Review Open Access
Copyright ©The Author(s) 2015. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Exp Med. Feb 20, 2015; 5(1): 1-10
Published online Feb 20, 2015. doi: 10.5493/wjem.v5.i1.1
Cytomegalovirus in human brain tumors: Role in pathogenesis and potential treatment options
Cecilia Söderberg-Nauclér, Experimental Cardiovascular Research Unit, Department of Medicine, Solna, Center for Molecular Medicine, Karolinska Institutet, Karolinska University Hospital, 171 76 Stockholm, Sweden
John Inge Johnsen, Childhood Cancer Research Unit, Department of Women’s and Children’s Health, Karolinska Institutet, S-171 76 Stockholm, Sweden
Author contributions: Söderberg-Nauclér C and Johnsen JI solely contributed to this paper.
Supported by Grants from Ragnar Söderbergs Foundation; The Swedish Children’s Cancer Foundation; BILTEMA Foundation; Family Ehring Perssons Foundation; Sten A Olssons Foundation; Stichting af Jochnicks Foundation; The Swedish Cancer Society, The Swedish Research Council, the Märta and Gunnar V Philipson Foundation; The Hans and Märit Rausing Charitable Fund; The Dämman Foundation; Swedish Society for Medical Research (SLS), Goljes Memory Foundation; Magnus Bergvalls Foundation; Swedish Society for Medical Research (SSMF) and Tore Nilsons Foundation.
Conflict-of-interest: The authors have no conflicting financial interests (although CS-N earlier held an independent grant from Roche supporting the clinical trial evaluating the efficacy and safety of valganciclovir treatment in glioblastoma patients).
Open-Access: This article is an open-access article which was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/
Correspondence to: Cecilia Söderberg-Nauclér, MD, PhD, Department of Medicine, Solna, Center for Molecular Medicine, Karolinska Institutet, Karolinska University Hospital, CMM L8:03, 171 76 Stockholm, Sweden. cecilia.naucler@ki.se
Telephone: +46-8-51779896 Fax: +46-8-313147
Received: October 1, 2014
Peer-review started: October 5, 2014
First decision: October 28, 2014
Revised: November 13, 2014
Accepted: December 29, 2014
Article in press: December 31, 2014
Published online: February 20, 2015

Abstract

During the last years increasing evidence implies that human cytomegalovirus (CMV) can be attributed to human malignancies arising from numerous tissues. In this perspective, we will review and discuss the potential mechanisms through which CMV infection may contribute to brain tumors by affecting tumor cell initiation, progression and metastasis formation. Recent evidence also suggests that anti-CMV treatment results in impaired tumor growth of CMV positive xenografts in animal models and potentially increased survival in CMV positive glioblastoma patients. Based on these observations and the high tumor promoting capacity of this virus, the classical and novel antiviral therapies against CMV should be revisited as they may represent a great promise for halting tumor progression and lower cancer deaths.

Key Words: Cytomegalovirus, Oncovirus, Glioblastoma, Medulloblastoma, Brain tumor

Core tip: Cytomegalovirus (CMV) has recently been detected in several human cancers. These findings have raised several concerns whether this virus is the cause or a passenger during oncogenesis. Here we discuss the pathogenesis behind CMV infection, its potential as an onco- or oncomodulatory-virus and possible modes of medical interventions.



HUMAN CYTOMEGALOVIRUS

Cytomegalovirus (CMV) is a common virus belonging to the herpesvirus family[1], which infects 70%-100% of the world′s population. After a primary infection that is generally mild or asymptomatic in the immunocompetent host, this virus establishes latency and persistence in myeloid lineage cells, and periodic asymptomatic reactivations are believed to occur during life[1] without clinical signs of infection. However, in individuals with a suppressed immune system, CMV may cause life-threatening infections. Despite a good prophylaxis and surveillance program for early detection of reactivation of CMV in transplant patients, these infections remain to be a clinical problem both as an acute infection as well as a cause of long-term complications. Patients who have had CMV infections in the post-transplant period are at higher risk of developing acute and chronic rejection, cardiovascular diseases, bacterial and fungal infections, post-transplant diabetes and some malignancies[2-5]. These conditions are mainly believed to be mediated by indirect effects of CMV[4], as the virus has been difficult to detect in affected organs at time of diagnosis[6]. However, more recently, the use of sensitive techniques for detection of CMV demonstrate that the presence of CMV in kidney grafts is associated with decreased organ function and graft survival[7,8], suggesting that this virus causes direct, in addition to indirect effects, in the graft.

Emerging evidences also suggest that CMV is highly prevalent in patients with breast, colon, and prostate cancer, rhabdomyosarcoma, hepatocellular cancer, salivary gland tumors, neuroblastoma and brain tumors[9-25]. Over 90% of these tumors have been shown to be positive for CMV proteins and nucleic acids as determined by methods including in situ hybridisation, PCR, electron microscopy, DNA and RNA sequencing, immunostaining of tissue specimens, flow cytometry analyses of tumor cells from surgical resections and western blot analysis. Also most neoplastic cells in sentinel lymph nodes of > 90% of breast cancers have been shown to be CMV positive[26] as well as 98% of brain metastases of colon and breast cancers contain CMV proteins and/or nucleic acids[27]. In sharp contrast, CMV proteins are not detected in healthy tissues surrounding CMV positive tumors or metastases. These observations suggest that this virus does not represent an epiphenomenon of CMV positive tumors, but rather that it may directly aid in tumor progression or even in the initiation and cancer development. As CMV proteins are mainly confined to metastatic cells in metastases of breast and colon cancer[26,27], CMV may be maintained within cells that initiate the development of the metastasis. However, the role of CMV in tumor initiation, progression and metastasis formation needs to be further thoroughly examined in depth since some researchers have failed to detect CMV in tumors[28-33]. This causes confusion in the field and voices have been raised that the presence of CMV in tumors is false and based on artefact data. However, stimulation of dendritic cells with glioblastoma tumor lysates leads to expansion of CMV specific T cells in glioblastoma patients and CMV pp65 specific T cells can kill autologous glioblastoma cells in vitro. This suggests that CMV epitopes are present in glioblastoma tumors[34,35]. The inconsistency in the detection of CMV in tumors samples between laboratories is most likely due to differences in the sample preparations, control specimens and sensitivity of the methods employed for the detection of CMV nucleic acids or proteins. Therefore studies to search for CMV in tumors should be based on techniques that have been developed and proven to work in tumor specimens.

HUMAN MALIGNANT BRAIN TUMORS; GLIOBLASTOMA AND MEDULLOBLASTOMA

Human brain tumors are diverse neoplasms that currently include more than 120 different clinopathological entities[36]. Glioblastomas and medulloblastomas are the most frequently occurring malignant brain tumors in adults and children, respectively. Gliomas accounts for approximately 75% of all primary brain tumors, whereas medulloblastoma is the most common solid childhood tumor[37,38]. The median age of diagnosis of gliomas is 64 years whereas for medulloblastomas about 99% of the tumors are detected in early childhood or adolescence[36,39].

Gliomas are histopathologically classified as ependy-nomas, astrocytomas, oligodendrogliomas and mixed oligo-astrocytomas and graded with regard to malignancy as Grade I-IV. Grade III (anaplastic astrocytomas) and IV gliomas [glioblastoma (GBM); glioblastoma multiforme] account for approximately 82% of cases and are considered malignant or high-grade gliomas[39]. The current standard of care for malignant gliomas includes surgical resection, radiation therapy with concomitant and adjuvant chemotherapy[40]. Despite aggressive treatment approaches, the median survival for patients diagnosed with Grade IV glioblastoma is 16 to 19 mo with 25%-30% of the patients alive 2 years after treatment[41].

Malignant gliomas are heterogenous tumors with different clinical and molecular signatures. Glioblastomas have been divided into four molecular subtypes charac-terized by abnormalities in PDGFRA/IDH1, EGFR, and NF1 that represent the proneural, classical and mechencymal subtypes, respectively[42]. In addition, a fourth transcriptional subtype termed neural subtype is characterized by high expression of neural markers like NEFL, GABRA1, SYT1, and SLC12A5[42]. Many of the molecular abnormalities overlap between the different glioblastoma subtypes and additional rare mutations and chromosomal aberrations have been described, adding to the heterogeneity of glioblastomas. Cells of origin for glioma are not clearly identified but neural stem cells and oligodendrocyte precursor cells that both derive from a neuroepithelial cell, have been suggested[43].

