Skip to main content
Erschienen in: Journal of Hematology & Oncology 1/2020

Open Access 01.12.2020 | Review

Oncolytic viruses for cancer immunotherapy

verfasst von: Otto Hemminki, João Manuel dos Santos, Akseli Hemminki

Erschienen in: Journal of Hematology & Oncology | Ausgabe 1/2020

Abstract

In this review, we discuss the use of oncolytic viruses in cancer immunotherapy treatments in general, with a particular focus on adenoviruses. These serve as a model to elucidate how versatile viruses are, and how they can be used to complement other cancer therapies to gain optimal patient benefits. Historical reports from over a hundred years suggest treatment efficacy and safety with adenovirus and other oncolytic viruses. This is confirmed in more contemporary patient series and multiple clinical trials. Yet, while the first viruses have already been granted approval from several regulatory authorities, room for improvement remains.
As good safety and tolerability have been seen, the oncolytic virus field has now moved on to increase efficacy in a wide array of approaches. Adding different immunomodulatory transgenes to the viruses is one strategy gaining momentum. Immunostimulatory molecules can thus be produced at the tumor with reduced systemic side effects. On the other hand, preclinical work suggests additive or synergistic effects with conventional treatments such as radiotherapy and chemotherapy. In addition, the newly introduced checkpoint inhibitors and other immunomodulatory drugs could make perfect companions to oncolytic viruses. Especially tumors that seem not to be recognized by the immune system can be made immunogenic by oncolytic viruses. Logically, the combination with checkpoint inhibitors is being evaluated in ongoing trials. Another promising avenue is modulating the tumor microenvironment with oncolytic viruses to allow T cell therapies to work in solid tumors.
Oncolytic viruses could be the next remarkable wave in cancer immunotherapy.
Hinweise

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
5-FC
5-fluorocytosine
ADCC
Antibody-dependent cell-mediated cytotoxicity
ATAP
Advancer Therapy Access Program
BCG
Bacillus Calmette–Guérin
CAR
Chimeric antigen receptor
CART
Chimeric antigen receptor T cells
CEA
Carcinoembryonic antigen
CXADR
Coxsackie- and adenovirus receptor
DAMP
Damage-associated molecular pattern
DCs
|Dendritic cells
DNA
Deoxyribonucleic acid
EGFR
Epidermal growth factor receptor
EMA
European Medicines Agency
FAP
Familial adenomatous polyposis
FDA
Food and Drug Administration
GMCSF
Granulocyte macrophage colony stimulating factor
hTERT
Human telomerase reverse transcriptase
IFN
Interferon
IL-2
Interleukin-2
MDSCs
Myeloid-derived suppressor cells
NK
Natural killer cells
PAMP
Pathogen-associated molecular pattern
PBMCs
Peripheral blood mononuclear cells
PSA
Prostate-specific antigen
RGD
Arginylglycylaspartic acid
TAMs
Tumor-associated macrophages
TCR
T cell receptor
TGFb
Transforming growth factor beta
TIL
Tumor-infiltrating lymphocyte
TNFa
Tumor necrosis factor alpha
Treg
Tegulatory T cell

Background

Cancer is becoming a leading cause of death globally. Eighteen million new cancers are diagnosed every year causing almost 10 million deaths (International Agency for Research on Cancer (IARC), 2019). There has been significant progress in the prevention and diagnosis of cancer, but the incidence and mortality are still increasing [13]. Conventional therapies, such as surgery, chemotherapy, hormonal therapies, targeted therapies or radiation therapy, deliver limited durable responses in a great majority of patients with advanced cancers [4, 5]. Hematological and testicular malignancies are some of the few exceptions where current therapies can be curative even in metastatic cases [68].
Cancer is a genetic disease and it represents a range of manifestations. The principles of tumorigenesis are however similar across different tumors and relatively well characterized. In brief, frequent mutations occur during cell divisions or due to exogenous factors such as radiation or other carcinogens. Most of these mutations are corrected by specialized intracellular proteins. If such mechanisms are unsuccessful, mutated cells are generally cleared by apoptosis. The vast majority of mutations do not help the cell to gain cancerous properties (passenger mutations). In contrast, driver mutations provide exclusive abilities to tumor cells, such as cell death resistance or metastatic capacity, for example [9, 10]. Most of these mutated cells are, however, recognized by our immune system and destroyed before clinical detection. Accumulating evidence supports the notion that a dysfunctional immune system is intimately associated with tumor development, progression, and recurrence [11]. Also known as immunosuppression, this phenomenon is actively propagated by cancer cells either directly or through the tumor microenvironment [12]. This understanding has galvanized the interest in the development of immunotherapies, which aims at modifying and activating immune cells to attack cancer cells. The approach is rational as our immune system has been trained to detect, destroy, and memorize non-self patterns. By definition, all cancer cells have multiple mutations causing non-self structures that can potentially be detected by our immune system [9].
The concept of immunotherapy has been acknowledged already for centuries (Table 1). The relationship between microbial infections and spontaneous tumor regressions has been reported several times in the literature [13]. Probably the first evidence was the Ebers papyrus (1550 bc), one of the oldest and most important medical documents of ancient Egypt. The physicians of Egyptian pharaoh Imhotep (2600 bc) used poultice, followed by incision, for the treatment of tumors. This facilitated the development of infections which helped to cause regression of tumors [14]. In 1320, Peregrine Laziozi was affected by cancer of the tibia requiring amputation. Unfortunately, local recurrence and progression were later observed and the tumor finally grew through his skin causing infection. Later, to everyone’s astonishment, the tumor disappeared and no relapse was observed. This phenomenon is today known as the St. Peregrine tumor [15].
Table 1
A timeline including some key steps in development of cancer treatments
Other cancer treatments*
 
Cancer Immunotherapy
Surgery
2600 BCE
Use of poultice (pharaoh Imhotep’s physicians)
Surgery under ether anesthesia
1840s CE
Purposeful infection of tumors
Radiotherapy
1890s
Coley’s toxins (deactivated bacteria) were injected to tumor
Hormonal therapy (estrogen, castration), chemotherapy (nitrogen mustard, antifolates)
1900–1940s
Case reports of tumor regression after natural viral infections
Linear accelerator for radiotherapy, combination chemotherapy
1950s–1970s
Hundreds of case series treating cancer with multiple viruses (e.g., varicella, measles, vaccinia, West Nile, adenovirus, mumps)
BCG adopted in bladder cancer
Stereotactic radiotherapy, antiestrogens
1980s
Adoptive T cell transfer, cytokine therapies (e.g., IFN-alpha and IL-2)
Mini-invasive surgery, monoclonal antibodies (rituximab, trastuzumab)
1990s
HD-IL-2 approved by the FDA
Antiangiogenic therapies (bevacizumab), kinase inhibitors (imatinib)
2000s
First oncolytic adenovirus (H101) approved in China
Small molecular inhibitors of various proteins
2010–
Cellular immunotherapy (sipuleucel-T, TCR, CART), six different checkpoint inhibitors, oncolytic virus (T-vec)
*Many treatments in this column have also immunological properties (e.g., rituximab, trastuzumab, chemotherapy, and radiation therapy)
In the seventeenth and eighteenth centuries, various forms of immunotherapy became widely used. In the eighteenth and nineteenth centuries, septic dressings enclosing ulcerative tumors were sometimes used for cancer treatments. Surgical wounds could be deliberately left open to enable the development of infection as purulent infections were suggested helpful [15]. One of the more detailed case series, indicating several responses, was reported by surgeon William B. Coley. He treated cancer patients with a bacterial lysate (heat-killed Streptococcus pyogenes and Serratia marcescens), also known as the Coley’s toxin [16].
Reports of viruses having therapeutic benefits in cancer started appearing early last century with multiple reports of leukemia patients becoming disease-free after viral infections [17]. Typically, the reported patients were young and the remissions were short-lived lasting for 1 or 2 months [18]. These observations did not go unnoticed by the medical community, who subsequently begun utilizing viruses for the treatment of cancer. Especially during the 1950s and 1960s, multiple wild type viruses (e.g., hepatitis, Epstein-Barr, West Nile, Uganda, dengue, yellow fewer) were used to treat different cancers in hundreds of case series. Results were variable and occasionally poorly documented [17]. However, during this time, it was becoming clear that most wild type viruses lacked efficacy or safety. Some of the more promising results with tolerable side effects were associated with adenoidal-pharyngeal-conjunctival virus [1922], nowadays known as the adenovirus. For example, in 1956, 30 women with advanced epidermoid carcinoma of the cervix were treated with adenovirus. Intra-arterial, intravenous, and intratumoral administration was used. Within 10 days, two-thirds of the patients showed necrosis in their tumors and most remarkably, it appeared to be restricted to the cancerous tissue. No safety problems were reported in the use of this wild type virus, suggesting a degree of natural tumor tropism.
Generally, however, virotherapy received little attention and in the 1970s and 1980s, the regulatory aspects of clinical trials with living pathogens became stricter. In turn, chemotherapies, radiation therapy, hormonal therapy, targeted, and antiangiogenic therapies all became mainstream. It took more than three decades before viruses re-emerged, this time as “oncolytic viruses” [23, 24]. An oncolytic virus is a virus that infects and lyses (breaks down) cancer cells but not normal cells. Oncolytic viruses can occur naturally or can be made in the laboratory by modifying natural viruses. These modifications started a new era of less toxic cancer targeted virus-based therapies [25].
Rapid increases in molecular biotechnology techniques provided means to develop novel strategies to harness the immune system for cancer therapy. Currently, a number of approaches, including adoptive cell therapies, monoclonal antibodies, checkpoint inhibitors, and oncolytic viruses constitute the most prominent advancements in cancer treatment due to the capacity to provide durable and effective clinical responses in cancer patients [13]. Conventional treatments such as radiation and chemotherapy treatments also seem to have immunomodulatory effects not recognized before [26, 27].
However, it is essential to note that presently therapeutic benefits are restricted to a limited fraction of patients treated with immunotherapy. In particular, solid cancers generally contain a suppressive tumor microenvironment that inhibits T cell activity and supports tumor progression [28]. In addition, new immunotherapy treatments have led to the occurrence of new immunological adverse events, including cytokine storm and autoimmune events. Considering these challenges, further alterations to these therapeutic strategies are needed. In addition to new immunological treatment strategies, we also need better understanding of individual immune environments to provide maximal patient benefit.
Our aim in this review is to present the current possibilities that oncolytic viruses have to offer. We concentrate on adenoviruses that are the most widely studied virus type. We start by describing general adenovirus biology and describe then some typical modifications that can be used to generate better anti-cancer capabilities. Finally, we describe clinical studies with oncolytic adenoviruses and describe three different types of oncolytic viruses that are already regulatory approved to treat cancer. The aim is to provide a general overview on the field and we acknowledge that many viral constructs, especially with preclinical studies only, are not covered. However, recruiting or completed oncolytic adenovirus trials (source clinicaltrials.​org) give a good general view of where the field is today. Finally, we conclude by addressing some future perspectives.

Main text

Structure, function, and immunogenic cell death: adenovirus as an example

Adenovirus biology has been investigated in detail and it is well understood. Adenoviruses can be used as an example virus when describing oncolytic viruses as a whole. While there are naturally many differences with different viruses, many factors are similar to some degree, however not all viruses are naturally oncolytic.
To date, 57 distinct adenovirus serotypes are described and classified into 7 subgroups: A to G [29, 30]. While adenoviruses trigger common-flu type infections, these represent one of the most versatile platforms for cancer therapy. In particular, serotype 5 (group C) is the most commonly used backbone for oncolytic virus design [29]. Its structure encompasses an icosahedral shaped capsid (composed mainly of hexon, penton, and fiber proteins) surrounding a non-enveloped double-stranded DNA [31]. Adenoviruses have the ability to infect cells independently of their division status.
Infection of tumor cells initiates with the attachment of the virus fiber knob to receptors located on the surface of tumor cells. This interaction is mediated by different receptors depending on the serotype of the virus. For example, serotype 5 adenoviruses bind preferentially to the coxsackie- and adenovirus receptor CXADR [32], while serotype 3 adenoviruses bind desmoglein-2, CD46 or CD80/86 [3335]. Some of these receptors are frequently found on cancer cells, while unfortunately some, such as CXADR, are downregulated in many advanced tumors [30]. A second interaction between the virus’ penton proteins and tumor cell integrins occurs, resulting in virus internalization [36]. Adenoviruses can use also several other receptors as shown in Table 2. A multistep process takes the virus DNA into the nucleus, where transcription of early viral proteins (E1–E4) starts. Following expression of late proteins, thousands of new viral progeny emerges, disrupting the cell membrane after a few days and the newly formed viruses infect new cells, until the immune system eventually stops this process [50].
Table 2
Entry receptors for adenoviruses [37]
Receptor
Ligand
Reference
Primary receptors
 CXADR (coxsackie-adenovirus receptor)
Ad5 knob (and other groups A, C, D, E, and F)
[38]
 CD46
Ad3 knob (and other groups B and D)
[39]
 Desmoglein-2
Ad3 knob (and other group B)
[40, 41]
 CD80/CD86
Ad3 knob (and other group B)
[33, 37]
Other receptors
 Cellular integrins
Ad capsid (penton base) or Arg-Gly-Asp (RGD) modifications
[42]
 Sialic acid
Ad 37, 8, 19a (group D)
[43]
 MHC-I (major histocompatibility complex class I)
Ad5
[44]
 Vascular adhesion molecule 1 (VCAM-1)
Ad5
[45]
 Heparan sulfate glycosaminoglycans (HS-GAGs)
Ad5
[46]
 Dipalmitoyl phosphatidylcholine (DPPC)
Ad5 hexon
[47]
 Blood coagulation factor F(X)
Ad 5 hexon
[48, 49]
Although not discovered until the treatment of patients, such cell lysis is a highly immunogenic process [51]. This factor is of pivotal importance considering that most cancers seem to be able to hide from our immune system. Immunogenic cell death reveals multiple tumor-associated antigens for presentation to the immune system via activated mature dendritic cells. High numbers of virus genomes activate immunological danger signaling through damage-associated molecular pattern (DAMP) and pathogen-associated molecular pattern (PAMP) receptors. These processes form a recipe that retargets the adaptive immune system, including cytotoxic CD8+ T cells and helper CD4+ T cells, towards the tumor, thus lifting local immunosuppression [52]. Of note, anti-adenoviral T cell immunity is polyfunctional rendering increased quality to the overall ongoing antitumor response [53]. Simultaneously, adenovirus infection also mediates activation of natural killer cells further contributing for the antitumor immune response [54]. The generation of antiviral antibodies equally stimulates the response by triggering antibody-dependent cytotoxicity [55] (Fig. 1).

