Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-23T06:13:57.247Z Has data issue: false hasContentIssue false

Immunological Aspects of Malignant Gliomas

Published online by Cambridge University Press:  13 April 2016

Or Cohen-Inbar
Affiliation:
Department of Neurological Surgery, Rambam Health Care Center, Haifa, Israel Molecular Immunology Laboratory, Haifa, Israel Faculty of Medicine, Technion Israel Institute of Technology, Haifa, Israel Department of Neurological Surgery and Gamma-Knife Radiosurgical Center, University of Virginia Health Care Center, USA.
Menashe Zaaroor
Affiliation:
Department of Neurological Surgery, Rambam Health Care Center, Haifa, Israel Molecular Immunology Laboratory, Haifa, Israel Faculty of Medicine, Technion Israel Institute of Technology, Haifa, Israel
Rights & Permissions [Opens in a new window]

Abstract

Glioblastoma Multiforme (GBM) is the most common malignant primary brain neoplasm having a mean survival time of <24 months. This figure remains constant, despite significant progress in medical research and treatment. The lack of an efficient anti-tumor immune response and the micro-invasive nature of the glioma malignant cells have been explained by a multitude of immune-suppressive mechanisms, proven in different models. These immune-resistant capabilities of the tumor result in a complex interplay this tumor shares with the immune system. We present a short review on the immunology of GBM, discussing the different unique pathological and molecular features of GBM, current treatment modalities, the principles of cancer immunotherapy and the link between GBM and melanoma. Current knowledge on immunological features of GBM, as well as immunotherapy past and current clinical trials, is discussed in an attempt to broadly present the complex and formidable challenges posed by GBM.

Résumé

Aspects immunologiques des gliomes malins. Le glioblastome multiforme (GBM) est le plus fréquent des néoplasmes cérébraux primaires malins. La survie moyenne est de moins de 24 mois et demeure inchangée malgré les progrès importants réalisés par la recherche médicale et les essais thérapeutiques. L’absence de réponse immunitaire anti-tumeur efficace et la nature micro-invasive des cellules malignes du gliome ont été expliquées par une multitude de mécanismes immunosuppresseurs, démontrés dans différents modèles expérimentaux. L’immunorésistance de la tumeur donne lieu à une interaction complexe entre la tumeur et le système immunitaire. Nous présentons une courte revue de l’immunologie du GBM et nous discutons de ses caractéristiques anatomopathologiques et moléculaires uniques, des modalités actuelles de traitement, des principes de l’immunothérapie du cancer et du lien entre le GBM et le mélanome. Nous exposons les connaissances actuelles sur les caractéristiques immunologiques du GBM ainsi que les essais thérapeutiques antérieurs et actuels d’immunothérapie, afin d’esquisser quels sont les défis considérables et complexes que pose le GBM.

Type
Review Articles
Copyright
Copyright © The Canadian Journal of Neurological Sciences Inc. 2016 

Pathological and Molecular Features of GBM

Astrocytoma (a subtype of gliomas – glial cell originated tumors) is graded on a scale ranging from the indolent benign grade-I astrocytoma to low grade (II) and to malignant astrocytoma grade III-IV. Grade-IV Astrocytoma or Glioblastoma Multiforme (GBM) is the most malignant astrocytoma in adults. Glioblastoma Multiforme is highly age specific; the incidence for the population younger than 14 years of age is 0.2.Reference Mao, Desmeules, Semenciw, Hill, Gaudette and Wigle 1 This formidable entity is both the most common and the most malignant glioma in adults, suspected to arise from astrocytes. It usually presents as a unilateral, solitary tumor of the cerebral hemispheres. Its necrotic center may be partially delineated at gross examination, but infiltrating glioma cells can be easily identified microscopically well beyond the apparent gross tumor boundaries. Glioblastoma Multiforme is characterized histologically, like other high grade malignancies, by cellular atypia, mitoses, microvascular proliferation and necrosis. Primary GBM typically develops in patients in their sixth decade and beyond, while secondary GBM tends to arise in younger people (<45 years) through malignant progression from a diffuse astrocytoma World Health Organization (WHO) grade II-III.

The classical histological characterization of GBM resulted in somewhat of a wastebasket category. We now recognize multiple molecular subsets of GBM (Table 1). Current 2007 WHO classification recognizes three major GBM variants, of which the classical form is further distinguished by DNA-chip expression patterns (based on genetic differences) and immunohistochemical (IHC) features (Table 1). Emerging GBM variants having very different prognostic horizons is further proof to the heterogeneity of the GBM definition and obsolete past classification tools. To further add to the confusion, different genetic, and IHC prognostic markers such as the bRAF status, epidermal growth factor receptor (EGF-R) expression pattern, methylation pattern etc., impose further classification onto existing systems. Table 1, presenting the WHO classical forms and emerging variants, does not take into account these markers, which are very important in determining prognosis, and thus is incomplete. The dissection into this obscure general term, identifying clinical entities with very different clinical behavior and prognosis will undoubtly result in redefining and reclassifying this disease in the near future.Reference Karsy 2 - Reference Habberstad, Lind-Landström, Sundstrøm and Torp 3

Table 1 Characteristics of Established GBM Tumor Variants

WHO – World Health Organization, OS – overall survival, PNET - primitive neuroectodermal tumor, GFAP - glial fibrillary acidic protein, EMA -Epithelial Membrane Antigen, CAM - cell adhesion molecule, pCEA - polyclonal antibody against carcinoembryonic antigen, NFP – neurofilament protein, HGG - high-grade glioma.Reference Karsy 2 - Reference Habberstad, Lind-Landström, Sundstrøm and Torp 3

Current Treatment and Prognosis of GBM

The most prevalent brain tumor, GBM is associated with a dismal median overall survival of one to two years and a five-year survival rate of less than 10%.Reference Buckner 4 - Reference DeAngelis 6 In the 1930s, Walter Dandy reported recurrence of contralateral gliomas even after a hemispherectomy, thus illustrating how infiltrative these tumors are. Current day, first-line treatment for GBM patients includes a combination of debulking surgery, chemotherapy and radiotherapy.Reference Stupp, Mason and van den Bent 5 A multimodality regimen of radiotherapy, Temozolomide and Gliadel (carmustine-loaded biodegradable polymers) shows the highest mean overall survival of 20 monthsReference Chaichana, Zaidi and Pendleton 7 - Reference McGirt, Than and Weingart 8 and established current day clinical treatment of GBM. Yet, even this combination has only modestly improved overall survival.Reference Nagasawa, Chow, Yew, Kim, Cremer and Yang 9

O 6-Methylguanine-DNA methyltransferase (MGMT) is an important repair enzyme that contributes to the resistance of tumors against alkylating agents such as carmustine or Temodar. O 6-Methylguanine-DNA methyltransferase promoter methylation silences the gene, thus decreasing DNA repair activity and increasing the susceptibility of the tumor cells to carmustine or Temodar. Temodar (temozolomide) puts methyl groups to guanine, a process halted in part by the repair enzyme product of the MGMT gene. Temodar is currently used for the treatment of GBM regardless of its MGMT promoter methylation status. Consequently, glioblastoma ultimately relapses in almost all patients, and none of the current treatments can effectively prolong survival after relapse.Reference Weller, van den Bent and Hopkins 10 This dismal prognosis, resulted in considerable interest directed towards the development of new therapeutic approaches for this disease.Reference Preusser, Lim, Hafler, Reardon and Sampson 11

Immunotherapy has come to the fore of anti-cancer therapy with the Food and Drug Administraton (FDA) approval of cytokine based therapy with interleukin-2 in 1998 for malignant melanoma, the introduction of sipuleucel-T in 2010 as the first antigen-specific vaccine for treatment resistant prostate cancer, the approval of Ipilimumab, the first checkpoint inhibitor for advanced melanoma in 2011.Reference Drake 12 - Reference Lipson and Drake 13 While some cancer types have been amenable to immunotherapeutic approaches, GBM has not received similar clinical successes, likely due to its poor immunogenicity, few characterized cancer antigens, and for its location in the immunologically distinct central nervous system (CNS).Reference Patel and Pardoll 14 We present a short review at the obstacles and ongoing attempts in developing a potent immunotherapeutic approach and tool to battle GBM.

