] [PubMed] [Google Scholar]
  • 280.Galluzzi L, Kepp O, Kroemer G. Mitochondria: master regulators of danger signalling. Nat Rev Mol Cell Biol. 2012;13:780–8. doi: 10.1038/nrm3479. [DOI] [PubMed] [Google Scholar]
  • 281.Aranda F, Vacchelli E, Eggermont A, Galon J, Sautès-Fridman C, Tartour E, Zitvogel L, Kroemer G, Galluzzi L. Trial Watch: Peptide vaccines in cancer therapy. Oncoimmunology. 2013;2:e26621. doi: 10.4161/onci.26621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 282.Rommelfanger DM, Compte M, Diaz RM, Ilett E, Alvarez-Vallina L, Thompson JM, Kottke TJ, Melcher A, Vile RG. The efficacy versus toxicity profile of combination virotherapy and TLR immunotherapy highlights the danger of administering TLR agonists to oncolytic virus-treated mice. Mol Ther. 2013;21:348–57. doi: 10.1038/mt.2012.204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.Vacchelli E, Senovilla L, Eggermont A, Fridman WH, Galon J, Zitvogel L, Kroemer G, Galluzzi L. Trial watch: Chemotherapy with immunogenic cell death inducers. Oncoimmunology. 2013;2:e23510. doi: 10.4161/onci.23510. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Hasegawa N, Abei M, Yokoyama KK, Fukuda K, Seo E, Kawashima R, Nakano Y, Yamada T, Nakade K, Hamada H, et al. Cyclophosphamide enhances antitumor efficacy of oncolytic adenovirus expressing uracil phosphoribosyltransferase (UPRT) in immunocompetent Syrian hamsters. Int J Cancer. 2013;133:1479–88. doi: 10.1002/ijc.28132. [DOI] [PubMed] [Google Scholar]
  • 285.Ziccheddu G, Proietti E, Moschella F. The Janus face of cyclophosphamide: A sterile inflammatory response that potentiates cancer immunotherapy. Oncoimmunology. 2013;2:e25789. doi: 10.4161/onci.25789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 286.Kottke T, Chester J, Ilett E, Thompson J, Diaz R, Coffey M, Selby P, Nuovo G, Pulido J, Mukhopadhyay D, et al. Precise scheduling of chemotherapy primes VEGF-producing tumors for successful systemic oncolytic virotherapy. Mol Ther. 2011;19:1802–12. doi: 10.1038/mt.2011.147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 287.Vacchelli E, Prada N, Kepp O, Galluzzi L. Current trends of anticancer immunochemotherapy. Oncoimmunology. 2013;2:e25396. doi: 10.4161/onci.25396. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Oncoimmunology are provided here courtesy of Taylor & Francis

RESOURCES