6,609
Views
88
CrossRef citations to date
0
Altmetric
Review

Oncolytic viruses: From bench to bedside with a focus on safety

, , &
Pages 1573-1584 | Received 16 Jan 2015, Accepted 29 Mar 2015, Published online: 06 Jul 2015

Abstract

Oncolytic viruses are a relatively new class of anti-cancer immunotherapy agents. Several viruses have undergone evaluation in clinical trials in the last decades, and the first agent is about to be approved to be used as a novel cancer therapy modality. In the current review, an overview is presented on recent (pre)clinical developments in the field of oncolytic viruses that have previously been or currently are being evaluated in clinical trials. Special attention is given to possible safety issues like toxicity, environmental shedding, mutation and reversion to wildtype virus.

Introduction

Oncolytic viruses (OVs), reported first halfway the previous century, have undergone a tremendous evolution from anecdotal experimental and clinical efficacy to state-of-the-art clinical trials employing recombinant viruses in the last decade. With the advent of reverse genetics techniques, modifications attributing to efficacy and safety have marked the introduction of new generations of recombinant OVs. Most recent developments have focused on conditional replication in tumor cells, expression of (therapeutic) transgenes as well as targeting and/or delivery of OVs.

In this review, an overview is given of the current state of affairs concerning OVs that are being developed toward clinical trials or that are used in clinical trials. Besides information on (pre)clinical efficacy, special focus is given to the safety of these new agents, specifically toxicity, environmental shedding and mutation rates or reversion to wild type virus. A summarizing overview is presented in .

Table 1. Characteristics of oncolytic virus platforms

Safety of OVs primarily relates to the toxicity of the administration, especially when thinking about high dose intravenous application. In addition, environmental shedding of infectious viruses is also of concern, not only for perceived safety, but also for regulatory purposes. To this end, in general, OVs should be generated which preferably are tumor-specific and have low to no shedding upon (systemic) administration.

Family Herpesviridae: Herpes Simplex Virus 1 (HSV)

HSV-1 causes herpes labialis (cold sores) in humans. HSV-1 was one of the first viruses to be developed into a recombinant oncolytic virotherapeutic vector. The large HSV genome is easy to manipulate and allows insertions of multiple additional transgenes. Furthermore, HSV infects and replicates in most tumor cell types and spreads throughout the tumor. If needed, viral replication can be hampered with anti-HSV medication (Acyclovir). Because HSV is neurotropic and causes a latent infection, most genetic modifications of oncolytic HSV (oHSV) have first focused on this potential safety issue.

To increase safety, tumor-specific oHSVs have been generated using 3 main strategies, as reviewed earlier by Campadelli-Fiume et al.Citation1 These strategies include first attenuation by means of conditional replication in tumor cells through deletion of viral genes that are essential for viral replication in non-dividing cells (UL39), counteract the PKR response (γ1-34.5) or contribute to immune evasion (α47). Secondly, to increase the oncolytic efficacy (and often co-incidentally improve safety), oHSVs have been armed with immune stimulatory genes to boost local cytotoxic immune responses or other therapy enhancing transgenes. Thirdly, targeting by tropism or transcription specificity has been applied to limit virus infection even further to only tumor cells.

First generation oHSVs harboring the aforementioned genetic deletions have shown to be safe regarding toxicity based on their attenuation in normal cells. However, they are also attenuated in tumor cells and thus less cytotoxic. Recent strategies have focused on improving the targeting of less attenuated oHSVs by changing tropism or transcription specificity. Glycoprotein D is the receptor binding protein of HSV-1, and fusion of this glycoprotein with a heterologous ligand can retarget the virus to the tumor-specific receptor of choice. This tumor specific targeting is enhanced by detargeting the normal receptor.Citation2,3 Examples include IL‐13Rα2, HER2, and EGF-R.Citation4-6 Using transcriptional targeting, tumor specificity has been enhanced by placing viral genes under the control of tumor-specific promoters. A promising example is rQNestin34.5, which has the expression of the γ1-34.5 gene under the control of the glioma-specific nestin promoter, which restores viral replication and cytotoxicity only in glioma tumors.Citation7

Several early generation oHSVs (talimogene laherparepvec, HSV1716, NV1020, G207, M032, rRp450 and others) have already been evaluated in clinical trials.Citation1 Most of these trials have demonstrated a good safety profile and treatment benefits were also observed. Talimogene laherparepvec (oHSV with deletions in γ1-34.5 and α47, armed with GM-CSF) has recently undergone evaluation in a phase III clinical trial in patients with advanced or metastatic melanoma. Probably this will be the first oncolytic virus to obtain FDA approval, while a marketing authorization application for the European Union has just been submitted to the EMA.Citation8

Preclinical evaluation employing intra-organ (brain or prostate) injection with oHSV in non-human primates demonstrated no shedding of virus, which points to limitation of oHSV replication to injection sites.Citation9,10 This was confirmed by early clinical trials in patients injected intratumorally (glioma) with oHSVs: sporadic shedding of HSV in saliva was noted, but this was shown to be wildtype virus as opposed to the injected oHSV.Citation11,12 Other studies have observed limited leakage of oHSV from injection sites up to 2 weeks post treatment, without other excreta containing viable oHSV.Citation13-15 Intra-arterial hepatic injection did not result in detectable environmental shedding either.Citation16 Thus oHSV fulfills the criteria for safety with regards to shedding. A possible concern is that an oHSV recombines with a wildtype endogenous virus. If the oHSV carries heterologous genes, the recombinant would have to arise by illegitimate recombination – an extremely rare event that cannot be replicated in vitro.Citation17 Spontaneous reversion of oHSVs with deleted viral genes to wildtype virus is not possible. However, compensatory mutations can arise, which can compromise safety, but so far resulting HSV mutants have been highly attenuated.Citation18 These compensatory mutations can be of concern when evaluating the safety of newer generation oHSVs, because more virulent oHSVs could arise.

Family Adenoviridae: Human Mastadenovirus (HAdV)

HAdVs can be associated with different diseases in humans: (upper) respiratory tract infection (mainly species B and C), conjunctivitis (species B and D) and gastroenteritis (species F and G). Sporadically, HAdVs can cause viral meningitis, encephalitis or hemorrhagic cystitis.Citation19

Because of its association with mild disease and a relatively easy to manipulate genome (as compared to other HAdV types/species), most work on HAdV as vector for (cancer) gene therapy has been done with serotype 5 of species C. HAdVs have distinct advantages as a gene transfer vector, including high transfection efficiency of cells irrespective of their growth status. The genome of HAdVs is easy to manipulate for retargeting and insertion of transgenes, and efficient production of high titer virus stocks is possible. Disadvantages include high immunogenicity of prevalent serotypes with pre-existing immunity, and transient expression of the transgene due to dilution of replication-deficient HAdV (rdHAdV) episomes upon cellular division.Citation20

In case of oncolytic HAdV vectors, replication of the virus is thought to be advantageous because of direct cancer cell killing induced by viral replication, and due to which the number of administrations needed for effective treatment can be reduced. Efforts to improve safety have been made in developing conditionally replicating HAdVs (crHAdVs), with specific and higher replication in cancer cells. Early examples are ONYX-015 (dl1520) and H101, which have a deletion of E1B-55kDa (and a deletion in E3 for H101), normally responsible for p53 binding and inactivation.Citation21 The tumor specific replication of ONYX-015 was later shown to be due to loss of E1B-55kDa-mediated late viral RNA export, rather than p53-inactivation.Citation22 In a similar approach, newer crHAdVs have been created exploiting the defects in Rb pathways in cancer cells by deleting the Rb-binding E1A-CR2 region, creating dl922–947, also known as Delta24.Citation23 Additional modifications in Delta24 have been created and successfully evaluated for oncolytic efficacy,Citation24,25 as well as oncolytic crHAdVs which target cells with an (hyper)active KRAS pathway,Citation26 or with YB-1 overexpression,Citation27 limiting crHAdV replication to cancer cells. A different approach for creating crHAdVs is using cancer- or tissue-specific promoters to limit expression of essential early HAdV genes to specific celltypes and/or tissues.Citation28,29

Like other oncolytic viruses that have undergone extensive development, crHAdVs have also been armed with transgenes, often under the control of a tissue/cancer-specific promoter. Examples include i.e. immunomodulatory, pro-apoptotic or pro-drug converting enzyme genes.Citation30-33

Despite their capacity to achieve tumor infection in animal models and in clinical trials, the therapeutic efficacy of rdHAdVs in clinical trials has been disappointing; Advexin and Cerepro have not been approved by the FDA and EMA, although a similar agent called Gendicine has been approved for cancer therapy in China.Citation34,35 The discrepancy between preclinical and clinical studies using HAdV-5 could be explained by the differences in expression of CAR in primary tumors compared to established laboratory cell lines.Citation36 In addition, off-target toxicity by sequestration in mainly the liver is a serious concern, even when HAdVs are blinded for CAR, because this can lead to serious liver damage.Citation37-39 Hexon mutations or even complete exchange of hexons have been shown to reduce liver sequestration and transduction dramatically.Citation40,41 Other strategies used to circumvent liver sequestration include PEGylation or polymer/dendrimer coating of HAdV virions, and cell-based or magnetic/liposomal nanoparticle delivery. To circumvent the limitation of low CAR expression in (tumor) cells, retargeting has also been applied to HAdVs, permitting CAR-independent infection.Citation42 The retargeting strategy can also circumvent existing humoral immunity for HAdV-5 in the general population, and contributes to the prevention of liver sequestration as described above. Other examples include conjugation with anti-knob or anti-penton/hexon antibodies or adapters with retargeting ligands, pseudotyping or xenotyping with (chimeric) fiber knobs or capsids, peptide presentation (RGD or other), Affibody targeting, knob-less HAdVs and genetically modified capsids and/or fiber knobs.Citation20 More recently, efforts have also been made to develop HAdVs fully based on other serotypes, most notably HAdV-3, or even non-human AdVs.Citation43-45 Using ‘directed evolution’ or ‘accelerated evolution’ strategies, ColoAd1 and other crHAdVs have been generated which appear to be more potent than their parental HAdV-5 based vectors.Citation46,47

A total of 458 clinical trials employing HAdV-mediated gene therapy have been reported to date. ONYX-015, H101 (Oncorine) and other first-generation oncolytic crHAdVs have gone through several phase I/II trials without relevant signs of high grade toxicity but also without significant therapeutic effects, resulting in discontinuation of further trails.Citation48 More recent clinical trials employing new generations of crHAdVs like RGD retargeted oncolytic crHAdVs,Citation20,49,50 crHAdV-5/3 chimeric vectors,Citation32,51-54 ColoAd1,Citation55 hTERT-promoter driven crHAdV-5 vector Telomelysin,Citation56 E2F-1-promoter driven CG007033, Rb-targeted crHAdV expressing hyaluronidase (VCN-01)Citation57 and crHAdV vectors expressing immunomodulating genes have shown safety (low toxicity) with some promising preliminary results.

