Next Article in Journal
Donor Lymphocyte Infusion in the Treatment of Post-Transplant Relapse of Acute Myeloid Leukemias and Myelodysplastic Syndromes Significantly Improves Overall Survival: A French–Italian Experience of 134 Patients
Previous Article in Journal
The Application of Long-Read Sequencing to Cancer
Previous Article in Special Issue
Decrease in Heparan Sulphate Binding in Tropism-Retargeted Oncolytic Herpes Simplex Virus (ReHV) Delays Blood Clearance and Improves Systemic Anticancer Efficacy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Immunotherapeutic Strategies for the Treatment of Glioblastoma: Current Challenges and Future Perspectives

by
Ilaria Salvato
1,2,3 and
Antonio Marchini
2,4,*,†
1
NORLUX Neuro-Oncology Laboratory, Department of Cancer Research, Luxembourg Institute of Health (LIH), L-1210 Luxembourg, Luxembourg
2
Laboratory of Oncolytic Virus Immuno-Therapeutics (LOVIT), Department of Cancer Research, Luxembourg Institute of Health (LIH), L-1210 Luxembourg, Luxembourg
3
Department of Life Sciences and Medicine, Faculty of Science, Technology and Medicine (FSTM), University of Luxembourg, L-4367 Belvaux, Luxembourg
4
Laboratory of Oncolytic Virus Immuno-Therapeutics, German Cancer Research Center, 69120 Heidelberg, Germany
*
Author to whom correspondence should be addressed.
Current address: European Commission, Joint Research Centre (JRC), 2440 Geel, Belgium.
Cancers 2024, 16(7), 1276; https://doi.org/10.3390/cancers16071276
Submission received: 28 February 2024 / Revised: 14 March 2024 / Accepted: 21 March 2024 / Published: 25 March 2024
(This article belongs to the Collection Advances and Future Prospects in Oncolytic Virus Immunotherapy)

Abstract

:

Simple Summary

Glioblastoma (GBM) poses a formidable challenge as a central nervous system tumor with extremely limited responsiveness to conventional treatments. While immunotherapeutic approaches have shown success in treating other solid tumors, their effectiveness against GBM is limited. Our review systematically addresses the intrinsic features of GBM that hinder the success of both standard therapies and immunotherapies. Furthermore, we comprehensively analyze all the immune-based approaches currently undergoing clinical evaluation for GBM, both as standalone treatments and in combination with standard therapy or other immunotherapies.

Abstract

Despite decades of research and the best up-to-date treatments, grade 4 Glioblastoma (GBM) remains uniformly fatal with a patient median overall survival of less than 2 years. Recent advances in immunotherapy have reignited interest in utilizing immunological approaches to fight cancer. However, current immunotherapies have so far not met the anticipated expectations, achieving modest results in their journey from bench to bedside for the treatment of GBM. Understanding the intrinsic features of GBM is of crucial importance for the development of effective antitumoral strategies to improve patient life expectancy and conditions. In this review, we provide a comprehensive overview of the distinctive characteristics of GBM that significantly influence current conventional therapies and immune-based approaches. Moreover, we present an overview of the immunotherapeutic strategies currently undergoing clinical evaluation for GBM treatment, with a specific emphasis on those advancing to phase 3 clinical studies. These encompass immune checkpoint inhibitors, adoptive T cell therapies, vaccination strategies (i.e., RNA-, DNA-, and peptide-based vaccines), and virus-based approaches. Finally, we explore novel innovative strategies and future prospects in the field of immunotherapy for GBM.

1. Introduction

Glioblastoma (GBM) is the most aggressive primary brain tumor, accounting for nearly 50% of all the primary central nervous system malignancies [1,2]. GBMs develop spontaneously within the brain (de novo) and typically infiltrate nearby brain tissues without spreading to distant organs [3]. Its incidence is 3.23 per 100,000 persons in the United States, with a slightly higher occurrence in males compared to females [4]. It is a fast-growing tumor occurring in patients with an average age at diagnosis of 65 years and a median overall survival (OS) of only 15 to 16 months after tumor diagnosis [4]. Long-term survival is uncommon, with fewer than 5% of patients on average surviving for five years or more after being diagnosed (source: Central Brain Tumor Registry of the United State from 2014 to 2018) [4].
Based on the new guidelines released in 2021 by the World Health Organization (WHO), GBM is classified as a grade 4 adult-type diffuse glioma based on its molecular and histopathological features. From a molecular point of view, GBM can be distinguished from other types of diffuse gliomas, such as astrocytomas and oligodendrogliomas, by its isocitrate dehydrogenase (IDH) wild-type status, intact chromosome arms 1p and 19q, retained expression of nuclear Alpha thalassemia/mental retardation X-linked syndrome (ATRX), and the absence of mutations in histone H3 genes. Furthermore, GBM is commonly characterized by histological features such as microvascular proliferation and necrosis, along with key molecular alterations, including the telomerase reverse transcriptase (TERT) promoter mutation, epidermal growth factor receptor (EGFR) amplification, and the +7/−10 cytogenetic signature [1].
In this review, we present a detailed overview of the current treatment options for patients with GBM, alongside an exploration of the underlying factors contributing to the failure of many anti-GBM therapies (both conventional and immune-based approaches). Furthermore, we provide an in-depth examination of the most promising immunotherapies targeting GBM, with a special emphasis on those that have already advanced to phase 3 clinical trials.