Medulloblastomas are also a compilation of mole-cular and clinical diverse tumor types that arise either in the cerebellum or brainstem[44]. These tumors are highly malignant and current treatment of patients with medulloblastoma consists of surgery, whole brain and spinal cord radiation (in patients above 3 years) and aggressive chemotherapy, sometimes followed by stem cell transplantation[44]. Even though long-term survival among medulloblastoma patients is 60%-70%, the patients frequently experience disease or treatment related complications including developmental, neurological, neuroendocrine, and psychosocial deficits[45]. Although medulloblastomas contain significantly less mutations than adult cancers[46], specific subsets have been identified and aberrant expression of key molecules regulating developmental signaling cascades, oncogenic drivers and mutations in tumor suppressor genes, have been described, such as alterations of the Shh and Wnt signaling pathways, overexpression of MYC, MYCN and activation of growth factor independent 1 family proto-oncogenes, GFI1 and GFI1B by enhancer hijacking[44,47-49]. However, since many medulloblastoma tumors have no apparent mutations of any known cancer gene, it has been suggested that epigenetic changes or other etiological factors including infections also may be responsible for tumor initiation and progression[9,44,46,50].

Medulloblastomas have been linked to disordered mechanisms of normal development and medulloblastoma cells retain many features resembling precursor cells of the embryonic brain. Approximately half of these tumors contain abnormal activation of the developmental signalling cascades, Shh and Wnt[51,52]. Moreover, activation of the PI3K/Akt signalling pathway has been shown to be important for initiation and proliferation of medulloblastoma[53-55]. Molecular analysis have shown that there are four major medulloblastoma subgroups (Wnt, Shh, Group 3 and Group 4[44]). These subgroups are distinct in tumor cell histology and biology and exhibit divergent clinical phenotypes. Different subtypes of medulloblastoma are also believed to have distinct cellular origins. One subtype originates from cerebellar granule neural precursor cells located in the external granular layer of the cerebellum as a result of aberrant Shh signalling[56,57]. A subpopulation of cells from these tumors is positive for the progenitor markers Math1 and CD15[58]. A different medulloblastoma subtype arises outside the cerebellum, likely from cells of the dorsal brainstem and is dependent on Wnt signalling. These tumors contain aberrantly proliferating Zic (+) precursor cells[59]. Finally, evidence of a third medulloblastoma subtype deriving from CD133-positive (Prom1) cerebellar stem cells has also been proposed. These tumors contain elevated Myc expression[60,61].

Many malignant gliomas and medulloblastomas are hence thought to initiate from precursors cells within the CNS by sequential and cumulative genetic alterations or developmental errors[43,49,62]. Rare hereditary syndromes including Cowden, Turcot, Li-Fraumeni, tuberous sclerosis, neurofibromatosis and schwannomatosis have been associated with increased risk of glioma, whereas increased risk of medulloblastomas have been observed in individuals with Turcot, Gorlin and Li-Fraumeni syndromes[62]. However, the etiology behind the vast majority of gliomas and medulloblastomas still remain largely unknown and no fundamental environmental factors, except ionizing radiation that is associated with increased risk for glioma development, has been convincingly demonstrated[37]. On the other hand, it seems to be a strong inverse relationship between atopic diseases and glioma[63]. Both glioblastomas and medulloblastomas express high levels of cyclooxygenase-2 (COX-2) that catalyzes conversion of arachidonic acid to prostaglandins and other eicosanoids with concomitant secretion of proinflammatory prostaglandin E2 (PGE2)[64-66]. In gliomas, the level of COX-2 expression is directly correlated to glioma grade and associated with shorter survival in glioblastoma patients[65]. Non-steroidal anti-inflammatory drugs, capable of inhibiting cyclooxygenase, significantly suppress the growth of glioblastoma and medulloblastoma in preclinical models[9,64,66]. Also, short-term use (< 10 years) of anti-inflammatory medication is associated with a protective effect against glioblastoma[67]. Taken together, this suggests that brain tumors are at least partly dependent on an inflammatory microenvironment in order to proliferate and progress.

An inflammatory microenvironment can be induced directly by tumor cells through activation of oncogenes that activate transcriptional programs leading to the production of pro-inflammatory eicosanoids, cytokines and chemokines that attract different immunological cells to the surrounding tumor microenvironment. Inflammation in the tumor microenvironment can also be caused indirectly by viral and microbial infections, autoimmune diseases, and dietary products[68]. Tumor-related inflammation is hence important for tumor cells to sustain a proliferative state, escape apoptosis and enhance angiogenesis, metastasis and suppression of the immune system[68].

CMV'S POTENTIAL ROLE IN BRAIN TUMORS: CMV PROTEINS CONFER BOTH ONCOMODULATORY AND ONCOGENIC FUNCTIONS

In the light of the above description of the phenotypic and molecular diversity of gliomas and medulloblastomas, and their different sites of origin in the brain, it is interesting to note that most of glioblastomas and medulloblastomas appear to be CMV positive. The presence of CMV proteins in medulloblastoma and glioblastoma hence raise questions whether this virus plays an important role in tumor initiation and/or progression of these tumors. CMV is not a typical oncogenic virus, but CMV proteins provide many mechanisms that can promote tumor biology relevant mechanisms. During the evolution, there has been a strong evolutionary pressure on CMV to cope with and survive the attacks by the immune system and to create efficient virus factories. CMV was believed to encode for approximately 180 proteins, of which only about 45 have been estimated to be essential for virus replication[69-72]. A more recent study based on ribosomal profiling suggests that 751 unique CMV proteins are translated in infected cells[73]; if true, this virus is far more complex than previously appreciated. However, regardless of the exact number of CMV proteins that are encoded by this virus, the vast majority of CMV proteins must confer other functions during the virus life cycle than ensuring replication and formation of new virus particles. Several CMV encoded proteins can under certain circumstances initiate cellular transformation or through other ways aid in tumour development and provide mechanisms representing the cornerstones of hallmarks of cancer[74].

The concept of a role of CMV in cancer is not new. Already in the 1970,s, Fred Rapp’s group reported the frequent presence of CMV in prostate cancer, and isolated a virus strain from tumors that was oncogenic in vitro and in immunodeficient mice[75]. However, in several later studies, CMV failed to transform normal human cells, wherefore this virus was not considered to be oncogenic. The classical view implies that oncoviruses encode gene products that can induce cellular transformation under certain circumstances, e.g., HPV, SV40, EBV, Hepatitis B and adenoviruses. For CMV, the term oncomodulation has instead been proposed to describe the indirect influence of CMV on tumorigenesis (reviewed in[6,76,77]). Oncomodulation is defined as the ability to promote, in an appropriate genetic environment supplied by tumor cells, an oncogenic process characterized by disruptions in intracellular signalling pathways, transcription factors and tumor suppressor proteins. For example, CMV proteins control the cell cycle, induce telomerase activity, inhibits apoptosis, induce angiogenesis and cellular migration and hence provide oncomodulatory mechanisms[76-79]. Furthermore, CMV proteins can promote stemness by blocking cellular differentiation and interact with the DNA damage response pathway to alter the cell cycle (reviewed in[77]). CMV proteins induce expression of oncogenes, control expression of tumor supressors, induce specific chromosomal breaks and p53 mutations, inhibit DNA repair mechanisms, control epigenetic functions and cellular proliferation[80-82], and provide immune evasion strategies[83-85].