Modern oncolytic viruses

Today, adenoviruses, herpes viruses, measles viruses, coxsackie viruses, polioviruses, reoviruses, poxviruses, and Newcastle disease viruses, among others, are some of the oncolytic viruses under preclinical and clinical development for cancer therapy [56]. Tumor-preferential replication can be “natural” given the defective viral sensing mechanisms of most cancer cells [57]. Some of the cancer cells also harbor increased expression of viral entry receptors, and some viruses do not appear to need specific receptors for entry [58]. Abnormal function of intracellular signaling pathways such as interferon can be exploited by some viruses [58]. Like many other viruses, adenoviruses are naturally prone to replicate aggressively in tumor cells and, their wild type versions could in theory be used in cancer treatments, as was done in historical series [17, 1922]. However, better patient outcomes are expected by rational design of viruses rendering them tumor selective (“oncolytic”). In recent years, adenoviruses have been extensively altered to merge high antitumor potency with minimal toxicity [25].
Existing molecular biology techniques allow us to (a) select entry receptors highly expressed on tumors, (b) refine safety by restricting replication to cancer cells, and (c) insert specific therapeutic transgenes for increased efficacy. These approaches are discussed below and summarized in Table 3.
Table 3
Examples of viral modifications in oncolytic adenoviruses
Modification
Reference
Enhancing tumor tropism
 RGD modification
[5961]
 5/3 chimerism
[62, 63]
 Fully serotype 3, Ad3/Ad11p
[30, 64, 65]
Safety, restricting replication to cancer cells
 E1A gene 24-base pair deletion
[61, 66, 67]
 hTERT promoter
[64, 68]
 p53 promoter
[69]
 CEA promoter
[70]
 PSA promoter
[71]
 E2F promotor
[72]
 Cox2l promoter
[73]
Transgenes, enhancing efficacy
 Cytosine deaminase and thymidine kinase
[74]
 Ganciclovir and/or 5-fluorocytosine (5-FC) prodrugs
[75]
 GMCSF
[66, 67, 72, 76, 77]
 CD40L
[78, 79]
 hNIS
[80]
 TNFalpha and interleukin 2
NCT04217473
 CD40L and 41BBL
NCT03225989
 PH20 hyaluronidase
NCT03284268
 Anti-CTLA4
[81]
 IL-12 and decorin
[82]
 OX40L
[83]
 EGFR
[84]
 FR-a
[85]
 FAP
[10]
 CD44v6
[86]

Enhancing tumor tropism

Effective entry of oncolytic viruses into tumor cells is a prerequisite for subsequent oncolysis. Conversely, low receptor expression can be a limiting factor. To avoid the conundrum of low expression of CXADR on tumors cells, serotype 5 adenoviruses can be modified to contain arginylglycylaspartic acid (RGD) peptides in their fiber knobs. Such modification has been shown useful to increase efficacy and reduce the toxicity of adenoviruses [59, 60]. Clinical implementation of such modified oncolytic adenoviruses, such as DNX-2401 (an serotype 5 adenovirus with an RGD modification), has shown promising results in a phase I clinical study where 20% of glioma patients showed durable responses [61].
Likewise, replacement of the serotype 5 fiber knob with one belonging to the serotype 3 has provided substantial improvements in antitumor efficacy, while retaining the appealing systemic kinetics and safety of the ubiquitous serotype 5 [62, 63]. Capsid modification (e.g., 5/3 chimerism) allows partial overcoming of pre-existing neutralizing antibodies against Ad5 [72]. Of note, the antibody question is complex, and in fact baseline neutralizing antibody titers have not prevented antitumor efficacy in humans [76, 87]. The large amounts of virus that are produced by tumors into the bloodstream might simply overcome neutralizing antibodies producible into blood. In wild type adenovirus infections, only small amounts of adenovirus enter the blood. Neutralizing antibodies are designed to block such virus, not the huge numbers produced by tumors infected with oncolytic virus. Nevertheless, in epidemiological analysis, lack of antibodies at baseline impacted survival statistically significantly [76], but not in a clinically meaningful manner, as responses and long survival could be seen regardless of baseline antibody titers. Interestingly, sequential intravenous treatments by changing the virus or serotype might make a difference [88].
As expression of the Ad5 receptor CXADR appears limiting for efficacy in the context of advanced tumors, fully serotype 3-based oncolytic adenoviruses have been constructed [64]. This virus enters through non-CXADR-mediated mechanisms including desmoglein 2, which is highly expressed in advanced solid tumors [30, 35, 64, 78]. Good safety and signs of efficacy were seen in patients treated with a fully serotype 3 adenovirus [64]. Furthermore, a direct evolution method [65] was used to generate an Ad3/Ad11p chimeric virus which has been used in multiple trials under the name ColoAd1 [89].

Safety, restricting replication to cancer cells

To achieve efficient dissemination of input virus and to minimize virus-related adverse events, adenoviruses have been modified to achieve increased tumor selectivity. The strategies employed include transcriptional control of adenovirus early proteins such as the E1A or E1B. An E1A gene 24-base pair deletion produces a mutated E1A protein which cannot bind to the retinoblastoma protein, thus preventing healthy cells from entering in synthesis (“S”) phase. This blocks adenovirus DNA replication in quiescent normal tissues [61, 66]. In contrast, the replicative potential remains intact in tumor cells because ubiquitous defects in the p16/Rb pathway ensure that cancer cells permanently stay in synthesis phase [90].
The fact that tumor cells contain several active oncogenes led to the realization that their resulting proteins could be harnessed to control transcription of adenovirus DNA. For example, telomerase activity is a known feature of cancer cells, while activity in healthy cells is minimal [30, 91]. Therefore, adenovirus replication has been successfully placed under the control of a human telomerase reverse transcriptase (hTERT) promoter showing antitumor efficacy in advanced cancers [64, 68]. Similarly, p53 [69], carcinoembryonic antigen (CEA) [70] and prostate-specific antigen (PSA) [71] have been utilized to control expression of early adenovirus proteins. Due to Rb/p16 pathway defects, cancer cells feature high levels of intracellular free E2F, which can be used for tumor specificity when the E2F promotor is inserted to control viral replication [72].

Transgenes, enhancing efficacy

The nature of viruses allows them to hijack the host cell to produce virus proteins. This allows therapeutic exploitation with the insertion of therapeutic transgenes into the adenovirus genome. Before the recognition of oncolytic adenoviruses as immunotherapy, one of the most common modifications was the insertion of the cytosine deaminase and herpes simplex-derived thymidine kinase [74]. The combined administration of oncolytic adenoviruses with ganciclovir and/or 5-fluorocytosine (5-FC) prodrugs caused tumor cell death due to their conversion into cytotoxic compounds by transduced tumor cells [75].
More recently, the increased recognition of the immune system as an important component in the efficacy of oncolytic viruses led researchers to perceive oncolytic adenoviruses as potent vehicles for immune factors. Adding a granulocyte macrophage colony stimulating factor (GMCSF) cytokine transgene into the adenoviral genome is a commonly used modification. In this approach, virus replication is accompanied by GMCSF production, which results in the recruitment and maturation of dendritic cells (DCs), and subsequent priming of T cells with tumor-associated antigens released by oncolysis [92]. CGTG-102 (previously Ad5/3-D24-GMCSF, currently ONCOS-102), is an oncolytic adenovirus expressing GMCSF. Patient data confirms this notion, with reported increases in peripheral levels of T cells against tumor-associated antigens [66]. These finding suggest dendritic cell priming in humans as predicted by the established mechanism of action of GMCSF [93]. The increased CD8+ T cell infiltration found in tumor biopsies after treatment of advanced cancer patients with ONCOS-102 underlines the immunological potency of this approach [94]. However, the pleiotropic effects of GMCSF may endanger antitumor immunity as the cytokine can unwantedly stimulate myeloid-derived suppressor cells (MDSCs) and tumor-associated macrophages (TAMs), both known to inhibit T cell and natural killer cell (NK) activity [95]. However, emerging human data suggests that GMCSF producing viruses might be safe and effective [76, 77].
Beyond GMCSF, combined expression of IL-12 and decorin in an oncolytic adenovirus allowed the recovery of antitumor immunity in a poorly immunogenic murine breast cancer model, via cytotoxic T cell infiltration and transforming growth factor beta (TGFb) reduction [82]. Coexpression of CD40L and 4-1BBL by an oncolytic adenovirus has also shown promising results, due to its ability to promote the destruction of pancreatic tumors, through repolarization of the tumor microenvironment. Such polarization enabled release of T cell attractants and immune stimulatory cytokines, allowing potent antitumor T cell responses [96]. Production of OX40L mediated by an oncolytic adenovirus (d24-RGDOX) promoted increased tumor control via highly functional effector T cells and epitope spreading [83].
Also, antibodies can be inserted as transgenes to enhance the efficacy of oncolytic virotherapy. For example, anti-CTLA4, a checkpoint inhibitor, has been successfully inserted in an oncolytic adenovirus platform. Its usage in murine models and ex vivo cultures of cancer patient peripheral blood mononuclear cells (PBMCs) resulted in increased antitumor activity of T cells [81]. More recently, dual targeted antibodies targeting T cells and cancer-specific cell surface antigens such as epidermal growth factor receptor (EGFR) [84], FR-a [85], familial adenomatous polyposis (FAP) [10] and CD44v6 [86], have demonstrated promising preclinical results [84].
Moreover, also other approaches have been studied. These include arming with fusogenic molecules, antibodies, T cell engagers, and ion channels capable of concentrating radioiodine. While in these cases, the transgenes are not necessarily immunologically active, and the oncolytic platform results in immunostimulation. It is important to note that clinical data suggests that oncolytic adenovirus single-agent efficacy has often been somewhat limited. Several barriers that affect oncolytic adenovirus therapies have been suggested. These include antiviral interferons, which can be produced by the tumor stroma even if the cancer cells themselves lack such ability [50]. Other reasons include stromal barriers, hypoxia, hyperbaric, necrotic, and acidic areas [97, 98]. However, some of these hurdles have been addressed in redesigned adenoviruses conditionally replicating in response to hypoxic factors or acidic tumor microenvironments [99].
Alternatively, oncolytic adenoviruses have been armed with hyaluronidase [100], an enzyme that degrades hyaluronic acid which hampers virus dissemination. Notably, treatment of a number of preclinical in vivo tumor models allowed increased antitumor efficacy. Neutralizing antibodies remain a concern for oncolytic immunotherapies. However, the use of bispecific adapters to retarget antiviral neutralizing antibodies can offer an attractive approach to increase the effectiveness of oncolytic adenovirus therapy [101]. The coating of oncolytic adenoviruses with tumor derivatives has been reported to allow for successful delivery of particles into the tumor with potent antitumor responses [102].

Advanced Therapy Access Program (ATAP)

Between 2007 and 2012, 290 advanced cancer patients were treated with 10 different oncolytic viruses (Table 4) totaling 821 treatments. A long-term follow-up of these patients has been published [5]. Treatments were given in the context of an individualized treatment program under the EU Advanced Therapies directive [5]. While many objective responses were seen, no definite conclusions regarding overall survival benefit can be drawn as no reliable control group was available. However, some case-control analyses that were performed suggest survival benefit [88].
Table 4
Viruses used in ATAP
 
Serotype
Main target receptor
Tumor specificity
Arming
Ad5-D24-GMCSF [92]
5
Coxsackie virus and adenovirus receptor
24 bp deletion in E1A1)
GMCSF
Ad5-RGD-D24 [103]
5
Alpha-v-beta integrins
24 bp deletion in E1A1)
No
Ad5-RGD-D24-GMCSF [103]
5
Alpha-v-beta integrins
24 bp deletion in E1A1)
GMCSF
ICOVIR-7 [104]
5
Alpha-v-beta integrins
E2F1 promoter and 24 bp deletion in E1A1)
No
Ad5/3-Cox2L-D24 [73]
5
Desmoglein-2
Cox2L promoter and 24 bp deletion in E1A1)
No
Ad5/3-D24-GMCSF [88]*
5
Desmoglein-2
24 bp deletion in E1A1)
GMCSF
Ad5/3-hTERT-hCD40L [73]
5
Desmoglein-2
hTERT promoter2)
CD40L
Ad5/3-E2F1-D24-GMCSF [105]
5
Desmoglein-2
E2F1 promoter and 24 bp deletion in E1A 1)
GMCSF
Ad5/3-D24-hNIS [80]
5
Desmoglein-2
24 bp deletion in E1A1)
hNIS
Ad3-hTERT-E1A [64]
3
Desmoglein-2
hTERT promoter2)
No
*Ad5/3-D24-GMCSF, also known as CGTG-102, and later renamed ONCOS-102, has been subsequently used in several phase 1 and phase 2 clinical trials (www.​targovax.​com)
1)Replication in cells with a deficient Rb/p16 pathway (a hallmark of cancer)
2)Replication in cells with active telomerase (a hallmark of cancer)
While taking into account the limitations of nonrandomized data, some interesting findings emerged [5, 51, 64, 66, 80, 88, 92, 103, 104, 106]. One of the most important observations was that all of the administered viruses appeared quite safe in patients with advanced cancer. Good tolerability was seen across different serotype viruses, including various capsid modifications, and different immunological arming devices (i.e., GMCSF or CD40L). Concomitant low-dose cyclophosphamide and temozolomide were also well tolerated [107]. The former was used to reduce regulatory T cells. The latter aimed at increased induction of autophagy in infected cancer cells, as this appears to be an important mediator of oncolytic cell death. Moreover, virus replication could be increased by concomitant calcium channel blockage [108]. No treatment-related patient deaths were observed [63]. Typical flu-like symptoms, such as fever and fatigue, were observed in most patients a few days after treatment. These findings were confirmed in multiple subsequent clinical trials [61, 94, 109]. Flu-like symptoms and fever could be effectively reduced with acetaminophen (paracetamol).

Systemic delivery of oncolytic virus

Biodistribution studies done with adenoviruses in rodents are unreliable as most animals lack entry receptors or their organ distribution is different from humans. From ATAP patients, we have been able to collect important information about virus biodistribution in humans [63]. Many humans have neutralizing antibodies against different adenovirus serotypes, although often at low baseline titer [64]. However, as hundreds of billions of viruses are given in a typical treatment, pre-existing antibodies may be unable to completely block intravenous delivery. In subsequent intravenous treatments with the same serotype, the situation is more unclear, which is one of the reasons intratumoral delivery is commonly used with oncolytic viruses. Of note, it has been proposed that antiviral immunity helps generate antitumor immunity [66].
It has been established in humans that adenovirus is able to travel through blood to metastases despite neutralizing antibodies [63]. For some viruses, the mechanism appears to relate to binding to blood cells [110]. Interestingly, adenoviruses in blood (qPCR data) were most often found from blood clots, while some patients had significant amount of virus in the serum compartment [64]. Also cancer patient’s antibody response varied [64]. Treatment responses or long survival are seen regardless of neutralizing antibody titers, although it should be noted that most patients were treated intratumorally [76].
Interestingly, we treated seven patients with the serotype 3 adenovirus using only intravenous administration. Signs suggesting virus replication were seen, including prolonged and/or rising virus titers in the blood. This was seen also with patients who had pre-existing antibodies against the virus. Also, 5 of the 6 evaluable patients showed signs of possible benefit. This data indicates that viruses might be able to enter tumors also via the intravenous route [64]. This was later confirmed in an autopsy study where non-injected tumors were shown to have oncolytic adenovirus [63].