In general, the use of early generation oncolytic crHAdVs appears to be reasonably safe with low toxicity when administered locally and at lower doses systemically. However, the development of newer generations of crHAdVs expressing transgenes, having altered capsids or different promoters can dramatically alter this perceived safety. Shedding of crHAdVs from injection sites and patient excretions, although not always reported, has been observed in several (pre) clinical trials, and increases with higher doses and systemic administration.Citation49,58-64 Shedding of HAdV vectors could result in homologous recombination between AdVs of the same subgroup, which occurs with high efficiency during growth in co-infected cultured cells, and there is evidence of recombination events in humans as well.Citation65-67 Theoretically, homologous recombination between wildtype AdVs and recombinant crHAdVs could lead to new wildtype AdVs that e.g. possess transgenes, or worse, have expanded tissue tropism due to retargeting strategies. However, to date such recombination has never been detected in any clinical trial.

Family Paramyxoviridae: Measles Virus (MeV)

MeV is highly contagious via the respiratory route and is responsible for high morbidity and mortality rates in immune naive subjects.Citation68 Large-scale vaccination programs with live-attenuated MeV have been very successful. An extensive safety record has been established for the use of vaccine strains of MeV in humans.Citation69

Most (pre) clinical research with oncolytic MeVs have used the attenuated vaccine Edmonston strain, which is perceived to be very safe in terms of toxicity.Citation70 The cancer selectivity of MeV stems from overexpression of the MeV receptor CD46 on malignant cells.Citation71 Recombinant MeV can accommodate and maintain large sizes of foreign genetic material with good genetic stability in vitro and in vivo, and MeVs expressing transgene(s) have shown good genetic stability upon passaging. Both arming and targeting strategies have been used to improve efficacy of MeV in a wide array of malignancies.Citation70

Completed and ongoing clinical trials in patients with T cell lymphoma, ovarian cancer or glioblastoma multiforme have first used wild type MeV and later recombinant MeV expressing marker genes CEA and NIS.Citation72-74 Intratumoral, intraperitoneal and intravenous administration have been reported using doses up to 109 infectious viral particles without dose limiting toxicity or MeV induced immunosuppression.Citation72-74 Although wildtype MeV can cause potentially serious disease, attenuated MeV vaccine strains like Edmonston have an excellent safety record.Citation69 In clinical trials with rMeV-CEA, no evidence was seen of shedding in sputum and urine samples of patients who were intraperitoneally injected.Citation73 Spread of oncolytic MeV in the general population is highly unlikely since most individuals in industrialized countries are immunized, although herd immunity is currently waning with declining vaccination percentages. As noted, the oncolytic MeV of choice to date has been of the Edmonston strain, which has a good safety profile without capability of causing overt disease.

Family Paramyxoviridae: Newcastle Disease Virus (NDV)

NDV is an avian virus, and as such causes no serious disease in humans.Citation75 NDV strains are categorized in 3 different groups based on the severity of the disease they cause in birds: lentogenic, mesogenic and velogenic.Citation76 NDV has been shown to be very safe with regards to toxicity in tumor models using mice or rats, even when used in high dose and injected intravenously, and NDV also appears to be safe for high dose administration in humans, with no serious adverse events noted in early clinical trials.Citation77 Several wildtype NDV strains have shown (limited) antitumor activity without major side effects in phase I–II clinical trials for patients with various types of solid cancer.Citation77

Using recombinant NDVs, the oncolytic efficacy has been improved by increasing the virulence of the virus and the expression of immunomodulating or apoptotic transgenes. In addition, tumor cells are targeted with modified attachment proteins and combinations with other treatment modalities, most recently immune checkpoint blockade.Citation78-85 Clinical trials with these improved viruses have not yet been described, but pre-clinical data indicate efficacy with low toxicity in multiple tumor models for several solid malignancies, including pancreatic adenocarcinoma.Citation78-86

Virulent NDV strains pose an environmental risk, as birds (specifically poultry) are very susceptible to infection with mesogenic or velogenic strains. A preclinical study evaluating lentogenic and mesogenic oncolytic NDV injected intravenously in non-human primates showed i.v. administration of the virus to be safe without relevant toxicity of high dosages, although relevant shedding was noted.Citation87

Family Rhabdoviridae: Vesicular Stomatitis Virus (VSV)

VSV is the causative agent of vesicular stomatitis in cattle, causing a mild fever and the formation of blister‐like lesions on the inside of the mouth, the lips, nose, hooves and udder.Citation88

Compared with other oncolytic viruses, VSV has some distinct advantages: first of all a well-studied biology with relative replication independency of cell cycle and a specific receptor. Secondly, VSV produces high virus yields in a wide range of cell types, it replicates intracytoplasmatic without risk of genomic integration, it harbors a small and easy to manipulate genome, and there is no pre-existing immunity in humans.Citation88 VSV infection in humans is generally asymptomatic and limited to people having direct contact with VSV.Citation88 A single case of VSV strain Indiana related encephalitis in humans has been reported.Citation89 VSV oncoselectivity is based largely on defective or reduced type I IFN responses in tumor cells,Citation90 although abnormal translation machinery and other cellular proteins have also been shown to play a role.Citation91,92 All 3 strategies previously described (conditional replication, arming, and targeting) have been employed to increase efficacy of VSV.Citation88 Furthermore, combination therapy has been described with different other therapies. Finally, optimizing delivery and distribution of oncolytic VSVs has been evaluated using cell-based carriers and aptamer or PEGylation of virions. Hastie & Grdzelishvili published an excellent overview of abovementioned strategies and resulting oncolytic VSVs in 2012.Citation88

A recent study in purpose-bred beagle dogs showed that a dose up to 1010 TCID50 of VSV-hIFNβ was well tolerated, with mild adverse events with the exception of one dog that received 10Citation11 TCID50 which developed severe hepatotoxicity and shock leading to euthanasia.Citation93 A following study testing VSV-hIFNβ on rhesus macaques via intrahepatic injection did not result in neurological symptoms and is considered to be safe enough to proceed into phase I clinical trials, which are currently ongoing in humans and pet dogs.Citation94,95 With regards to shedding, no VSV RNA was detected in buccal swabs taken from non-human primates after intrahepatic injection with VSV-hIFNβ.Citation94 Theoretically, VSV mutants harboring mutations in their M or G gene (making them oncoselective and abolishing neurotropism) could revert to wildtype VSV upon passaging. Also, VSVs expressing attenuating transgenes like hIFNβ can acquire mutations in this transgene, which has been shown in several studies.Citation96-98 Furthermore, oncolytic VSVs have been shown to optimize targeting to glycoproteins upon passaging in tumor cells,Citation99 and to mutate expressed transgenes to optimize replication.Citation100 These examples have strangely not been perceived as a safety problem, but should be taken into account in future (pre)clinical trials.

Family Picornaviridae: Coxsackievirus (CVA)

Coxsackieviruses can be divided into 2 groups (A and B) based on their pathogenicity in mice. The best known example of CVA-related human disease is hand, foot and mouth disease, caused by CVA-16.Citation101 CVA-21 causes upper respiratory tract infections in humans, and it is considered one of the ‘common cold’ viruses.Citation102 Similarly to rhinoviruses, CVA-21 binds to ICAM-1 and additionally needs DAF-attachment for productive viral infection.Citation103 Given that ICAM-1 and DAF are overexpressed in melanoma cells, efforts to evaluate the oncolytic potential of wildtype CVA-21 (and other coxsackieviruses) have mainly focused on this disease.Citation104

Currently ongoing phase I/II clinical trials employing intratumoral injection of wildtype CVA-21 (CAVATAK) in Australian patients with advanced melanoma are showing promising preliminary results.Citation105 All (pre) clinical work so far has been conducted with wildtype CVA, while no progress has been made regarding conditional replication, arming or targeting.

Clinical trials thus far have not led to serious adverse events. No information is available regarding shedding. When considering non- or low-pathogenic coxsackieviruses for oncolytic virotherapy, environmental risks are considered to be low. However, when using viruses that do cause (severe) disease in humans, care should be taken to evaluate and/or attenuate these new vectors.