2. Standard of Care for GBM Patients

The established gold standard of care (SOC) for patients with newly diagnosed GBM is known as the “Stupp protocol” and comprises surgical resection, radiotherapy, and concomitant and adjuvant chemotherapy with the alkylating agent temozolomide (TMZ) [5]. If feasible, GBM interventions begin with maximal surgical resection, which eliminates most of the tumor. Surgical resection or biopsies also provide indispensable tumor material for a correct histological diagnosis and molecular testing. The extent of the tumor removed during surgery is a prognostic indicator, and according to the 2021 EANO guidelines, it should be evaluated using MRI within the first 24–48 h after the procedure [6]. Surgical resection is followed by six weeks of radiotherapy (60 Gray [Gy] in 2-Gy fractions) and concomitant daily TMZ (75 mg/m2), followed by six cycles of adjuvant TMZ (150–200 mg/m2) [5,6]. TMZ induces base methylations (i.e., N7-methylguanine, N3-methyladenine and O6-methylguanine) that, in the absence of an effective DNA damage repair system, ultimately lead to tumor cell death [7]. Of note, TMZ treatment is mostly beneficial in patients with a hypermethylated, and therefore epigenetically silent, O6-methylguanine-DNA methyltransferase (MGMT) gene. The enzyme MGMT is involved in DNA repair by removing the O6-methyl group from DNA and, if absent, enables effective chemotherapy and confers a survival advantage [5,8,9].
The Stupp protocol has remained unchanged over the last 18 years and typically provides an overall survival of less than two years to the patients. Thus, many clinical trials have been launched with the goal of finding new treatments to expand the life of individuals with GBM. Among various treatments, the use of tumor-treating fields (TTFs), namely low-intensity alternating electric fields delivered to the scalp of GBM individuals to induce tumor cell mitosis, has emerged as a novel modality able to ameliorate patient survival [10,11,12]. Despite the efficacy shown in a phase 3 clinical trial [11] and FDA approval, TTFs have not been yet incorporated into GBM SOC due to concerns about the unblinded nature of TTF clinical trials, as well as questions related to high cost, skin toxicity, and patient compliance [13,14].
Despite these first-line treatments, GBM virtually always recurs (median OS at recurrence = 2–9 months; median PFS at recurrence = 1.5–6 months) [15,16,17]. Once the tumor relapses, treatment options are very limited and, depending on the patient’s conditions, include second surgery, chemo-radiotherapy, and experimental treatments. As recently reviewed by Vaz-Salgato et al. (2023) [18], various second-line chemotherapeutics have been tested for the treatment of GBM, including anti-vascular endothelial growth factor (VEGF) [19,20,21], anti-transforming growth factor β (TGFβ)-receptor-I [22], anti-receptor tyrosine kinase [23], anti-protein kinase C [24], anti-EGFR [25], and anti-tyrosine kinase [26]. Although showing great promise at the preclinical level, these drugs failed to significantly improve the overall survival of GBM patients when tested in randomized clinical trials.

3. Therapeutic Challenges for GBM Therapies

The development of effective treatments targeting GBM could plausibly be hampered by GBM’s unique traits, including its challenging anatomical location protected by the blood–brain barrier (BBB), its invasiveness, the complexity of tumor variations within and between patients, and the immunosuppressive nature of the tumor microenvironment (TME) (Figure 1).

3.1. Anatomical Location

The brain is an essential organ of the human body’s governing motility, senses, emotions, cognition, memory, and survival instincts—in essence, many of the fundamental processes that regulate our body and mind. Surgical resection is therefore applicable only when GBM lies within non-critical regions of the brain that do not affect movement, speech, vision, or memory. As stated in the 2021 EANO guidelines [6], surgeons need to prioritize patients’ quality of life over extent of resection to prevent permanent neurological deficits. As recently reviewed in Bonosi et al. (2023) [27,28], there are multiple pre-operative (i.e., functional MRI imaging, magnetoencephalography, navigated transcranial magnetic stimulation, and diffusion tensor imaging) and intra-operative (i.e., ultrasonography, electrostimulation, cerebral perfusion measurements, and 5-aminolevulinic [5-ALA] tumor labeling) techniques that facilitate surgery and minimize the damages to the healthy brain tissue. As an example, patients operated with 5-ALA fluorescence-guided surgery showed a 6-month increase in progression-free survival (PFS) compared to patients operated via classical method [29,30].