Experimental data also suggest that CMV can be oncogenic. A gene region of the CMV genome, the transforming region II (mtrII), a 980-bp sequence, was first shown to transform rodent fibroblasts[86-90]. Expression of the CMV proteins IE72 or IE86 together with the adenovirus E1A protein can induce cellular transformation through a “hit and run” mechanism[91]. The CMV IE proteins can bind to p53, Rb and degrade p21, and thereby modulate cell cycle regulation, induce telomerase activity[78] and downregulate tumor suppressor proteins, which may aid in oncogenic transformation[77]. The CMV protein US28, a G coupled chemokine receptor homologue, has several characteristics resembling a viral oncoprotein[92-95]. Expression of US28 in NIH3T3 cells renders them tumorigenic upon injection in nude mice[94,95], which involves induced COX-2 expression and VEGF production[95]. Furthermore, transgenic mice expressing US28 only in intestinal epithelial cells developed intestinal adenomas and adenocarcinomas[92], by inhibiting glycogen synthase 3β (GSK-3β) activity, resulting in an accumulation of β-catenin and increased expression of Wnt target genes involved in the control of cell proliferation[92]. US28 also induces STAT3 phosphorylation through IL-6 production, which correlates with poor survival in GBM patients[93]. Analysis of clinical GBM samples in situ showed co-localization of US28 with phosphorylated STAT3, COX-2, VEGF and e-NOS, and US28 can induce cellular migration in vitro, which suggests that US28 may contribute to tumour invasiveness and angiogenesis in vivo[77,93,96]. CMV protein expression in mucoepidermoid cancer also correlated with activation of known oncogenic pathways such as EGFR, ERK and amphiregulin, and protein expression was related to severity[97]. These experimental data suggest a direct molecular link between the expression of US28 and tumorigenesis. In addition, US28 has also been shown to activate the transcription factor nuclear factor B (NF-Kβ), a critical regulator of immunity, stress responses, apoptosis, cellular differentiation and migration[96].

More recently, the microenvironment at the tumor site and the potential close connection to inflammation has received increasing attention, and there seems to be a close link between inflammation and tumor development. COX-2 and PGE2 are over-expressed in a number of different cancers and high COX-2 expression is often correlated with poor prognosis[66,98]. CMV infection induces COX-2 and 5-lipoxygenase (5-LO) expression and mediate production of PGE2 and leukotrienes that are both potent inflammatory mediators[99,100]. PGE2 also induces cellular proliferation, angiogenesis, inhibition of apoptosis and stimulation of invasion, and can contribute to the generation of a tumour promoting inflammatory microenvironment[98]. Interestingly, we observed a clear association between CMV protein expression and COX-2 expression in medulloblastoma, suggesting that CMV may control COX-2 expression in these tumors[9]. Viruses could also by their sole presence induce an immune response through expression of non-self peptides to T cells and create an inflammatory microenvironment.

Epithelial cells can undergo a transition into mesen-chymal cells [epithelial to mesenchymal transition (EMT)], involving a series of events resulting in the loss of cell-to-cell contacts and dramatic remodelling of the cytoskeleton. In addition to its role in normal physiological development, recent data implicates a role for EMT and mesenchymal to epithelial transition (EndoMT) in tumor pathology, particularly in regards to metastatic capacity of epithelial tumors. A major factor that regulates the EMT process is transforming growth factor beta (TGFβ). CMVs ability to induce TGFβ provides a role of this virus to facilitate the EMT process[101]. In support of this hypothesis, CMV infected epithelial cells treated with TGFβin vitro, were shown to undergo morphologic and transcriptional changes similar of EMT; this also occurred in uninfected cells[102]. CMV infected epithelial cells can also activate extracellular latent TGFβ1 through induction of metalloproteinase 2 (MMP-2), which was proposed to be mediated by the CMV proteins CMV IE72 or IE86[102]. Induced MMP-2 activity could also in theory mediate degradation of the extra cellular matrix[103], which would further aid in the formation of a metastasis. In addition, CMV US28 can interfere with the activity of expression of GSK3β, which is known to phosphorylate and control the stability of key oncogenic transcription factors such as the Smads and Snail that can trigger an EMT program. Furthermore, virus induced COX-2 expression, and activation of Ras/Erk and PI3K/AKT signalling pathways may further induce and maintain a viscous paracrine loop leading to possible cellular invasion into surrounding stroma.

POTENTIAL TREATMENT OPTIONS TARGETING CMV IN BRAIN TUMORS

We found CMV DNA and proteins in 92% of primary medulloblastoma tumors and in 99% of glioblastomas and also detected the virus in eight of eight examined medulloblastoma cell lines, grown in culture for decades[9]. When the medulloblastoma cells were implanted subcutaneously in immunodeficient mice, all tumors were CMV protein positive; a majority of the tumor cells expressed CMV IE proteins[9]. We found CMV proteins in medulloblastoma cells in culture expressing CD133 and CD15, which are proposed markers of medulloblastoma stem cells. The cancer stem cell hypothesis states that only a subpopulation of cancer cells have self-renewing ability and the capacity to give rise to tumours[104]. Such cancer stem cells or tumour initiating cells (TICs) exhibit an immature phenotype, e.g., expression of pluri-/multipotency associated transcription factors[105], a slower proliferation rate and increased resistance to cancer therapy relative to more differentiated cancer cells[106]. In theory, TICs could either be directly infected by CMV as immature cells or represent a dedifferentiated mature cell with a potential ability to affect tumorigenesis and EMT. CMV infected tumor cells undergoing EMT may detach from adjacent cells and potentially enter the circulation via the lymphatic system or the blood stream. It is possible that CMV positive tumor cells in primary breast, prostate and colon tumors can undergo EMT to obtain stem cell characteristics, i.e., a potential TIC/ EMT cell will circulate, undergo the reverse process of EndoMT as metastases are developing in lymph nodes or distant organs such as the brain. If CMV resides in TICs, it would explain why all xenografted tumors were virus positive, although only a minority of medulloblastoma tumor cells in culture expressed CMV proteins[9]. If this holds true, drugs targeting CMV infection may not only be beneficial to inhibit primary tumor growth but may also reduce the capacity of the tumour to become invasive as it may selectively kill the TICs. Such scenario implies that most stages of tumor development, primary growth, migration, invasion, intravasation and potentially metastasis may be sensitive to anti-viral drugs making CMV an ideal target for therapeutic intervention.

Hence, regardless if CMV plays a role in the develop-ment of tumors, CMV has been detected in brain tumors such as glioblastoma and medulloblastoma and virus infected tumor cells may therefore represent a new target of therapy. In support of this hypothesis, we showed that animals carrying CMV positive human medulloblastoma or neuroblastoma tumors that were treated with anti-CMV drugs had significantly smaller tumors than placebo treated animals[9,107]. A synergistic effect was observed with a COX-2 inhibitor, which resulted in a 72%-97% reduced medulloblastoma tumor growth in vitro and in vivo[9]. Both COX-1 and COX-2 inhibitors are efficient anti-viral drugs that prevent the replication of CMV[99,108]. Thus, antiviral drugs and COX-2 inhibitors may act synergistically to affect the growth of CMV positive tumors[9]. Interestingly, a number of recent studies demonstrate that Aspirin, a non-selective COX- inhibitor, significantly prevents cancer development and metastases[109-113]. With current data at hand, it cannot be excluded that some of the preventive effect of COX-2 inhibitors involves CMV mediated tumor mechanisms.

Treatment of GBM xenograft tumors with the anti-CMV drug Cidofovir also reduced tumor growth, although a CMV independent mechanism was observed[114]. Importantly, treatment of 50 glioblastoma patients who received anti-CMV treatment as an add-on to standard therapy at Karolinska University Hospital as adjuvant treatment demonstrate a remarkably high survival: the 2 year survival was 70% among 40 patients receiving 6 mo of anti-viral therapy and as high as 90% among patients with continuous treatment (n = 25) compared with 18% in contemporary controls (n = 137); median OS was 56.4 mo compared with 13.5 mo in controls in the latter group (P < 0.0001[115]). These observations call for a deeper understanding of CMVs role in cancer and whether this virus is a novel target in anti-cancer therapy. Also, anti-CMV drugs used as an add-on therapy should be further analysed in larger glioblastoma patient populations to give robust statistical data in randomised trials to confirm or dismiss the use of valganciclovir in these patients. The presence of CMV in glioblastoma would also imply that immunotherapy protocols that target CMV epitopes expressed in the tumor can be exploited as cancer therapy[116-118]. Several immunotherapy protocols are currently under evaluation in clinical trials to evaluate different CMV based protocols for glioblastoma patients; these need to also consider the immunosuppressive state of glioblastoma patients.