Family Picornaviridae: Poliovirus (PV)

The vast majority of PV infections remain asymptomatic in humans, but in 1–2% of cases infection results in neurological complications. Clinical polio syndrome is dominated by flaccid paralysis, due to cell tropism of PV for lower motor neurons in the spinal cord and brainstem expressing CD155/Necl5.Citation106 Overexpression of CD155 has also been shown in (neuro) ectodermal tumors, and transcriptional upregulation has been linked to signaling pathways commonly affected in malignancy, including Raf-Erk-Mnk signaling.Citation107,108 The neuropathogenicity of PV is dependent on the neuronal cell-type-specific function of its IRES element, which assures initiation of translation in a 5' end- and cap-independent manner. Mutations in the IRES genomic region or exchange with other viral IRES counterparts result in markedly neuro-attenuation in CD155-transgenic mice and non-human primates, without reducing the cytopathogenicity in malignant cell types that express CD155.Citation109

Most preclinical research has been performed with PVS-RIPO, a recombinant PV type 1 (Sabin vaccine) strain with the IRES element of human rhinovirus type 2. PVS-RIPO has shown oncolytic efficacy in immune-deficient xenograft rat and mouse models of malignant glioma.Citation110

Currently a phase I clinical trial is ongoing with intratumoral infusion of PVS-RIPO in patients with recurrent GBM showing durable responses.Citation111 Extensive evaluation in non-human primates has shown PVS-RIPO to be safe for intraspinal and intrathalamic injection, without observations of extraneural replication or shedding.Citation109,112 No serious adverse events have been observed so far in an ongoing phase I clinical trial.Citation111 One of the biggest concerns with PV is the inherent genomic instability of picornaviruses and thus the possible reversion to wildtype pathogenicity. PVS-RIPO has been evaluated extensively for genomic instability by e.g., serial passaging in vitro and in vivo and it was shown that escape mutants reverting to neuropathogenic virulence in the CD155-transgenic mouse model do arise.Citation113 Similar mutants have not been observed in other animal models, which makes it unclear what the importance of this preclinical finding is in relation to currently ongoing clinical trials in humans.

Family Picornaviridae: Seneca Valley Virus (SVV)

SVV was first isolated at Genetic Therapy Inc. (Gaithersburg, MD) as a contaminant from cell culture media and is presumed to be introduced via bovine serum or porcine trypsin source.Citation114,115 Serum samples taken from different farm animal populations indicated that (healthy) pigs and other animals are exposed to SVV. However, attempts to infect pigs with SVV isolates failed to demonstrate any specific disease. SVV does not infect humans but does propagate in tumor cells with neuroendocrine features, giving the virus a safe profile for use in virotherapy.Citation114,115 Since its introduction as an oncolytic virus in 2007, SVV has shown preclinical efficacy in nude mice xenograft models for several malignancies.Citation114-118

In a phase I clinical trial employing an intravenous dose escalation in patients with neuroendocrine tumors, SVV had (marginal) treatment benefits without causing serious adverse events when administered even in high dose (1011 viral particles/kg).Citation119 A phase II RCT in patients with extensive stage NSCLC and a phase I dose escalation trial in pediatric patients with neuroblastoma, rhabdomyosarcoma or rare tumors with neuroendocrine features are currently underway.Citation120

Recent reports indicated that, although the natural host is still uncertain, this virus seems safe with regards to toxicity for use as oncolytic virotherapy in (pediatric) patients.Citation119 Analysis of samples obtained from researchers in close contact with phase I clinical trial patients revealed the absence of neutralizing antibody titers, which implicates that imposed hygiene policies were effective.Citation119 However, detailed evaluation of shedding was not performed, and should be determined in future clinical trials.

Family Poxviridae: Vaccinia Virus (VV)

VV infection induces a strong cytotoxic T lymphocyte response and neutralizing antibodies without causing significant disease in humans.Citation121 As an oncolytic virus, VV has the advantage of fast replication and cell lysis with a broad cell/tumor tropism. Furthermore, it lacks genomic integration, and shields extracellular enveloped VV virions from host immunity resulting in capability of (systemic) spreading between tumors. Lastly, it also harbors a large genome packaging accommodation.Citation122

Several strategies have been described to target oncolytic VV specifically to tumor cells. The VV protein VGF is homologous to cellular growth factor EGF and transforming growth factor α (TGFα) and can stimulate the cell for enhanced viral replication through EGF-R. Deletion of the VGF gene will result in a VV that is targeted to cells with inherent EGF-R pathway activity, which is often observed in cancer cells.Citation123 J2R gene (encoding for viral tk) deletion similarly results in a VV that is dependent on overexpression of cellular tk, which is also often observed in cancer cells.Citation124 The combination of VGF and tk gene deletion is known as vvDD and results in an even more selective oncolytic VV, adding to the safety profile.Citation123 VV gene B18R binds to the IFN receptor and can thereby inhibit the cellular antiviral innate immune response. Deletion of B18R thus leads to selectivity for IFN-deficient cells.Citation125 A56R gene encodes for HA and deletion results in severe (neuro)-attenuation.Citation126

Arming of VV has also been described, e.g. with immune stimulators, apoptotic proteins, anti-angiogenic antibodies/proteins, ECM proteases and prodrug-converting enzymes.

Early clinical trials employing non-recombinant vaccine strains of VV have shown safety when injected superficially into melanoma tumors, while local control of bladder cancer was also noted.Citation127,128 JX-594 (tk gene deleted, GM-CSF expressing VV Wyeth; Pexa-Vec)Citation129 has been evaluated in phase I-II clinical trials for patients with metastatic melanoma, (primary) liver tumors, lung, colorectal and various other solid cancer types. GLV-1h68 (GL-ONC1) is currently being investigated in several phase I clinical trials.Citation130,131

Clinical trials with oncolytic VV have thus far reported good safety with regards to toxicity with minor side effects like transient low-grade fever and local pain. Commonly, live vaccinia virus is shed from skin injection sites after vaccination.Citation132 Also, in clinical trials, live JX-594 was detected in throat swabs and skin pustules of patients up to one week after administration.Citation133 Theoretically, recombination between oncolytic recombinant VV and wildtype VV is possible, however, since VV vaccination is not practiced on a large scale anymore, this is highly unlikely. In addition, spontaneous mutation rates for VV have been shown to be very low.Citation134

Family Reoviridae: Mammalian Orthoreovirus (mORV)

mORV is a ubiquitous pathogen with high seropositivity in humans, and has been isolated from sewage, stagnant and river water throughout the world.Citation135,136 mORV is not associated with a named disease in humans, although it can cause mild flu-like upper respiratory or gastrointestinal tract symptoms.Citation135 Three serotypes of mORV can be distinguished: type 1 Lang, type 2 Jones and type 3 Abney or Dearing (mORV-T3D). mORV-T3D was isolated from the intestinal tract of a child with diarrhea, and is used most in (pre)clinical oncolytic virus research.Citation137

mORV-T3D replicates in cells with dysfunctional cell signaling cascades, most importantly (but not exclusively) KRAS-overexpression and subsequent PKR inhibition, making it an inherent oncolytic virus, meaning that it is tumor-specific contributing to safety.Citation138,139 A multitude of cancer types have been shown to respond to mORV-T3D treatment in (animal) models. Cellular immunity has been found to be important for increasing anti-tumor efficacy.Citation137 The absence or inaccessibility of the JAM-A/1 receptor is perceived as a possible limitation for mORV-T3D infection of tumor cells.Citation140 As such, bio-selection through passaging has been attempted to retarget mORV-T3D to other receptors, although this strategy is probably limited by the quasispecies presence in mORV-T3D isolates.Citation141 Only one study using recombinant oncolytic mORV-T3D has been described thus far.Citation142 More studies with recombinant mORV-T3D can be expected in the near future, probably focusing on receptor retargeting and expression of therapeutic or imaging transgenes.

At this time, 16 clinical trials employing intratumoral or intravenous injection of wildtype mORV-T3D (REOLYSIN®: pelareorep) have been conducted and more are currently underway or planned to start in the near future. As excellently summarized by Harrington et al. and Maitra et al,Citation137,143 these trials have shown safety with regards to toxicity of mORV-T3D administration to patients with various solid tumors without dose limiting toxicities, while having some appreciable anti-tumor effects in phase II/III trials.

High mORV titers injected intravenously have been shown to be reasonably safe with low toxicity, even in combination with standard therapies like chemo- or radiotherapy, as well as in combination with transient immune suppression.Citation137,143 Limited mORV shedding has been observed in clinical trials in patient samples of urine, saliva and feces, mostly with high intravenous administrations.Citation137 As an RNA virus with a viral RNA polymerase, mORV genome replication is prone to errors which can lead to mutations in offspring. Furthermore, since wild-type isolates are in use, these probably represent several quasispecies.Citation144 Even so, since mORV-T3D does not seem to cause disease in human subjects, the relevance of this mutation rate is low.

Family Retroviridae: Murine leukemia Virus (MuLV)

MuLVs are widely distributed in domestic and feral mice. MuLVs induce leukemia in mice with latencies ranging from 2 to 18 months, depending on the strain of virus and mouse strain. MuLV is not known to cause a specific disease in humans.Citation145

MuLV development for cancer therapy has been focusing on non-replicating, as well as more recently, replication competent retroviral (RCR) oncolytic vectors. The capacity of MuLV and other retroviruses to integrate into the host genome of dividing cells carries the risk of insertional mutagenesis/oncogenesis. Reducing this risk has been an important goal in designing retroviral vectors. The replication capacity of RCR-MuLV is considered to be beneficial for optimizing gene expression in tumors. Recent RCR-MuLV vector genomes consist of an intact viral genome including an IRES-transgene immediately after the stop codon of the env gene, which results in more genetic stability, while retaining good replication capacity.Citation146,147 The fact that RCR-MuLVs can only infect and integrate in dividing cells results in an inherent onco-selectivity. In contrast to most other oncolytic viruses, the oncolytic activity of RCR-MuLVs depends solely on the transgene that is carried by the virus, since infection itself is not cytolytic. To date the transgene of choice has mostly been CD, which converts the antifungal drug 5-fluorocytosine into active chemotherapeutic agent 5-fluorouracil. Oncolytic activity of RCR-MuLV-CD (Toca 511) has been evaluated in preclinical (animal) models for breast cancer, GBM and mesothelioma.Citation148-150 Toca 511 has a modified backbone and a codon-optimized and heat-stabilized CD gene and has been shown to be highly genomically stable while maintaining oncolytic efficacy upon passaging.Citation145,149

Toca 511 is being investigated in clinical trials in the United States in subjects with recurrent high-grade glioma. Up to now, over 70 patients have been treated without dose limiting toxicity and with evidence of clinical oncolytic efficacy.Citation145 Since RCR-MuLV vectors are capable of genomic integration, germline transmission is a theoretical risk of these vectors, and should be taken into consideration when designing clinical trials.