3.2. Presence of the Blood–Brain Barrier

The brain is a highly vascularized organ and, to ensure proper neuronal functioning, needs to tightly control the trafficking of cells, molecules, and ions to and from the blood [31]. The blood–brain barrier (BBB) represents the most selective barrier of the human body, as it protects the brain from potentially harmful blood-borne agents and exogenous compounds (i.e., drugs, neurotoxins) that might damage the CNS [32,33]. It is constituted by endothelial cells of the capillaries located in the brain parenchyma, surrounded by pericytes and astrocytic endfeet, thereby isolating the brain from the bloodstream [32,34,35,36,37]. The BBB represents a major physical obstacle for the delivery of GBM therapeutics to the tumor, therefore limiting their clinical success. Indeed, a great amount of systemically administered chemotherapeutic agents failed to increase patient OS mainly due to their poor BBB penetration. An analysis of over 7000 chemotherapeutics found that only 1% of them could effectively cross the BBB and be active in the CNS [38,39]. In case of brain malignancies, including GBM, the BBB is partially disrupted leading to increased permeability, forming the so-called brain–tumor barrier (BTB). The disruption of the BBB in glioma exhibits heterogeneity, primarily manifesting within the tumor’s core while keeping its structure at the tumor rim, where invasive cells are located. The BTB stems from VEGF over-expression and increased angiogenesis in hypoxic zones, as well as the release of cytokines and chemical mediators, inducing the development of more immature and permeable vessels within the tumor [40,41,42,43,44]. Tumor-induced BBB leakage may enhance therapeutic delivery to the tumor core, yet the intact BBB beyond it can impede drug distribution. As outlined in a recent review by [45], brain drug delivery can be enhanced through surgical interventions such as intrathecal drug administration and convection-enhanced delivery (CED) and/or with the use of implantable pharmaceutical formulations, including biodegradable wafers or gels. Alternatively, researchers are focusing on improving drug penetration into the brain by enhancing drug liposolubility (e.g., using liposomes) or by modulating the BBB (e.g., through the modulation of efflux pumps, tight junctions, or the use of receptor agonists) [45]. Promising in terms of safety, these approaches require randomized clinical trials to thoroughly evaluate their effectiveness.

3.3. Tumor Heterogeneity and Plasticity

Another key GBM feature that can contribute to treatment failure is the high heterogeneity among (inter-tumoral) and within (intra-tumoral) tumors. Even when histologically similar, GBM tumors can differentially respond to treatments depending on their molecular profile. There are multiple signaling pathways that can be dysregulated in GBM, including p53, retinoblastoma (RB), and phosphoinositide 3-kinase (PI3K) signaling pathways [46,47]. The Cancer Genome Atlas network and subsequent studies tried to identify prognostically relevant GBM molecular subtypes based on large-scale genetic and epigenetic profiling. To date, three molecular subtypes have been proposed based on molecular analysis: proneural, mesenchymal, and classical [48,49]. They are meant to help clinicians diagnose and stratify GBM patients for potential personalized medicine [50]. However, to date, they have limited clinical relevance [51]. Moreover, researchers have recently focused on the identification of GBM subtypes by considering the characteristics and composition of the GBM tumor microenvironment. This classification system holds the potential to facilitate the implementation of precision immunotherapy approaches [52].
Inter-patient variability is further reinforced by intratumoral heterogeneity and plasticity. Within the tumor mass of an individual patient, there exists a complex, heterogeneous, and dynamic architecture of tumor cells that vary at the epigenetic, transcriptomic, protein, and metabolic levels [51,53]. Additionally, therapeutic approaches actively contribute to the phenotypic heterogeneity of GBM by modifying its tumor landscape [54]. This provides survival advantages to the tumor cells and may explain why drugs targeting the entire tumor may ultimately prove futile due to the rapid emergence of cell clones that are resistant to the specific treatment.