Immunology Background - The Role of Signal-1 and Signal-2 and the Immune Synapse

T cells that have graduated the thymic selection processes and successfully mounted a T-cell receptor (TCR) are defined as naïve T cells. To elicit their effector functions, they must undergo a multi-step priming process. The initial interaction during activation of a naïve T cell is the binding of the TCR with the MHC-peptide complex [MHC class I for cytotoxic T-lymphocytes (CTLs) or class II for T-helper cells].Reference Klein and Sato 15 This initial crucial step is than stabilized by the binding of CD4 or CD8 co-receptors molecules (in T helper cells or CTLs, respectively) to the MHC molecule (termed as Signal-1). Signal-1 events are accompanied by the interaction of several other receptors (termed as co-stimulatory and co-inhibitory signals) molecules (a process termed as signal-2). Signal-2 establishes the TCR signaling cascade and inducing differentiation and activation of the T-cell (Figure 1). A signal-1 interaction that lacks signal 2 will fail to elicit T-cell activation and differentiation. Different co-stimulatory (e.g. CD28) and co-inhibitory (e.g. CTLA-4) signal-2 molecules drive differentiation to different avenues, ranging from a potent Th1 based anti-tumor response (in case of CD28 activation), to anergy and even apoptosis (in case of CTLA-4 or Fas activation).Reference Schwartz 16

Figure 1 Normal T cell proliferation and mechanisms of glioma cell immunoresistance. Top Right: Normal T cell proliferation. T cell proliferation, differentiation, and cytokine release require two separate signal mechanisms. The first signal involves T cell receptor-mediated recognition of tumor antigen presented by MHC I, which is located on the tumor cell. A second costimulatory signal involves B7 ligand, located on the tumor cell, binding to CD28, a receptor on the T cell. Both of these signals stimulate a variety of intracellular signaling pathways, which lead to upregulated activity of regulator proteins such as nuclear factor-κB, BCl-2, and PI3K. These signals promote T cell activation. However, other ligand-receptor binding pairs can inhibit these cascades and restrict T cell activation. These inhibitory checkpoints include B7 binding to CTLA-4 and B7-H1 (PD-L1) binding to PD-1. Anti-CTLA-4 antibodies (ipilimumab) and anti-PD-1 antibodies facilitate T cell activation by obstructing inhibitory checkpoint processes. Bottom Left: Mechanisms of immunosuppression. glioma cells secrete factors leading to an immunosuppressive tumor microenvironment. Transforming growth factor B (TGFB) and prostaglandin E-2 downregulate the expression of MHC, restricting antigen presentation and T cell proliferation. Interleukin-6. interleukin-10 and vascular endothelial growth factor are potent STAT-3 activators, leading to the proliferation of immature dendritic cells (DCs) that are not able to function as APCs. These immature DCs also secrete TGFB which aid in the proliferation of immunosuppressive T-reg cells and STAT-3 positive TH17 cells. Glioma cells downregulate MHC on their surface leading to the decreased antigen presentation and decreased T cell proliferation. Downregulation of B7 works via a similar mechanism in that the Costimulatory signal is lost preventing T cell proliferation. Increased expression of B7-H1 and FasL act as proapoptotic signals for T cells.Reference Cohen-Inbar 69 , Reference Jackson, Ruzevick, Phallen, Belcaid and Lim 97 , Reference Caruso, Cohen-Inbar, Bilsky, Gerszten and Sheehan 98

The contact area between the T cell surface receptors and the antigen-presenting cells or target cell’s surface receptors has been termed as the immunological synapse (Figure 1). Several studies have estimated that T cell activation requires as little as a single MHC-peptide complex.Reference Irvine, Purbhoo, Krogsgaard and Davis 17 - Reference Stefanova, Dorfman and Germain 18 Recent evidence supports the concept that a single MHC-peptide complex can activate CD4+ T cells, whereas binding 10 MHC-peptide complexes are needed for synapse formation. The binding of more than 60 to 70 MHC-peptide complexes can activate the T cells without co-stimulatory receptors interaction (over-ruling the need for signal 2). Such sensitivity was also demonstrated for CTLs.Reference Davis, Krogsgaard and Huppa 19

The Principles of Cancer Immunotherapy

The categorization of the CNS as an immune privileged site has perhaps retarded the development of immunotherapies for brain tumours, with the temptation to anticipate ineffective immune function in the brain.Reference Walker, Calzascia and Dietrich 20 Different features of the CNS were identified that were proposed to explain this apparent lack of immune reactivity. These included the presence of the blood–brain barrier (BBB), low MHC expression in the brain parenchyma, the absence of organized lymphatic drainage and a lack of dendritic cells in the normal brain parenchyma. Nevertheless, it is now apparent that immune reactions can and do occur in the CNS: autoimmune diseases of the CNS,Reference Hohlfeld and Wekerle 21 immune responses to bacteria, neurotropic virusesReference Dörries 22 and parasitesReference Fischer, Bonifas and Reichmann 23 and anti-tumor responses, as discussed.

The first requirement for an effector T cell is that it must reach its target (i.e. the tumor). This problem of adequate tumor infiltration is relevant to all solid cancers, more-over in the case of tumors located in the brain parenchyma, in which case the T-cell must also penetrate through the blood brain barrier formed by the tight junctions between the endothelial cells, basement membrane and astrocytic podocytes.Reference Rubin and Staddon 24 Yet, for spontaneous malignant astrocytoma in humans, the integrity of the BBB was shown to be locally compromised and tumor-induced angiogenesis does not incorporate BBB characteristics.Reference Walker, Calzascia and Dietrich 20 In conclusion, the particular requirements for mounting an immune responses against tumors in the CNS include several steps: (1) Effector T cells must penetrate the brain parenchyma before reaching the tumor bed. (2) Target antigens must permit adequate discrimination of normal versus malignant tissue. (3) Tumor cells must express sufficient MHC molecule for direct specific CTL’s effector response. (4) Brain inflammation must be regulated. (5) T-cells must retain their anti-tumor effector function during migration through the parenchyma and its resident cells. (6) Effector cell functionality must be retained during the encounter with the tumor cell.