Viruses that have already received regulatory approval for the treatment of cancer

From a clinical standpoint, the use of viruses for cancer treatment in the modern era is in its infancy. Initially, wild type viruses were used, but this approach could result in adverse events caused by virus replication in normal tissues. Nevertheless, Rigvir (an ECHO-7 virus) [111], an oncolytic picornavirus with some innate tumor selectivity, was the first approved oncolytic virotherapy product for cancer approved 2004 in Latvia and later in a few other countries. The second oncolytic virus was rationally designed for tumor selectivity. Named H101 (Oncorine), this adenovirus has been used in China since 2005 for the treatments of solid tumors [112]. Of note, both of these viruses lack arming devices.
The acknowledgement that repurposing the immune system to exert antitumor functions could provide a promising approach to treat cancer enabled scientists to employ the immunological capabilities of oncolytic viruses [113]. For example, the addition of immunological transgenes such as the GMCSF has been a popular approach. Talimogene laherparepvec (also known as T-vec, Imlygic®) is a herpes simplex-1 virus encoding for GMCSF, and was one of the first oncolytic viruses designed to provide an immunological boost. Its clinical application eventually led to a randomized phase III clinical trial (OPTiM). In this trial stage IIIB/C and IV metastatic, unresectable melanoma patients receiving intratumoral T-vec had a 19.3% durable response rate of which more than 80% were complete responses [114]. The fact that subcutaneous administration of GM-CSF offered inferior efficacy (1.4% durable response rate, 0.7% complete response) led to approval by the Food and Drug Administration (FDA) in 2015, followed by European Medicines Agency (EMA) [114].
This landmark approval in Western countries encouraged optimism in the medical community to continue developing and improving oncolytic viruses for cancer therapy, including adenoviruses. Later, in a similar patient population, pembrolizumab was combined with the virus leading to responses in 62% of the patients, of which 33% were complete. As expected, a high presence of cytotoxic T cell infiltration was observed in the tumors following treatment [115]. A recent phase II clinical trial in advanced melanoma patients demonstrated that T-vec increased the response rate of ipilimumab as compared to ipilimumab alone (38% vs 18%, respectively) [116]. Of note, and in contrast to combinations of checkpoint inhibitors [117], adverse events were not compounded. This suggests that oncolytic viruses can be combined with checkpoint inhibition without a problematic decrease in safety.

Oncolytic adenovirus trials

At the time of writing this review (March 2020), we did a search on clinicaltrials.​org, resulting in 101 trial results. Limiting the search to “oncolytic adenovirus” phases I–II trials, we came up with 41 results; of these, 10 were completed and 15 recruiting. Sixteen different oncolytic viruses were used in these completed or recruiting trials. No active phase III clinical trials were found. Interestingly, 6 out of the 16 viruses have been posted recently, during 2019–2020, indicating growing interest and available funding for oncolytic adenovirus trials (Table 5).
Table 5
Sixteen oncolytic adenoviruses used in phase I–II trials that have been completed or recruiting (Mar-2020 clinicaltrials.​org).
First posted
Last update
Oncolytic adenovirus (transgene)
Study title(s)
Indication(s)
NCT# identifier
2012
2019
ONCOS-102a (GMCSF)
Completed: ONCOS-102 (Previously CGTG-102) for Therapy of Advanced Cancers
Recruiting: (1) A Pilot Study of Sequential ONCOS-102, an Engineered Oncolytic Adenovirus Expressing GMCSF, and Pembrolizumab in Patients With Advanced or Unresectable Melanoma Progressing After Programmed Cell Death Protein 1 (PD1) Blockade (2) A Phase I/II, Safety Clinical Trial of DCVAC/PCa and ONCOS-102 in Men With Metastatic Castration-resistant Prostate Cancer
Solid tumors, melanoma, and prostate cancer
NCT03003676, NCT03514836
2012
2015
Delta-24-rgd
Completed: Safety Study of Replication-competent Adenovirus (Delta-24-rgd) in Patients With Recurrent Glioblastoma
Glioblastoma
NCT01582516
2013
2020
DNX-2401b
Completed: (1) Oncolytic Adenovirus, DNX-2401, for Naive Diffuse Intrinsic Pontine Gliomas (2) DNX-2401 With Interferon Gamma (IFN-γ) for Recurrent Glioblastoma or Gliosarcoma Brain Tumors (TARGET-I) (3) Virus DNX2401 and Temozolomide in Recurrent Glioblastoma (D24GBM)
Recruiting: Oncolytic Adenovirus DNX-2401 in Treating Patients With Recurrent High-Grade Glioma
Gliomas
NCT03178032, NCT02197169, NCT01956734, NCT03896568
2013
2016
CELYVIR
Completed: Safety and Efficacy of Repeated Infusion of CELYVIR in Children and Adults With Metastatic and Refractory Tumors.
Solid tumors
NCT01844661
2013
2017
ICOVIR-5
Completed: Phase I Endovenous Administration of Oncolytic Adenovirus ICOVIR-5 in Patients With Advanced or Metastatic Melanoma
Melanoma
NCT01864759
2014
2019
VCN-01 (PH20 hyaluronidase)
Completed: A phase I Dose Escalation Study of Intratumoral VCN-01 Injections With Gemcitabine and Abraxane in Patients With Advanced Pancreatic Cancer
Recruiting: Safety, Tolerability, and Efficacy of VCN-01 With Durvalumab in R/M Head and Neck Squamous Cell Carcinoma
Pancreatic cancer and head and neck
NCT03284268, NCT03799744
2014
2020
ColoAd1c
Completed: Mechanism of Action Trial of ColoAd1 (MOA)
Recruiting: Chemoradiation With Enadenotucirev as a Radiosensitiser in Locally Advanced Rectal Cancer (CEDAR)
Colon, non-small cell lung, bladder, and renal cell
NCT02053220
2015
2019
CG0070
(GMCSF)
Completed: Safety and Efficacy of CG0070 Oncolytic Virus Regimen for High Grade NMIBC After BCG Failure (BOND2)
Superficial bladder cancer
NCT02365818
2020
2020
TILT-123d
(TNFalpha and IL-2)
Recruiting: TNFalpha and Interleukin 2 Coding Oncolytic Adenovirus TILT-123 During TIL Treatment of Advanced Melanoma
Melanoma
NCT04217473
2016
2020
LOAd703
(CD40L and 41BBL)
Recruiting: (1) Trial Investigating an Immunostimulatory Oncolytic Adenovirus for Cancer (2) LOAd703 Oncolytic Virus Therapy for Pancreatic Cancer (3) A Phase I/II Trial Investigating LOAd703 in Combination With Atezolizumab in Malignant Melanoma
Pancreatic, ovarian, biliary, colorectal, and melanoma
NCT03225989, NCT02705196, NCT04123470
2019
2020
ORCA-010
Recruiting: First in Man Clinical Study to Evaluate Safety and Tolerability of an Oncolytic Adenovirus in Prostate Cancer Patients.
Prostate cancer
NCT04097002
2017
2017
Ad5-yCD/mutTKSR39rep-hIL12 (IL-12)
Recruiting: Phase 1 Trial of Interleukin 12 Gene Therapy for Metastatic Pancreatic Cancer
Pancreatic cancer
NCT03281382
2017
2019
NSC-CRAd-Survivin-pk7
Recruiting: Neural Stem Cell Based Virotherapy of Newly Diagnosed Malignant Glioma (neural stem cells loaded with NSC-CRAd-Survivin-pk7)
Glioma
NCT03072134
2019
2020
NG-641 (CXCL9/CXCL10/IFNα)
Recruiting: First in Human Study With NG-641, an Oncolytic Transgene Expressing Adenoviral Vector
FAP-TAc antibody together with an immune enhancer module (CXCL9/CXCL10/IFNα).
Epithelial tumors
NCT04053283
2016
2019
ADV/HSV-tk (HSV-tk)
Recruiting: SBRT and Oncolytic Virus Therapy Before Pembrolizumab for Metastatic TNBC and NSCLC (STOMP)
Breast and lung
NCT03004183
2019
2020
NG-350A (anti-CD40 ab)
Recruiting: First in Human Study of NG-350A (an Oncolytic Adenoviral Vector Which Expresses an Anti-CD40 Antibody) (FORTITUDE)
Epithelial tumors
NCT03852511
aONCOS-102 was previously used in our ATAP treatment series as Ad5/3-D24-GMCSF and CGTG-102 [88]
bPreviously known as Delta-24-rgd
cAlso knowns as enadenotucirev
dTILT-123 is a double-armed virus designed for activation of T cells. It was designed based on human data from ATAP

Use of oncolytic viruses in hematologic malignancies

Observations of hematological cancer regressions following virus infection were seen early last century [17, 18]. Of note, response evaluation was possible for hematological malignancies, by microscopy and cell counting, while this was not true for most solid tumors at the time. This might have caused overrepresentation of hematological cancer in early literature, due to observational bias. However, it cannot be denied that many reports suggested regression of hematological tumors after viral infections. As such, there is strong rationale to believe that oncolytic viruses could be used also in hematological malignancies. However, not all viruses are suitable for treatment of blood cell tumors. For example, adenovirus does not appear to be able to lyse white blood cells [118].
According to recent publications, growing interest towards oncolytic viruses is present also in the hematological field, as several viruses are being studied in preclinical settings [119]. However, only a few trials have been published [120]. A search from clinicaltrials.​gov did not reveal any results for trials on “hematologic neoplasm” and “oncolytic viruses” (March 2020). However, a search for “multiple myeloma” and “oncolytic viruses” found two studies.
To conclude, to date, there have been few trials with oncolytic viruses in hematological cancers. Some noteworthy efforts include early phase trials with reovirus [121], measles [122], and vesicular stomatitis virus (VSV) in multiple myeloma [120]. The currently recruiting VSV trial (NCT03017820) also includes patients with relapsed acute myeloid leukemia and T cell lymphoma. No hematological trials with oncolytic adenovirus were found.

Limitations and newer strategies to improve efficacy of oncolytic viruses

Arming with immunostimulatory cytokines has been one popular method to generate immunological synergy with the effects of oncolysis. Clinical benefit of this approach was seen in the phase III OPTiM trial where 1 in 6 patients achieved complete responses with the oncolytic virus talimogene laherparepvec. The median duration of these complete responses in the virus group was not reached and 8 out of 9 patients survived over 5 years [114].
Combining oncolytic viruses with T cell activating checkpoint inhibition can improve antitumor efficacy of oncolytic adenovirus therapy. Especially tumors with low amount of immunological cells—“cold tumors”—can be efficiently immune activated by oncolytic adenoviruses. This makes tumors “hot” and promotes the effects of checkpoint inhibitors [94, 123]. Clinical proof-of-concept for the efficacy of combining oncolytic viruses with checkpoint inhibitors has been presented [116, 124]. In a randomized phase II study (n = 198) a checkpoint inhibitor (ipilimumab) was combined with the oncolytic virus talimogene laherparepvec, showing objective responses in 39% of the patients, compared to 18% in the checkpoint inhibitor-only arm [116, 125].
In addition to melanoma, recent publications suggest clinical activity also in other cancers such as metastatic sarcoma where the same combination resulted in a 30% objective response rate in a single-arm trial [126]. Furthermore, the safety profile of these treatments has been good and oncolytic viruses do not seem to increase the rate of serious adverse events.
Although these combinations seem to yield increased potency and long-term benefits to some patients, not all benefit, and there is clearly a role for further improvement. Combining chemotherapy or radiation therapy in a rational way to improve treatment benefits, and even these conventional therapies seem to have an immunological component [27, 127]. However, clear clinical proof in support of this approach is currently limited.
Combining other therapies such as adoptive cell therapy or targeted therapies might also result in better treatments. However, we are still lacking much information about the immune effects in individual tumors. This insufficiency of knowledge makes it hard to understand which patients would benefit most of what kind of treatment combinations. This could be the greatest challenge in the field at the moment. Classic trial design is not well suited for understanding mechanisms on an individual tumor and patient level.

TILT-123 studies

An ideal cancer treatment should be so good that most patients would clearly benefit while side effects should be tolerable. Our own contribution to this quest is a novel oncolytic adenovirus, designed specifically with T cells in mind. Ad5/3-E2F-D24-hTNFa-IRES-hIL2 (TILT-123) is based on the well understood and safe adenovirus serotype 5, but its fiber knob has been changed to a serotype 3 knob for enhanced penetrance to tumor cells [128]. The replication of the virus is strictly limited to cancer cells by dual control (E2F promoter and D24 deletion) and the potency of the virus is optimized by two transgenes, which were selected in a data-driven manner [95, 129]. Chimeric 5/3 adenoviruses also represent the best native T cell stimulator, among clinically relevant oncolytic adenoviruses [130].
The combination that emerged as the best approach for recruiting and activating T cells was interleukin-2 (IL-2) and tumor necrosis factor alpha (TNFa) [129]. IL-2 is required for T cell growth and survival and TNFa is a potent inducer of T cell trafficking and tumor apoptosis [129, 131133]. In preclinical models, administration of cytokine-coding adenoviruses increased the antitumor efficacy of three forms of adoptive T cell therapy: T cell receptor (TCR)-engineered T cells [129], CAR T cell therapy [134], and tumor-infiltrating lymphocyte (TIL) therapy [128, 135].
This outcome resulted from the following: (1) improved infiltration of transferred cells induced by TNFa, (2) improved activity of cytotoxic T cells induced by IL-2, and (3) overall decrease of immune suppressive subsets including regulatory T cells (Tregs), MDSCs, and M2 macrophages in the tumor microenvironment [129]. Hamsters bearing pancreatic tumors treated with TILT-123 showed signs of improved antitumor efficacy as compared to animals receiving TIL therapy or TILT-123 alone [128]. Combination of IL-2 and TNFa coding adenoviruses and anti-PD-1 therapy also fostered long-term 100% survival in preclinical models bearing a melanoma tumor model [136].
Further studies with the virus demonstrated abscopal effect in non-injected tumors, showing the systemic nature of the immune response generated by local treatment [128]. Moreover, TILT-123 therapy in TIL-treated animals demonstrated protective immunity towards tumor re-challenge [128]. Importantly, this data demonstrates the potential of TILT-123 to fine-tune and overcome challenges of T cell-based adoptive cell therapy. The first trial with this advanced oncolytic adenovirus is now ongoing (NCT04217473).

Conclusions

In summary, while preclinical data for oncolytic adenoviruses is impressive and clinical data have shown efficacy, more potent treatment strategies are needed to achieve long-term tumor control in patients. Thus, it is logical that several approaches are being utilized for improving the efficacy of oncolytic viruses. Hopefully, the efforts of the scientific and medical communities during multiple decades will lead to better treatment options in the near future. Better understanding of individual patient-tumor immune status will likely emerge and we will be able to treat advanced cancer patients better. It appears that oncolytic viruses will be part of future multimodality approaches. New treatments are urgently needed to help the 10 million cancer patients that are dying each year.