Several studies indicate that GBM patients exhibit functional impairments of their T cell functions[119,120]. However, polyfunctional CMV specific T cells can be restored by in vitro stimulation with CMV antigens and gamma C cytokines. It was recently demonstrated the CMV pp65 specific T cells can kill autologous glioblastoma cells in vitro[35], and that immunotherapy using CMV specific T cells was associated with prolonged survival in a single patient[18]. This suggests that adoptive therapy of in vitro expanded T cells may be a preferred protocol to be used to overcome the problem of unresponsive T cells in vivo, although T cell activation may be possible to overcome by the right stimulus also in vivo. Today two clinical studies are open for enrolment of GBM patients; one is testing genetically modified CMV specific cytotoxic cells for recurrent GBM; a chimeric antigen receptor recognizes human epidermal growth factor receptor 2 coupled to CD28. The other is based on DC vaccination together with a monoclonal antibody against CD25 aimed to inhibit IL-2 signalling. Results from these trials are expected in 2015 and 2016, respectively, and currently results from two other studies that are closed for recruitment are also awaited in the near future. Most likely, future protocols will have to evaluate the effect of immunotherapy in combination with anti-viral therapy for CMV to obtain optimal anti-CMV effects in cancer patients.

CONCLUSION

In summary, emerging data suggest that CMV may play a pathogenic role in cancers of epithelial and neuronal origin. Under such circumstances, anti-viral treatment strategies may provide new options in cancer therapy of CMV positive tumors and metastases to improve patient outcome.

ACKNOWLEDGMENTS

We apologize to our colleagues whose work we were unable to cite due to space limitations and to the specific focus of this review.

Footnotes

P- Reviewer: Kawasaki H, Nevels M, Tang Q S- Editor: Song XX L- Editor: A E- Editor: Wu HL