In general terms, immunotherapy can be either active or passive. Active immunotherapy can be specific or non-specific. Passive immunity can be antibody based or cellular effector based. Active non-specific approaches have yielded potent immunological anti-tumor responses. One important example is the Bacillus Calmette –Guérin (BCG) vaccine. Examples of active specific approaches include immunization with specific tumor antigens or with peptide pulsed dendritic cells. These approaches are known to achieve meaningful clinical responses in a variety of tumors. Tumor antigens have been detected using T lymphocytes isolated from patients or by using SEREX (screening for auto-antibodies). Vaccines can be based on whole or lysed tumor cells, defined protein or peptide antigens, nucleic acids, heat shock proteins (Table 2).

Table 2 Selected Immunotherapeutic Trials for Malignant Gliomas

LAKs – Lymphokine activated killer cells, CTLs – Cytotoxic T-lymphocytes, HSP – heat shock proteins, EGFR – epidermal growth factor receptor, TGF – transforming growth factor, IL – interleukin, IFN – interferon, C -____?___, MS – median survival, PFS - progression free survival, TTP – time to progression, CR – complete response, PR – partial response, MR – minimal response, OS – overall survival.

The use of monoclonal antibodies (mAbs) and recombinant antibody fragments for passive specific or non-specific immunotherapy are in widespread use. Several mAbs were approved by the FDA for clinical use either as a monotherapy or in combination with other chemotherapy agents. Targeted therapies using mAbs include both unconjugated and conjugated monoclonal antibodies. Unconjugated antibodies can induce recruitment and activation of effector cells by different mechanisms [antibody dependent cellular cytotoxicity (ADCC), cell dependent cytotoxicity (CDC)], blocking the interaction between receptor-ligand or by inducing apoptosis. Conjugated antibodies are used for the delivery of radioisotopes, enzymes, toxins or drugs to the tumor.Reference Harris 25 - Reference Zafir-Lavie, Michaeli and Reiter 26

Passive cellular effector based immunotherapy include techniques such as adoptive cell transfer, which includes isolation of tumor infiltrating lymphocytes (TIL's), their ex-vivo expansion and administration back to the patient. This simple approach is showing promising results in melanoma treatment. These cells can be given together with systemic cytokines, such as interleukin-2, or transfected with cytokine genes, such as tumour necrotic factor. Another variation to this approach involves activation of peripheral blood mononuclear cells ex vivo [lymphokine-activated killer (LAK)] cells prior to reinjection. Evolving tetramer technology has made it possible to select CD8+ T cells from the peripheral blood having high affinity for some particular peptide epitopes.Reference Davis 27

The Link between Glioblastoma and Melanoma

Glial cells and melanocytes are both originated from the neuro-ectoderm and, as such, share embryonic antigens. Although melanoma associated antigens (MAA) are well characterized, glial derived antigens are less studied. Melanogenesis related proteins are characteristic of melanoma yet are found also in neuroectodermal originated organs such as the cochlea, the eye, and glial cells. This antigen resemblance was also proven between melanomas and GBM.Reference Princiotta, Finzi and Qian 28 - Reference Liu, Ying, Zeng, Wheeler, Black and Yu 29 In addition, melanoma associated differentiation antigens (DA) such as MART-1, Gp-100, Tyrosinase, Gage, and MAGE have been proven to be present in glioblastoma and other neuro ectodermal originated cell lines and tumors.Reference Zhang, Eguchi and Kruse 30 Glioblastoma multiforme specific antigens are poorly characterized. Most studies characterizing GBM associated antigens have used detection methods such as detecting messenger RNA levels, intracellular protein levels and T-cell activity assays,Reference Zhang, Eguchi and Kruse 30 - Reference Prins, Odesa and Liau 32 proving only the mere presence of intracellular antigens, or T-cell activity against a certain examined peptide, derived from intracellular antigens. The presence of an intracellular protein was shown not to correlate with its HLA antigenic presentation, which is dependent, among other, on stability and intracellular processing.Reference Zhang, Eguchi and Kruse 30 - Reference Prins, Odesa and Liau 32 These features, inter alia, explain both the rational and the relative success in implementing anti-melanoma immunotherapeutic approaches (such as utilizing the immune-stimulating effect of anti-CTLA4 and anti-PD1 antibodies Ipilimumab and pembrolizumab) in the treatment of high grade gliomas. It also sets the rational for employing targeted therapy (monoclonal antibodies, MHC-tetramers etc.) against melanoma associated differentiation antigens (DA) for GBM patients.

The Immunology of GBM

As with systemic neoplasia, the cause of death from GBM is not always clear. Although some patients succumb to the mass effect and subsequent cerebral herniation, others die without a clear evidence of significant mass effect.Reference Silbergeld, Rostomily and Alvord 33 Because these tumors rarely metastasize outside the CNS, the typical progressive neurological deterioration and eventual demise is most likely caused by injury to the brain mediated by unknown mechanisms. Although these tumors remain rapidly fatal, some long-term survivors have been reported (Table 1).

Patients with GBM exhibit impaired antitumor immunity and impaired systemic immunity leading to bacterial infections. The former relies on local cellular immunity mediated by the Th1 subset of helper T cells, while the latter relies on systemic humoral immunity mediated by the Th2 subset of helper T cells.Reference Aghi, Batchelor, Louis, Barker and Curry 34 - Reference Parsa, Waldron and Panner 35 Past empiric observations reported that patients with GBM who experienced a postoperative cranial wound infection exhibited longer survival. This observation was initially attributed to bacterial lipopolysaccharide (LPS) eliciting a nonspecific immune response which also targets the tumor. Since these observations a plethora of studies focused on characterizing the immunological roles GBM have been conducted.

Patients harboring GBM were shown to exhibit low number of circulating T-cells, rendering them prone to viral infections, abnormal delayed type hypersensitivity and impaired cytotoxic T-cell reaction.Reference Hao, Parney, Roa, Turner, Petruk and Ramsay 36 Fecci et alReference Fecci, Mitchell and Whitesides 37 reported a greater than 2.5-fold increase in the proportion of circulating regulatory T cells (CD4+FoxP+ T cells) in patients with newly-diagnosed or recurrent GBM. This effect was noted not be related to the co-administration of glucocorticoids.Reference Fecci, Mitchell and Whitesides 37 The frequency of T regulatory cells (T-regs) was shown to correlate directly with in vitro suppression of T cell activation.Reference El Andaloussi and Lesniak 38 - Reference See, Parker and Waziri 39 In contrast to the evidence documenting the expansion of T-regs within the peripheral T cell compartment, it remains a matter of some debate as to whether T-regs are found at increased frequency within TIL from patients with GBM.Reference Fecci, Mitchell and Whitesides 37 - Reference See, Parker and Waziri 39

The immunosuppressive effect the tumor imparts can be demonstrated both locally and systemically. A multitude of immunosuppressive mechanisms were suggested and proven in different models, trying to capture and define the complex interplay this tumor shares with the immune system. Multiple genetic pathways are known to be altered in GBM, including p16/pRb/CDK4, p53/ MDM2/ p14ARF, EGF-R [with unique variants like the EGFRvIII emerging in some tumors, responsive to Erlotinib (EGF-receptor tyrosine kinase inhibitor)], platelet-derived growth factor receptor (PDGF), and PI3-kinase/PTEN.Reference Choe, Horvath and Cloughesy 40 - Reference Louis, Holland and Cairncross 42 The most common genetic alteration in GBM tumors is the loss of heterozygosity of chromosome #10, occurring in 80~95% of these tumors.Reference Aghi, Batchelor, Louis, Barker and Curry 34 - Reference Parsa, Waldron and Panner 35 , Reference Choe, Horvath and Cloughesy 40 - Reference Louis, Holland and Cairncross 42 Interestingly, disruption of two of these tumor suppressor genes found on chromosome 10q may mediate a decrease in tumor cell immunogenicity (i.e. DBMT1, PTEN). DBMT1 is a candidate tumor suppressor for brain, gastrointestinal, and lung cancer. Phosphatase and tensin (PTEN) homologue deleted in chromosome 10q23.3) is an inhibitor of the PI3 kinase-signalling pathway, whose disruption may increase expression of immunosuppressive protein B7-H1 and also increases Th2 type cytokines release,Reference Choe, Horvath and Cloughesy 40 - Reference Louis, Holland and Cairncross 42 supporting the evolution of anergy and tolerance to the tumor (Figure 1).