Acknowledgements

Not applicable
Not applicable
Not applicable

Competing interests

OH and AH are shareholders in TILT Biotherapeutics Ltd. JMS and AH are employees of TILT Biotherapeutics Ltd, and TILT-123 is a product of this company. AH and OH are also shareholders in Targovax ASA, and ONCOS-102 is a product of this company.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

e.Med Innere Medizin

Kombi-Abonnement

Mit e.Med Innere Medizin erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes Innere Medizin, den Premium-Inhalten der internistischen Fachzeitschriften, inklusive einer gedruckten internistischen Zeitschrift Ihrer Wahl.

Literatur
1.
Zurück zum Zitat Lippman SM, Hawk ET. Cancer prevention: from 1727 to milestones of the past 100 years. Cancer Res. 2009;69(13):5269–84.PubMedCrossRef Lippman SM, Hawk ET. Cancer prevention: from 1727 to milestones of the past 100 years. Cancer Res. 2009;69(13):5269–84.PubMedCrossRef
2.
Zurück zum Zitat Segovia-Siapco G, Sabate J. Health and sustainability outcomes of vegetarian dietary patterns: a revisit of the EPIC-Oxford and the Adventist Health Study-2 cohorts. Eur J Clin Nutr. 2019;72(Suppl 1):60–70.PubMedCrossRef Segovia-Siapco G, Sabate J. Health and sustainability outcomes of vegetarian dietary patterns: a revisit of the EPIC-Oxford and the Adventist Health Study-2 cohorts. Eur J Clin Nutr. 2019;72(Suppl 1):60–70.PubMedCrossRef
3.
Zurück zum Zitat Bray F, Ferlay J, Soerjomataram I, Siegel RL, Torre LA, Jemal A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2018;68(6):394–424.CrossRefPubMed Bray F, Ferlay J, Soerjomataram I, Siegel RL, Torre LA, Jemal A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2018;68(6):394–424.CrossRefPubMed
4.
Zurück zum Zitat Cerullo V, Capasso C, Vaha-Koskela M, Hemminki O, Hemminki A. Cancer-targeted oncolytic adenoviruses for modulation of the immune system. Curr Cancer Drug Targets. 2018;18(2):124–38.PubMedCrossRef Cerullo V, Capasso C, Vaha-Koskela M, Hemminki O, Hemminki A. Cancer-targeted oncolytic adenoviruses for modulation of the immune system. Curr Cancer Drug Targets. 2018;18(2):124–38.PubMedCrossRef
5.
Zurück zum Zitat Hemminki O, Oksanen M, Taipale K, Liikanen I, Koski A, Joensuu T, et al. Oncograms visualize factors influencing long-term survival of cancer patients treated with adenoviral oncolytic immunotherapy. Mol Ther Oncolytics. 2018;9:41–50.PubMedPubMedCentralCrossRef Hemminki O, Oksanen M, Taipale K, Liikanen I, Koski A, Joensuu T, et al. Oncograms visualize factors influencing long-term survival of cancer patients treated with adenoviral oncolytic immunotherapy. Mol Ther Oncolytics. 2018;9:41–50.PubMedPubMedCentralCrossRef
6.
Zurück zum Zitat Gratwohl A, Baldomero H, Aljurf M, Pasquini MC, Bouzas LF, Yoshimi A, et al. Hematopoietic stem cell transplantation: a global perspective. JAMA. 2010;303(16):1617–24.PubMedPubMedCentralCrossRef Gratwohl A, Baldomero H, Aljurf M, Pasquini MC, Bouzas LF, Yoshimi A, et al. Hematopoietic stem cell transplantation: a global perspective. JAMA. 2010;303(16):1617–24.PubMedPubMedCentralCrossRef
7.
Zurück zum Zitat Zhang L, Yu H, Hemminki O, Forsti A, Sundquist K, Hemminki K. Familial associations in testicular cancer with other cancers. Sci Rep. 2018;8(1):10880.PubMedPubMedCentralCrossRef Zhang L, Yu H, Hemminki O, Forsti A, Sundquist K, Hemminki K. Familial associations in testicular cancer with other cancers. Sci Rep. 2018;8(1):10880.PubMedPubMedCentralCrossRef
8.
Zurück zum Zitat Zhang L, Hemminki O, Chen T, Yu H, Zheng G, Chattopadhyay S, et al. Second cancers and causes of death in patients with testicular cancer in Sweden. PLoS One. 2019;14(3):e0214410.PubMedPubMedCentralCrossRef Zhang L, Hemminki O, Chen T, Yu H, Zheng G, Chattopadhyay S, et al. Second cancers and causes of death in patients with testicular cancer in Sweden. PLoS One. 2019;14(3):e0214410.PubMedPubMedCentralCrossRef
9.
10.
Zurück zum Zitat Freedman JD, Hagel J, Scott EM, Psallidas I, Gupta A, Spiers L, et al. Oncolytic adenovirus expressing bispecific antibody targets T-cell cytotoxicity in cancer biopsies. EMBO Mol Med. 2017;9(8):1067–87.PubMedPubMedCentralCrossRef Freedman JD, Hagel J, Scott EM, Psallidas I, Gupta A, Spiers L, et al. Oncolytic adenovirus expressing bispecific antibody targets T-cell cytotoxicity in cancer biopsies. EMBO Mol Med. 2017;9(8):1067–87.PubMedPubMedCentralCrossRef
11.
Zurück zum Zitat Vinay DS, Ryan EP, Pawelec G, Talib WH, Stagg J, Elkord E, et al. Immune evasion in cancer: mechanistic basis and therapeutic strategies. Semin Cancer Biol. 2015;35(Suppl):S185–S98.PubMedCrossRef Vinay DS, Ryan EP, Pawelec G, Talib WH, Stagg J, Elkord E, et al. Immune evasion in cancer: mechanistic basis and therapeutic strategies. Semin Cancer Biol. 2015;35(Suppl):S185–S98.PubMedCrossRef
12.
Zurück zum Zitat Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144(5):646–74.PubMedCrossRef Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144(5):646–74.PubMedCrossRef
13.
Zurück zum Zitat Lizee G, Overwijk WW, Radvanyi L, Gao J, Sharma P, Hwu P. Harnessing the power of the immune system to target cancer. Annu Rev Med. 2013;64:71–90.PubMedCrossRef Lizee G, Overwijk WW, Radvanyi L, Gao J, Sharma P, Hwu P. Harnessing the power of the immune system to target cancer. Annu Rev Med. 2013;64:71–90.PubMedCrossRef
15.
Zurück zum Zitat Kucerova P, Cervinkova M. Spontaneous regression of tumour and the role of microbial infection–possibilities for cancer treatment. Anti-Cancer Drugs. 2016;27(4):269–77.PubMedPubMedCentralCrossRef Kucerova P, Cervinkova M. Spontaneous regression of tumour and the role of microbial infection–possibilities for cancer treatment. Anti-Cancer Drugs. 2016;27(4):269–77.PubMedPubMedCentralCrossRef
16.
Zurück zum Zitat Coley WB. The treatment of malignant tumors by repeated inoculations of erysipelas. With a report of ten original cases. 1893. Clin Orthop Relat Res. 1991;(262):3–11. Coley WB. The treatment of malignant tumors by repeated inoculations of erysipelas. With a report of ten original cases. 1893. Clin Orthop Relat Res. 1991;(262):3–11.
17.
Zurück zum Zitat Kelly E, Russell SJ. History of oncolytic viruses: genesis to genetic engineering. Mol Ther. 2007;15(4):651–9.PubMedCrossRef Kelly E, Russell SJ. History of oncolytic viruses: genesis to genetic engineering. Mol Ther. 2007;15(4):651–9.PubMedCrossRef
19.
Zurück zum Zitat Huebner RJ, Bell JA, Rowe WP, Ward TG, Suskind RG, Hartley JW, et al. Studies of adenoidal-pharyngeal-conjunctival vaccines in volunteers. J Am Med Assoc. 1955;159(10):986–9.PubMedCrossRef Huebner RJ, Bell JA, Rowe WP, Ward TG, Suskind RG, Hartley JW, et al. Studies of adenoidal-pharyngeal-conjunctival vaccines in volunteers. J Am Med Assoc. 1955;159(10):986–9.PubMedCrossRef
20.
Zurück zum Zitat Huebner RJ, Rowe WP, Schatten WE, Smith RR, Thomas LB. Studies on the use of viruses in the treatment of carcinoma of the cervix. Cancer. 1956;9(6):1211–8.PubMedCrossRef Huebner RJ, Rowe WP, Schatten WE, Smith RR, Thomas LB. Studies on the use of viruses in the treatment of carcinoma of the cervix. Cancer. 1956;9(6):1211–8.PubMedCrossRef
21.
Zurück zum Zitat Zielinski T, Jordan E. Remote results of clinical observation of the oncolytic action of adenoviruses on cervix cancer. Nowotwory. 1969;19(3):217–21.PubMed Zielinski T, Jordan E. Remote results of clinical observation of the oncolytic action of adenoviruses on cervix cancer. Nowotwory. 1969;19(3):217–21.PubMed
22.
Zurück zum Zitat Georgiades J, Zielinski T, Cicholska A, Jordan E. Research on the oncolytic effect of APC viruses in cancer of the cervix uteri; preliminary report. Biuletyn Instytutu Medycyny Morskiej w Gdansku. 1959;10:49–57.PubMed Georgiades J, Zielinski T, Cicholska A, Jordan E. Research on the oncolytic effect of APC viruses in cancer of the cervix uteri; preliminary report. Biuletyn Instytutu Medycyny Morskiej w Gdansku. 1959;10:49–57.PubMed
23.
Zurück zum Zitat Bischoff JR, Kirn DH, Williams A, Heise C, Horn S, Muna M, et al. An adenovirus mutant that replicates selectively in p53-deficient human tumor cells. Science. 1996;274(5286):373–6.PubMedCrossRef Bischoff JR, Kirn DH, Williams A, Heise C, Horn S, Muna M, et al. An adenovirus mutant that replicates selectively in p53-deficient human tumor cells. Science. 1996;274(5286):373–6.PubMedCrossRef
24.
Zurück zum Zitat Andreansky S, Soroceanu L, Flotte ER, Chou J, Markert JM, Gillespie GY, et al. Evaluation of genetically engineered herpes simplex viruses as oncolytic agents for human malignant brain tumors. Cancer Res. 1997;57(8):1502–9.PubMed Andreansky S, Soroceanu L, Flotte ER, Chou J, Markert JM, Gillespie GY, et al. Evaluation of genetically engineered herpes simplex viruses as oncolytic agents for human malignant brain tumors. Cancer Res. 1997;57(8):1502–9.PubMed
25.
Zurück zum Zitat Hemminki O, Hemminki A. A century of oncolysis evolves into oncolytic immunotherapy. Oncoimmunology. 2016;5(2):e1074377.PubMedCrossRef Hemminki O, Hemminki A. A century of oncolysis evolves into oncolytic immunotherapy. Oncoimmunology. 2016;5(2):e1074377.PubMedCrossRef
26.
Zurück zum Zitat Sun X, Gan L, Na A, Ge L, Chen B, Liu J. Combination with stereotactic body radiotherapy offers a promising strategy to overcome resistance to immunotherapy in advanced renal cell cancer. J Oncol. 2019;2019:1483406.PubMedPubMedCentral Sun X, Gan L, Na A, Ge L, Chen B, Liu J. Combination with stereotactic body radiotherapy offers a promising strategy to overcome resistance to immunotherapy in advanced renal cell cancer. J Oncol. 2019;2019:1483406.PubMedPubMedCentral
27.
28.
Zurück zum Zitat Salmon H, Remark R, Gnjatic S, Merad M. Host tissue determinants of tumour immunity. Nat Rev Cancer. 2019;19(4):215–27.PubMedPubMedCentral Salmon H, Remark R, Gnjatic S, Merad M. Host tissue determinants of tumour immunity. Nat Rev Cancer. 2019;19(4):215–27.PubMedPubMedCentral
30.
Zurück zum Zitat Hemminki O, Bauerschmitz G, Hemmi S, Lavilla-Alonso S, Diaconu I, Guse K, et al. Oncolytic adenovirus based on serotype 3. Cancer Gene Ther. 2011;18(4):288–96.PubMedCrossRef Hemminki O, Bauerschmitz G, Hemmi S, Lavilla-Alonso S, Diaconu I, Guse K, et al. Oncolytic adenovirus based on serotype 3. Cancer Gene Ther. 2011;18(4):288–96.PubMedCrossRef
31.
Zurück zum Zitat Cerullo V, Diaconu I, Romano V, Hirvinen M, Ugolini M, Escutenaire S, et al. An oncolytic adenovirus enhanced for toll-like receptor 9 stimulation increases antitumor immune responses and tumor clearance. Mol Ther. 2012;20(11):2076–86.PubMedPubMedCentralCrossRef Cerullo V, Diaconu I, Romano V, Hirvinen M, Ugolini M, Escutenaire S, et al. An oncolytic adenovirus enhanced for toll-like receptor 9 stimulation increases antitumor immune responses and tumor clearance. Mol Ther. 2012;20(11):2076–86.PubMedPubMedCentralCrossRef
32.
Zurück zum Zitat Bergelson JM, Cunningham JA, Droguett G, Kurt-Jones EA, Krithivas A, Hong JS, et al. Isolation of a common receptor for Coxsackie B viruses and adenoviruses 2 and 5. Science. 1997;275(5304):1320–3.PubMedCrossRef Bergelson JM, Cunningham JA, Droguett G, Kurt-Jones EA, Krithivas A, Hong JS, et al. Isolation of a common receptor for Coxsackie B viruses and adenoviruses 2 and 5. Science. 1997;275(5304):1320–3.PubMedCrossRef
33.
Zurück zum Zitat Short JJ, Vasu C, Holterman MJ, Curiel DT, Pereboev A. Members of adenovirus species B utilize CD80 and CD86 as cellular attachment receptors. Virus Res. 2006;122(1-2):144–53.PubMedPubMedCentralCrossRef Short JJ, Vasu C, Holterman MJ, Curiel DT, Pereboev A. Members of adenovirus species B utilize CD80 and CD86 as cellular attachment receptors. Virus Res. 2006;122(1-2):144–53.PubMedPubMedCentralCrossRef
34.
Zurück zum Zitat Sirena D, Lilienfeld B, Eisenhut M, Kalin S, Boucke K, Beerli RR, et al. The human membrane cofactor CD46 is a receptor for species B adenovirus serotype 3. J Virol. 2004;78(9):4454–62.PubMedPubMedCentralCrossRef Sirena D, Lilienfeld B, Eisenhut M, Kalin S, Boucke K, Beerli RR, et al. The human membrane cofactor CD46 is a receptor for species B adenovirus serotype 3. J Virol. 2004;78(9):4454–62.PubMedPubMedCentralCrossRef
35.
Zurück zum Zitat Wang H, Li ZY, Liu Y, Persson J, Beyer I, Moller T, et al. Desmoglein 2 is a receptor for adenovirus serotypes 3, 7, 11 and 14. Nat Med. 2011;17(1):96–104.PubMedCrossRef Wang H, Li ZY, Liu Y, Persson J, Beyer I, Moller T, et al. Desmoglein 2 is a receptor for adenovirus serotypes 3, 7, 11 and 14. Nat Med. 2011;17(1):96–104.PubMedCrossRef
36.
Zurück zum Zitat Wickham TJ, Mathias P, Cheresh DA, Nemerow GR. Integrins alpha v beta 3 and alpha v beta 5 promote adenovirus internalization but not virus attachment. Cell. 1993;73(2):309–19.PubMedCrossRef Wickham TJ, Mathias P, Cheresh DA, Nemerow GR. Integrins alpha v beta 3 and alpha v beta 5 promote adenovirus internalization but not virus attachment. Cell. 1993;73(2):309–19.PubMedCrossRef
37.
Zurück zum Zitat Hemminki O. Cancer Immunotherapy with a Serotype 3 Oncolytic Adenovirus: LAP LAMBERT Academic Publishing; 2016. Hemminki O. Cancer Immunotherapy with a Serotype 3 Oncolytic Adenovirus: LAP LAMBERT Academic Publishing; 2016.
38.
Zurück zum Zitat Cohen CJ, Shieh JT, Pickles RJ, Okegawa T, Hsieh JT, Bergelson JM. The coxsackievirus and adenovirus receptor is a transmembrane component of the tight junction. Proc Natl Acad Sci U S A. 2001;98(26):15191–6.PubMedPubMedCentralCrossRef Cohen CJ, Shieh JT, Pickles RJ, Okegawa T, Hsieh JT, Bergelson JM. The coxsackievirus and adenovirus receptor is a transmembrane component of the tight junction. Proc Natl Acad Sci U S A. 2001;98(26):15191–6.PubMedPubMedCentralCrossRef
39.
Zurück zum Zitat Fleischli C, Sirena D, Lesage G, Havenga MJ, Cattaneo R, Greber UF, et al. Species B adenovirus serotypes 3, 7, 11 and 35 share similar binding sites on the membrane cofactor protein CD46 receptor. J Gen Virol. 2007;88(Pt 11):2925–34.PubMedCrossRef Fleischli C, Sirena D, Lesage G, Havenga MJ, Cattaneo R, Greber UF, et al. Species B adenovirus serotypes 3, 7, 11 and 35 share similar binding sites on the membrane cofactor protein CD46 receptor. J Gen Virol. 2007;88(Pt 11):2925–34.PubMedCrossRef
40.
Zurück zum Zitat Wang H, Li Z, Yumul R, Lara S, Hemminki A, Fender P, et al. Multimerization of adenovirus serotype 3 fiber knob domains is required for efficient binding of virus to desmoglein 2 and subsequent opening of epithelial junctions. J Virol. 2011;85(13):6390–402.PubMedPubMedCentralCrossRef Wang H, Li Z, Yumul R, Lara S, Hemminki A, Fender P, et al. Multimerization of adenovirus serotype 3 fiber knob domains is required for efficient binding of virus to desmoglein 2 and subsequent opening of epithelial junctions. J Virol. 2011;85(13):6390–402.PubMedPubMedCentralCrossRef
41.