Discussion

The field of oncolytic virus research has seen a tremendous progression of several first and second generation vectors toward clinical trials. Most current strategies used in oncolytic virotherapy focus on the use of second and third generation of more virulent conditionally replicating viruses, armed with immune stimulating, anti-tumor or tracking transgenes. Also, immune evasion is still sought after to optimize vector delivery. With the first oncolytic virus talimogene laherparepvec now on the break of FDA and EMA approval, we can expect an even greater interest for this relatively young field of oncologic research in the near future.

The newer generation of oncolytic viruses has been evaluated extensively for their efficacy in preclinical trials, and they have shown to be more effective than first generation vectors on many occasions. Also, ample evidence has been gathered regarding their safety in terms of toxicity in laboratory animals. However, studies focusing on environmental shedding and possible recombination of these new oncolytic agents with wildtype viruses are scarce. This subject seems to be of less interest to oncolytic virus researchers. However, a good oncolytic virus should also be evaluated for environmental safety. This holds true not only from a scientific point of view, but also from a regulatory and public health point of view. Without a thorough environmental risk assessment, new agents will not be accepted by the regulatory agencies like EMA and FDA for marketing as new therapies. Especially with the newer oncolytic agents becoming more virulent and with the possibility of expressing transgenes that alter the nature of the virus, any possibility of environmental shedding and recombination with wild type virus should be excluded. More studies evaluating the environmental safety of promising oncolytic viruses should therefore be conducted and reported. With proper safety evaluations, oncolytic virotherapy is ready to make the next step toward clinical applications.

Disclosure of Potential Conflicts of Interest

No potential conflicts of interest were disclosed.

Funding

This study was supported by the Dutch Committee on Genetic Modification (COGEM).