References
1.  Mocarski E, Shenk TR, P . Cytomegaloviruses. Fields Virology. 5th ed. Philadelphia: Lippincott Williams and Wilkins 2007; 2701-2772.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Rubin RH. The indirect effects of cytomegalovirus infection on the outcome of organ transplantation. JAMA. 1989;261:3607-3609.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 145]  [Cited by in F6Publishing: 142]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
3.  Hodson EM, Jones CA, Webster AC, Strippoli GF, Barclay PG, Kable K, Vimalachandra D, Craig JC. Antiviral medications to prevent cytomegalovirus disease and early death in recipients of solid-organ transplants: a systematic review of randomised controlled trials. Lancet. 2005;365:2105-2115.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 275]  [Cited by in F6Publishing: 287]  [Article Influence: 15.1]  [Reference Citation Analysis (0)]
4.  Freeman RB. The ‘indirect’ effects of cytomegalovirus infection. Am J Transplant. 2009;9:2453-2458.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 159]  [Cited by in F6Publishing: 161]  [Article Influence: 10.7]  [Reference Citation Analysis (0)]
5.  Razonable R. Direct and indirect effects of cytomegalovirus: can we prevent them? Enferm Infecc Microbiol Clin. 2010;28:1-5.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 30]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
6.  Söderberg-Nauclér C. Does cytomegalovirus play a causative role in the development of various inflammatory diseases and cancer? J Intern Med. 2006;259:219-246.  [PubMed]  [DOI]  [Cited in This Article: ]
7.  Helanterä I, Koskinen P, Finne P, Loginov R, Kyllönen L, Salmela K, Grönhagen-Riska C, Lautenschlager I. Persistent cytomegalovirus infection in kidney allografts is associated with inferior graft function and survival. Transpl Int. 2006;19:893-900.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 65]  [Cited by in F6Publishing: 65]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
8.  Dzabic M, Rahbar A, Yaiw KC, Naghibi M, Religa P, Fellström B, Larsson E, Söderberg-Nauclér C. Intragraft cytomegalovirus protein expression is associated with reduced renal allograft survival. Clin Infect Dis. 2011;53:969-976.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 49]  [Cited by in F6Publishing: 50]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
9.  Baryawno N, Rahbar A, Wolmer-Solberg N, Taher C, Odeberg J, Darabi A, Khan Z, Sveinbjörnsson B, FuskevÅg OM, Segerström L. Detection of human cytomegalovirus in medulloblastomas reveals a potential therapeutic target. J Clin Invest. 2011;121:4043-4055.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 144]  [Cited by in F6Publishing: 154]  [Article Influence: 11.8]  [Reference Citation Analysis (0)]
10.  Cobbs CS, Harkins L, Samanta M, Gillespie GY, Bharara S, King PH, Nabors LB, Cobbs CG, Britt WJ. Human cytomegalovirus infection and expression in human malignant glioma. Cancer Res. 2002;62:3347-3350.  [PubMed]  [DOI]  [Cited in This Article: ]
11.  Scheurer ME, Bondy ML, Aldape KD, Albrecht T, El-Zein R. Detection of human cytomegalovirus in different histological types of gliomas. Acta Neuropathol. 2008;116:79-86.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 165]  [Cited by in F6Publishing: 168]  [Article Influence: 10.5]  [Reference Citation Analysis (0)]
12.  Rahbar A, Stragliotto G, Orrego A, Peredo I, Taher C, Willems J, Söderberg-Naucler C. Low levels of Human Cytomegalovirus Infection in Glioblastoma multiforme associates with patient survival; -a case-control study. Herpesviridae. 2012;3:3.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 57]  [Cited by in F6Publishing: 57]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
13.  Harkins L, Volk AL, Samanta M, Mikolaenko I, Britt WJ, Bland KI, Cobbs CS. Specific localisation of human cytomegalovirus nucleic acids and proteins in human colorectal cancer. Lancet. 2002;360:1557-1563.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 272]  [Cited by in F6Publishing: 271]  [Article Influence: 12.3]  [Reference Citation Analysis (0)]
14.  Harkins LE, Matlaf LA, Soroceanu L, Klemm K, Britt WJ, Wang W, Bland KI, Cobbs CS. Detection of human cytomegalovirus in normal and neoplastic breast epithelium. Herpesviridae. 2010;1:8.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Samanta M, Harkins L, Klemm K, Britt WJ, Cobbs CS. High prevalence of human cytomegalovirus in prostatic intraepithelial neoplasia and prostatic carcinoma. J Urol. 2003;170:998-1002.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 216]  [Cited by in F6Publishing: 217]  [Article Influence: 10.3]  [Reference Citation Analysis (0)]
16.  Ranganathan P, Clark PA, Kuo JS, Salamat MS, Kalejta RF. Significant association of multiple human cytomegalovirus genomic Loci with glioblastoma multiforme samples. J Virol. 2012;86:854-864.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 113]  [Cited by in F6Publishing: 120]  [Article Influence: 9.2]  [Reference Citation Analysis (0)]
17.  Bhattacharjee B, Renzette N, Kowalik TF. Genetic analysis of cytomegalovirus in malignant gliomas. J Virol. 2012;86:6815-6824.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 82]  [Cited by in F6Publishing: 82]  [Article Influence: 6.8]  [Reference Citation Analysis (0)]
18.  Crough T, Beagley L, Smith C, Jones L, Walker DG, Khanna R. Ex vivo functional analysis, expansion and adoptive transfer of cytomegalovirus-specific T-cells in patients with glioblastoma multiforme. Immunol Cell Biol. 2012;90:872-880.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 59]  [Cited by in F6Publishing: 59]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
19.  Ghazi A, Ashoori A, Hanley PJ, Brawley VS, Shaffer DR, Kew Y, Powell SZ, Grossman R, Grada Z, Scheurer ME. Generation of polyclonal CMV-specific T cells for the adoptive immunotherapy of glioblastoma. J Immunother. 2012;35:159-168.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 53]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
20.  Price RL, Bingmer K, Harkins L, Iwenofu OH, Kwon CH, Cook C, Pelloski C, Chiocca EA. Cytomegalovirus infection leads to pleomorphic rhabdomyosarcomas in Trp53+/- mice. Cancer Res. 2012;72:5669-5674.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 31]  [Cited by in F6Publishing: 33]  [Article Influence: 2.8]  [Reference Citation Analysis (0)]
21.  Dziurzynski K, Chang SM, Heimberger AB, Kalejta RF, McGregor Dallas SR, Smit M, Soroceanu L, Cobbs CS. Consensus on the role of human cytomegalovirus in glioblastoma. Neuro Oncol. 2012;14:246-255.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 206]  [Cited by in F6Publishing: 216]  [Article Influence: 18.0]  [Reference Citation Analysis (0)]
22.  Bianchi E, Roncarati P, Hougrand O, Guérin-El Khourouj V, Boreux R, Kroonen J, Martin D, Robe P, Rogister B, Delvenne P. Human cytomegalovirus and primary intracranial tumors: frequency of tumor infection and lack of correlation with systemic immune anti-viral responses. Neuropathol Appl Neurobiol. 2014;Jul 20; Epub ahead of print.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 16]  [Cited by in F6Publishing: 17]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
23.  Lucas KG, Bao L, Bruggeman R, Dunham K, Specht C. The detection of CMV pp65 and IE1 in glioblastoma multiforme. J Neurooncol. 2011;103:231-238.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 93]  [Cited by in F6Publishing: 103]  [Article Influence: 7.4]  [Reference Citation Analysis (0)]
24.  Mitchell DA, Xie W, Schmittling R, Learn C, Friedman A, McLendon RE, Sampson JH. Sensitive detection of human cytomegalovirus in tumors and peripheral blood of patients diagnosed with glioblastoma. Neuro Oncol. 2008;10:10-18.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 270]  [Cited by in F6Publishing: 280]  [Article Influence: 16.5]  [Reference Citation Analysis (0)]
25.  Lepiller Q, Tripathy MK, Di Martino V, Kantelip B, Herbein G. Increased HCMV seroprevalence in patients with hepatocellular carcinoma. Virol J. 2011;8:485.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 28]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
26.  Taher C, de Boniface J, Mohammad AA, Religa P, Hartman J, Yaiw KC, Frisell J, Rahbar A, Söderberg-Naucler C. High prevalence of human cytomegalovirus proteins and nucleic acids in primary breast cancer and metastatic sentinel lymph nodes. PLoS One. 2013;8:e56795.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 94]  [Cited by in F6Publishing: 101]  [Article Influence: 9.2]  [Reference Citation Analysis (0)]
27.  Taher C, Frisk G, Fuentes S, Religa P, Costa H, Assinger A, Vetvik KK, Bukholm IR, Yaiw KC, Smedby KE. High prevalence of human cytomegalovirus in brain metastases of patients with primary breast and colorectal cancers. Transl Oncol. 2014;7:732-740.  [PubMed]  [DOI]  [Cited in This Article: ]
28.  Forslund O, Holmquist Mengelbier L, Gisselsson D. Regarding human cytomegalovirus in neuroblastoma. Cancer Med. 2014;3:1038-1040.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5]  [Cited by in F6Publishing: 5]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
29.  Wick W, Wick A, Platten M. Challenging cytomegalovirus data in glioblastoma. Neuro Oncol. 2014;16:165.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 10]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
30.  Tang KW, Alaei-Mahabadi B, Samuelsson T, Lindh M, Larsson E. The landscape of viral expression and host gene fusion and adaptation in human cancer. Nat Commun. 2013;4:2513.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 221]  [Cited by in F6Publishing: 220]  [Article Influence: 22.0]  [Reference Citation Analysis (0)]