Zurück zum Zitat Lu ZZ, Wang H, Zhang Y, Cao H, Li Z, Fender P, et al. Penton-dodecahedral particles trigger opening of intercellular junctions and facilitate viral spread during adenovirus serotype 3 infection of epithelial cells. PLoS Pathog. 2013;9(10):e1003718.PubMedPubMedCentralCrossRef Lu ZZ, Wang H, Zhang Y, Cao H, Li Z, Fender P, et al. Penton-dodecahedral particles trigger opening of intercellular junctions and facilitate viral spread during adenovirus serotype 3 infection of epithelial cells. PLoS Pathog. 2013;9(10):e1003718.PubMedPubMedCentralCrossRef
42.
Zurück zum Zitat Arnberg N. Adenovirus receptors: implications for targeting of viral vectors. Trends Pharmacol Sci. 2012;33(8):442–8.PubMedCrossRef Arnberg N. Adenovirus receptors: implications for targeting of viral vectors. Trends Pharmacol Sci. 2012;33(8):442–8.PubMedCrossRef
44.
Zurück zum Zitat Hong SS, Karayan L, Tournier J, Curiel DT, Boulanger PA. Adenovirus type 5 fiber knob binds to MHC class I alpha2 domain at the surface of human epithelial and B lymphoblastoid cells. EMBO J. 1997;16(9):2294–306.PubMedPubMedCentralCrossRef Hong SS, Karayan L, Tournier J, Curiel DT, Boulanger PA. Adenovirus type 5 fiber knob binds to MHC class I alpha2 domain at the surface of human epithelial and B lymphoblastoid cells. EMBO J. 1997;16(9):2294–306.PubMedPubMedCentralCrossRef
45.
Zurück zum Zitat Chu Y, Heistad D, Cybulsky MI, Davidson BL. Vascular cell adhesion molecule-1 augments adenovirus-mediated gene transfer. Arterioscler Thromb Vasc Biol. 2001;21(2):238–42.PubMedCrossRef Chu Y, Heistad D, Cybulsky MI, Davidson BL. Vascular cell adhesion molecule-1 augments adenovirus-mediated gene transfer. Arterioscler Thromb Vasc Biol. 2001;21(2):238–42.PubMedCrossRef
46.
Zurück zum Zitat Smith TA, Idamakanti N, Rollence ML, Marshall-Neff J, Kim J, Mulgrew K, et al. Adenovirus serotype 5 fiber shaft influences in vivo gene transfer in mice. Hum Gene Ther. 2003;14(8):777–87.PubMedCrossRef Smith TA, Idamakanti N, Rollence ML, Marshall-Neff J, Kim J, Mulgrew K, et al. Adenovirus serotype 5 fiber shaft influences in vivo gene transfer in mice. Hum Gene Ther. 2003;14(8):777–87.PubMedCrossRef
47.
Zurück zum Zitat Balakireva L, Schoehn G, Thouvenin E, Chroboczek J. Binding of adenovirus capsid to dipalmitoyl phosphatidylcholine provides a novel pathway for virus entry. J Virol. 2003;77(8):4858–66.PubMedPubMedCentralCrossRef Balakireva L, Schoehn G, Thouvenin E, Chroboczek J. Binding of adenovirus capsid to dipalmitoyl phosphatidylcholine provides a novel pathway for virus entry. J Virol. 2003;77(8):4858–66.PubMedPubMedCentralCrossRef
48.
Zurück zum Zitat Hemminki O, Immonen R, Narvainen J, Kipar A, Paasonen J, Jokivarsi KT, et al. In vivo magnetic resonance imaging and spectroscopy identifies oncolytic adenovirus responders. Int J Cancer. 2014;134(12):2878–90.PubMedCrossRef Hemminki O, Immonen R, Narvainen J, Kipar A, Paasonen J, Jokivarsi KT, et al. In vivo magnetic resonance imaging and spectroscopy identifies oncolytic adenovirus responders. Int J Cancer. 2014;134(12):2878–90.PubMedCrossRef
49.
Zurück zum Zitat Koski A, Karli E, Kipar A, Escutenaire S, Kanerva A, Hemminki A. Mutation of the fiber shaft heparan sulphate binding site of a 5/3 chimeric adenovirus reduces liver tropism. PLoS One. 2013;8(4):e60032.PubMedPubMedCentralCrossRef Koski A, Karli E, Kipar A, Escutenaire S, Kanerva A, Hemminki A. Mutation of the fiber shaft heparan sulphate binding site of a 5/3 chimeric adenovirus reduces liver tropism. PLoS One. 2013;8(4):e60032.PubMedPubMedCentralCrossRef
50.
Zurück zum Zitat Liikanen I, Monsurro V, Ahtiainen L, Raki M, Hakkarainen T, Diaconu I, et al. Induction of interferon pathways mediates in vivo resistance to oncolytic adenovirus. Mol Ther. 2011;19(10):1858–66.PubMedPubMedCentralCrossRef Liikanen I, Monsurro V, Ahtiainen L, Raki M, Hakkarainen T, Diaconu I, et al. Induction of interferon pathways mediates in vivo resistance to oncolytic adenovirus. Mol Ther. 2011;19(10):1858–66.PubMedPubMedCentralCrossRef
51.
Zurück zum Zitat Liikanen I, Koski A, Merisalo-Soikkeli M, Hemminki O, Oksanen M, Kairemo K, et al. Serum HMGB1 is a predictive and prognostic biomarker for oncolytic immunotherapy. Oncoimmunology. 2015;4(3):e989771.PubMedPubMedCentralCrossRef Liikanen I, Koski A, Merisalo-Soikkeli M, Hemminki O, Oksanen M, Kairemo K, et al. Serum HMGB1 is a predictive and prognostic biomarker for oncolytic immunotherapy. Oncoimmunology. 2015;4(3):e989771.PubMedPubMedCentralCrossRef
52.
Zurück zum Zitat Ranki T, Joensuu T, Jager E, Karbach J, Wahle C, Kairemo K, et al. Local treatment of a pleural mesothelioma tumor with ONCOS-102 induces a systemic antitumor CD8+ T-cell response, prominent infiltration of CD8+ lymphocytes and Th1 type polarization. Oncoimmunology. 2014;3(10):e958937.PubMedPubMedCentralCrossRef Ranki T, Joensuu T, Jager E, Karbach J, Wahle C, Kairemo K, et al. Local treatment of a pleural mesothelioma tumor with ONCOS-102 induces a systemic antitumor CD8+ T-cell response, prominent infiltration of CD8+ lymphocytes and Th1 type polarization. Oncoimmunology. 2014;3(10):e958937.PubMedPubMedCentralCrossRef
53.
Zurück zum Zitat Seder RA, Darrah PA, Roederer M. T-cell quality in memory and protection: implications for vaccine design. Nat Rev Immunol. 2008;8(4):247–58.PubMedCrossRef Seder RA, Darrah PA, Roederer M. T-cell quality in memory and protection: implications for vaccine design. Nat Rev Immunol. 2008;8(4):247–58.PubMedCrossRef
54.
Zurück zum Zitat Leung EYL, Ennis DP, Kennedy PR, Hansell C, Dowson S, Farquharson M, et al. NK cells augment oncolytic adenovirus cytotoxicity in ovarian cancer. Mol Ther Oncolytics. 2020;16:289–301.PubMedPubMedCentralCrossRef Leung EYL, Ennis DP, Kennedy PR, Hansell C, Dowson S, Farquharson M, et al. NK cells augment oncolytic adenovirus cytotoxicity in ovarian cancer. Mol Ther Oncolytics. 2020;16:289–301.PubMedPubMedCentralCrossRef
55.
Zurück zum Zitat Massari I, Donnini A, Argentati K, Straino S, Mangoni A, Gaetano C, et al. 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(6):823–31.PubMedCrossRef Massari I, Donnini A, Argentati K, Straino S, Mangoni A, Gaetano C, et al. 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(6):823–31.PubMedCrossRef
57.
Zurück zum Zitat de Queiroz N, Xia T, Konno H, Barber GN. Ovarian cancer cells commonly exhibit defective STING signaling which affects sensitivity to viral oncolysis. Mol Cancer Res. 2019;17(4):974–86.PubMedCrossRef de Queiroz N, Xia T, Konno H, Barber GN. Ovarian cancer cells commonly exhibit defective STING signaling which affects sensitivity to viral oncolysis. Mol Cancer Res. 2019;17(4):974–86.PubMedCrossRef
58.
Zurück zum Zitat Jhawar SR, Thandoni A, Bommareddy PK, Hassan S, Kohlhapp FJ, Goyal S, et al. Oncolytic viruses-natural and genetically engineered cancer immunotherapies. Front Oncol. 2017;7:202.PubMedPubMedCentralCrossRef Jhawar SR, Thandoni A, Bommareddy PK, Hassan S, Kohlhapp FJ, Goyal S, et al. Oncolytic viruses-natural and genetically engineered cancer immunotherapies. Front Oncol. 2017;7:202.PubMedPubMedCentralCrossRef
59.
Zurück zum Zitat Bayo-Puxan N, Gimenez-Alejandre M, Lavilla-Alonso S, Gros A, Cascallo M, Hemminki A, et al. Replacement of adenovirus type 5 fiber shaft heparan sulfate proteoglycan-binding domain with RGD for improved tumor infectivity and targeting. Hum Gene Ther. 2009;20(10):1214–21.PubMedCrossRef Bayo-Puxan N, Gimenez-Alejandre M, Lavilla-Alonso S, Gros A, Cascallo M, Hemminki A, et al. Replacement of adenovirus type 5 fiber shaft heparan sulfate proteoglycan-binding domain with RGD for improved tumor infectivity and targeting. Hum Gene Ther. 2009;20(10):1214–21.PubMedCrossRef
60.
Zurück zum Zitat Kanerva A, Wang M, Bauerschmitz GJ, Lam JT, Desmond RA, Bhoola SM, et al. Gene transfer to ovarian cancer versus normal tissues with fiber-modified adenoviruses. Mol Ther. 2002;5(6):695–704.PubMedCrossRef Kanerva A, Wang M, Bauerschmitz GJ, Lam JT, Desmond RA, Bhoola SM, et al. Gene transfer to ovarian cancer versus normal tissues with fiber-modified adenoviruses. Mol Ther. 2002;5(6):695–704.PubMedCrossRef
61.
Zurück zum Zitat Lang FF, Conrad C, Gomez-Manzano C, Yung WKA, Sawaya R, Weinberg JS, et al. Phase I Study of DNX-2401 (Delta-24-RGD) oncolytic adenovirus: replication and immunotherapeutic effects in recurrent malignant glioma. J Clin Oncol. 2018;36(14):1419–27.PubMedPubMedCentralCrossRef Lang FF, Conrad C, Gomez-Manzano C, Yung WKA, Sawaya R, Weinberg JS, et al. Phase I Study of DNX-2401 (Delta-24-RGD) oncolytic adenovirus: replication and immunotherapeutic effects in recurrent malignant glioma. J Clin Oncol. 2018;36(14):1419–27.PubMedPubMedCentralCrossRef
62.
Zurück zum Zitat Kanerva A, Mikheeva GV, Krasnykh V, Coolidge CJ, Lam JT, Mahasreshti PJ, et al. Targeting adenovirus to the serotype 3 receptor increases gene transfer efficiency to ovarian cancer cells. Clin Cancer Res. 2002;8(1):275–80.PubMed Kanerva A, Mikheeva GV, Krasnykh V, Coolidge CJ, Lam JT, Mahasreshti PJ, et al. Targeting adenovirus to the serotype 3 receptor increases gene transfer efficiency to ovarian cancer cells. Clin Cancer Res. 2002;8(1):275–80.PubMed
63.
Zurück zum Zitat Koski A, Bramante S, Kipar A, Oksanen M, Juhila J, Vassilev L, et al. Biodistribution analysis of oncolytic adenoviruses in patient autopsy samples reveals vascular transduction of noninjected tumors and tissues. Mol Ther. 2015;23(10):1641–52.PubMedPubMedCentralCrossRef Koski A, Bramante S, Kipar A, Oksanen M, Juhila J, Vassilev L, et al. Biodistribution analysis of oncolytic adenoviruses in patient autopsy samples reveals vascular transduction of noninjected tumors and tissues. Mol Ther. 2015;23(10):1641–52.PubMedPubMedCentralCrossRef
64.
Zurück zum Zitat Hemminki O, Diaconu I, Cerullo V, Pesonen SK, Kanerva A, Joensuu T, et al. Ad3-hTERT-E1A, a fully serotype 3 oncolytic adenovirus, in patients with chemotherapy refractory cancer. Mol Ther. 2012;20(9):1821–30.PubMedPubMedCentralCrossRef Hemminki O, Diaconu I, Cerullo V, Pesonen SK, Kanerva A, Joensuu T, et al. Ad3-hTERT-E1A, a fully serotype 3 oncolytic adenovirus, in patients with chemotherapy refractory cancer. Mol Ther. 2012;20(9):1821–30.PubMedPubMedCentralCrossRef
65.
Zurück zum Zitat Kuhn I, Harden P, Bauzon M, Chartier C, Nye J, Thorne S, et al. Directed evolution generates a novel oncolytic virus for the treatment of colon cancer. PLoS One. 2008;3(6):e2409.PubMedPubMedCentralCrossRef Kuhn I, Harden P, Bauzon M, Chartier C, Nye J, Thorne S, et al. Directed evolution generates a novel oncolytic virus for the treatment of colon cancer. PLoS One. 2008;3(6):e2409.PubMedPubMedCentralCrossRef
66.
Zurück zum Zitat Kanerva A, Nokisalmi P, Diaconu I, Koski A, Cerullo V, Liikanen I, et al. Antiviral and antitumor T-cell immunity in patients treated with GM-CSF-coding oncolytic adenovirus. Clin Cancer Res. 2013;19(10):2734–44.PubMedCrossRef Kanerva A, Nokisalmi P, Diaconu I, Koski A, Cerullo V, Liikanen I, et al. Antiviral and antitumor T-cell immunity in patients treated with GM-CSF-coding oncolytic adenovirus. Clin Cancer Res. 2013;19(10):2734–44.PubMedCrossRef
67.
Zurück zum Zitat Bramante S, Koski A, Kipar A, Diaconu I, Liikanen I, Hemminki O, et al. Serotype chimeric oncolytic adenovirus coding for GM-CSF for treatment of sarcoma in rodents and humans. Int J Cancer. 2014;135(3):720–30.PubMedCrossRef Bramante S, Koski A, Kipar A, Diaconu I, Liikanen I, Hemminki O, et al. Serotype chimeric oncolytic adenovirus coding for GM-CSF for treatment of sarcoma in rodents and humans. Int J Cancer. 2014;135(3):720–30.PubMedCrossRef
68.
Zurück zum Zitat Chang J, Zhao X, Wu X, Guo Y, Guo H, Cao J, et al. A Phase I study of KH901, a conditionally replicating granulocyte-macrophage colony-stimulating factor: armed oncolytic adenovirus for the treatment of head and neck cancers. Cancer Biol Ther. 2009;8(8):676–82.PubMedCrossRef Chang J, Zhao X, Wu X, Guo Y, Guo H, Cao J, et al. A Phase I study of KH901, a conditionally replicating granulocyte-macrophage colony-stimulating factor: armed oncolytic adenovirus for the treatment of head and neck cancers. Cancer Biol Ther. 2009;8(8):676–82.PubMedCrossRef
69.
Zurück zum Zitat Wang X, Su C, Cao H, Li K, Chen J, Jiang L, et al. A novel triple-regulated oncolytic adenovirus carrying p53 gene exerts potent antitumor efficacy on common human solid cancers. Mol Cancer Ther. 2008;7(6):1598–603.PubMedCrossRef Wang X, Su C, Cao H, Li K, Chen J, Jiang L, et al. A novel triple-regulated oncolytic adenovirus carrying p53 gene exerts potent antitumor efficacy on common human solid cancers. Mol Cancer Ther. 2008;7(6):1598–603.PubMedCrossRef
70.
Zurück zum Zitat Xu C, Sun Y, Wang Y, Yan Y, Shi Z, Chen L, et al. CEA promoter-regulated oncolytic adenovirus-mediated Hsp70 expression in immune gene therapy for pancreatic cancer. Cancer Lett. 2012;319(2):154–63.PubMedCrossRef Xu C, Sun Y, Wang Y, Yan Y, Shi Z, Chen L, et al. CEA promoter-regulated oncolytic adenovirus-mediated Hsp70 expression in immune gene therapy for pancreatic cancer. Cancer Lett. 2012;319(2):154–63.PubMedCrossRef
71.
Zurück zum Zitat DeWeese TL, van der Poel H, Li S, Mikhak B, Drew R, Goemann M, et al. A phase I trial of CV706, a replication-competent, PSA selective oncolytic adenovirus, for the treatment of locally recurrent prostate cancer following radiation therapy. Cancer Res. 2001;61(20):7464–72.PubMed DeWeese TL, van der Poel H, Li S, Mikhak B, Drew R, Goemann M, et al. A phase I trial of CV706, a replication-competent, PSA selective oncolytic adenovirus, for the treatment of locally recurrent prostate cancer following radiation therapy. Cancer Res. 2001;61(20):7464–72.PubMed
72.
Zurück zum Zitat Hemminki O, Parviainen S, Juhila J, Turkki R, Linder N, Lundin J, et al. Immunological data from cancer patients treated with Ad5/3-E2F-Delta24-GMCSF suggests utility for tumor immunotherapy. Oncotarget. 2015;6(6):4467–81.PubMedPubMedCentralCrossRef Hemminki O, Parviainen S, Juhila J, Turkki R, Linder N, Lundin J, et al. Immunological data from cancer patients treated with Ad5/3-E2F-Delta24-GMCSF suggests utility for tumor immunotherapy. Oncotarget. 2015;6(6):4467–81.PubMedPubMedCentralCrossRef
73.
Zurück zum Zitat Bauerschmitz GJ, Guse K, Kanerva A, Menzel A, Herrmann I, Desmond RA, et al. Triple-targeted oncolytic adenoviruses featuring the cox2 promoter, E1A transcomplementation, and serotype chimerism for enhanced selectivity for ovarian cancer cells. Mol Ther. 2006;14(2):164–74.PubMedCrossRef Bauerschmitz GJ, Guse K, Kanerva A, Menzel A, Herrmann I, Desmond RA, et al. Triple-targeted oncolytic adenoviruses featuring the cox2 promoter, E1A transcomplementation, and serotype chimerism for enhanced selectivity for ovarian cancer cells. Mol Ther. 2006;14(2):164–74.PubMedCrossRef
74.