References

  • Campadelli-Fiume G, De Giovanni C, Gatta V, Nanni P, Lollini PL, Menotti L. Rethinking herpes simplex virus: the way to oncolytic agents. Revi Med Virol 2011; 21:213-26; PMID:21626603; http://dx.doi.org/10.1002/rmv.691
  • Zhou G, Roizman B. Construction and properties of a herpes simplex virus 1 designed to enter cells solely via the IL-13alpha2 receptor. Proc Natl Acad Sci U S A 2006; 103:5508-13; PMID:16554374; http://dx.doi.org/10.1073/pnas.0601258103
  • Menotti L, Cerretani A, Hengel H, Campadelli-Fiume G. Construction of a fully retargeted herpes simplex virus 1 recombinant capable of entering cells solely via human epidermal growth factor receptor 2. J Virol 2008; 82:10153-61; PMID:18684832; http://dx.doi.org/10.1128/JVI.01133-08
  • Zhou G, Ye GJ, Debinski W, Roizman B. Engineered herpes simplex virus 1 is dependent on IL13Ralpha 2 receptor for cell entry and independent of glycoprotein D receptor interaction. Proc Natl Acad Sci U S A 2002; 99:15124-9; PMID:12417744; http://dx.doi.org/10.1073/pnas.232588699
  • Menotti L, Cerretani A, Campadelli-Fiume G. A herpes simplex virus recombinant that exhibits a single-chain antibody to HER2/neu enters cells through the mammary tumor receptor, independently of the gD receptors. J Virol 2006; 80:5531-9; PMID:16699034; http://dx.doi.org/10.1128/JVI.02725-05
  • Grandi P, Fernandez J, Szentirmai O, Carter R, Gianni D, Sena-Esteves M, Breakefield XO. Targeting HSV-1 virions for specific binding to epidermal growth factor receptor-vIII-bearing tumor cells. Cancer Gene Ther 2010; 17:655-63; PMID:20508670; http://dx.doi.org/10.1038/cgt.2010.22
  • Kambara H, Okano H, Chiocca EA, Saeki Y. An oncolytic HSV-1 mutant expressing ICP34.5 under control of a nestin promoter increases survival of animals even when symptomatic from a brain tumor. Cancer Res 2005; 65:2832-9; PMID:15805284; http://dx.doi.org/10.1158/0008-5472.CAN-04-3227
  • Amgen. http://www.amgen.com/media/media_pr_detail.jsp?releaseID=1962767
  • Varghese S, Newsome JT, Rabkin SD, McGeagh K, Mahoney D, Nielsen P, Todo T, Martuza RL. Preclinical safety evaluation of G207, a replication-competent herpes simplex virus type 1, inoculated intraprostatically in mice and nonhuman primates. Hum Gene Ther 2001; 12:999-1010; PMID:11387063; http://dx.doi.org/10.1089/104303401750195944
  • Wolfe D, Niranjan A, Trichel A, Wiley C, Ozuer A, Kanal E, Kondziolka D, Krisky D, Goss J, Deluca N, et al. Safety and biodistribution studies of an HSV multigene vector following intracranial delivery to non-human primates. Gene Ther 2004; 11:1675-84; PMID:15306839; http://dx.doi.org/10.1038/sj.gt.3302336
  • 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; PMID:10845725; http://dx.doi.org/10.1038/sj.gt.3301205
  • 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; PMID:18957964; http://dx.doi.org/10.1038/mt.2008.228
  • Hu JC, Coffin RS, Davis CJ, Graham NJ, Groves N, Guest PJ, Harrington KJ, James ND, Love CA, McNeish I, et al. A phase I study of OncoVEXGM-CSF, a second-generation oncolytic herpes simplex virus expressing granulocyte macrophage colony-stimulating factor. Clin Cancer Res 2006; 12:6737-47; PMID:17121894; http://dx.doi.org/10.1158/1078-0432.CCR-06-0759
  • Senzer NN, Kaufman HL, Amatruda T, Nemunaitis M, Reid T, Daniels G, Gonzalez R, Glaspy J, Whitman E, Harrington K, et al. Phase II clinical trial of a granulocyte-macrophage colony-stimulating factor-encoding, second-generation oncolytic herpesvirus in patients with unresectable metastatic melanoma. J Clin Oncol 2009; 27:5763-71; PMID:19884534; http://dx.doi.org/10.1200/JCO.2009.24.3675
  • Harrington KJ, Hingorani M, Tanay MA, Hickey J, Bhide SA, Clarke PM, Renouf LC, Thway K, Sibtain A, McNeish IA, et al. Phase I/II study of oncolytic HSV GM-CSF in combination with radiotherapy and cisplatin in untreated stage III/IV squamous cell cancer of the head and neck. Clin Cancer Res 2010; 16:4005-15; PMID:20670951; http://dx.doi.org/10.1158/1078-0432.CCR-10-0196
  • Geevarghese SK, Geller DA, de Haan HA, Horer 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; PMID:20486770; http://dx.doi.org/10.1089/hum.2010.020
  • Umene K. Mechanism and application of genetic recombination in herpesviruses. Rev Med Virol 1999; 9:171-82; PMID:10479778; http://dx.doi.org/10.1002/(SICI)1099-1654(199907/09)9:3%3c171::AID-RMV243%3e3.0.CO;2-A
  • Cassady KA, Gross M, Roizman B. The second-site mutation in the herpes simplex virus recombinants lacking the gamma134.5 genes precludes shutoff of protein synthesis by blocking the phosphorylation of eIF-2alpha. J Virol 1998; 72:7005-11; PMID:9696792
  • Lenaerts L, De Clercq E, Naesens L. Clinical features and treatment of adenovirus infections. Rev Med Virol 2008; 18:357-74; PMID:18655013; http://dx.doi.org/10.1002/rmv.589
  • Pesonen S, Kangasniemi L, Hemminki A. Oncolytic adenoviruses for the treatment of human cancer: focus on translational and clinical data. Mol Pharm 2011; 8:12-28; PMID:21126047; http://dx.doi.org/10.1021/mp100219n
  • Bischoff JR, Kirn DH, Williams A, Heise C, Horn S, Muna M, Ng L, Nye JA, Sampson-Johannes A, Fattaey A, et al. An adenovirus mutant that replicates selectively in p53-deficient human tumor cells. Science 1996; 274:373-6; PMID:8832876; http://dx.doi.org/10.1126/science.274.5286.373
  • O'Shea CC, Johnson L, Bagus B, Choi S, Nicholas C, Shen A, Boyle L, Pandey K, Soria C, Kunich J, et al. Late viral RNA export, rather than p53 inactivation, determines ONYX-015 tumor selectivity. Cancer Cell 2004; 6:611-23; PMID:15607965; http://dx.doi.org/10.1016/j.ccr.2004.11.012
  • Fueyo J, Gomez-Manzano C, Alemany R, Lee PS, McDonnell TJ, Mitlianga P, Shi YX, Levin VA, Yung WK, Kyritsis AP. A mutant oncolytic adenovirus targeting the Rb pathway produces anti-glioma effect in vivo. Oncogene 2000; 19:2-12; PMID:10644974; http://dx.doi.org/10.1038/sj.onc.1203251
  • Johnson L, Shen A, Boyle L, Kunich J, Pandey K, Lemmon M, Hermiston T, Giedlin M, McCormick F, Fattaey A. Selectively replicating adenoviruses targeting deregulated E2F activity are potent, systemic antitumor agents. Cancer Cell 2002; 1:325-37; PMID:12086848; http://dx.doi.org/10.1016/S1535-6108(02)00060-0
  • Kim J, Cho JY, Kim JH, Jung KC, Yun CO. Evaluation of E1B gene-attenuated replicating adenoviruses for cancer gene therapy. Cancer Gene Ther 2002; 9:725-36; PMID:12189522; http://dx.doi.org/10.1038/sj.cgt.7700494
  • Cascallo M, Capella G, Mazo A, Alemany R. Ras-dependent oncolysis with an adenovirus VAI mutant. Cancer Res 2003; 63:5544-50; PMID:14500393
  • Mantwill K, Naumann U, Seznec J, Girbinger V, Lage H, Surowiak P, Beier D, Mittelbronn M, Schlegel J, Holm PS. YB-1 dependent oncolytic adenovirus efficiently inhibits tumor growth of glioma cancer stem like cells. J Transl Med 2013; 11:216; PMID:24044901; http://dx.doi.org/10.1186/1479-5876-11-216
  • Heise C, Kirn DH. Replication-selective adenoviruses as oncolytic agents. J Clin Invest 2000; 105:847-51; PMID:10749561; http://dx.doi.org/10.1172/JCI9762
  • Nettelbeck DM. Virotherapeutics: conditionally replicative adenoviruses for viral oncolysis. Anticancer Drugs 2003; 14:577-84; PMID:14501378; http://dx.doi.org/10.1097/00001813-200309000-00001
  • Li X, Liu Y, Wen Z, Li C, Lu H, Tian M, Jin K, Sun L, Gao P, Yang E, et al. Potent anti-tumor effects of a dual specific oncolytic adenovirus expressing apoptin in vitro and in vivo. Mol Cancer 2010; 9:10; PMID:20085660; http://dx.doi.org/10.1186/1476-4598-9-10
  • Gao Q, Chen C, Ji T, Wu P, Han Z, Fang H, Li F, Liu Y, Hu W, Gong D, et al. A systematic comparison of the anti-tumoural activity and toxicity of the three Adv-TKs. PLoS One 2014; 9:e94050; PMID:24722669; http://dx.doi.org/10.1371/journal.pone.0094050
  • Koski A, Kangasniemi L, Escutenaire S, Pesonen S, Cerullo V, Diaconu I, Nokisalmi P, Raki M, Rajecki M, Guse K, et al. Treatment of cancer patients with a serotype 5/3 chimeric oncolytic adenovirus expressing GMCSF. Mol Ther 2010; 18:1874-84; PMID:20664527; http://dx.doi.org/10.1038/mt.2010.161
  • 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; PMID:23088985; http://dx.doi.org/10.1016/j.juro.2012.07.097
  • European Medicines Agency (EMA). Withdrawal assessment report for Advexin. EMEA/692328/2008 2008
  • European Medicines Agency (EMA). Withdrawal assessment report for Cerepro. EMEA/203243/2008 2007
  • Kim M, Zinn KR, Barnett BG, Sumerel LA, Krasnykh V, Curiel DT, Douglas JT. The therapeutic efficacy of adenoviral vectors for cancer gene therapy is limited by a low level of primary adenovirus receptors on tumour cells. Eur J Cancer 2002; 38:1917-26; PMID:12204675; http://dx.doi.org/10.1016/S0959-8049(02)00131-4
  • Tao N, Gao GP, Parr M, Johnston J, Baradet T, Wilson JM, Barsoum J, Fawell SE. Sequestration of adenoviral vector by Kupffer cells leads to a nonlinear dose response of transduction in liver. Mol Ther 2001; 3:28-35; PMID:11162308; http://dx.doi.org/10.1006/mthe.2000.0227
  • Alemany R, Curiel DT. CAR-binding ablation does not change biodistribution and toxicity of adenoviral vectors. Gene Ther 2001; 8:1347-53; PMID:11571572; http://dx.doi.org/10.1038/sj.gt.3301515
  • Smith T, Idamakanti N, Kylefjord H, Rollence M, King L, Kaloss M, Kaleko M, Stevenson SC. In vivo hepatic adenoviral gene delivery occurs independently of the coxsackievirus-adenovirus receptor. Mol Ther 2002; 5:770-9; PMID:12027562; http://dx.doi.org/10.1006/mthe.2002.0613