3.4. Infiltrative Nature

As with other malignant gliomas, GBM is characterized by a high invasive capacity that is associated with treatment resistance, recurrence, and poor OS. Brain tumor cells modify and degrade the extracellular matrix (ECM), enabling their invasive behavior through processes involving glutamate release and Ca2+ signaling pathways [55]. Within a GBM tumor, there are various levels of invasiveness reflecting the intratumoral heterogeneity of this cancer type. While tumor core cells have a higher tendency to proliferate, cells at the periphery of the tumor tend to be more invasive, allowing them to penetrate into the surrounding normal brain tissue [56]. Invasive GBM cells can move as individual cells [57] or in groups [58,59] and preferentially migrate along preexisting structures such us the brain parenchyma, white matter tracts, blood vessels, and subarachnoid spaces [60,61]. GBM cells can move along the brain tissue by remodeling the extracellular matrix and their own cytoskeleton, as well as their energy metabolism [61,62,63]. Differently from other cancer types, GBM cells rarely enter into circulation and thus do not normally metastasize to distant organs/tissues [64,65,66]. GBM cells’ invasive nature hinders complete surgical resection, and the remaining resistant clones lead to tumor recurrence [67]. As outlined in a recent review by [55], researchers have explored various approaches to inhibit invasion, including blocking Ca2+ channels (Mibefradil) [68], α V integrins (Cilengitide) [69], matrix metalloproteinases (MMP) [70,71], AMPA receptors (Talampanel) [72,73], and the PI3K/Akt pathway [74]. Overall, these interventions have had limited success in GBM patients.