Zurück zum Zitat Barton KN, Paielli D, Zhang Y, Koul S, Brown SL, Lu M, et al. Second-generation replication-competent oncolytic adenovirus armed with improved suicide genes and ADP gene demonstrates greater efficacy without increased toxicity. Mol Ther. 2006;13(2):347–56.PubMedCrossRef Barton KN, Paielli D, Zhang Y, Koul S, Brown SL, Lu M, et al. Second-generation replication-competent oncolytic adenovirus armed with improved suicide genes and ADP gene demonstrates greater efficacy without increased toxicity. Mol Ther. 2006;13(2):347–56.PubMedCrossRef
75.
Zurück zum Zitat Boucher PD, Im MM, Freytag SO, Shewach DS. A novel mechanism of synergistic cytotoxicity with 5-fluorocytosine and ganciclovir in double suicide gene therapy. Cancer Res. 2006;66(6):3230–7.PubMedCrossRef Boucher PD, Im MM, Freytag SO, Shewach DS. A novel mechanism of synergistic cytotoxicity with 5-fluorocytosine and ganciclovir in double suicide gene therapy. Cancer Res. 2006;66(6):3230–7.PubMedCrossRef
76.
Zurück zum Zitat Taipale K, Liikanen I, Koski A, Heiskanen R, Kanerva A, Hemminki O, et al. Predictive and prognostic clinical variables in cancer patients treated with adenoviral oncolytic immunotherapy. Mol Ther. 2016;24(7):1323–32.PubMedPubMedCentralCrossRef Taipale K, Liikanen I, Koski A, Heiskanen R, Kanerva A, Hemminki O, et al. Predictive and prognostic clinical variables in cancer patients treated with adenoviral oncolytic immunotherapy. Mol Ther. 2016;24(7):1323–32.PubMedPubMedCentralCrossRef
77.
Zurück zum Zitat Andtbacka RH, Ross M, Puzanov I, Milhem M, Collichio F, Delman KA, et al. Patterns of clinical response with talimogene laherparepvec (T-VEC) in patients with melanoma treated in the OPTiM phase III clinical trial. Ann Surg Oncol. 2016;23(13):4169–77.PubMedPubMedCentralCrossRef Andtbacka RH, Ross M, Puzanov I, Milhem M, Collichio F, Delman KA, et al. Patterns of clinical response with talimogene laherparepvec (T-VEC) in patients with melanoma treated in the OPTiM phase III clinical trial. Ann Surg Oncol. 2016;23(13):4169–77.PubMedPubMedCentralCrossRef
78.
Zurück zum Zitat Zafar S, Parviainen S, Siurala M, Hemminki O, Havunen R, Tahtinen S, et al. Intravenously usable fully serotype 3 oncolytic adenovirus coding for CD40L as an enabler of dendritic cell therapy. Oncoimmunology. 2017;6(2):e1265717.PubMedCrossRef Zafar S, Parviainen S, Siurala M, Hemminki O, Havunen R, Tahtinen S, et al. Intravenously usable fully serotype 3 oncolytic adenovirus coding for CD40L as an enabler of dendritic cell therapy. Oncoimmunology. 2017;6(2):e1265717.PubMedCrossRef
79.
Zurück zum Zitat Zafar S, Sorsa S, Siurala M, Hemminki O, Havunen R, Cervera-Carrascon V, et al. CD40L coding oncolytic adenovirus allows long-term survival of humanized mice receiving dendritic cell therapy. Oncoimmunology. 2018;7(10):e1490856.PubMedPubMedCentralCrossRef Zafar S, Sorsa S, Siurala M, Hemminki O, Havunen R, Cervera-Carrascon V, et al. CD40L coding oncolytic adenovirus allows long-term survival of humanized mice receiving dendritic cell therapy. Oncoimmunology. 2018;7(10):e1490856.PubMedPubMedCentralCrossRef
80.
Zurück zum Zitat Rajecki M, Kangasmaki A, Laasonen L, Escutenaire S, Hakkarainen T, Haukka J, et al. Sodium iodide symporter SPECT imaging of a patient treated with oncolytic adenovirus Ad5/3-Delta24-hNIS. Mol Ther. 2011;19(4):629–31.PubMedPubMedCentralCrossRef Rajecki M, Kangasmaki A, Laasonen L, Escutenaire S, Hakkarainen T, Haukka J, et al. Sodium iodide symporter SPECT imaging of a patient treated with oncolytic adenovirus Ad5/3-Delta24-hNIS. Mol Ther. 2011;19(4):629–31.PubMedPubMedCentralCrossRef
81.
Zurück zum Zitat Dias JD, Hemminki O, Diaconu I, Hirvinen M, Bonetti A, Guse K, et al. Targeted cancer immunotherapy with oncolytic adenovirus coding for a fully human monoclonal antibody specific for CTLA-4. Gene Ther. 2012;19(10):988–98.PubMedCrossRef Dias JD, Hemminki O, Diaconu I, Hirvinen M, Bonetti A, Guse K, et al. Targeted cancer immunotherapy with oncolytic adenovirus coding for a fully human monoclonal antibody specific for CTLA-4. Gene Ther. 2012;19(10):988–98.PubMedCrossRef
82.
Zurück zum Zitat Oh E, Choi IK, Hong J, Yun CO. Oncolytic adenovirus coexpressing interleukin-12 and decorin overcomes Treg-mediated immunosuppression inducing potent antitumor effects in a weakly immunogenic tumor model. Oncotarget. 2017;8(3):4730–46.PubMedCrossRef Oh E, Choi IK, Hong J, Yun CO. Oncolytic adenovirus coexpressing interleukin-12 and decorin overcomes Treg-mediated immunosuppression inducing potent antitumor effects in a weakly immunogenic tumor model. Oncotarget. 2017;8(3):4730–46.PubMedCrossRef
83.
Zurück zum Zitat Jiang H, Shin DH, Nguyen TT, Fueyo J, Fan X, Henry V, et al. Localized treatment with oncolytic adenovirus Delta-24-RGDOX induces systemic immunity against disseminated subcutaneous and intracranial melanomas. Clin Cancer Res. 2019;25(22):6801–14.PubMedPubMedCentralCrossRef Jiang H, Shin DH, Nguyen TT, Fueyo J, Fan X, Henry V, et al. Localized treatment with oncolytic adenovirus Delta-24-RGDOX induces systemic immunity against disseminated subcutaneous and intracranial melanomas. Clin Cancer Res. 2019;25(22):6801–14.PubMedPubMedCentralCrossRef
84.
Zurück zum Zitat Fajardo CA, Guedan S, Rojas LA, Moreno R, Arias-Badia M, de Sostoa J, et al. Oncolytic adenoviral delivery of an EGFR-targeting T-cell engager improves antitumor efficacy. Cancer Res. 2017;77(8):2052–63.PubMedCrossRef Fajardo CA, Guedan S, Rojas LA, Moreno R, Arias-Badia M, de Sostoa J, et al. Oncolytic adenoviral delivery of an EGFR-targeting T-cell engager improves antitumor efficacy. Cancer Res. 2017;77(8):2052–63.PubMedCrossRef
85.
Zurück zum Zitat Wing A, Fajardo CA, Posey AD Jr, Shaw C, Da T, Young RM, et al. Improving CART-cell therapy of solid tumors with oncolytic virus-driven production of a bispecific T-cell engager. Cancer Immunol Res. 2018;6(5):605–16.PubMedPubMedCentralCrossRef Wing A, Fajardo CA, Posey AD Jr, Shaw C, Da T, Young RM, et al. Improving CART-cell therapy of solid tumors with oncolytic virus-driven production of a bispecific T-cell engager. Cancer Immunol Res. 2018;6(5):605–16.PubMedPubMedCentralCrossRef
86.
Zurück zum Zitat Porter CE, Rosewell Shaw A, Jung Y, Yip T, Castro PD, Sandulache VC, et al. Oncolytic adenovirus armed with BiTE, cytokine, and checkpoint inhibitor enables CAR T cells to control the growth of heterogeneous tumors. Mol Ther. 2020. Porter CE, Rosewell Shaw A, Jung Y, Yip T, Castro PD, Sandulache VC, et al. Oncolytic adenovirus armed with BiTE, cytokine, and checkpoint inhibitor enables CAR T cells to control the growth of heterogeneous tumors. Mol Ther. 2020.
87.
Zurück zum Zitat Koski A, Kangasniemi L, Escutenaire S, Pesonen S, Cerullo V, Diaconu I, et al. Treatment of cancer patients with a serotype 5/3 chimeric oncolytic adenovirus expressing GMCSF. Mol Ther. 2010;18(10):1874–84.PubMedPubMedCentralCrossRef Koski A, Kangasniemi L, Escutenaire S, Pesonen S, Cerullo V, Diaconu I, et al. Treatment of cancer patients with a serotype 5/3 chimeric oncolytic adenovirus expressing GMCSF. Mol Ther. 2010;18(10):1874–84.PubMedPubMedCentralCrossRef
88.
Zurück zum Zitat Kanerva A, Koski A, Liikanen I, Oksanen M, Joensuu T, Hemminki O, et al. Case-control estimation of the impact of oncolytic adenovirus on the survival of patients with refractory solid tumors. Mol Ther. 2015;23(2):321–9.PubMedCrossRef Kanerva A, Koski A, Liikanen I, Oksanen M, Joensuu T, Hemminki O, et al. Case-control estimation of the impact of oncolytic adenovirus on the survival of patients with refractory solid tumors. Mol Ther. 2015;23(2):321–9.PubMedCrossRef
89.
Zurück zum Zitat Machiels JP, Salazar R, Rottey S, Duran I, Dirix L, Geboes K, et al. A phase 1 dose escalation study of the oncolytic adenovirus enadenotucirev, administered intravenously to patients with epithelial solid tumors (EVOLVE). J Immunother Cancer. 2019;7(1):20.PubMedPubMedCentralCrossRef Machiels JP, Salazar R, Rottey S, Duran I, Dirix L, Geboes K, et al. A phase 1 dose escalation study of the oncolytic adenovirus enadenotucirev, administered intravenously to patients with epithelial solid tumors (EVOLVE). J Immunother Cancer. 2019;7(1):20.PubMedPubMedCentralCrossRef
90.
Zurück zum Zitat Kent LN, Leone G. The broken cycle: E2F dysfunction in cancer. Nat Rev Cancer. 2019;19(6):326–38.PubMedCrossRef Kent LN, Leone G. The broken cycle: E2F dysfunction in cancer. Nat Rev Cancer. 2019;19(6):326–38.PubMedCrossRef
92.
Zurück zum Zitat Cerullo V, Pesonen S, Diaconu I, Escutenaire S, Arstila PT, Ugolini M, et al. Oncolytic adenovirus coding for granulocyte macrophage colony-stimulating factor induces antitumoral immunity in cancer patients. Cancer Res. 2010;70(11):4297–309.PubMedCrossRef Cerullo V, Pesonen S, Diaconu I, Escutenaire S, Arstila PT, Ugolini M, et al. Oncolytic adenovirus coding for granulocyte macrophage colony-stimulating factor induces antitumoral immunity in cancer patients. Cancer Res. 2010;70(11):4297–309.PubMedCrossRef
93.
Zurück zum Zitat Simmons AD, Li B, Gonzalez-Edick M, Lin C, Moskalenko M, Du T, et al. GM-CSF-secreting cancer immunotherapies: preclinical analysis of the mechanism of action. Cancer Immunol Immunother. 2007;56(10):1653–65.PubMedCrossRef Simmons AD, Li B, Gonzalez-Edick M, Lin C, Moskalenko M, Du T, et al. GM-CSF-secreting cancer immunotherapies: preclinical analysis of the mechanism of action. Cancer Immunol Immunother. 2007;56(10):1653–65.PubMedCrossRef
94.
Zurück zum Zitat Ranki T, Pesonen S, Hemminki A, Partanen K, Kairemo K, Alanko T, et al. Phase I study with ONCOS-102 for the treatment of solid tumors - an evaluation of clinical response and exploratory analyses of immune markers. J Immunother Cancer. 2016;4:17.PubMedPubMedCentralCrossRef Ranki T, Pesonen S, Hemminki A, Partanen K, Kairemo K, Alanko T, et al. Phase I study with ONCOS-102 for the treatment of solid tumors - an evaluation of clinical response and exploratory analyses of immune markers. J Immunother Cancer. 2016;4:17.PubMedPubMedCentralCrossRef
95.
Zurück zum Zitat Tahtinen S, Kaikkonen S, Merisalo-Soikkeli M, Gronberg-Vaha-Koskela S, Kanerva A, Parviainen S, et al. Favorable alteration of tumor microenvironment by immunomodulatory cytokines for efficient T-cell therapy in solid tumors. PLoS One. 2015;10(6):e0131242.PubMedPubMedCentralCrossRef Tahtinen S, Kaikkonen S, Merisalo-Soikkeli M, Gronberg-Vaha-Koskela S, Kanerva A, Parviainen S, et al. Favorable alteration of tumor microenvironment by immunomodulatory cytokines for efficient T-cell therapy in solid tumors. PLoS One. 2015;10(6):e0131242.PubMedPubMedCentralCrossRef
96.
Zurück zum Zitat Eriksson E, Milenova I, Wenthe J, Stahle M, Leja-Jarblad J, Ullenhag G, et al. Shaping the tumor stroma and sparking immune activation by CD40 and 4-1BB signaling induced by an armed oncolytic virus. Clin Cancer Res. 2017;23(19):5846–57.PubMedCrossRef Eriksson E, Milenova I, Wenthe J, Stahle M, Leja-Jarblad J, Ullenhag G, et al. Shaping the tumor stroma and sparking immune activation by CD40 and 4-1BB signaling induced by an armed oncolytic virus. Clin Cancer Res. 2017;23(19):5846–57.PubMedCrossRef
98.
Zurück zum Zitat Pipiya T, Sauthoff H, Huang YQ, Chang B, Cheng J, Heitner S, et al. Hypoxia reduces adenoviral replication in cancer cells by downregulation of viral protein expression. Gene Ther. 2005;12(11):911–7.PubMedCrossRef Pipiya T, Sauthoff H, Huang YQ, Chang B, Cheng J, Heitner S, et al. Hypoxia reduces adenoviral replication in cancer cells by downregulation of viral protein expression. Gene Ther. 2005;12(11):911–7.PubMedCrossRef
99.
Zurück zum Zitat Cherry T, Longo SL, Tovar-Spinoza Z, Post DE. Second-generation HIF-activated oncolytic adenoviruses with improved replication, oncolytic, and antitumor efficacy. Gene Ther. 2010;17(12):1430–41.PubMedPubMedCentralCrossRef Cherry T, Longo SL, Tovar-Spinoza Z, Post DE. Second-generation HIF-activated oncolytic adenoviruses with improved replication, oncolytic, and antitumor efficacy. Gene Ther. 2010;17(12):1430–41.PubMedPubMedCentralCrossRef
100.
Zurück zum Zitat Rodriguez-Garcia A, Gimenez-Alejandre M, Rojas JJ, Moreno R, Bazan-Peregrino M, Cascallo M, et al. Safety and efficacy of VCN-01, an oncolytic adenovirus combining fiber HSG-binding domain replacement with RGD and hyaluronidase expression. Clin Cancer Res. 2015;21(6):1406–18.PubMedCrossRef Rodriguez-Garcia A, Gimenez-Alejandre M, Rojas JJ, Moreno R, Bazan-Peregrino M, Cascallo M, et al. Safety and efficacy of VCN-01, an oncolytic adenovirus combining fiber HSG-binding domain replacement with RGD and hyaluronidase expression. Clin Cancer Res. 2015;21(6):1406–18.PubMedCrossRef
101.
Zurück zum Zitat Niemann J, Woller N, Brooks J, Fleischmann-Mundt B, Martin NT, Kloos A, et al. Molecular retargeting of antibodies converts immune defense against oncolytic viruses into cancer immunotherapy. Nat Commun. 2019;10(1):3236.PubMedPubMedCentralCrossRef Niemann J, Woller N, Brooks J, Fleischmann-Mundt B, Martin NT, Kloos A, et al. Molecular retargeting of antibodies converts immune defense against oncolytic viruses into cancer immunotherapy. Nat Commun. 2019;10(1):3236.PubMedPubMedCentralCrossRef
102.
Zurück zum Zitat Fusciello M, Fontana F, Tahtinen S, Capasso C, Feola S, Martins B, et al. Artificially cloaked viral nanovaccine for cancer immunotherapy. Nat Commun. 2019;10(1):5747.PubMedPubMedCentralCrossRef Fusciello M, Fontana F, Tahtinen S, Capasso C, Feola S, Martins B, et al. Artificially cloaked viral nanovaccine for cancer immunotherapy. Nat Commun. 2019;10(1):5747.PubMedPubMedCentralCrossRef
103.
Zurück zum Zitat Pesonen S, Diaconu I, Cerullo V, Escutenaire S, Raki M, Kangasniemi 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(8):1937–47.PubMedCrossRef Pesonen S, Diaconu I, Cerullo V, Escutenaire S, Raki M, Kangasniemi 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(8):1937–47.PubMedCrossRef
104.
Zurück zum Zitat Nokisalmi P, Pesonen S, Escutenaire S, Sarkioja M, Raki M, Cerullo V, et al. Oncolytic adenovirus ICOVIR-7 in patients with advanced and refractory solid tumors. Clin Cancer Res. 2010;16(11):3035–43.PubMedCrossRef Nokisalmi P, Pesonen S, Escutenaire S, Sarkioja M, Raki M, Cerullo V, et al. Oncolytic adenovirus ICOVIR-7 in patients with advanced and refractory solid tumors. Clin Cancer Res. 2010;16(11):3035–43.PubMedCrossRef
105.
Zurück zum Zitat Ranki T. Preclinical and clinical evaluation of oncolytic immunotherapy with Ad5/3-E2F1-Δ24-GMCSF (CGTG-602), a GM-CSF producing adenovirus targeted to tumors on four levels ASGCT 2012. Philadelphia; 2012. Ranki T. Preclinical and clinical evaluation of oncolytic immunotherapy with Ad5/3-E2F1-Δ24-GMCSF (CGTG-602), a GM-CSF producing adenovirus targeted to tumors on four levels ASGCT 2012. Philadelphia; 2012.
106.