31.  Yamashita Y, Ito Y, Isomura H, Takemura N, Okamoto A, Motomura K, Tsujiuchi T, Natsume A, Wakabayashi T, Toyokuni S. Lack of presence of the human cytomegalovirus in human glioblastoma. Mod Pathol. 2014;27:922-929.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 41]  [Cited by in F6Publishing: 41]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
32.  Lau SK, Chen YY, Chen WG, Diamond DJ, Mamelak AN, Zaia JA, Weiss LM. Lack of association of cytomegalovirus with human brain tumors. Mod Pathol. 2005;18:838-843.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 86]  [Cited by in F6Publishing: 91]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
33.  Tang KW, Hellstrand K, Larsson E. Absence of cytomega-lovirus in high-coverage DNA sequencing of human glioblastoma multiforme. Int J Cancer. 2015;136:977-981.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 31]  [Cited by in F6Publishing: 32]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
34.  Prins RM, Cloughesy TF, Liau LM. Cytomegalovirus immunity after vaccination with autologous glioblastoma lysate. N Engl J Med. 2008;359:539-541.  [PubMed]  [DOI]  [Cited in This Article: ]
35.  Nair SK, De Leon G, Boczkowski D, Schmittling R, Xie W, Staats J, Liu R, Johnson LA, Weinhold K, Archer GE. Recognition and killing of autologous, primary glioblastoma tumor cells by human cytomegalovirus pp65-specific cytotoxic T cells. Clin Cancer Res. 2014;20:2684-2694.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 62]  [Cited by in F6Publishing: 65]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
36.  Louis DN, Ohgaki H, Wiestler OD, Cavenee WK, Burger PC, Jouvet A, Scheithauer BW, Kleihues P. The 2007 WHO classification of tumours of the central nervous system. Acta Neuropathol. 2007;114:97-109.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 7079]  [Cited by in F6Publishing: 7740]  [Article Influence: 455.3]  [Reference Citation Analysis (0)]
37.  Bondy ML, Scheurer ME, Malmer B, Barnholtz-Sloan JS, Davis FG, Il’yasova D, Kruchko C, McCarthy BJ, Rajaraman P, Schwartzbaum JA. Brain tumor epidemiology: consensus from the Brain Tumor Epidemiology Consortium. Cancer. 2008;113:1953-1968.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 587]  [Cited by in F6Publishing: 572]  [Article Influence: 35.8]  [Reference Citation Analysis (0)]
38.  Johnsen JI, Kogner P, Albihn A, Henriksson MA. Embryonal neural tumours and cell death. Apoptosis. 2009;14:424-438.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 43]  [Cited by in F6Publishing: 44]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
39.  Thakkar JP, Dolecek TA, Horbinski C, Ostrom QT, Lightner DD, Barnholtz-Sloan JS, Villano JL. Epidemiologic and molecular prognostic review of glioblastoma. Cancer Epidemiol Biomarkers Prev. 2014;23:1985-1996.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 648]  [Cited by in F6Publishing: 773]  [Article Influence: 77.3]  [Reference Citation Analysis (0)]
40.  Stupp R, Mason WP, van den Bent MJ, Weller M, Fisher B, Taphoorn MJ, Belanger K, Brandes AA, Marosi C, Bogdahn U. Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma. N Engl J Med. 2005;352:987-996.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 14033]  [Cited by in F6Publishing: 14415]  [Article Influence: 758.7]  [Reference Citation Analysis (0)]
41.  Omuro A, DeAngelis LM. Glioblastoma and other malignant gliomas: a clinical review. JAMA. 2013;310:1842-1850.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1419]  [Cited by in F6Publishing: 1525]  [Article Influence: 138.6]  [Reference Citation Analysis (0)]
42.  Verhaak RG, Hoadley KA, Purdom E, Wang V, Qi Y, Wilkerson MD, Miller CR, Ding L, Golub T, Mesirov JP. Integrated genomic analysis identifies clinically relevant subtypes of glioblastoma characterized by abnormalities in PDGFRA, IDH1, EGFR, and NF1. Cancer Cell. 2010;17:98-110.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5060]  [Cited by in F6Publishing: 5155]  [Article Influence: 368.2]  [Reference Citation Analysis (0)]
43.  Swartling FJ, Hede SM, Weiss WA. What underlies the diversity of brain tumors? Cancer Metastasis Rev. 2013;32:5-24.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 14]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
44.  Northcott PA, Jones DT, Kool M, Robinson GW, Gilbertson RJ, Cho YJ, Pomeroy SL, Korshunov A, Lichter P, Taylor MD. Medulloblastomics: the end of the beginning. Nat Rev Cancer. 2012;12:818-834.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 476]  [Cited by in F6Publishing: 466]  [Article Influence: 38.8]  [Reference Citation Analysis (0)]
45.  Mueller S, Chang S. Pediatric brain tumors: current treatment strategies and future therapeutic approaches. Neurotherapeutics. 2009;6:570-586.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 123]  [Cited by in F6Publishing: 127]  [Article Influence: 8.5]  [Reference Citation Analysis (0)]
46.  Parsons DW, Li M, Zhang X, Jones S, Leary RJ, Lin JC, Boca SM, Carter H, Samayoa J, Bettegowda C. The genetic landscape of the childhood cancer medulloblastoma. Science. 2011;331:435-439.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 589]  [Cited by in F6Publishing: 558]  [Article Influence: 42.9]  [Reference Citation Analysis (0)]
47.  Northcott PA, Lee C, Zichner T, Stütz AM, Erkek S, Kawauchi D, Shih DJ, Hovestadt V, Zapatka M, Sturm D. Enhancer hijacking activates GFI1 family oncogenes in medulloblastoma. Nature. 2014;511:428-434.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 414]  [Cited by in F6Publishing: 431]  [Article Influence: 43.1]  [Reference Citation Analysis (0)]
48.  Baryawno N, Sveinbjörnsson B, Eksborg S, Chen CS, Kogner P, Johnsen JI. Small-molecule inhibitors of phosphatidylinositol 3-kinase/Akt signaling inhibit Wnt/beta-catenin pathway cross-talk and suppress medulloblastoma growth. Cancer Res. 2010;70:266-276.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 109]  [Cited by in F6Publishing: 120]  [Article Influence: 8.0]  [Reference Citation Analysis (0)]
49.  Baryawno N, Sveinbjörnsson B, Kogner P, Johnsen JI. Medulloblastoma: a disease with disorganized developmental signaling cascades. Cell Cycle. 2010;9:2548-2554.  [PubMed]  [DOI]  [Cited in This Article: ]
50.  Hovestadt V, Jones DT, Picelli S, Wang W, Kool M, Northcott PA, Sultan M, Stachurski K, Ryzhova M, Warnatz HJ. Decoding the regulatory landscape of medulloblastoma using DNA methylation sequencing. Nature. 2014;510:537-541.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 314]  [Cited by in F6Publishing: 315]  [Article Influence: 31.5]  [Reference Citation Analysis (0)]
51.  Hambardzumyan D, Squatrito M, Carbajal E, Holland EC. Glioma formation, cancer stem cells, and akt signaling. Stem Cell Rev. 2008;4:203-210.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 65]  [Cited by in F6Publishing: 64]  [Article Influence: 4.0]  [Reference Citation Analysis (0)]
52.  Hambardzumyan D, Becher OJ, Holland EC. Cancer stem cells and survival pathways. Cell Cycle. 2008;7:1371-1378.  [PubMed]  [DOI]  [Cited in This Article: ]
53.  Hambardzumyan D, Becher OJ, Rosenblum MK, Pandolfi PP, Manova-Todorova K, Holland EC. PI3K pathway regulates survival of cancer stem cells residing in the perivascular niche following radiation in medulloblastoma in vivo. Genes Dev. 2008;22:436-448.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 360]  [Cited by in F6Publishing: 347]  [Article Influence: 21.7]  [Reference Citation Analysis (0)]
54.  Hartmann W, Digon-Söntgerath B, Koch A, Waha A, Endl E, Dani I, Denkhaus D, Goodyer CG, Sörensen N, Wiestler OD. Phosphatidylinositol 3’-kinase/AKT signaling is activated in medulloblastoma cell proliferation and is associated with reduced expression of PTEN. Clin Cancer Res. 2006;12:3019-3027.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 104]  [Cited by in F6Publishing: 111]  [Article Influence: 6.2]  [Reference Citation Analysis (0)]
55.  Rao G, Pedone CA, Del Valle L, Reiss K, Holland EC, Fults DW. Sonic hedgehog and insulin-like growth factor signaling synergize to induce medulloblastoma formation from nestin-expressing neural progenitors in mice. Oncogene. 2004;23:6156-6162.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 183]  [Cited by in F6Publishing: 434]  [Article Influence: 22.8]  [Reference Citation Analysis (0)]
56.  Schüller U, Heine VM, Mao J, Kho AT, Dillon AK, Han YG, Huillard E, Sun T, Ligon AH, Qian Y. Acquisition of granule neuron precursor identity is a critical determinant of progenitor cell competence to form Shh-induced medulloblastoma. Cancer Cell. 2008;14:123-134.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 510]  [Cited by in F6Publishing: 476]  [Article Influence: 29.8]  [Reference Citation Analysis (0)]
57.  Yang ZJ, Ellis T, Markant SL, Read TA, Kessler JD, Bourboulas M, Schüller U, Machold R, Fishell G, Rowitch DH. Medulloblastoma can be initiated by deletion of Patched in lineage-restricted progenitors or stem cells. Cancer Cell. 2008;14:135-145.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 529]  [Cited by in F6Publishing: 496]  [Article Influence: 31.0]  [Reference Citation Analysis (0)]
58.  Read TA, Fogarty MP, Markant SL, McLendon RE, Wei Z, Ellison DW, Febbo PG, Wechsler-Reya RJ. Identification of CD15 as a marker for tumor-propagating cells in a mouse model of medulloblastoma. Cancer Cell. 2009;15:135-147.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 270]  [Cited by in F6Publishing: 263]  [Article Influence: 17.5]  [Reference Citation Analysis (0)]