Another mechanism suggested is the tumor’s ability to down-regulate or express low levels of class-I MHC,Reference Yang, Ng and Lillehei 43 hiding its existence from the cellular arm of the immune system, as well as the aberrant expression of non-classical MHC class I molecule (class Ib) termed HLA-G, structurally related to classical MHC class Ia (HLA-A, -B, -C), that has been attributed functions as an antigen-presenting molecule but also immune regulatory functions. HLA-G expression, render cells highly resistant to direct alloreactive lysis, inhibits the alloproliferative response, and prevents efficient priming of cytotoxic T cells. The inhibitory effects of HLA-G are directed against CD8 expressing and CD4 expressing T cells but appear to be NK (natural killer) cell independent.Reference Pistoia, Morandi, Wang and Ferrone 44 - Reference Wiendl, Mitsdoerffer and Weller 45

Another mechanism involves the upregulation of anti-apoptotic proteins (IAP’s – inhibitors of apoptosis), such as the survivin proteinReference Das, Tan, Teo and Smith 46 by the GBM tumor cells, rending these cells immortal. Finally, parallel and in consort with all mechanisms mentioned, the tumor microenvironment is characterized by its immunosuppressive nature.Reference Gomez and Kruse 47 The cytokines secreted, be it directly or indirectly by the GBM tumor cells, mediate immune-anergy and tumor proliferation (Interferon-γ, IL-10, TGFβ)Reference Hao, Parney, Roa, Turner, Petruk and Ramsay 36 , Reference Zisakis, Piperi and Themistocleous 48 The interplay between the different mechanisms stated is complex and largely unknown. A schematic representation of key mechanisms is presented in Figure 1.

GBM Immunotherapy – Past trials

Many immunotherapeutic approaches have been tested in order to facilitate cure or a meaningful clinical response. Clinically speaking, such approaches are incorporated into current treatment algorithms. The need for tissue diagnosis and for surgical debulking will likely always be a component of treatment, as will be some form of focused radiation (radiosurgery, intensity-modulated radiation therapy, external beam radiation therapy, etc.). Table 2 depicts some examples of clinical trials, based on immunological mediators and mechanisms, highlighting the preliminary response. One should note that, although different strategies have been employed through the years (Table 2), early studies did not employ combined treatments. Current surgical, radiotherapeutic and medical alkylating agents were not available early on and their failure may not necessarily imply conceptual failure. As described previously and presented in Figure 1, most immunological avenues have been harnessed in an attempt to circumvent the GBM induced anti-tumor immunosuppression. Dendritic cell vaccination,Reference Yamanaka 49 autologous formalin-fixed tumor vaccines,Reference Muragaki, Maruyama and Iseki 50 cytokine gene therapy,Reference Palu, Cavaggioni and Calvi 51 adoptive cell transfer based therapies,Reference Merchant, Grant, Merchant and Young 52 cytokine modulation, lymphokine activated killer (LAK) cells based approaches and CTL based approaches previously discussed, are noted. Heat shock proteins based vaccines concept is based on evidence that these proteins are implicated in the activation of both innate and adaptive immune systems. Vaccines formulated from heat shock protein-peptide complexes, derived from autologous tumor, have been applied to the field of immunotherapy for glioblastoma.Reference Parsa, Crane and Wilson 53 - Reference Ampie, Choy, Lamano, Fakurnejad, Bloch and Parsa 54 In general, although some encouraging clinical trials have been conducted, these are limited.

In the past five years, immunotherapy with immune checkpoint inhibitors has shown promise and clinical results in fighting tumors resistant to conventional therapies, such as melanoma and lung adenocarcinoma (non-small cell lung carcinoma, (NSCLC)).Reference Topalian, Hodi and Brahmer 55 - Reference Rizvi, Mazières and Planchard 59 These compounds facilitate effective antineoplastic immune response by antibody mediated suppression of co-inhibitory receptors and pathways (co-inhibitory signal-2). This inhibition tilts the scales in favor of a potent CTL’s activation, and not anergy as discussed. A key point in these modalities is the finding that immune checkpoint inhibitors can induce a deep and durable remission with an acceptable safety and side-effect profile.Reference Topalian, Hodi and Brahmer 55 - Reference Rizvi, Mazières and Planchard 59 The FDA recently approved the two checkpoint inhibitors that target programmed cell death protein 1 (PD1) in late 2014 (pembrolizumab and nivolumab for metastatic melanoma), and nivolumab for non small-cell lung carcinoma (NSCLC) in early 2015. 60 - 61 The first large phase-III trial of nivolumab in GBM patients (NCT02017717) was initiated in 2014 and initial results seem promising.Reference Preusser, Lim, Hafler, Reardon and Sampson 11

Future Directions

Immunotherapy for GBM, be it passive or active and specific or non-specific, is being actively studied in preclinical models and translated to clinical trials. Combining immunotherapy modalities or treatment regimens involving both standard therapies and immunotherapies show promise as powerful anti-cancer therapies in GBM. A phase I clinical trial studying the effects of anti-PD-1 and anti-CTLA-4 combination therapy is currently recruiting for recurrent GBM and a number of studies of dendritic cell vaccines in recurrent and newly diagnosed GBM are also underway. A phase-2 study of concurrent radiation therapy, temozolomide and the histone deacetylase inhibitor (HDAC) valproic acid showed improved outcomes and merits further investigation.Reference Krauze, Myrehaug and Chang 62 The combined use of multiple checkpoint inhibitors with other intracellular enzymes whose expression correlate to malignancy and checkpoint inhibition such as the Indoleamine 2,3 dioxygenase 1 (IDO) catabolic enzyme, has recently shown a survival advantage in a mouse model.Reference Wainwright, Chang and Dey 63 - Reference Castro, Baker and Lowenstein 64

Pioneering work done with TCR-like antibodies,Reference Cohen, Hoffmann, Farago, Hoogenboom, Eisenbach and Reiter 65 - Reference Denkberg, Klechevsky and Reiter 68 MMP targeting,Reference Cohen-Inbar 69 scorpion and spider toxin used targeting,Reference Cohen-Inbar 69 in which a modular targeting moiety, consisting of a single chain antibody fragment (ScFv) or another targeting peptide, linked via a flexible linker to an effector moiety holds great promise. The effector domain is an HLA-A2 molecule (human leukocyte antigen) bearing a highly immunogenic, viral peptide. This strategy enables recruitment and redirection of previously formed, highly potent memory CTLs to the tumor’s milieu, thus presenting tumor cells as viral infected to the immune system.