Zurück zum Zitat Taipale K, Liikanen I, Juhila J, Turkki R, Tahtinen S, Kankainen M, et al. Chronic activation of innate immunity correlates with poor prognosis in cancer patients treated with oncolytic adenovirus. Mol Ther. 2016;24(1):175–83.PubMedCrossRef Taipale K, Liikanen I, Juhila J, Turkki R, Tahtinen S, Kankainen M, et al. Chronic activation of innate immunity correlates with poor prognosis in cancer patients treated with oncolytic adenovirus. Mol Ther. 2016;24(1):175–83.PubMedCrossRef
107.
Zurück zum Zitat Liikanen I, Ahtiainen L, Hirvinen ML, Bramante S, Cerullo V, Nokisalmi P, et al. Oncolytic adenovirus with temozolomide induces autophagy and antitumor immune responses in cancer patients. Mol Ther. 2013;21(6):1212–23.PubMedPubMedCentralCrossRef Liikanen I, Ahtiainen L, Hirvinen ML, Bramante S, Cerullo V, Nokisalmi P, et al. Oncolytic adenovirus with temozolomide induces autophagy and antitumor immune responses in cancer patients. Mol Ther. 2013;21(6):1212–23.PubMedPubMedCentralCrossRef
108.
Zurück zum Zitat Koski A, Raki M, Nokisalmi P, Liikanen I, Kangasniemi L, Joensuu T, et al. Verapamil results in increased blood levels of oncolytic adenovirus in treatment of patients with advanced cancer. Mol Ther. 2012;20(1):221–9.PubMedCrossRef Koski A, Raki M, Nokisalmi P, Liikanen I, Kangasniemi L, Joensuu T, et al. Verapamil results in increased blood levels of oncolytic adenovirus in treatment of patients with advanced cancer. Mol Ther. 2012;20(1):221–9.PubMedCrossRef
109.
Zurück zum Zitat Garcia M, Moreno R, Gil-Martin M, Cascallo M, de Olza MO, Cuadra C, et al. A phase 1 trial of oncolytic adenovirus ICOVIR-5 administered intravenously to cutaneous and uveal melanoma patients. Hum Gene Ther. 2019;30(3):352–64.PubMedCrossRef Garcia M, Moreno R, Gil-Martin M, Cascallo M, de Olza MO, Cuadra C, et al. A phase 1 trial of oncolytic adenovirus ICOVIR-5 administered intravenously to cutaneous and uveal melanoma patients. Hum Gene Ther. 2019;30(3):352–64.PubMedCrossRef
110.
Zurück zum Zitat Adair RA, Roulstone V, Scott KJ, Morgan R, Nuovo GJ, Fuller M, et al. Cell carriage, delivery, and selective replication of an oncolytic virus in tumor in patients. Sci Transl Med. 2012;4(138):138ra77.PubMedPubMedCentralCrossRef Adair RA, Roulstone V, Scott KJ, Morgan R, Nuovo GJ, Fuller M, et al. Cell carriage, delivery, and selective replication of an oncolytic virus in tumor in patients. Sci Transl Med. 2012;4(138):138ra77.PubMedPubMedCentralCrossRef
111.
Zurück zum Zitat Alberts P, Tilgase A, Rasa A, Bandere K, Venskus D. The advent of oncolytic virotherapy in oncology: the Rigvir(R) story. Eur J Pharmacol. 2018;837:117–26.PubMedCrossRef Alberts P, Tilgase A, Rasa A, Bandere K, Venskus D. The advent of oncolytic virotherapy in oncology: the Rigvir(R) story. Eur J Pharmacol. 2018;837:117–26.PubMedCrossRef
112.
Zurück zum Zitat Liang M. Oncorine, the world first oncolytic virus medicine and its update in China. Curr Cancer Drug Targets. 2018;18(2):171–6.PubMedCrossRef Liang M. Oncorine, the world first oncolytic virus medicine and its update in China. Curr Cancer Drug Targets. 2018;18(2):171–6.PubMedCrossRef
113.
Zurück zum Zitat Santos JM, Havunen R, Hemminki A. Modulation of the tumor microenvironment with an oncolytic adenovirus for effective T-cell therapy and checkpoint inhibition. Methods Enzymol. 2020;635:205–30.PubMedCrossRef Santos JM, Havunen R, Hemminki A. Modulation of the tumor microenvironment with an oncolytic adenovirus for effective T-cell therapy and checkpoint inhibition. Methods Enzymol. 2020;635:205–30.PubMedCrossRef
114.
Zurück zum Zitat Andtbacka RHI, Collichio F, Harrington KJ, Middleton MR, Downey G, Ӧhrling K, et al. Final analyses of OPTiM: a randomized phase III trial of talimogene laherparepvec versus granulocyte-macrophage colony-stimulating factor in unresectable stage III-IV melanoma. J Immunother Cancer. 2019;7(1):145.PubMedPubMedCentralCrossRef Andtbacka RHI, Collichio F, Harrington KJ, Middleton MR, Downey G, Ӧhrling K, et al. Final analyses of OPTiM: a randomized phase III trial of talimogene laherparepvec versus granulocyte-macrophage colony-stimulating factor in unresectable stage III-IV melanoma. J Immunother Cancer. 2019;7(1):145.PubMedPubMedCentralCrossRef
115.
Zurück zum Zitat Ribas A, Dummer R, Puzanov I, VanderWalde A, Andtbacka RHI, Michielin O, et al. Oncolytic virotherapy promotes intratumoral T cell infiltration and improves anti-PD-1 immunotherapy. Cell. 2018;174(4):1031–2.PubMedCrossRef Ribas A, Dummer R, Puzanov I, VanderWalde A, Andtbacka RHI, Michielin O, et al. Oncolytic virotherapy promotes intratumoral T cell infiltration and improves anti-PD-1 immunotherapy. Cell. 2018;174(4):1031–2.PubMedCrossRef
116.
Zurück zum Zitat Chesney J, Puzanov I, Collichio F, Singh P, Milhem MM, Glaspy J, et al. Randomized, open-label phase II study evaluating the efficacy and safety of talimogene laherparepvec in combination with ipilimumab versus ipilimumab alone in patients with advanced, unresectable melanoma. J Clin Oncol. 2018;36(17):1658–67.PubMedCrossRef Chesney J, Puzanov I, Collichio F, Singh P, Milhem MM, Glaspy J, et al. Randomized, open-label phase II study evaluating the efficacy and safety of talimogene laherparepvec in combination with ipilimumab versus ipilimumab alone in patients with advanced, unresectable melanoma. J Clin Oncol. 2018;36(17):1658–67.PubMedCrossRef
117.
Zurück zum Zitat Hodi FS, Chiarion-Sileni V, Gonzalez R, Grob JJ, Rutkowski P, Cowey CL, et al. Nivolumab plus ipilimumab or nivolumab alone versus ipilimumab alone in advanced melanoma (CheckMate 067): 4-year outcomes of a multicentre, randomised, phase 3 trial. Lancet Oncol. 2018;19(11):1480–92.PubMedCrossRef Hodi FS, Chiarion-Sileni V, Gonzalez R, Grob JJ, Rutkowski P, Cowey CL, et al. Nivolumab plus ipilimumab or nivolumab alone versus ipilimumab alone in advanced melanoma (CheckMate 067): 4-year outcomes of a multicentre, randomised, phase 3 trial. Lancet Oncol. 2018;19(11):1480–92.PubMedCrossRef
118.
Zurück zum Zitat Bauerschmitz GJ, Kanerva A, Wang M, Herrmann I, Shaw DR, Strong TV, et al. Evaluation of a selectively oncolytic adenovirus for local and systemic treatment of cervical cancer. Int J Cancer. 2004;111(2):303–9.PubMedCrossRef Bauerschmitz GJ, Kanerva A, Wang M, Herrmann I, Shaw DR, Strong TV, et al. Evaluation of a selectively oncolytic adenovirus for local and systemic treatment of cervical cancer. Int J Cancer. 2004;111(2):303–9.PubMedCrossRef
119.
Zurück zum Zitat Angelova AL, Witzens-Harig M, Galabov AS, Rommelaere J. The oncolytic virotherapy era in cancer management: prospects of applying H-1 parvovirus to treat blood and solid cancers. Front Oncol. 2017;7:93.PubMedPubMedCentralCrossRef Angelova AL, Witzens-Harig M, Galabov AS, Rommelaere J. The oncolytic virotherapy era in cancer management: prospects of applying H-1 parvovirus to treat blood and solid cancers. Front Oncol. 2017;7:93.PubMedPubMedCentralCrossRef
120.
Zurück zum Zitat Calton CM, Kelly KR, Anwer F, Carew JS, Nawrocki ST. Oncolytic viruses for multiple myeloma therapy. Cancers (Basel). 2018;10(6). Calton CM, Kelly KR, Anwer F, Carew JS, Nawrocki ST. Oncolytic viruses for multiple myeloma therapy. Cancers (Basel). 2018;10(6).
121.
Zurück zum Zitat Sborov DW, Nuovo GJ, Stiff A, Mace T, Lesinski GB, Benson DM Jr, et al. A phase I trial of single-agent reolysin in patients with relapsed multiple myeloma. Clin Cancer Res. 2014;20(23):5946–55.PubMedPubMedCentralCrossRef Sborov DW, Nuovo GJ, Stiff A, Mace T, Lesinski GB, Benson DM Jr, et al. A phase I trial of single-agent reolysin in patients with relapsed multiple myeloma. Clin Cancer Res. 2014;20(23):5946–55.PubMedPubMedCentralCrossRef
122.
Zurück zum Zitat Dispenzieri A, Tong C, LaPlant B, Lacy MQ, Laumann K, Dingli D, et al. Phase I trial of systemic administration of Edmonston strain of measles virus genetically engineered to express the sodium iodide symporter in patients with recurrent or refractory multiple myeloma. Leukemia. 2017;31(12):2791–8.PubMedPubMedCentralCrossRef Dispenzieri A, Tong C, LaPlant B, Lacy MQ, Laumann K, Dingli D, et al. Phase I trial of systemic administration of Edmonston strain of measles virus genetically engineered to express the sodium iodide symporter in patients with recurrent or refractory multiple myeloma. Leukemia. 2017;31(12):2791–8.PubMedPubMedCentralCrossRef
123.
Zurück zum Zitat Ribas A, Dummer R, Puzanov I, VanderWalde A, Andtbacka RHI, Michielin O, et al. Oncolytic virotherapy promotes intratumoral T cell infiltration and improves anti-PD-1 immunotherapy. Cell. 2017;170(6):1109–19 e10.PubMedCrossRefPubMedCentral Ribas A, Dummer R, Puzanov I, VanderWalde A, Andtbacka RHI, Michielin O, et al. Oncolytic virotherapy promotes intratumoral T cell infiltration and improves anti-PD-1 immunotherapy. Cell. 2017;170(6):1109–19 e10.PubMedCrossRefPubMedCentral
124.
Zurück zum Zitat Sun L, Funchain P, Song JM, Rayman P, Tannenbaum C, Ko J, et al. Talimogene laherparepvec combined with anti-PD-1 based immunotherapy for unresectable stage III-IV melanoma: a case series. J Immunother Cancer. 2018;6(1):36.PubMedPubMedCentralCrossRef Sun L, Funchain P, Song JM, Rayman P, Tannenbaum C, Ko J, et al. Talimogene laherparepvec combined with anti-PD-1 based immunotherapy for unresectable stage III-IV melanoma: a case series. J Immunother Cancer. 2018;6(1):36.PubMedPubMedCentralCrossRef
125.
Zurück zum Zitat Chesney J, Puzanov I, Collichio F, Milhem MM, Hauschild A, Chen L, et al. Patterns of response with talimogene laherparepvec in combination with ipilimumab or ipilimumab alone in metastatic unresectable melanoma. Br J Cancer. 2019;121(5):417–20.PubMedPubMedCentralCrossRef Chesney J, Puzanov I, Collichio F, Milhem MM, Hauschild A, Chen L, et al. Patterns of response with talimogene laherparepvec in combination with ipilimumab or ipilimumab alone in metastatic unresectable melanoma. Br J Cancer. 2019;121(5):417–20.PubMedPubMedCentralCrossRef
126.
Zurück zum Zitat Kelly CM, Antonescu CR, Bowler T, Munhoz R, Chi P, Dickson MA, et al. Objective response rate among patients with locally advanced or metastatic sarcoma treated with talimogene laherparepvec in combination with pembrolizumab: a phase 2 clinical trial. JAMA Oncol. 2020. Kelly CM, Antonescu CR, Bowler T, Munhoz R, Chi P, Dickson MA, et al. Objective response rate among patients with locally advanced or metastatic sarcoma treated with talimogene laherparepvec in combination with pembrolizumab: a phase 2 clinical trial. JAMA Oncol. 2020.
127.
Zurück zum Zitat Rajecki M, af Hallstrom T, Hakkarainen T, Nokisalmi P, Hautaniemi S, Nieminen AI, et al. Mre11 inhibition by oncolytic adenovirus associates with autophagy and underlies synergy with ionizing radiation. Int J Cancer. 2009;125(10):2441–9.PubMedCrossRef Rajecki M, af Hallstrom T, Hakkarainen T, Nokisalmi P, Hautaniemi S, Nieminen AI, et al. Mre11 inhibition by oncolytic adenovirus associates with autophagy and underlies synergy with ionizing radiation. Int J Cancer. 2009;125(10):2441–9.PubMedCrossRef
128.
Zurück zum Zitat Havunen R, Siurala M, Sorsa S, Gronberg-Vaha-Koskela S, Behr M, Tahtinen S, et al. Oncolytic adenoviruses armed with tumor necrosis factor alpha and interleukin-2 enable successful adoptive cell therapy. Mol Ther Oncolytics. 2017;4:77–86.PubMedCrossRef Havunen R, Siurala M, Sorsa S, Gronberg-Vaha-Koskela S, Behr M, Tahtinen S, et al. Oncolytic adenoviruses armed with tumor necrosis factor alpha and interleukin-2 enable successful adoptive cell therapy. Mol Ther Oncolytics. 2017;4:77–86.PubMedCrossRef
129.
Zurück zum Zitat Siurala M, Havunen R, Saha D, Lumen D, Airaksinen AJ, Tahtinen S, et al. Adenoviral delivery of tumor necrosis factor-alpha and interleukin-2 enables successful adoptive cell therapy of immunosuppressive melanoma. Mol Ther. 2016;24(8):1435–43.PubMedPubMedCentralCrossRef Siurala M, Havunen R, Saha D, Lumen D, Airaksinen AJ, Tahtinen S, et al. Adenoviral delivery of tumor necrosis factor-alpha and interleukin-2 enables successful adoptive cell therapy of immunosuppressive melanoma. Mol Ther. 2016;24(8):1435–43.PubMedPubMedCentralCrossRef
130.
Zurück zum Zitat Cervera-Carrascon V, Quixabeira DCA, Havunen R, Santos JM, Kutvonen E, Clubb JHA, et al. Comparison of clinically relevant oncolytic virus platforms for enhancing T-cell therapy of solid tumors. Mol Ther Oncolytics. 2020. Cervera-Carrascon V, Quixabeira DCA, Havunen R, Santos JM, Kutvonen E, Clubb JHA, et al. Comparison of clinically relevant oncolytic virus platforms for enhancing T-cell therapy of solid tumors. Mol Ther Oncolytics. 2020.
131.
Zurück zum Zitat Lee S, Margolin K. Cytokines in cancer immunotherapy. Cancers (Basel). 2011;3(4):3856–93.CrossRef Lee S, Margolin K. Cytokines in cancer immunotherapy. Cancers (Basel). 2011;3(4):3856–93.CrossRef
132.
Zurück zum Zitat Josephs SF, Ichim TE, Prince SM, Kesari S, Marincola FM, Escobedo AR, et al. Unleashing endogenous TNF-alpha as a cancer immunotherapeutic. J Transl Med. 2018;16(1):242.PubMedPubMedCentralCrossRef Josephs SF, Ichim TE, Prince SM, Kesari S, Marincola FM, Escobedo AR, et al. Unleashing endogenous TNF-alpha as a cancer immunotherapeutic. J Transl Med. 2018;16(1):242.PubMedPubMedCentralCrossRef
134.
Zurück zum Zitat Watanabe K, Luo Y, Da T, Guedan S, Ruella M, Scholler J, et al. Pancreatic cancer therapy with combined mesothelin-redirected chimeric antigen receptor T cells and cytokine-armed oncolytic adenoviruses. JCI Insight. 2018;3(7). Watanabe K, Luo Y, Da T, Guedan S, Ruella M, Scholler J, et al. Pancreatic cancer therapy with combined mesothelin-redirected chimeric antigen receptor T cells and cytokine-armed oncolytic adenoviruses. JCI Insight. 2018;3(7).
135.
Zurück zum Zitat Santos JM, Cervera-Carrascon V, Havunen R, Zafar S, Siurala M, Sorsa S, et al. Adenovirus Coding for interleukin-2 and tumor necrosis factor alpha replaces lymphodepleting chemotherapy in adoptive T cell therapy. Mol Ther. 2018;26(9):2243–54.PubMedPubMedCentralCrossRef Santos JM, Cervera-Carrascon V, Havunen R, Zafar S, Siurala M, Sorsa S, et al. Adenovirus Coding for interleukin-2 and tumor necrosis factor alpha replaces lymphodepleting chemotherapy in adoptive T cell therapy. Mol Ther. 2018;26(9):2243–54.PubMedPubMedCentralCrossRef
136.
Zurück zum Zitat Cervera-Carrascon V, Siurala M, Santos JM, Havunen R, Tahtinen S, Karell P, et al. TNFa and IL-2 armed adenoviruses enable complete responses by anti-PD-1 checkpoint blockade. Oncoimmunology. 2018;7(5):e1412902.PubMedPubMedCentralCrossRef Cervera-Carrascon V, Siurala M, Santos JM, Havunen R, Tahtinen S, Karell P, et al. TNFa and IL-2 armed adenoviruses enable complete responses by anti-PD-1 checkpoint blockade. Oncoimmunology. 2018;7(5):e1412902.PubMedPubMedCentralCrossRef
Metadaten
Titel
Oncolytic viruses for cancer immunotherapy
verfasst von
Otto Hemminki
João Manuel dos Santos
Akseli Hemminki
Publikationsdatum
01.12.2020
Verlag
BioMed Central
Erschienen in
Journal of Hematology & Oncology / Ausgabe 1/2020
Elektronische ISSN: 1756-8722
DOI
https://doi.org/10.1186/s13045-020-00922-1