3.5. Systemic and Local Immunosuppression

While historically considered “immune privileged”, the brain may be now better described as “immunologically distinct”, meaning with unique immune characteristics compared to other body parts. The brain possesses a specialized lymphatic drainage system called the “glymphatic system”, which plays a role in immunosurveillance, as it drains the cerebrospinal fluid (CSF), carrying immune cells and solutes, from the CNS into deep cervical lymph nodes [75,76]. While classical antigen presenting cells (APCs) are normally not detected in the healthy brain parenchyma, they can be found in adjacent vascular-rich tissues such as the choroid plexus and meninges [77]. They have access to the CSF and can detect brain parenchymal antigens. Moreover, in inflammatory conditions, APCs rapidly migrate towards the brain parenchyma through afferent lymphatics or endothelial venules to survey for antigens [77]. They then leave the brain and reach the deep cervical lymph nodes, where they can present brain-derived antigens and prime T and B lymphocytes, promoting adaptive immune responses [76,78]. T cells have also been observed in the brain parenchyma and CSF of healthy individuals, albeit in very low numbers, carrying out immune surveillance of the CNS and deep cervical lymph nodes [79].
As outlined in Zhang et al. (2022) [80], GBM patients often experience pronounced immunosuppression, affecting both their overall immune system (systemic) and the immune responses within the tumor environment (local). GBM patients have smaller secondary lymphoid organs and lower MHC-II expression levels in peripheral blood monocytes and are characterized by T cell lymphopenia compared to healthy individuals [81,82,83]. The decline in size and function of the thymus gland, known as thymic involution, results in decreased T cell production and, therefore, in reduced T cell availability for anti-GBM immunity [84]. T cells are majorly sequestered in the BM, due to the loss of surface sphingosine-1-phosphate receptor 1 (S1P1). S1P1 is responsible for the egress of T cells from the thymus and secondary lymphoid organs [85], but in GBM patients, the missing S1P1 receptor prevents T cells from leaving the bone marrow and entering the bloodstream [83]. Interestingly, in vitro studies revealed that serum isolated from GBM tumor-bearing mice impairs immune cell activation [86]. Circulating cytokines produced by the tumor as well as immunosuppressive treatment with corticosteroids and TMZ may further contribute to the systemic immunosuppression observed in GBM patients [81,87].
This systemic immunosuppression is further reinforced locally. In GBM, the BBB is disrupted and displays increased permeability, allowing for the influx of immune cells that are normally scarce in the brain parenchyma [88,89]. The GBM TME is highly heterogeneous and consists of various components, including GBM cancer cells, various signaling molecules, the extracellular matrix, vasculature, brain-resident non-immune cells (such as astrocytes and neurons), and lymphoid and myeloid immune cells. Despite the potential of immune responses to eliminate neoplastic cells or hinder their growth, GBM cancer cells have developed multiple mechanisms to evade immune surveillance and to shape the TME in their favor to allow for tumor development and progression. The communication between GBM cells and the TME occurs via cell-to-cell contact, soluble molecules [90,91,92], and via extracellular vesicles [93,94].
(i)
Soluble molecules: Secreted by various cellular players of the GBM microenvironment, the TME contains various growth factors and cytokines, such as (i) tumor-promoting cytokines, including interleukin (IL)-1, and basic fibroblast growth factor (bFGF) and (ii) immunosuppressive chemical mediators, including TGF-β, IL-10, IL-6 and prostaglandin E-2 (PGE2) [95,96]. While IL-1 and bFGF promote tumorigenesis, TGF-β, IL-10, IL-6, and PGE2 generally shift the immune response from an inflammatory response to a pro-tumoral and wound-healing one. This alteration leads to a reduced ability of immune cells to efficiently eliminate tumor cells. Moreover, the GBM TME is characterized by high levels of CC Chemokine Ligand 2 (CCL2), a very potent chemoattractant essential for the recruitment of regulatory T cells (Tregs) and myeloid cells [97].
(ii)
Extracellular matrix (ECM): In GBM, ECM composition is altered due to an overexpression and increased secretion of laminin, collagen, and fibronectin, and this physically results in elevated overall density and tumor stiffness [98]. This contributes to limiting the ability of chemotherapeutic drugs to diffuse and penetrate the tumor, reducing their effectiveness. Moreover, high levels of fibronectin and hyaluronic acid, along with surrounding ECM degradation via metalloproteinases, increases the mobility and invasiveness of glioma cells [99].
(iii)
Vasculature: The GBM TME is characterized by abnormal vasculature, and the central areas of the tumor experience poor blood flow, leading to a decrease in oxygen delivery [100]. This hypoxic microenvironment increases the expression of hypoxia-inducible factor 1-α, promoting angiogenesis and tumor cell invasion [100]. HIF1-α upregulates immunomodulatory surface ligands such as cytotoxic T-lymphocyte-associated protein 4 (CTLA-4) and programmed death-ligand 1 (PD-L1), inhibiting efficient anti-tumor immune responses [101].
(iv)
Healthy brain cells: In response to CNS injury, astrocytes normally secrete growth factors and cytokines to facilitate tissue repair in a process known as astrogliosis [102]. However, in GBM, this process is exploited to promote tumor growth. In particular, the TME promotes crosstalk between astrocytes and neighboring microglia, resulting in the activation of the JAK/STAT and PD-L1 pathways within astrocytes. This activation triggers an elevated production of anti-inflammatory cytokines like IL-10, TGF-β, and STAT3, thereby fostering an immunosuppressive milieu [103]. Moreover, neurons play a role in facilitating GBM tumor progression by upregulating neuroligin-3. This leads to the activation of the PI3K signaling pathway, promoting the proliferative activity of glioma cells [104].
(v)
Tumor-associated myeloid cells: Tumor-associated microglia and macrophages (TAMs) are the main infiltrating population in GBM, attracted towards the tumors in response to high concentrations of various chemoattractants secreted by glioma cells, including CCL2 [105,106,107]. Within the TME, they adopt immunosuppressive and tumor-supportive phenotypes [108]. Activation of the mTOR signaling pathway leads to increased STAT3 phosphorylation and suppression of the NF-κB pathway, resulting in the upregulation of anti-inflammatory cytokines such as IL-6, and IL-10 [109]. TAMs exhibit a decreased expression of surface MHC class II molecules and costimulatory molecules (CD40, CD80, and CD86), impairing antigen presentation and activation of T cells [110,111,112]. Myeloid-derived suppressor cells (MDSCs) suppress the immune system through multiple mechanisms. They express arginase, which reduces L-arginine levels necessary for TCR expression and function. They also secrete nitric oxide and reactive oxygen species, further inhibiting T cell activity. Additionally, MDSCs express PD-L1, promoting T cell exhaustion [113,114].
(vi)
Tumor-infiltrating lymphocytes (TILs): In GBM, TILs often exhibit dysfunction and exhaustion caused by factors released by glioma and microenvironmental cells, including TGF-β, IL-10, and CCL2, which recruit Tregs, MDSCs, and TAMs to the tumor site [115]. In response to TGF-β, CD4+ T cells upregulate FoxP3 and differentiate into Tregs. They account for 25% of TILs and are associated with a poor prognosis in GBM [116]. Through IL-10 and TGF-β signaling, Tregs promote the transition of other T cells into regulatory ones, exert an immunosuppressive function over natural killer (NK) and CD8+ T cells, help to generate MDSCs, and impair the antigen presentation capability of DCs [117]. TGF-β1 leads to a reduction in the expression of the activating receptor natural killer group 2 (NKG2D) on the surface of both CD8+ T cells and NK cells, thereby hindering their cytotoxic effects on GBM cells [118]. Moreover, Tregs highly express immune checkpoint molecules such as PD-1 and CTLA-4 that, via interaction with their respective receptors on the surface of T cells, suppress their effector functions [119]. Glioma cells further suppress lymphocyte activity through molecules such as FasL, PD-L1, PD-L2, CD70, and ganglioside [120,121,122]. The scarcity of TILs and accumulation of exhausted T cells in the tumor microenvironment contribute to immunotherapy resistance and relapse.