It seems that in the fight against GBM, this challenge will be met through the use of a multidisciplinary combined treatment approach, utilizing the advantages offered by surgical debulking, focused radiation, alkylating agents, pro-inflammatory agents (such as checkpoint blockers) and T-cell recruitment to the tumor milieu, each harnessing different anti-tumor mechanism and working synergistically.

Conflict of Interest Disclosure

The author has no personal or institutional financial interest in drugs or materials in relation to this paper. Or Cohen-Inbar and Menashe Zaaroor do not have anything to disclose.

References

1. Mao, YM, Desmeules, M, Semenciw, RM, Hill, G, Gaudette, L, Wigle, DT. Increasing brain cancer rates in Canada. CMAJ. 1991;145:1583-1591.Google ScholarPubMed
2. Karsy, M. Mechanisms of gliomagenesis and glioblastoma multiforme variants. Folia Neuropathol. 2012;50:301-321.CrossRefGoogle Scholar
3. Habberstad, AH, Lind-Landström, T, Sundstrøm, S, Torp, SH. Primary human glioblastomas - prognostic value of clinical and histopathological parameters. Clin Neuropathol. 2012;31:361-368.10.5414/NP300439CrossRefGoogle ScholarPubMed
4. Buckner, JC. Factors influencing survival in high-grade gliomas. Semin Oncol. 2003;30:10-14.10.1053/j.seminoncol.2003.11.031CrossRefGoogle ScholarPubMed
5. Stupp, R, Mason, WP, van den Bent, MJ, et al. Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma. N Engl J Med. 2005;352:987-996.10.1056/NEJMoa043330CrossRefGoogle ScholarPubMed
6. DeAngelis, LM. Brain tumors. N Engl J Med. 2001;344:114-123.10.1056/NEJM200101113440207CrossRefGoogle ScholarPubMed
7. Chaichana, KL, Zaidi, H, Pendleton, C, et al. The efficacy of carmustine wafers for older patients with glioblastoma multiforme: prolonging survival. Neurol Res. 2011;33:759-764.CrossRefGoogle ScholarPubMed
8. McGirt, MJ, Than, KD, Weingart, JD, et al. Gliadel (BCNU) wafer plus concomitant temozolomide therapy after primary resection of glioblastoma multiforme. J Neurosurg. 2009;110:583-588.10.3171/2008.5.17557CrossRefGoogle ScholarPubMed
9. Nagasawa, DT, Chow, F, Yew, A, Kim, W, Cremer, N, Yang, I. Temozolomide and other potential agents for the treatment of glioblastoma multiforme. Neurosurg Clin N Am. 2012;23:307-322.CrossRefGoogle ScholarPubMed
10. Weller, M, van den Bent, M, Hopkins, K, et al. EANO guideline for the diagnosis and treatment of anaplastic gliomas and glioblastoma. Lancet Oncol. 2014;s15:e395-e403.CrossRefGoogle Scholar
11. Preusser, M, Lim, M, Hafler, DA, Reardon, DA, Sampson, JH. Prospects of immune checkpoint modulators in the treatment of glioblastoma. Nat Rev Neurol. 2015, Sep;11(9):504–14.10.1038/nrneurol.2015.139CrossRefGoogle ScholarPubMed
12. Drake, CG. Prostate cancer as a model for tumour immunotherapy. Nat Rev Immunol. 2010;10:580-593.10.1038/nri2817CrossRefGoogle Scholar
13. Lipson, EJ, Drake, CG. Ipilimumab: an anti-CTLA-4 anti-body for metastatic melanoma. Clin Cancer Res. 2011;17:6958-6962.10.1158/1078-0432.CCR-11-1595CrossRefGoogle Scholar
14. Patel, MA, Pardoll, DM. Concepts of immunotherapy for glioma. J Neurooncol. 2015;123:323-330.10.1007/s11060-015-1810-5CrossRefGoogle ScholarPubMed
15. Klein, J, Sato, A. The HLA system. First of two parts. N Engl J Med. 2000;343:702-709.CrossRefGoogle ScholarPubMed
16. Schwartz, RH. T cell anergy. Annu Rev Immunol. 2003;21:305-334.10.1146/annurev.immunol.21.120601.141110CrossRefGoogle ScholarPubMed
17. Irvine, DJ, Purbhoo, MA, Krogsgaard, M, Davis, MM. Direct observation of ligand recognition by T cells. Nature. 2002;419:845-849.10.1038/nature01076CrossRefGoogle ScholarPubMed
18. Stefanova, I, Dorfman, JR, Germain, RN. Self-recognition promotes the foreign antigen sensitivity of naive T lymphocytes. Nature. 2002;420:429-434.CrossRefGoogle ScholarPubMed
19. Davis, MM, Krogsgaard, M, Huppa, JB, et al. Dynamics of cell surface molecules during T cell recognition. Annu Rev Biochem. 2003;72:717-742.10.1146/annurev.biochem.72.121801.161625CrossRefGoogle ScholarPubMed
20. Walker, PR, Calzascia, T, Dietrich, PY. All in the head: obstacles for immune rejection of brain tumours. Immunology. 2002;107:28-38.CrossRefGoogle ScholarPubMed
21. Hohlfeld, R, Wekerle, H. Immunological update on multiple sclerosis. Curr Opin Neurol. 2001;14:299-304.CrossRefGoogle ScholarPubMed
22. Dörries, R. The role of T-cell-mediated mechanisms in virus infections of the nervous system. Curr Top Microbiol Immunol. 2001;253:219-245.Google ScholarPubMed
23. Fischer, HG, Bonifas, U, Reichmann, G. Phenotype and functions of brain dendritic cells emerging during chronic infection of mice with Toxoplasma gondii. J Immunol. 2000;164:4826-4834.10.4049/jimmunol.164.9.4826CrossRefGoogle ScholarPubMed
24. Rubin, LL, Staddon, JM. The cell biology of the blood–brain barrier. Annu Rev Neurosci. 1999;22:11-28.CrossRefGoogle ScholarPubMed
25. Harris, M. Monoclonal antibodies as therapeutic agents for cancer. Lancet Oncol. 2004;5:292-302.10.1016/S1470-2045(04)01467-6CrossRefGoogle ScholarPubMed
26. Zafir-Lavie, I, Michaeli, Y, Reiter, Y. Novel antibodies as anticancer agents. Oncogene. 2007;26:3714-3733.10.1038/sj.onc.1210372CrossRefGoogle ScholarPubMed
27. Davis, ID. An overview of cancer immunotherapy. Immunol Cell Biol. 2000;78:179-195.10.1046/j.1440-1711.2000.00906.xCrossRefGoogle ScholarPubMed
28. Princiotta, MF, Finzi, D, Qian, SB, et al. Quantitating protein synthesis, degradation, and endogenous antigen processing, Immunity. 2003;18:343-354.Google Scholar
29. Liu, G, Ying, H, Zeng, G, Wheeler, CJ, Black, KL, Yu, JS. HER-2, gp100, and MAGE-1 are expressed in human glioblastoma and recognized by cytotoxic T cells. Cancer Res. 2004;64:4980-4986.CrossRefGoogle ScholarPubMed
30. Zhang, JG, Eguchi, J, Kruse, CA, et al. Antigenic profiling of glioma cells to generate allogeneic vaccines or dendritic cell-based therapeutics. Clin Cancer Res. 2007;13:566-575.10.1158/1078-0432.CCR-06-1576CrossRefGoogle ScholarPubMed
31. Saikali, S, Avril, T, Collet, B, et al. Expression of nine tumour antigens in a series of human glioblastoma multiforme: interest of EGFRvIII, IL-13Ralpha2, gp100 and TRP-2 for immunotherapy. J Neurooncol. 2007;81:139-148.10.1007/s11060-006-9220-3CrossRefGoogle Scholar
32. Prins, RM, Odesa, SK, Liau, LM. Immunotherapeutic targeting of shared melanoma-associated antigens in a murine glioma model. Cancer Res. 2003;63:8487-8491.Google Scholar
33. Silbergeld, DL, Rostomily, RC, Alvord, EC Jr. The cause of death in patients with glioblastoma is multifactorial: clinical factors and autopsyfindings in 117 cases of supratentorial glioblastoma in adults. J Neurooncol. 1991;10:179-185.10.1007/BF00146880CrossRefGoogle Scholar
34. Aghi, MK, Batchelor, TT, Louis, DN, Barker, FG, Curry, WT Jr. Decreased rate of infection in glioblastoma patients with allelic loss of chromosome 10q. J Neurooncol. 2009;93:115-120.10.1007/s11060-009-9826-3CrossRefGoogle ScholarPubMed
35. Parsa, AT, Waldron, JS, Panner, A, et al. Loss of tumor suppressor PTEN function increases B7-H1 expression and immunoresistance in glioma. Nat Med. 2007;13:84-88.10.1038/nm1517CrossRefGoogle ScholarPubMed
36. Hao, C, Parney, IF, Roa, WH, Turner, J, Petruk, KC, Ramsay, DA. Cytokine and cytokine receptor mRNA expression in human glioblastomas: evidence of Th1, Th2 and Th3 cytokine dysregulation. Acta Neuropathol. 2002;103:171-178.Google ScholarPubMed