  • Alba R, Bradshaw AC, Coughlan L, Denby L, McDonald RA, Waddington SN, Buckley SM, Greig JA, Parker AL, Miller AM, et al. Biodistribution and retargeting of FX-binding ablated adenovirus serotype 5 vectors. Blood 2010; 116:2656-64; PMID:20610817; http://dx.doi.org/10.1182/blood-2009-12-260026
  • Short JJ, Rivera AA, Wu H, Walter MR, Yamamoto M, Mathis JM, Curiel DT. Substitution of adenovirus serotype 3 hexon onto a serotype 5 oncolytic adenovirus reduces factor × binding, decreases liver tropism, and improves antitumor efficacy. Mol Cancer Ther 2010; 9:2536-44; PMID:20736345; http://dx.doi.org/10.1158/1535-7163.MCT-10-0332
  • Coughlan L, Alba R, Parker AL, Bradshaw AC, McNeish IA, Nicklin SA, Baker AH. Tropism-modification strategies for targeted gene delivery using adenoviral vectors. Viruses 2010; 2:2290-355; PMID:21994621; http://dx.doi.org/10.3390/v2102290
  • Hemminki O, Diaconu I, Cerullo V, Pesonen SK, Kanerva A, Joensuu T, Kairemo K, Laasonen L, Partanen K, Kangasniemi L, et al. Ad3-hTERT-E1A, a fully serotype 3 oncolytic adenovirus, in patients with chemotherapy refractory cancer. Mol Ther 2012; 20:1821-30; PMID:22871667; http://dx.doi.org/10.1038/mt.2012.115
  • Belousova N, Mikheeva G, Xiong C, Soghomonian S, Young D, Le Roux L, Naff K, Bidaut L, Wei W, Li C, et al. Development of a targeted gene vector platform based on simian adenovirus serotype 24. J Virol 2010; 84:10087-101; PMID:20631120; http://dx.doi.org/10.1128/JVI.02425-09
  • Sharma A, Bangari DS, Tandon M, Pandey A, HogenEsch H, Mittal SK. Comparative analysis of vector biodistribution, persistence and gene expression following intravenous delivery of bovine, porcine and human adenoviral vectors in a mouse model. Virology 2009; 386:44-54; PMID:19211122; http://dx.doi.org/10.1016/j.virol.2009.01.008
  • 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; PMID:18560559; http://dx.doi.org/10.1371/journal.pone.0002409
  • Miura Y, Yamasaki S, Davydova J, Brown E, Aoki K, Vickers S, Yamamoto M. Infectivity-selective oncolytic adenovirus developed by high-throughput screening of adenovirus-formatted library. Mol Ther 2013; 21:139-48; PMID:23032977; http://dx.doi.org/10.1038/mt.2012.205
  • Reid T, Galanis E, Abbruzzese J, Sze D, Wein LM, Andrews J, Randlev B, Heise C, Uprichard M, Hatfield M, et al. Hepatic arterial infusion of a replication-selective oncolytic adenovirus (dl1520): phase II viral, immunologic, and clinical endpoints. Cancer Res 2002; 62:6070-9; PMID:12414631
  • Kimball KJ, Preuss MA, Barnes MN, Wang M, Siegal GP, Wan W, Kuo H, Saddekni S, Stockard CR, Grizzle WE, et al. A phase I study of a tropism-modified conditionally replicative adenovirus for recurrent malignant gynecologic diseases. Clin Cancer Res 2010; 16:5277-87; PMID:20978148; http://dx.doi.org/10.1158/1078-0432.CCR-10-0791
  • 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; PMID:21630267; http://dx.doi.org/10.1002/ijc.26216
  • Pesonen S, Nokisalmi P, Escutenaire S, Sarkioja M, Raki M, Cerullo V, Kangasniemi L, Laasonen L, Ribacka C, Guse K, et al. Prolonged systemic circulation of chimeric oncolytic adenovirus Ad5/3-Cox2L-D24 in patients with metastatic and refractory solid tumors. Gene Ther 2010; 17:892-904; PMID:20237509; http://dx.doi.org/10.1038/gt.2010.17
  • Raki M, Sarkioja M, Escutenaire S, Kangasniemi L, Haavisto E, Kanerva A, Cerullo V, Joensuu T, Oksanen M, Pesonen S, et al. Switching the fiber knob of oncolytic adenoviruses to avoid neutralizing antibodies in human cancer patients. J Gene Med 2011; 13:253-61; PMID:21520358; http://dx.doi.org/10.1002/jgm.1565
  • 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. 2014 Aug 1; 135(3):720-30.
  • 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-Delta24, a novel serotype-chimeric, infectivity-enhanced, conditionally-replicative adenovirus (CRAd), in patients with recurrent ovarian cancer. Gynecol Oncol 2013; 130:518-24; PMID:23756180; http://dx.doi.org/10.1016/j.ygyno.2013.06.003
  • Blanc C, Calvo E, Martin MG, Machiels JP, Rottey S, Carbonero RG, McNeish IA, Ellis C, Fisher K, Beadle J. Review of the current status of phase I clinical studies of Enadenotucirev (ColoAd1), an Ad11/Ad3 chimeric group B adenovirus, in patients with metastatic epithelial solid tumors. 8th International Conference on Oncolytic Virus Therapeutics 2014
  • Nemunaitis J, Tong AW, Nemunaitis M, Senzer N, Phadke AP, Bedell C, Adams N, Zhang YA, Maples PB, Chen S, et al. A phase I study of telomerase-specific replication competent oncolytic adenovirus (telomelysin) for various solid tumors. Mol Ther 2010; 18:429-34; PMID:19935775; http://dx.doi.org/10.1038/mt.2009.262
  • Gimenez-Alejandre M, Rodriguez-Garcia A, Moreno-Olie R, Condom i Mundó E, Nadal M, Bazan-Peregrino M, Alemany R, Cascallo M. Chemosensitization to gemcitabine by VCN-01, a tumor-targeted oncolytic adenovirus armed with hyaluronidase: from bench to bedside. 8th International Conference on Oncolytic Virus Therapeutics 2014
  • Chiocca EA, Smith KM, McKinney B, Palmer CA, Rosenfeld S, Lillehei K, Hamilton A, DeMasters BK, Judy K, Kirn D. A phase I trial of Ad.hIFN-beta gene therapy for glioma. Mol Ther 2008; 16:618-26; PMID:18180770; http://dx.doi.org/10.1038/sj.mt.6300396
  • Keedy V, Wang W, Schiller J, Chada S, Slovis B, Coffee K, Worrell J, Thet LA, Johnson DH, Carbone DP. Phase I study of adenovirus p53 administered by bronchoalveolar lavage in patients with bronchioloalveolar cell lung carcinoma: ECOG 6597. J Clin Oncol 2008; 26:4166-71; PMID:18757331; http://dx.doi.org/10.1200/JCO.2007.15.6927
  • Tian G, Liu J, Zhou JS, Chen W. Multiple hepatic arterial injections of recombinant adenovirus p53 and 5-fluorouracil after transcatheter arterial chemoembolization for unresectable hepatocellular carcinoma: a pilot phase II trial. Anticancer Drugs 2009; 20:389-95; PMID:19287305; http://dx.doi.org/10.1097/CAD.0b013e32832a2df9
  • Xu F, Li S, Li XL, Guo Y, Zou BY, Xu R, Liao H, Zhao HY, Zhang Y, Guan ZZ, et al. Phase I and biodistribution study of recombinant adenovirus vector-mediated herpes simplex virus thymidine kinase gene and ganciclovir administration in patients with head and neck cancer and other malignant tumors. Cancer Gene Ther 2009; 16:723-30; PMID:19363470; http://dx.doi.org/10.1038/cgt.2009.19
  • Page JG, Tian B, Schweikart K, Tomaszewski J, Harris R, Broadt T, Polley-Nelson J, Noker PE, Wang M, Makhija S, et al. Identifying the safety profile of a novel infectivity-enhanced conditionally replicative adenovirus, Ad5-delta24-RGD, in anticipation of a phase I trial for recurrent ovarian cancer. Am J Obstet Gynecol 2007; 196:389.e1-9; discussion .e9–10; PMID:17403430; http://dx.doi.org/10.1016/j.ajog.2006.12.016
  • Griscelli F, Opolon P, Saulnier P, Mami-Chouaib F, Gautier E, Echchakir H, Angevin E, Le Chevalier T, Bataille V, Squiban P, et al. Recombinant adenovirus shedding after intratumoral gene transfer in lung cancer patients. Gene Ther 2003; 10:386-95; PMID:12601393; http://dx.doi.org/10.1038/sj.gt.3301928
  • Schenk-Braat EA, van Mierlo MM, Wagemaker G, Bangma CH, Kaptein LC. An inventory of shedding data from clinical gene therapy trials. J Gene Med 2007; 9:910-21; PMID:17880045; http://dx.doi.org/10.1002/jgm.1096
  • Robinson CM, Rajaiya J, Walsh MP, Seto D, Dyer DW, Jones MS, Chodosh J. Computational analysis of human adenovirus type 22 provides evidence for recombination among species D human adenoviruses in the penton base gene. J Virol 2009; 83:8980-5; PMID:19553309; http://dx.doi.org/10.1128/JVI.00786-09
  • Walsh MP, Chintakuntlawar A, Robinson CM, Madisch I, Harrach B, Hudson NR, Schnurr D, Heim A, Chodosh J, Seto D, et al. Evidence of molecular evolution driven by recombination events influencing tropism in a novel human adenovirus that causes epidemic keratoconjunctivitis. PLoS One 2009; 4:e5635; PMID:19492050; http://dx.doi.org/10.1371/journal.pone.0005635
  • Singh G, Robinson CM, Dehghan S, Jones MS, Dyer DW, Seto D, Chodosh J. Homologous recombination in E3 genes of human adenovirus species D. J Virol 2013; 87:12481-8; PMID:24027303; http://dx.doi.org/10.1128/JVI.01927-13
  • Duke T, Mgone CS. Measles: not just another viral exanthem. Lancet 2003; 361:763-73; PMID:12620751; http://dx.doi.org/10.1016/S0140-6736(03)12661-X
  • Griffin DE, Pan CH. Measles: old vaccines, new vaccines. Curr Top Microbiol Immunol 2009; 330:191-212; PMID:19203111
  • Msaouel P, Iankov ID, Dispenzieri A, Galanis E. Attenuated oncolytic measles virus strains as cancer therapeutics. Curr Pharm Biotechnol 2012; 13:1732-41; PMID:21740361; http://dx.doi.org/10.2174/138920112800958896
  • Blechacz B, Russell SJ. Measles virus as an oncolytic vector platform. Curr Gene Ther 2008; 8:162-75; PMID:18537591; http://dx.doi.org/10.2174/156652308784746459
  • Msaouel P, Dispenzieri A, Galanis E. Clinical testing of engineered oncolytic measles virus strains in the treatment of cancer: an overview. Curr Opin Mol Ther 2009; 11:43-53; PMID:19169959
  • Galanis E, Hartmann LC, Cliby WA, Long HJ, Peethambaram PP, Barrette BA, Kaur JS, Haluska PJ, Jr., Aderca I, Zollman PJ, et al. Phase I trial of intraperitoneal administration of an oncolytic measles virus strain engineered to express carcinoembryonic antigen for recurrent ovarian cancer. Cancer Res 2010; 70:875-82; PMID:20103634; http://dx.doi.org/10.1158/0008-5472.CAN-09-2762
  • Russell SJ, Federspiel MJ, Peng KW, Tong C, Dingli D, Morice WG, Lowe V, O'Connor MK, Kyle RA, Leung N, et al. Remission of disseminated cancer after systemic oncolytic virotherapy. Mayo Clinic proceedings 2014