4. Immunotherapeutic Strategies for the Treatment of GBM

Immunotherapy has revolutionized the field of oncology by aiming to re-activate the cells of the immune system to react against the tumor, rather than directly targeting the cancer cells. Immune-based approaches have shown sustained clinical benefit and, in some instances, full remission of solid tumors, thus becoming part of their standard of care [123]. However, immune-based treatments have a different impact on each cancer type depending on tumor intrinsic features and level of immunosuppression. Regarding GBM tumors, current investigations into immunotherapeutic strategies encompass immune checkpoint inhibitors, adoptive T cell therapies, vaccination approaches, and virus-based therapies (Figure 2).
Figure 2 depicts the main immunotherapeutic strategies currently under evaluation in clinical trials for the treatment of GBM. These include (i) vaccination therapy, which aims to activate the patient’s adaptive immune system via the use of tumor-specific or tumor-associated antigens, delivered in the form of nucleic acids, peptides, or packaged into DCs; (ii) adoptive T cell therapy, involving the infusion of genetically modified (chimeric antigen receptor T cells [CAR-T cells]) or activated (tumor-infiltrating lymphocytes) autologous T cells to enhance their anti-GBM activity; (iii) immune checkpoint therapy, utilizing monoclonal antibodies to remove the “brakes” on the immune system’s response to GBM; and (iv) virus-based therapy, which explores the use of viruses either to selectively target and destroy GBM cells (oncolytic viruses) or to deliver therapeutic transgenes to the tumor (cancer gene therapy). Research on combining various immunotherapies holds great promise for the treatment of GBM. The image was created with BioRender (https://www.biorender.com/, accessed on 28 February 2024).

4.1. Immune Checkpoint Therapy

During prolonged antigenic exposure or tumor-T cell interaction, the effector T cells might gradually lose their tumor reactivity and become “exhausted”, a hypo-responsive state characterized by high levels of co-inhibitory molecules, also known as immune checkpoints (ICMs), decreased cytotoxicity, and reduced proliferation capacity [124]. ICMs are potent regulators of the immune system exploited by the TME to suppress immune responses towards malignant GBM cells. Over the last decades, several ICMs have been identified, including programmed cell death protein 1 (PD-1) and its ligand PD-L1, CTLA-4, Lymphocyte Activation Gene-3 (LAG-3), T cell immunoreceptor with immunoglobulin and ITIM domain (TIGIT), T cell immunoglobulin and mucin domain 3 (TIM-3), V-domain Ig suppressor of T cell activation (VISTA), and indoleamine 2,3-dioxygenase (IDO).
Being surface receptors, immune checkpoints can be inhibited by blocking monoclonal antibodies, known as immune checkpoint inhibitors (ICIs). The blockade of the PD-1/PD-L1 axis or CTLA-4 have shown remarkable success in the treatment of various solid tumors, including colorectal cancer, gastric cancer, hepatocellular carcinoma, melanoma, classic Hodgkin’s lymphoma, and non-small-cell lung carcinoma [125,126,127,128,129]. However, generally, minimal clinical benefit has been observed thus far for the treatment of GBM using these modalities, whether applied individually or in combination (Table 1).
Promising preclinical results [168,169] sparked three phase 3 clinical trials testing the efficacy of the anti-PD1 antibody Nivolumab for the treatment of GBM. The first study, checkmate 143 [161], evaluated the efficacy of Nivolumab and Ipilimumab in patients with recurrent GBM. Other studies, checkmate 548 [138,170] and 498 [137], instead tested Nivolumab in addition to radiation on MGMT methylated and unmethylated newly diagnosed GBM patients, respectively. All three studies failed to achieve the primary goal of ameliorating patient survival in comparison to standard treatments. Likewise, the anti-PD1 antibody Pembrolizumab, both as a monotherapy or in combination with bevacizumab, showed limited clinical benefit for recurrent GBM patients in phase 1 [139] and 2 clinical studies [149,152,171]. It is worth noting that neoadjuvant treatment with anti-PD-1 has shown promising outcomes in selected recurrent GBM patients during window-of-opportunity trials [131,172]. Another example of immune checkpoint therapy is Durvalumab, a human IgG1 monoclonal Ab against PD-L1. PD-L1 is expressed on the surface of nearly 90% of GBM cells [173]. Radiation-induced cell death may potentiate anti-PD1 and -PD-L1 therapies by releasing tumor antigens. A phase 2 multi-center study evaluating the combination of Durvalumab and standard radiotherapy in patients with unmethylated newly diagnosed GBM demonstrated favorable tolerability and potential efficacy, with one patient achieving a remarkable OS of 86 weeks [158].
As for the FDA-approved anti-CTLA4 antibody Ipilimumab, there are currently no published clinical data available of its use as a single therapy for GBM. As GBMs can rapidly adapt to ICI therapy by increasing the expression of alternative checkpoints following treatment [174], concluded and ongoing clinical studies rather focused on the combination of Ipilimumab with other agents, including anti-PD1 blocking antibodies (NCT02311920, NCT04606316, NCT03233152, NCT04817254, NCT04145115, NCT04396860), VEGF inhibitors [175], tumor-treating fields (NCT03430791), TMZ, and radiotherapy (NCT03367715). Unlike in melanoma [176,177], combining Ipilimumab and Nivolumab in GBM yielded no added benefit and actually increased immune toxicity compared to Nivolumab alone [178].
In addition to “classical” immune checkpoints, LAG-3, TIM-3, TIGIT, and IDO1 represent novel targets that are currently under investigation in GBM. NCT02658981 and NCT03493932 phase 1 studies investigated LAG-3 blockade (Relatlimab) either as a single agent or combined with anti-PD-1 therapy in patients with recurrent GBM or newly diagnosed GBM, respectively [166,179]. Results of the treatment efficacy are awaited. NCT03961971 is currently testing the inhibition of TIM-3 (Sabatolimab) and PD-1 (Spartalizumab) together with stereotactic radiosurgery in recurrent GBM. NCT04656535 phase 0/1 study is currently recruiting recurrent GBM patients for testing the combination of Domvanalimab (targeting TIGIT) and Zimberelimab (targeting PD-1). Instead, IDO is currently under investigation in combination with other therapies (i.e., radiotherapy, TMZ) in newly diagnosed GBM patients (NCT04047706, NCT02052648) [130,164].
As recently reviewed by Arrieta et al. (2023) [180], the failure of ICI treatment in GBM can be attributed to various factors, including the low mutational burden of GBM, elevated tumor heterogeneity, limited T cell infiltration, intratumoral downregulation of MHC-I/MHC-II molecules, and insufficient drug penetration across the blood–brain barrier [112,181]. Researchers are currently focusing on combining laser interstitial thermal therapy (LITT) with ICIs, which may benefit recurrent GBM patients, as LITT ablates tumor tissue and has been shown to enhance drug penetration through the BBB breakdown [142,182,183]. Understanding of the safety and efficacy of this approach will be gained from the active ongoing NCT02311582 phase 1/2 clinical trial and from the recruiting NCT03277638 phase 1/2 clinical trial.