37. Fecci, PE, Mitchell, DA, Whitesides, JF, et al. Increased regulatory T-cell fraction amidst a diminished CD4 compartment explains cellular immune defects in patients with malignant glioma. Cancer Res. 2006;66:3294-3302.10.1158/0008-5472.CAN-05-3773CrossRefGoogle ScholarPubMed
38. El Andaloussi, A, Lesniak, MS. An increase in CD4+CD25+FOXP3+ regulatory T cells in tumor-infiltrating lymphocytes of human glioblastoma multiforme. Neurooncol. 2006;8:234-243.Google ScholarPubMed
39. See, AP, Parker, JJ, Waziri, A. The role of regulatory T cells and microglia in Glioblastoma-associated immunosuppression. J Neurooncol. 2015;123:405-412.10.1007/s11060-015-1849-3CrossRefGoogle ScholarPubMed
40. Choe, G, Horvath, S, Cloughesy, TF, et al. Analysis of the phosphatidylinositol 3’-kinase signaling pathway in glioblastoma patients in vivo. Cancer Res. 2003;63:2742-2746.Google ScholarPubMed
41. Ermoian, RP, Furniss, CS, Lamborn, KR, et al. Dysregulation of PTEN and protein kinase B is associated with glioma histology and patient survival. Clin Cancer Res. 2002;8:1100-1106.Google ScholarPubMed
42. Louis, DN, Holland, EC, Cairncross, JG. Glioma classification: a molecular reappraisal. Am J Pathol. 2001;159:779-786.10.1016/S0002-9440(10)61750-6CrossRefGoogle ScholarPubMed
43. Yang, L, Ng, KY, Lillehei, KO. Cell-mediated immunotherapy: a new approach to the treatment of malignant glioma. Cancer Control. 2003;10:138-147.Google Scholar
44. Pistoia, V, Morandi, F, Wang, X, Ferrone, S. Soluble HLA-G: Are they clinically relevant? Semin Cancer Biol. 2007;17:469-479.CrossRefGoogle ScholarPubMed
45. Wiendl, H, Mitsdoerffer, M, Weller, M. Hide-and-seek in the brain: a role for HLA-G mediating immune privilege for glioma cells. Semin Cancer Biol. 2003;13:343-351.10.1016/S1044-579X(03)00025-7CrossRefGoogle Scholar
46. Das, A, Tan, WL, Teo, J, Smith, DR. Expression of survivin in primary glioblastomas. J Cancer Res Clin Oncol. 2002;128:302-306.CrossRefGoogle ScholarPubMed
47. Gomez, GG, Kruse, CA. Mechanisms of malignant glioma immune resistance and sources of immunosuppression. Gene Ther Mol Biol. 2006;10:133-146.Google ScholarPubMed
48. Zisakis, A, Piperi, C, Themistocleous, MS, et al. Comparative analysis of peripheral and localised cytokine secretion in glioblastoma patients. Cytokine. 2007;39:99-105.CrossRefGoogle ScholarPubMed
49. Yamanaka, R. Dendritic-cell- and peptide-based vaccination strategies for glioma, Neurosurg Rev. 2009;32:265-273.Google Scholar
50. Muragaki, Y, Maruyama, T, Iseki, H, et al. Phase I/IIa trial of autologous formalin-fixed tumor vaccine concomitant with fractionated radiotherapy for newly diagnosed glioblastoma: clinical article. J Neurosurg. 2011;115:248-255.CrossRefGoogle ScholarPubMed
51. Palu, G, Cavaggioni, A, Calvi, P, et al. Gene therapy of glioblastoma multiforme via combined expression of suicide and cytokine genes: a pilot study in humans. Gene Ther. 1999;6:330-337.CrossRefGoogle Scholar
52. Merchant, RE, Grant, AJ, Merchant, LH, Young, HF. Adoptive immunotherapy for recurrent glioblastoma multiforme using lymphokine activated killer cells and recombinant interleukin-2. Cancer. 1988;62:665-671.3.0.CO;2-O>CrossRefGoogle ScholarPubMed
53. Parsa, A, Crane, C, Wilson, S, et al. Autologous tumor derived gp96 evokes a tumor specific immune response in recurrent glioma patients that correlates with clinical response to therapy. In Proceedings of the AACE-NCI-EORTC International Conference Molecular Targets and Cancer Therapeutics 2007.Google Scholar
54. Ampie, L, Choy, W, Lamano, JB, Fakurnejad, S, Bloch, O, Parsa, AT. Heat shock protein vaccines against glioblastoma: from bench to bedside. J Neurooncol. 2015;123:441-448.10.1007/s11060-015-1837-7CrossRefGoogle ScholarPubMed
55. Topalian, SL, Hodi, FS, Brahmer, JR, et al. Safety, activity, and immune correlates of anti-PD-1 antibody in cancer. N Engl J Med. 2012;366:2443-2454.CrossRefGoogle ScholarPubMed
56. Wolchok, JD, Kluger, H, Callahan, MK, et al. Nivolumab plus ipilimumab in advanced melanoma. N Engl J Med. 2013;369:122-133.10.1056/NEJMoa1302369CrossRefGoogle ScholarPubMed
57. Ansell, SM, Lesokhin, AM, Borrello, I, et al. PD-1 blockade with nivolumab in relapsed or refractory Hodgkin’s lymphoma. N Engl J Med. 2015;372:311-319.CrossRefGoogle ScholarPubMed
58. Motzer, RJ, Rini, BI, McDermott, DF, et al. Nivolumab for metastatic renal cell carcinoma: results of a randomized phase II trial. J Clin Oncol. 2015;33:1430-1437.CrossRefGoogle ScholarPubMed
59. Rizvi, NA, Mazières, J, Planchard, D, et al. Activity and safety of nivolumab, an anti-PD-1 immune checkpoint inhibitor, for patients with advanced, refractory squamous non-small-cell lung cancer (CheckMate 063): a phase 2, single-arm trial. Lancet Oncol. 2015;16:257-265.10.1016/S1470-2045(15)70054-9CrossRefGoogle ScholarPubMed
60. Opdivo (nivolumab) [prescribing information] http://www.opdivo.bmscustomer connect.com/gateway (Bristol-Myers Squibb Company, 2014).Google Scholar
61. Keytruda (pembrolizumab) [prescribing information] http://www.merck.com/product/usa/pi_circulars/k/keytruda/keytruda_pi.pdf (Merck & Co., Inc. 2015).Google Scholar
62. Krauze, AV, Myrehaug, SD, Chang, MG, et al. A Phase 2 Study of Concurrent Radiation Therapy, Temozolomide, and the Histone Deacetylase Inhibitor Valproic Acid for Patients With Glioblastoma. Int J Radiat Oncol Biol Phys. 2015;92:986-992.10.1016/j.ijrobp.2015.04.038CrossRefGoogle ScholarPubMed
63. Wainwright, DA, Chang, AL, Dey, M, et al. Durable therapeutic efficacy utilizing combinatorial blockade against IDO, CTLA-4, and PD-L1 in mice with brain tumors. Clin Cancer Res. 2014;20:5290-5301.10.1158/1078-0432.CCR-14-0514CrossRefGoogle ScholarPubMed
64. Castro, MG, Baker, GJ, Lowenstein, PR. Blocking immunosuppressive checkpoints for glioma therapy: the more the Merrier! Clin Cancer Res. 2014;20:5147-5149.CrossRefGoogle Scholar
65. Cohen, CJ, Hoffmann, N, Farago, M, Hoogenboom, HR, Eisenbach, L, Reiter, Y. Direct detection and quantitation of a distinct T-cell epitope derived from tumor-specific epithelial cell-associated mucin using human recombinant antibodies endowed with the antigen-specific, major histocompatibility complex-restricted specificity of T cells. Cancer Res. 2002;62:5835-5844.Google ScholarPubMed
66. Denkberg, G, Cohen, CJ, Lev, A, Chames, P, Hoogenboom, HR, Reiter, Y. Direct visualization of distinct T cell epitopes derived from a melanoma tumor-associated antigen by using human recombinant antibodies with MHC- restricted T cell receptor-like specificity. Proc Natl Acad Sci. 2002;99:9421-9426.CrossRefGoogle ScholarPubMed
67. Cohen, CJ, Sarig, O, Yamano, Y, Tomaru, U, Jacobson, S, Reiter, Y. Direct phenotypic analysis of human MHC class I antigen presentation: visualization, quantitation, and in situ detection of human viral epitopes using peptide-specific, MHC-restricted human recombinant antibodies. J Immunol. 2003;170:4349-4361.CrossRefGoogle ScholarPubMed
68. Denkberg, G, Klechevsky, E, Reiter, Y. Modification of a tumor-derived peptide at an HLA-A2 anchor residue can alter the conformation of the MHC-peptide complex: probing with TCR-like recombinant antibodies. J Immunol. 2002;169:4399-4407.10.4049/jimmunol.169.8.4399CrossRefGoogle ScholarPubMed