59.  Gibson P, Tong Y, Robinson G, Thompson MC, Currle DS, Eden C, Kranenburg TA, Hogg T, Poppleton H, Martin J. Subtypes of medulloblastoma have distinct developmental origins. Nature. 2010;468:1095-1099.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 584]  [Cited by in F6Publishing: 589]  [Article Influence: 42.1]  [Reference Citation Analysis (0)]
60.  Kawauchi D, Robinson G, Uziel T, Gibson P, Rehg J, Gao C, Finkelstein D, Qu C, Pounds S, Ellison DW. A mouse model of the most aggressive subgroup of human medulloblastoma. Cancer Cell. 2012;21:168-180.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 220]  [Cited by in F6Publishing: 217]  [Article Influence: 18.1]  [Reference Citation Analysis (0)]
61.  Pei Y, Moore CE, Wang J, Tewari AK, Eroshkin A, Cho YJ, Witt H, Korshunov A, Read TA, Sun JL. An animal model of MYC-driven medulloblastoma. Cancer Cell. 2012;21:155-167.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 232]  [Cited by in F6Publishing: 225]  [Article Influence: 18.8]  [Reference Citation Analysis (0)]
62.  Hottinger AF, Khakoo Y. Update on the management of familial central nervous system tumor syndromes. Curr Neurol Neurosci Rep. 2007;7:200-207.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 16]  [Cited by in F6Publishing: 17]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
63.  Linos E, Raine T, Alonso A, Michaud D. Atopy and risk of brain tumors: a meta-analysis. J Natl Cancer Inst. 2007;99:1544-1550.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 191]  [Cited by in F6Publishing: 176]  [Article Influence: 10.4]  [Reference Citation Analysis (0)]
64.  Joki T, Heese O, Nikas DC, Bello L, Zhang J, Kraeft SK, Seyfried NT, Abe T, Chen LB, Carroll RS. Expression of cyclooxygenase 2 (COX-2) in human glioma and in vitro inhibition by a specific COX-2 inhibitor, NS-398. Cancer Res. 2000;60:4926-4931.  [PubMed]  [DOI]  [Cited in This Article: ]
65.  Shono T, Tofilon PJ, Bruner JM, Owolabi O, Lang FF. Cyclooxygenase-2 expression in human gliomas: prognostic significance and molecular correlations. Cancer Res. 2001;61:4375-4381.  [PubMed]  [DOI]  [Cited in This Article: ]
66.  Baryawno N, Sveinbjörnsson B, Eksborg S, Orrego A, Segerström L, Oqvist CO, Holm S, Gustavsson B, Kågedal B, Kogner P. Tumor-growth-promoting cyclooxygenase-2 prostaglandin E2 pathway provides medulloblastoma therapeutic targets. Neuro Oncol. 2008;10:661-674.  [PubMed]  [DOI]  [Cited in This Article: ]
67.  Scheurer ME, Amirian ES, Davlin SL, Rice T, Wrensch M, Bondy ML. Effects of antihistamine and anti-inflammatory medication use on risk of specific glioma histologies. Int J Cancer. 2011;129:2290-2296.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 48]  [Article Influence: 3.7]  [Reference Citation Analysis (0)]
68.  Mantovani A, Allavena P, Sica A, Balkwill F. Cancer-related inflammation. Nature. 2008;454:436-444.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6912]  [Cited by in F6Publishing: 7713]  [Article Influence: 482.1]  [Reference Citation Analysis (0)]
69.  Davison AJ, Dolan A, Akter P, Addison C, Dargan DJ, Alcendor DJ, McGeoch DJ, Hayward GS. The human cytomegalovirus genome revisited: comparison with the chimpanzee cytomegalovirus genome. J Gen Virol. 2003;84:17-28.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 319]  [Cited by in F6Publishing: 322]  [Article Influence: 15.3]  [Reference Citation Analysis (0)]
70.  Chee MS, Bankier AT, Beck S, Bohni R, Brown CM, Cerny R, Horsnell T, Hutchison CA, Kouzarides T, Martignetti JA. Analysis of the protein-coding content of the sequence of human cytomegalovirus strain AD169. Curr Top Microbiol Immunol. 1990;154:125-169.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 170]  [Cited by in F6Publishing: 639]  [Article Influence: 18.8]  [Reference Citation Analysis (0)]
71.  Dunn W, Chou C, Li H, Hai R, Patterson D, Stolc V, Zhu H, Liu F. Functional profiling of a human cytomegalovirus genome. Proc Natl Acad Sci USA. 2003;100:14223-14228.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 511]  [Cited by in F6Publishing: 516]  [Article Influence: 24.6]  [Reference Citation Analysis (0)]
72.  Murphy E, Yu D, Grimwood J, Schmutz J, Dickson M, Jarvis MA, Hahn G, Nelson JA, Myers RM, Shenk TE. Coding potential of laboratory and clinical strains of human cytomegalovirus. Proc Natl Acad Sci USA. 2003;100:14976-14981.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 417]  [Cited by in F6Publishing: 411]  [Article Influence: 19.6]  [Reference Citation Analysis (0)]
73.  Stern-Ginossar N, Weisburd B, Michalski A, Le VT, Hein MY, Huang SX, Ma M, Shen B, Qian SB, Hengel H. Decoding human cytomegalovirus. Science. 2012;338:1088-1093.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 460]  [Cited by in F6Publishing: 456]  [Article Influence: 38.0]  [Reference Citation Analysis (0)]
74.  Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144:646-674.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39812]  [Cited by in F6Publishing: 42985]  [Article Influence: 3306.5]  [Reference Citation Analysis (4)]
75.  Geder L, Sanford EJ, Rohner TJ, Rapp F. Cytomegalovirus and cancer of the prostate: in vitro transformation of human cells. Cancer Treat Rep. 1997;61:139-146.  [PubMed]  [DOI]  [Cited in This Article: ]
76.  Cinatl J, Vogel JU, Kotchetkov R, Wilhelm Doerr H. Oncomodulatory signals by regulatory proteins encoded by human cytomegalovirus: a novel role for viral infection in tumor progression. FEMS Microbiol Rev. 2004;28:59-77.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 85]  [Cited by in F6Publishing: 89]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
77.  Soroceanu L, Cobbs CS. Is HCMV a tumor promoter? Virus Res. 2011;157:193-203.  [PubMed]  [DOI]  [Cited in This Article: ]
78.  Strååt K, Liu C, Rahbar A, Zhu Q, Liu L, Wolmer-Solberg N, Lou F, Liu Z, Shen J, Jia J. Activation of telomerase by human cytomegalovirus. J Natl Cancer Inst. 2009;101:488-497.  [PubMed]  [DOI]  [Cited in This Article: ]
79.  Matlaf LA, Harkins LE, Bezrookove V, Cobbs CS, Soroceanu L. Cytomegalovirus pp71 protein is expressed in human glioblastoma and promotes pro-angiogenic signaling by activation of stem cell factor. PLoS One. 2013;8:e68176.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 40]  [Cited by in F6Publishing: 41]  [Article Influence: 3.7]  [Reference Citation Analysis (0)]
80.  Michaelis M, Doerr HW, Cinatl J. The story of human cytomegalovirus and cancer: increasing evidence and open questions. Neoplasia. 2009;11:1-9.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 197]  [Cited by in F6Publishing: 208]  [Article Influence: 13.9]  [Reference Citation Analysis (0)]
81.  Price RL, Song J, Bingmer K, Kim TH, Yi JY, Nowicki MO, Mo X, Hollon T, Murnan E, Alvarez-Breckenridge C. Cytomegalovirus contributes to glioblastoma in the context of tumor suppressor mutations. Cancer Res. 2013;73:3441-3450.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 42]  [Cited by in F6Publishing: 45]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
82.  Esteki-Zadeh A, Karimi M, Strååt K, Ammerpohl O, Zeitelhofer M, Jagodic M, Mehrab-Mohseni M, Sjöholm L, Rahbar A, Söderberg-Nauclér C. Human cytomegalovirus infection is sensitive to the host cell DNA methylation state and alters global DNA methylation capacity. Epigenetics. 2012;7:585-593.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 31]  [Cited by in F6Publishing: 25]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
83.  Pandey JP. Immunoglobulin GM Genes, Cytomegalovirus Immunoevasion, and the Risk of Glioma, Neuroblastoma, and Breast Cancer. Front Oncol. 2014;4:236.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 13]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
84.  Soderberg-Naucler C. Human cytomegalovirus persists in its host and attacks and avoids elimination by the immune system. Crit Rev Immunol. 2006;26:231-264.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 33]  [Cited by in F6Publishing: 35]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
85.  Hanley PJ, Bollard CM. Controlling cytomegalovirus: helping the immune system take the lead. Viruses. 2014;6:2242-2258.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 53]  [Article Influence: 5.3]  [Reference Citation Analysis (0)]
86.  Thompson J, Inamdar A, Jahan N, Doniger J, Rosenthal LJ. Localization and sequence analysis of morphological transforming region III within human cytomegalovirus strain Towne. Intervirology. 1993;36:121-127.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1]  [Cited by in F6Publishing: 1]  [Article Influence: 0.0]  [Reference Citation Analysis (0)]
87.  Inamdar A, Thompson J, Kashanchi F, Doniger J, Brady JN, Rosenthal LJ. Identification of two promoters within human cytomegalovirus morphologic transforming region II. Intervirology. 1992;34:146-153.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 7]  [Cited by in F6Publishing: 7]  [Article Influence: 0.2]  [Reference Citation Analysis (0)]
88.  Razzaque A, Zhu F, Jones C. Functional analysis of human cytomegalovirus morphological transforming region II (mtrII). Virology. 1991;181:399-402.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 9]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
89.  Nelson JA, Fleckenstein B, Jahn G, Galloway DA, McDougall JK. Structure of the transforming region of human cytomegalovirus AD169. J Virol. 1984;49:109-115.  [PubMed]  [DOI]  [Cited in This Article: ]
90.  Kouzarides T, Bankier AT, Barrell BG. Nucleotide sequence of the transforming region of human cytomegalovirus. Mol Biol Med. 1983;1:47-58.  [PubMed]  [DOI]  [Cited in This Article: ]
91.  Shen Y, Zhu H, Shenk T. Human cytomagalovirus IE1 and IE2 proteins are mutagenic and mediate “hit-and-run” oncogenic transformation in cooperation with the adenovirus E1A proteins. Proc Natl Acad Sci USA. 1997;94:3341-3345.  [PubMed]  [DOI]  [Cited in This Article: ]