4.2. Vaccination Therapy

Cancer vaccines represent a form of active immunotherapy that seeks to activate the patient’s adaptive immune system in response to specific antigens. These vaccines are designed to incorporate either tumor-specific antigens (TSA), also known as neoantigens, meaning mutated proteins found exclusively on tumor cells, or tumor-associated antigens (TAA), which are found to be highly expressed in the tumor but also, to a lesser extent, in normal tissues and are mostly derived from the overexpression of self-antigens [184]. Once administered, antigens are presented by APCs in the lymph nodes to naive or memory T cells. Primed T cells then migrate to the tumor site, initiating an immune response against the GBM. The objective is to trigger tumor regression and elicit durable memory responses, thereby reducing the risk of tumor recurrence. Currently, various vaccination strategies are under investigation for the treatment of GBM, employing peptides, DNA, or RNA as sources of antigens. These vaccines are packaged into various vehicles, including DCs and heat shock proteins, and are administered via intravenous, intranodal, intradermal, or intramuscular routes [184] (Table 2). To enhance vaccine effectiveness, adjuvants like tetanus toxoid, granulocyte-macrophage colony-stimulating factor (GM-CSF), and poly-ICLC (polyinosinic–polycytidylic acid stabilized with polylysine and carboxymethylcellulose) are combined with the vaccine formulation. They either promote antigen presentation, induce the expression of co-stimulatory molecules, or favor the release of cytokines [185].
A major challenge in vaccination strategies targeting GBM antigens is the highly heterogeneous expression of antigens within and among GBM tumors, which limits treatment response and is compounded by antigen instability and loss over time. To overcome this, the concept of a single vaccine targeting multiple antigens has been proposed to generate more robust and durable anti-tumor immune responses and reduce the risk of tumor immune evasion. However, the limited availability of neoantigens, attributed to the low mutational burden in GBM, presents a challenge in pursuing this approach [181].