Weitere Artikel der Ausgabe 1/2020

Journal of Hematology & Oncology 1/2020 Zur Ausgabe

Adjuvante Immuntherapie verlängert Leben bei RCC

25.04.2024 Nierenkarzinom Nachrichten

Nun gibt es auch Resultate zum Gesamtüberleben: Eine adjuvante Pembrolizumab-Therapie konnte in einer Phase-3-Studie das Leben von Menschen mit Nierenzellkarzinom deutlich verlängern. Die Sterberate war im Vergleich zu Placebo um 38% geringer.

Alectinib verbessert krankheitsfreies Überleben bei ALK-positivem NSCLC

25.04.2024 NSCLC Nachrichten

Das Risiko für Rezidiv oder Tod von Patienten und Patientinnen mit reseziertem ALK-positivem NSCLC ist unter einer adjuvanten Therapie mit dem Tyrosinkinase-Inhibitor Alectinib signifikant geringer als unter platinbasierter Chemotherapie.

Bei Senioren mit Prostatakarzinom auf Anämie achten!

24.04.2024 DGIM 2024 Nachrichten

Patienten, die zur Behandlung ihres Prostatakarzinoms eine Androgendeprivationstherapie erhalten, entwickeln nicht selten eine Anämie. Wer ältere Patienten internistisch mitbetreut, sollte auf diese Nebenwirkung achten.

ICI-Therapie in der Schwangerschaft wird gut toleriert

Müssen sich Schwangere einer Krebstherapie unterziehen, rufen Immuncheckpointinhibitoren offenbar nicht mehr unerwünschte Wirkungen hervor als andere Mittel gegen Krebs.

Update Onkologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.