  • Evans AS. Pathogenicity and immunology of Newcastle disease virus (NVD) in man. Am J Public Health Nations Health 1955; 45:742-5; PMID:14376703; http://dx.doi.org/10.2105/AJPH.45.6.742
  • de Leeuw OS, Hartog L, Koch G, Peeters BP. Effect of fusion protein cleavage site mutations on virulence of Newcastle disease virus: non-virulent cleavage site mutants revert to virulence after one passage in chicken brain. J Gen Virol 2003; 84:475-84; PMID:12560582; http://dx.doi.org/10.1099/vir.0.18714-0
  • Sinkovics JG, Horvath JC. Newcastle disease virus (NDV): brief history of its oncolytic strains. J Clin Virol 2000; 16:1-15; PMID:10680736; http://dx.doi.org/10.1016/S1386-6532(99)00072-4
  • Vigil A, Park MS, Martinez O, Chua MA, Xiao S, Cros JF, Martinez-Sobrido L, Woo SL, Garcia-Sastre A. Use of reverse genetics to enhance the oncolytic properties of Newcastle disease virus. Cancer Res 2007; 67:8285-92; PMID:17804743; http://dx.doi.org/10.1158/0008-5472.CAN-07-1025
  • Zamarin D, Vigil A, Kelly K, Garcia-Sastre A, Fong Y. Genetically engineered Newcastle disease virus for malignant melanoma therapy. Gene Ther 2009; 16:796-804; PMID:19242529; http://dx.doi.org/10.1038/gt.2009.14
  • Zamarin D, Martinez-Sobrido L, Kelly K, Mansour M, Sheng G, Vigil A, Garcia-Sastre A, Palese P, Fong Y. Enhancement of oncolytic properties of recombinant newcastle disease virus through antagonism of cellular innate immune responses. Mol Ther 2009; 17:697-706; PMID:19209145; http://dx.doi.org/10.1038/mt.2008.286
  • Altomonte J, Marozin S, Schmid RM, Ebert O. Engineered newcastle disease virus as an improved oncolytic agent against hepatocellular carcinoma. Mol Ther 2010; 18:275-84; PMID:19809404; http://dx.doi.org/10.1038/mt.2009.231
  • Silberhumer GR, Brader P, Wong J, Serganova IS, Gonen M, Gonzalez SJ, Blasberg R, Zamarin D, Fong Y. Genetically engineered oncolytic Newcastle disease virus effectively induces sustained remission of malignant pleural mesothelioma. Mol Cancer Ther 2010; 9:2761-9; PMID:20858727; http://dx.doi.org/10.1158/1535-7163.MCT-10-0090
  • Li P, Chen CH, Li S, Givi B, Yu Z, Zamarin D, Palese P, Fong Y, Wong RJ. Therapeutic effects of a fusogenic newcastle disease virus in treating head and neck cancer. Head Neck 2011; 33:1394-9; PMID:21928411; http://dx.doi.org/10.1002/hed.21609
  • 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; PMID:24598590; http://dx.doi.org/10.1126/scitranslmed.3008095
  • Buijs P, van Nieuwkoop S, Van Eijck C, Fouchier R, Van den Hoogen B. Recombinant oncolytic Newcastle disease virus for treatment of pancreatic adenocarcinoma. 8th International Conference on Oncolytic Virus Therapeutics 2014
  • Buijs PR, van Eijck CH, Hofland LJ, Fouchier RA, van den Hoogen BG. Different responses of human pancreatic adenocarcinoma cell lines to oncolytic Newcastle disease virus infection. Cancer Gene Ther 2014; 21:24-30; PMID:24384773; http://dx.doi.org/10.1038/cgt.2013.78
  • Buijs PR, van Amerongen G, van Nieuwkoop S, Bestebroer TM, van Run PR, Kuiken T, Fouchier RA, van Eijck CH, van den Hoogen BG. Intravenously injected Newcastle disease virus in non-human primates is safe to use for oncolytic virotherapy. Cancer Gene Ther 2014; 21(11):463-71; PMID:25257305; http://dx.doi.org/10.1038/cgt.2014.51
  • Hastie E, Grdzelishvili VZ. Vesicular stomatitis virus as a flexible platform for oncolytic virotherapy against cancer. J Gen Virol 2012; 93:2529-45; PMID:23052398; http://dx.doi.org/10.1099/vir.0.046672-0
  • Quiroz E, Moreno N, Peralta PH, Tesh RB. A human case of encephalitis associated with vesicular stomatitis virus (Indiana serotype) infection. Am J Trop Med Hyg 1988; 39:312-4; PMID:2845825
  • 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; PMID:10888934; http://dx.doi.org/10.1038/77558
  • Barber GN. VSV-tumor selective replication and protein translation. Oncogene 2005; 24:7710-9; PMID:16299531; http://dx.doi.org/10.1038/sj.onc.1209042
  • Oliere S, Arguello M, Mesplede T, Tumilasci V, Nakhaei P, Stojdl D, Sonenberg N, Bell J, Hiscott J. Vesicular stomatitis virus oncolysis of T lymphocytes requires cell cycle entry and translation initiation. J Virol 2008; 82:5735-49; PMID:18417567; http://dx.doi.org/10.1128/JVI.02601-07
  • LeBlanc AK, Naik S, Galyon GD, Jenks N, Steele M, Peng KW, Federspiel MJ, Donnell R, Russell SJ. Safety studies on intravenous administration of oncolytic recombinant vesicular stomatitis virus in purpose-bred beagle dogs. Hum Gene Ther Clin Dev 2013; 24:174-81; PMID:24219832; http://dx.doi.org/10.1089/humc.2013.165
  • Jenks N, Myers R, Greiner SM, Thompson J, Mader EK, Greenslade A, Griesmann GE, Federspiel MJ, Rakela J, Borad MJ, et al. Safety studies on intrahepatic or intratumoral injection of oncolytic vesicular stomatitis virus expressing interferon-beta in rodents and nonhuman primates. Hum Gene Ther 2010; 21:451-62; PMID:19911974; http://dx.doi.org/10.1089/hum.2009.111
  • Naik S, LeBlanc AK, Galyon GD, Frazier S, Jenks N, Steele M, Peng KW, Federspiel M, Donnell R, Russell SJ. Early findings from a comparative oncology study evaluating systemic VSV therapy in client owned dogs with spontaneous hematologic malignancies. 8th International Conference on Oncolytic Virus Therapeutics 2014
  • Stanifer ML, Cureton DK, Whelan SP. A recombinant vesicular stomatitis virus bearing a lethal mutation in the glycoprotein gene uncovers a second site suppressor that restores fusion. J Virol 2011; 85:8105-15; PMID:21680501; http://dx.doi.org/10.1128/JVI.00735-11
  • Harouaka D, Wertz GW. Second-site mutations selected in transcriptional regulatory sequences compensate for engineered mutations in the vesicular stomatitis virus nucleocapsid protein. J Virol 2012; 86:11266-75; PMID:22875970; http://dx.doi.org/10.1128/JVI.01238-12
  • Quinones-Kochs MI, Schnell MJ, Buonocore L, Rose JK. Mechanisms of loss of foreign gene expression in recombinant vesicular stomatitis viruses. Virology 2001; 287:427-35; PMID:11531419; http://dx.doi.org/10.1006/viro.2001.1058
  • Gao Y, Whitaker-Dowling P, Watkins SC, Griffin JA, Bergman I. Rapid adaptation of a recombinant vesicular stomatitis virus to a targeted cell line. J Virol 2006; 80:8603-12; PMID:16912309; http://dx.doi.org/10.1128/JVI.00142-06
  • Dinh PX, Panda D, Das PB, Das SC, Das A, Pattnaik AK. A single amino acid change resulting in loss of fluorescence of eGFP in a viral fusion protein confers fitness and growth advantage to the recombinant vesicular stomatitis virus. Virology 2012; 432:460-9; PMID:22832124; http://dx.doi.org/10.1016/j.virol.2012.07.004
  • Sarma N. Hand, foot, and mouth disease: current scenario and Indian perspective. Indian J Dermatol Venereol Leprol 2013; 79:165-75; PMID:23442455; http://dx.doi.org/10.4103/0378-6323.107631
  • Hughes PJ, North C, Minor PD, Stanway G. The complete nucleotide sequence of coxsackievirus A21. J Gen Virol 1989; 70(Pt 11):2943-52; PMID:2584950; http://dx.doi.org/10.1099/0022-1317-70-11-2943
  • Shafren DR, Dorahy DJ, Ingham RA, Burns GF, Barry RD. Coxsackievirus A21 binds to decay-accelerating factor but requires intercellular adhesion molecule 1 for cell entry. J Virol 1997; 71:4736-43; PMID:9151867
  • 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; PMID:14734451; http://dx.doi.org/10.1158/1078-0432.CCR-0690-3
  • Andtbacka RHI, Au GG, Weisberg JI, Post L, Shafren DR. CAVATAK-mediated oncolytic immunotherapy in advanced melanoma patients. 8th International Conference on Oncolytic Virus Therapeutics 2014
  • Mendelsohn CL, Wimmer E, Racaniello VR. Cellular receptor for poliovirus: molecular cloning, nucleotide sequence, and expression of a new member of the immunoglobulin superfamily. Cell 1989; 56:855-65; PMID:2538245; http://dx.doi.org/10.1016/0092-8674(89)90690-9
  • Merrill MK, Bernhardt G, Sampson JH, Wikstrand CJ, Bigner DD, Gromeier M. Poliovirus receptor CD155-targeted oncolysis of glioma. Neuro Oncol 2004; 6:208-17; PMID:15279713; http://dx.doi.org/10.1215/S1152851703000577
  • Goetz C, Everson RG, Zhang LC, Gromeier M. MAPK signal-integrating kinase controls cap-independent translation and cell type-specific cytotoxicity of an oncolytic poliovirus. Mol Ther 2010; 18:1937-46; PMID:20648000; http://dx.doi.org/10.1038/mt.2010.145
  • Gromeier M, Bossert B, Arita M, Nomoto A, Wimmer E. Dual stem loops within the poliovirus internal ribosomal entry site control neurovirulence. J Virol 1999; 73:958-64; PMID:9882296
  • Dobrikova EY, Broadt T, Poiley-Nelson J, Yang X, Soman G, Giardina S, Harris R, Gromeier M. Recombinant oncolytic poliovirus eliminates glioma in vivo without genetic adaptation to a pathogenic phenotype. Mol Ther 2008; 16:1865-72; PMID:18766173; http://dx.doi.org/10.1038/mt.2008.184
  • Gromeier M, Dobrikova E, Dobrikov M, Brown M, Desjardins A, Friedman H, Sampson JH, Friedman A, Bigner DD. Oncolytic poliovirus immunotherapy of glioblastoma. 8th International Conference on Oncolytic Virus Therapeutics 2014
  • Dobrikova EY, Goetz C, Walters RW, Lawson SK, Peggins JO, Muszynski K, Ruppel S, Poole K, Giardina SL, Vela EM, et al. Attenuation of neurovirulence, biodistribution, and shedding of a poliovirus:rhinovirus chimera after intrathalamic inoculation in Macaca fascicularis. J Virol 2012; 86:2750-9; PMID:22171271; http://dx.doi.org/10.1128/JVI.06427-11