69. Cohen-Inbar, O. Recruitment of Immune Effector Cells against Astrocytoma by MHC-chlrotoxin Chimeric Protein. PhD dissertation, Technion Israel Institute of Technology, 2014.Google Scholar
70. Yu, JS, Wheeler, CJ, Zeltzer, PM, et al. Vaccination of malignant glioma patients with peptide-pulsed dendritic cells elicits systemic cytotoxicity and intracranial T-cell in-filtration. Cancer Res. 2001;61:842-847.Google Scholar
71. Liau, LM, Prins, RM, Kiertscher, SM, et al. Dendritic cell vaccination in glioblastoma patients induces systemic andintracranial T-cell responses modulated by the local central nervous system tumor microenvironment. Clin Cancer Res. 2005;11:5515-5525.10.1158/1078-0432.CCR-05-0464CrossRefGoogle Scholar
72. Sampson, JH, Heimberger, AB, Archer, GE, et al. Immunologic escape after prolonged progression-free survival with epidermal growth factor receptor variant III peptide vaccination in patients with newly diagnosed glioblastoma. J Clin Oncol. 2010;28:4722-4729.CrossRefGoogle ScholarPubMed
73. Schneider, T, Gerhards, R, Kirches, E, Firsching, R. Preliminary results of active specific immunization with modified tumor cell vaccine in glioblastoma multiforme. J Neuro-Oncol. 2001;53:39-46.CrossRefGoogle ScholarPubMed
74. Steiner, HH, Bonsanto, MM, Beckhove, P, et al. Antitumor vaccination of patients with glioblastoma multiforme: a pilot study to assess feasibility, safety, and clinical benefit. J Clin Oncol. 2004;22:4272-4281.10.1200/JCO.2004.09.038CrossRefGoogle ScholarPubMed
75. Ishikawa, E, Tsuboi, K, Yamamoto, T, et al. Clinical trial of autologous formalin-fixed tumor vaccine for Glioblastoma multiforme patients. Cancer Sci. 2007;98:1226-1233.10.1111/j.1349-7006.2007.00518.xCrossRefGoogle ScholarPubMed
76. Bogdahn, U, Hau, P, Stockhammer, G, et al. Targeted therapy for high-grade glioma with the TGF-beta2 inhibitor trabedersen: results of a randomized and controlled phase IIb study. Neuro-Oncol. 2011;13:132-142.10.1093/neuonc/noq142CrossRefGoogle ScholarPubMed
77. Merchant, RE, McVicar, DW, Merchant, LH, Young, HF. Treatment of recurrent malignant glioma by repeated intracerebral injections of human recombinant interleukin-2 alone or in combination with systemic interferon- α . Results of a phase I clinical trial. J Neuro-Oncol. 1992;12:75-83.CrossRefGoogle ScholarPubMed
78. Barba, D, Saris, SC, Holder, C, Rosenberg, SA, Oldfield, EH. Intratumoral, LAK cell and interleukin-2 therapy of human gliomas. J Neurosurgery. 1989;70:175-182.CrossRefGoogle ScholarPubMed
79. Colombo, F, Barzon, L, Franchin, E, et al. Combined HSV-TK/IL-2 gene therapy in patients with recurrent glioblastoma multiforme: biological and clinical results. Cancer Gene Ther. 2005;12:835-848.CrossRefGoogle ScholarPubMed
80. Farkkila, M, Jaaskelainen, J, Kallio, M, et al. Randomised, controlled study of intratumoral recombinant γ –interferon treatment in newly diagnosed glioblastoma. Br J Cancer. 1994;70:138-141.10.1038/bjc.1994.263CrossRefGoogle ScholarPubMed
81. Wolff, JE, Wagner, S, Reinert, C, et al. Maintenance treatment with interferon-gamma and low-dose cyclophosphamide for pediatric high-grade glioma. J Neuro-Oncol. 2006;79:315-321.CrossRefGoogle ScholarPubMed
82. Allen, J, Packer, R, Bleyer, A, Zeltzer, P, Prados, M, Nirenberg, A. Recombinant interferon beta: a phase I-II trial in children with recurrent brain tumors. J Clin Oncol. 1991;9:783-788.CrossRefGoogle ScholarPubMed
83. Fetell, MR, Housepian, EM, Oster, MW, et al. Intratumor administration of beta-interferon in recurrent malignant gliomas. A Phase I clinical and laboratory study. Cancer. 1990;65:78-83.3.0.CO;2-5>CrossRefGoogle ScholarPubMed
84. Mahaley, MS, Dropcho, EJ, Bertsch, L, Tirey, T, Gillespie, GY. Systemic beta-interferon therapy for recurrent gliomas: a brief report. J Neurosurgery. 1989;71:639-641.10.3171/jns.1989.71.5.0639CrossRefGoogle ScholarPubMed
85. Buckner, JC, Brown, LD, Kugler, JW, et al. Phase II evaluation of recombinant interferon alpha and BCNU in recurrent glioma. J Neurosurgery. 1995;82:430-435.CrossRefGoogle ScholarPubMed
86. Buckner, JC, Schomberg, PJ, McGinnis, WL, et al. A Phase III study of radiation therapy plus carmustine with or without recombinant interferon- α in the treatment of patients with newly diagnosed high-grade glioma. Cancer. 2001;92:420-433.10.1002/1097-0142(20010715)92:2<420::AID-CNCR1338>3.0.CO;2-33.0.CO;2-3>CrossRefGoogle ScholarPubMed
87. Olson, JJ, McKenzie, E, Skurski-Martin, M, Zhang, Z, Brat, D, Phuphanich, S. Phase I analysis of BCNU-impregnated biodegradable polymer wafers followed by systemic interferon alfa-2b in adults with recurrent glioblastoma multiforme. J Neuro-Oncol. 2008;90:293-299.10.1007/s11060-008-9660-zCrossRefGoogle ScholarPubMed
88. Okada, H, Lieberman, FS, Walter, KA, et al. Autologous glioma cell vaccine admixed with interleukin-4 gene transfected fibroblasts in the treatment of patients with malignant gliomas. J Transl Med. 2007;5:67.CrossRefGoogle ScholarPubMed
89. Rand, RW, Kreitman, RJ, Patronas, N, Varricchio, F, Pastan, I, Puri, RK. Intratumoral administration of recombinant circularly permuted interleukin-4-Pseudomonas exotoxin in patients with high-grade glioma. Clin Cancer Res. 2000;6:2157-2165.Google ScholarPubMed
90. Kikuchi, T, Akasaki, Y, Abe, T, et al. Vaccination of glioma patients with fusions of dendritic and glioma cells and recombinant human interleukin 12. J Immunother. 2004;27:452-459.CrossRefGoogle ScholarPubMed
91. Jacobs, SK, Wilson, DJ, Kornblith, PL, Grimm, EA. Interleukin-2 or autologous lymphokine-activated killer cell treatment of malignant glioma: phase I trial. Cancer Res. 1986;46:2101-2104.Google ScholarPubMed
92. Lillehei, KO, Mitchell, DH, Johnson, SD, McCleary, EL, Kruse, CA. Long-term follow-up of patients with recurrent malignant gliomas treated with adjuvant adoptive immunotherapy. Neurosurgery. 1991;28:16-23.CrossRefGoogle ScholarPubMed
93. Hayes, RL, Koslow, M, Hiesiger, EM, et al. Improved long term survival after intracavitary interleukin-2 and lymphokine-activated killer cells for adults with recurrent malignant glioma. Cancer. 1995;76:840-852.3.0.CO;2-R>CrossRefGoogle ScholarPubMed
94. Dillman, RO, Duma, CM, Schiltz, PM, et al. Intracavitary placement of autologous lymphokine-activated killer (LAK) cells after resection of recurrent glioblastoma. J Immunother. 2004;27:398-404.CrossRefGoogle ScholarPubMed
95. Tsurushima, H, Liu, SQ, Tuboi, K, et al. Reduction of end-stage malignant glioma by injection with autologous cytotoxic T lymphocytes. Jap J Cancer Res. 1999;90:536-545.10.1111/j.1349-7006.1999.tb00781.xCrossRefGoogle ScholarPubMed
96. Plautz, GE, Miller, DW, Barnett, GH, et al. T cell adoptive immunotherapy of newly diagnosed gliomas. Clin Cancer Res. 2000;6:2209-2218.Google ScholarPubMed
97. Jackson, C, Ruzevick, J, Phallen, J, Belcaid, Z, Lim, M. Challenges in Immunotherapy Presented by the Glioblastoma Multiforme Microenvironment. Clin Dev Immunol. 2011;2011:732413, doi: 10.1155/2011/732413. Epub 2011 Dec 10.CrossRefGoogle ScholarPubMed
98. Caruso, JP, Cohen-Inbar, O, Bilsky, MH, Gerszten, PC, Sheehan, JP. Stereotactic radiosurgery and immunotherapy for metastatic spinal melanoma. Neurosurg Focus. 2015;38:E6.10.3171/2014.11.FOCUS14716CrossRefGoogle ScholarPubMed
Figure 0