92.  Bongers G, Maussang D, Muniz LR, Noriega VM, Fraile-Ramos A, Barker N, Marchesi F, Thirunarayanan N, Vischer HF, Qin L. The cytomegalovirus-encoded chemokine receptor US28 promotes intestinal neoplasia in transgenic mice. J Clin Invest. 2010;120:3969-3978.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 86]  [Cited by in F6Publishing: 87]  [Article Influence: 6.2]  [Reference Citation Analysis (0)]
93.  Slinger E, Maussang D, Schreiber A, Siderius M, Rahbar A, Fraile-Ramos A, Lira SA, Söderberg-Nauclér C, Smit MJ. HCMV-encoded chemokine receptor US28 mediates proliferative signaling through the IL-6-STAT3 axis. Sci Signal. 2010;3:ra58.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 151]  [Cited by in F6Publishing: 165]  [Article Influence: 11.8]  [Reference Citation Analysis (0)]
94.  Maussang D, Langemeijer E, Fitzsimons CP, Stigter-van Walsum M, Dijkman R, Borg MK, Slinger E, Schreiber A, Michel D, Tensen CP. The human cytomegalovirus-encoded chemokine receptor US28 promotes angiogenesis and tumor formation via cyclooxygenase-2. Cancer Res. 2009;69:2861-2869.  [PubMed]  [DOI]  [Cited in This Article: ]
95.  Maussang D, Verzijl D, van Walsum M, Leurs R, Holl J, Pleskoff O, Michel D, van Dongen GA, Smit MJ. Human cytomegalovirus-encoded chemokine receptor US28 promotes tumorigenesis. Proc Natl Acad Sci USA. 2006;103:13068-13073.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 167]  [Cited by in F6Publishing: 172]  [Article Influence: 9.6]  [Reference Citation Analysis (0)]
96.  Streblow DN, Soderberg-Naucler C, Vieira J, Smith P, Wakabayashi E, Ruchti F, Mattison K, Altschuler Y, Nelson JA. The human cytomegalovirus chemokine receptor US28 mediates vascular smooth muscle cell migration. Cell. 1999;99:511-520.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 329]  [Cited by in F6Publishing: 334]  [Article Influence: 13.4]  [Reference Citation Analysis (0)]
97.  Melnick M, Sedghizadeh PP, Allen CM, Jaskoll T. Human cytomegalovirus and mucoepidermoid carcinoma of salivary glands: cell-specific localization of active viral and oncogenic signaling proteins is confirmatory of a causal relationship. Exp Mol Pathol. 2012;92:118-125.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 70]  [Cited by in F6Publishing: 77]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
98.  Wang D, Dubois RN. Eicosanoids and cancer. Nat Rev Cancer. 2010;10:181-193.  [PubMed]  [DOI]  [Cited in This Article: ]
99.  Zhu H, Cong JP, Yu D, Bresnahan WA, Shenk TE. Inhibition of cyclooxygenase 2 blocks human cytomegalovirus replication. Proc Natl Acad Sci USA. 2002;99:3932-3937.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 185]  [Cited by in F6Publishing: 187]  [Article Influence: 8.5]  [Reference Citation Analysis (0)]
100.  Qiu H, Strååt K, Rahbar A, Wan M, Söderberg-Nauclér C, Haeggström JZ. Human CMV infection induces 5-lipoxygenase expression and leukotriene B4 production in vascular smooth muscle cells. J Exp Med. 2008;205:19-24.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 54]  [Cited by in F6Publishing: 57]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
101.  Michelson S, Alcami J, Kim SJ, Danielpour D, Bachelerie F, Picard L, Bessia C, Paya C, Virelizier JL. Human cytomegalovirus infection induces transcription and secretion of transforming growth factor beta 1. J Virol. 1994;68:5730-5737.  [PubMed]  [DOI]  [Cited in This Article: ]
102.  Shimamura M, Murphy-Ullrich JE, Britt WJ. Human cytomegalovirus induces TGF-β1 activation in renal tubular epithelial cells after epithelial-to-mesenchymal transition. PLoS Pathog. 2010;6:e1001170.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 38]  [Cited by in F6Publishing: 40]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
103.  Reinhardt B, Winkler M, Schaarschmidt P, Pretsch R, Zhou S, Vaida B, Schmid-Kotsas A, Michel D, Walther P, Bachem M. Human cytomegalovirus-induced reduction of extracellular matrix proteins in vascular smooth muscle cell cultures: a pathomechanism in vasculopathies? J Gen Virol. 2006;87:2849-2858.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 21]  [Cited by in F6Publishing: 25]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
104.  Dirks PB. Brain tumor stem cells: the cancer stem cell hypothesis writ large. Mol Oncol. 2010;4:420-430.  [PubMed]  [DOI]  [Cited in This Article: ]
105.  Ben-Porath I, Thomson MW, Carey VJ, Ge R, Bell GW, Regev A, Weinberg RA. An embryonic stem cell-like gene expression signature in poorly differentiated aggressive human tumors. Nat Genet. 2008;40:499-507.  [PubMed]  [DOI]  [Cited in This Article: ]
106.  Bao S, Wu Q, McLendon RE, Hao Y, Shi Q, Hjelmeland AB, Dewhirst MW, Bigner DD, Rich JN. Glioma stem cells promote radioresistance by preferential activation of the DNA damage response. Nature. 2006;444:756-760.  [PubMed]  [DOI]  [Cited in This Article: ]
107.  Wolmer-Solberg N, Baryawno N, Rahbar A, Fuchs D, Odeberg J, Taher C, Wilhelmi V, Milosevic J, Mohammad AA, Martinsson T. Frequent detection of human cytomegalovirus in neuroblastoma: a novel therapeutic target? Int J Cancer. 2013;133:2351-2361.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 48]  [Cited by in F6Publishing: 51]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
108.  Speir E, Yu ZX, Ferrans VJ, Huang ES, Epstein SE. Aspirin attenuates cytomegalovirus infectivity and gene expression mediated by cyclooxygenase-2 in coronary artery smooth muscle cells. Circ Res. 1998;83:210-216.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 119]  [Cited by in F6Publishing: 114]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
109.  Rothwell PM, Fowkes FG, Belch JF, Ogawa H, Warlow CP, Meade TW. Effect of daily aspirin on long-term risk of death due to cancer: analysis of individual patient data from randomised trials. Lancet. 2011;377:31-41.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1088]  [Cited by in F6Publishing: 1058]  [Article Influence: 81.4]  [Reference Citation Analysis (0)]
110.  Rothwell PM, Price JF, Fowkes FG, Zanchetti A, Roncaglioni MC, Tognoni G, Lee R, Belch JF, Wilson M, Mehta Z. Short-term effects of daily aspirin on cancer incidence, mortality, and non-vascular death: analysis of the time course of risks and benefits in 51 randomised controlled trials. Lancet. 2012;379:1602-1612.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 622]  [Cited by in F6Publishing: 613]  [Article Influence: 51.1]  [Reference Citation Analysis (0)]
111.  Rothwell PM, Wilson M, Elwin CE, Norrving B, Algra A, Warlow CP, Meade TW. Long-term effect of aspirin on colorectal cancer incidence and mortality: 20-year follow-up of five randomised trials. Lancet. 2010;376:1741-1750.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 976]  [Cited by in F6Publishing: 947]  [Article Influence: 67.6]  [Reference Citation Analysis (0)]
112.  Rothwell PM, Wilson M, Price JF, Belch JF, Meade TW, Mehta Z. Effect of daily aspirin on risk of cancer metastasis: a study of incident cancers during randomised controlled trials. Lancet. 2012;379:1591-1601.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 698]  [Cited by in F6Publishing: 707]  [Article Influence: 58.9]  [Reference Citation Analysis (0)]
113.  Algra AM, Rothwell PM. Effects of regular aspirin on long-term cancer incidence and metastasis: a systematic comparison of evidence from observational studies versus randomised trials. Lancet Oncol. 2012;13:518-527.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 563]  [Cited by in F6Publishing: 583]  [Article Influence: 48.6]  [Reference Citation Analysis (0)]
114.  Hadaczek P, Ozawa T, Soroceanu L, Yoshida Y, Matlaf L, Singer E, Fiallos E, James CD, Cobbs CS. Cidofovir: a novel antitumor agent for glioblastoma. Clin Cancer Res. 2013;19:6473-6483.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 51]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
115.  Söderberg-Nauclér C, Rahbar A, Stragliotto G. Survival in patients with glioblastoma receiving valganciclovir. N Engl J Med. 2013;369:985-986.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 143]  [Cited by in F6Publishing: 145]  [Article Influence: 13.2]  [Reference Citation Analysis (0)]
116.  Schuessler A, Walker DG, Khanna R. Cellular immunotherapy directed against human cytomegalovirus as a novel approach for glioblastoma treatment. Oncoimmunology. 2014;3:e29381.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4]  [Cited by in F6Publishing: 6]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
117.  Schuessler A, Smith C, Beagley L, Boyle GM, Rehan S, Matthews K, Jones L, Crough T, Dasari V, Klein K. Autologous T-cell therapy for cytomegalovirus as a consolidative treatment for recurrent glioblastoma. Cancer Res. 2014;74:3466-3476.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 127]  [Cited by in F6Publishing: 131]  [Article Influence: 13.1]  [Reference Citation Analysis (0)]
118.  Nair SK, Sampson JH, Mitchell DA. Immunological targeting of cytomegalovirus for glioblastoma therapy. Oncoimmunology. 2014;3:e29289.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 17]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
119.  Yamanaka R. Novel immunotherapeutic approaches to glioma. Curr Opin Mol Ther. 2006;8:46-51.  [PubMed]  [DOI]  [Cited in This Article: ]
120.  Yamanaka R. Cell- and peptide-based immunotherapeutic approaches for glioma. Trends Mol Med. 2008;14:228-235.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 53]  [Cited by in F6Publishing: 57]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]