  • Jahan N, Wimmer E, Mueller S. A host-specific, temperature-sensitive translation defect determines the attenuation phenotype of a human rhinovirus/poliovirus chimera, PV1(RIPO). J Virol 2011; 85:7225-35; PMID:21561914; http://dx.doi.org/10.1128/JVI.01804-09
  • 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; PMID:17971529; http://dx.doi.org/10.1093/jnci/djm198
  • 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; PMID:23739064; http://dx.doi.org/10.1093/jnci/djt130
  • Morton CL, Houghton PJ, Kolb EA, Gorlick R, Reynolds CP, Kang MH, Maris JM, Keir ST, Wu J, Smith MA. Initial testing of the replication competent Seneca Valley virus (NTX-010) by the pediatric preclinical testing program. Pediatric Blood Cancer 2010; 55:295-303; PMID:20582972; http://dx.doi.org/10.1002/pbc.22535
  • Yu L, Baxter PA, Zhao X, Liu Z, Wadhwa L, Zhang Y, Su JM, Tan X, Yang J, Adesina A, et al. A single intravenous injection of oncolytic picornavirus SVV-001 eliminates medulloblastomas in primary tumor-based orthotopic xenograft mouse models. Neuro Oncol 2011; 13:14-27; PMID:21075780; http://dx.doi.org/10.1093/neuonc/noq148
  • Liu Z, Zhao X, Mao H, Baxter PA, Huang Y, Yu L, Wadhwa L, Su JM, Adesina A, Perlaky L, et al. Intravenous injection of oncolytic picornavirus SVV-001 prolongs animal survival in a panel of primary tumor-based orthotopic xenograft mouse models of pediatric glioma. Neuro Oncol 2013; 15:1173-85; PMID:23658322; http://dx.doi.org/10.1093/neuonc/not065
  • Rudin CM, Poirier JT, Senzer NN, Stephenson J, Jr., Loesch D, Burroughs KD, Reddy PS, Hann CL, Hallenbeck PL. Phase I clinical study of Seneca Valley Virus (SVV-001), a replication-competent picornavirus, in advanced solid tumors with neuroendocrine features. Clin Cancer Res 2011; 17:888-95; PMID:21304001; http://dx.doi.org/10.1158/1078-0432.CCR-10-1706
  • Burke MJ, Ahern C, Weigel BJ, Poirier JT, Rudin CM, Chen Y, Cripe TP, Bernhardt MB, Blaney SM. Phase I trial of seneca valley virus (NTX-010) in children with relapsed/refractory solid tumors: a report of the children's oncology group. Pediatr Blood Cancer. 2015 May; 62(5):743-50.
  • Miller JD, van der Most RG, Akondy RS, Glidewell JT, Albott S, Masopust D, Murali-Krishna K, Mahar PL, Edupuganti S, Lalor S, et al. Human effector and memory CD8+ T cell responses to smallpox and yellow fever vaccines. Immunity 2008; 28:710-22; PMID:18468462; http://dx.doi.org/10.1016/j.immuni.2008.02.020
  • Kirn DH, Wang Y, Liang W, Contag CH, Thorne SH. Enhancing poxvirus oncolytic effects through increased spread and immune evasion. Cancer Res 2008; 68:2071-5; PMID:18381410; http://dx.doi.org/10.1158/0008-5472.CAN-07-6515
  • Thorne SH, Hwang TH, O'Gorman WE, Bartlett DL, Sei S, Kanji F, Brown C, Werier J, Cho JH, Lee DE, et al. Rational strain selection and engineering creates a broad-spectrum, systemically effective oncolytic poxvirus, JX-963. J Clin Invest 2007; 117:3350-8; PMID:17965776; http://dx.doi.org/10.1172/JCI32727
  • Puhlmann M, Brown CK, Gnant M, Huang J, Libutti SK, Alexander HR, Bartlett DL. Vaccinia as a vector for tumor-directed gene therapy: biodistribution of a thymidine kinase-deleted mutant. Cancer Gene Ther 2000; 7:66-73; PMID:10678358; http://dx.doi.org/10.1038/sj.cgt.7700075
  • Kirn DH, Wang Y, Le Boeuf F, Bell J, Thorne SH. Targeting of interferon-beta to produce a specific, multi-mechanistic oncolytic vaccinia virus. PLoS Med 2007; 4:e353; PMID:18162040; http://dx.doi.org/10.1371/journal.pmed.0040353
  • Shida H, Hinuma Y, Hatanaka M, Morita M, Kidokoro M, Suzuki K, Maruyama T, Takahashi-Nishimaki F, Sugimoto M, Kitamura R, et al. Effects and virulences of recombinant vaccinia viruses derived from attenuated strains that express the human T-cell leukemia virus type I envelope gene. J Virol 1988; 62:4474-80; PMID:3184271
  • Roenigk HH, Jr., Deodhar S, St Jacques R, Burdick K. Immunotherapy of malignant melanoma with vaccinia virus. Arch Dermatol 1974; 109:668-73; PMID:4828533; http://dx.doi.org/10.1001/archderm.1974.01630050014003
  • Lee SS, Eisenlohr LC, McCue PA, Mastrangelo MJ, Lattime EC. Intravesical gene therapy: in vivo gene transfer using recombinant vaccinia virus vectors. Cancer Res 1994; 54:3325-8; PMID:8012943
  • Lusky M, Erbs P, Foloppe J, Acres RB. Oncolytic vaccinia virus: a silver bullet? Expert Rev Vaccines 2010; 9:1353-6; PMID:21105770; http://dx.doi.org/10.1586/erv.10.137
  • Lauer UM, Beil J, Berchtold S, Zimmermann M, Koppenhöfer U, Bitzer M, Malek NP, Glatzle J, Königsrainer A, Möhle R, et al. Tracking of tumor cell colonization, in-patient replication, and oncolysis by GL-ONC1 employed in a phase I/II virotherapy study on patients with peritoneal carcinomatosis. 8th International Conference on Oncolytic Virus Therapeutics 2014
  • Reinboth J, Ralph C, Sciagalla P, Yu YA, Agular J, Chen N, Szalay AA, Melcher AA, West EJ. Oncolytic vaccinia virus GL-ONC1 treatment of colorectal cancer patients. 8th International Conference on Oncolytic Virus Therapeutics 2014
  • Cono J, Casey CG, Bell DM. Smallpox vaccination and adverse reactions. Guidance for clinicians. MMWR Recomm Rep 2003; 52:1-28; PMID:12617510
  • 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; PMID:21886163; http://dx.doi.org/10.1038/nature10358
  • Qin L, Evans DH. Genome scale patterns of recombination between coinfecting vaccinia viruses. J Virol 2014; 88:5277-86; PMID:24574414; http://dx.doi.org/10.1128/JVI.00022-14
  • Rosen L, Evans HE, Spickard A. Reovirus infections in human volunteers. Am J Hyg 1963; 77:29-37; PMID:13974840
  • Tai JH, Williams JV, Edwards KM, Wright PF, Crowe JE, Jr., Dermody TS. Prevalence of reovirus-specific antibodies in young children in Nashville, Tennessee. J Infect Dis 2005; 191:1221-4; PMID:15776366; http://dx.doi.org/10.1086/428911
  • Maitra R, Ghalib MH, Goel S. Reovirus: a targeted therapeutic–progress and potential. Mol Cancer Res 2012; 10:1514-25; PMID:23038811; http://dx.doi.org/10.1158/1541-7786.MCR-12-0157
  • Shmulevitz M, Marcato P, Lee PW. Unshackling the links between reovirus oncolysis, Ras signaling, translational control and cancer. Oncogene 2005; 24:7720-8; PMID:16299532; http://dx.doi.org/10.1038/sj.onc.1209041
  • Lemay G, Tumilasci V, Hiscott J. Uncoating reo: uncovering the steps critical for oncolysis. Mol Ther 2007; 15:1406-7; PMID:17646836; http://dx.doi.org/10.1038/sj.mt.6300242
  • van Houdt WJ, Smakman N, van den Wollenberg DJ, Emmink BL, Veenendaal LM, van Diest PJ, Hoeben RC, Borel Rinkes IH, Kranenburg O. Transient infection of freshly isolated human colorectal tumor cells by reovirus T3D intermediate subviral particles. Cancer Gene Ther 2008; 15:284-92; PMID:18259212; http://dx.doi.org/10.1038/cgt.2008.2
  • van den Wollenberg DJ, Dautzenberg IJ, van den Hengel SK, Cramer SJ, de Groot RJ, Hoeben RC. Isolation of reovirus T3D mutants capable of infecting human tumor cells independent of junction adhesion molecule-A. PLoS One 2012; 7:e48064; PMID:23110175; http://dx.doi.org/10.1371/journal.pone.0048064
  • Van den Wollenberg DJ, Dautzenberg IJ, Van den Hengel SK, Ros W, Nadif S, Hoeben RC. Development of replication-competent, expanded-tropism oncolytic reovirus carrying a heterologous transgene. 8th International Conference on Oncolytic Virus Therapeutics 2014
  • Harrington KJ, Vile RG, Melcher A, Chester J, Pandha HS. Clinical trials with oncolytic reovirus: moving beyond phase I into combinations with standard therapeutics. Cytokine Growth Factor Rev 2010; 21:91-8; PMID:20223697; http://dx.doi.org/10.1016/j.cytogfr.2010.02.006
  • Chakrabarty R, Tran H, Fortin Y, Yu Z, Shen SH, Kolman J, Onions D, Voyer R, Hagerman A, Serl S, et al. Evaluation of homogeneity and genetic stability of REOLYSIN (pelareorep) by complete genome sequencing of reovirus after large scale production. Appl Microbiol Biotechnol 2014; 98:1763-70; PMID:24419798; http://dx.doi.org/10.1007/s00253-013-5499-0
  • Jolly DJ, Ibanez CE, Ostertag D, Robbins JM, Kasahara N, Pertschuk D, Gruber HE. Toca 511: update on the use of Retroviral Replicating Vector as an anti-tumor agent in preclinical and clinical trials. 8th International Conference on Oncolytic Virus Therapeutics 2014
  • Solly SK, Trajcevski S, Frisen C, Holzer GW, Nelson E, Clerc B, Abordo-Adesida E, Castro M, Lowenstein P, Klatzmann D. Replicative retroviral vectors for cancer gene therapy. Cancer Gene Ther 2003; 10:30-9; PMID:12489026; http://dx.doi.org/10.1038/sj.cgt.7700521
  • Logg CR, Logg A, Tai CK, Cannon PM, Kasahara N. Genomic stability of murine leukemia viruses containing insertions at the Env-3' untranslated region boundary. J Virol 2001; 75:6989-98; PMID:11435579; http://dx.doi.org/10.1128/JVI.75.15.6989-6998.2001
  • Hlavaty J, Jandl G, Liszt M, Petznek H, Konig-Schuster M, Sedlak J, Egerbacher M, Weissenberger J, Salmons B, Gunzburg WH, et al. Comparative evaluation of preclinical in vivo models for the assessment of replicating retroviral vectors for the treatment of glioblastoma. J Neurooncol 2011; 102:59-69; PMID:20623247; http://dx.doi.org/10.1007/s11060-010-0295-5
  • 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; PMID:22547150; http://dx.doi.org/10.1038/mt.2012.83
  • Kawasaki Y, Tamamoto A, Takagi-Kimura M, Maeyama Y, Yamaoka N, Terada N, Okamura H, Kasahara N, Kubo S. Replication-competent retrovirus vector-mediated prodrug activator gene therapy in experimental models of human malignant mesothelioma. Cancer Gene Ther 2011; 18:571-8; PMID:21660062; http://dx.doi.org/10.1038/cgt.2011.25

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.