Table 1 Characteristics of Established GBM Tumor Variants

Figure 1

Figure 1 Normal T cell proliferation and mechanisms of glioma cell immunoresistance. Top Right: Normal T cell proliferation. T cell proliferation, differentiation, and cytokine release require two separate signal mechanisms. The first signal involves T cell receptor-mediated recognition of tumor antigen presented by MHC I, which is located on the tumor cell. A second costimulatory signal involves B7 ligand, located on the tumor cell, binding to CD28, a receptor on the T cell. Both of these signals stimulate a variety of intracellular signaling pathways, which lead to upregulated activity of regulator proteins such as nuclear factor-κB, BCl-2, and PI3K. These signals promote T cell activation. However, other ligand-receptor binding pairs can inhibit these cascades and restrict T cell activation. These inhibitory checkpoints include B7 binding to CTLA-4 and B7-H1 (PD-L1) binding to PD-1. Anti-CTLA-4 antibodies (ipilimumab) and anti-PD-1 antibodies facilitate T cell activation by obstructing inhibitory checkpoint processes. Bottom Left: Mechanisms of immunosuppression. glioma cells secrete factors leading to an immunosuppressive tumor microenvironment. Transforming growth factor B (TGFB) and prostaglandin E-2 downregulate the expression of MHC, restricting antigen presentation and T cell proliferation. Interleukin-6. interleukin-10 and vascular endothelial growth factor are potent STAT-3 activators, leading to the proliferation of immature dendritic cells (DCs) that are not able to function as APCs. These immature DCs also secrete TGFB which aid in the proliferation of immunosuppressive T-reg cells and STAT-3 positive TH17 cells. Glioma cells downregulate MHC on their surface leading to the decreased antigen presentation and decreased T cell proliferation. Downregulation of B7 works via a similar mechanism in that the Costimulatory signal is lost preventing T cell proliferation. Increased expression of B7-H1 and FasL act as proapoptotic signals for T cells.69,97,98

Figure 2

Table 2 Selected Immunotherapeutic Trials for Malignant Gliomas