4.2.1. DNA/RNA Vaccines

The pioneering and extensive research by the Nobel Prize-winning Dr. Drew Weissman and Dr. Katalin Karikó on messenger RNA (mRNA) has played a pivotal role in the remarkable and swift development of mRNA-based vaccines for COVID-19. Deployed in at least 164 countries, these vaccines have been a lifeline, saving millions of lives during the global pandemic crisis, bringing considerable focus to nucleic acid vaccines in the context of cancer treatment. A notable benefit of nucleic acids is their applicability across all human HLA genotypes, enabling presentation on both MHC-I and MHC-II molecules for the activation of both CD8+ and CD4+ T cell responses [229,230].
DNA vaccines can be easily engineered, allowing for cost-effective production and purification. They also demonstrate remarkable stability and are considered safe for use. Moreover, the plasmids employed in DNA vaccines serve as potent “danger signals”, activating various DNA-sensing innate immune receptors that facilitate the development of effective adaptive immune responses [229]. However, DNA vaccines have shown a limited response in clinical trials, partly due to low in vivo transfection efficiency. By contrast, RNA vaccines provide even more advantages in terms of safety, such as the absence of risk for insertional mutagenesis, inability to self-replicate, and rapid degradation through proteases [230]. The main drawback of RNA-based therapies lies in the RNA inherent instability and limited ability to effectively penetrate cells. To increase their preservation and facilitate their delivery, RNA molecules are loaded within cells, virus-like capsid, or nanoparticles [230]. Conclusive results on the effectiveness of DNA and RNA vaccines for GBM treatment are still pending, as these vaccines have not yet undergone phase 3 clinical trials. The ongoing NCT03491683 phase 1/2 trial is investigating the combination of two DNA vaccines with a PD-1 inhibitor in newly diagnosed GBM patients. The first vaccine, named INO-5401, encodes for Wilms Tumor-1 (WT1), prostate-specific membrane antigen (PSMA), and hTERT. The second vaccine, named INO-9012, encodes for IL-12. Both vaccines are administered intramuscularly with subsequent electroporation. The latter is used as a delivery system, applying high-intensity electricity to increase membrane permeability [231]. Interim analysis shows promising results in terms of safety, immunological effectiveness, and potential survival advantage [213,232]. A phase 1 study (NCT04015700) is in progress to evaluate the efficacy of INO-9012 together with a personalized DNA vaccine, and electroporation delivery. As for RNA vaccines, a phase 1/2 study (NCT04573140) is currently investigating the intravenous administration of autologous tumor messenger RNA (mRNA) in GBM patients using lipid particles.

4.2.2. Peptide Vaccines

Peptide-based vaccines use short synthetic peptides mimicking antigenic epitopes that can trigger potent and highly targeted responses. Peptide vaccines have been shown to predominantly induce humoral immunity but can also trigger cell-mediated immunity against the desired antigen [233]. So far, peptide vaccines have not demonstrated significant clinical benefit in the cure of GBM patients. This is partially due to the inherent instability and limited immunogenicity of peptides. As reviewed by Frederico et al. (2021), five main GBM-targeting peptide vaccines are currently under investigation: rindopepimut, SurVaxM, IMA950, heat shock protein–peptide complexes 96 (HSPPC-96)-specific vaccine, and personalized neoantigens vaccines [184]. Rindopepimut is a 13 aa peptide vaccine based on EGFRvIII. Despite promising results in phase 2 clinical trials [196,197,198], rindopepimut plus standard chemotherapy failed to improve the OS of newly diagnosed GBM patients in a randomized phase 3 clinical study (ACT-IV) [200]. However, trial data demonstrated increased humoral immune responses in the treatment arm compared to the control arm [200]. More than half of the trial patients, regardless of receiving rindopepimut, experienced a loss of EGFRvIII expression at relapse. This antigenic loss (~50% loss rate at relapse) reduces the number of eligible patients who can benefit from rindopepimut. Biopsy confirmation of EGFRvIII expression is therefore a crucial factor for clinical trial enrollment.
The SurVaxM vaccine specifically targets survivin, an anti-apoptotic protein that exhibits high expression in GBM while being undetectable in normal brain tissue [206,234]. Currently, a phase 2 study (NCT02455557) is actively investigating the efficacy of TMZ and the SurVaxM vaccine in treating newly diagnosed GBM patients [207,208]. Preliminary results have demonstrated the safety and tolerability of the vaccine, along with elevated levels of survivin antibodies and CD8+ T cells post-vaccination, leading to improved PFS and OS compared to historical controls [207,208].
The multi-peptide treatment IMA950, consisting of 11 TAAs commonly found in GBM tumors, has shown promising results in phase 1 and 2 clinical trials. Administered intradermally to newly diagnosed GBM patients treated with radiochemotherapy, IMA950 elicited CD8+ T cell responses to both single and multiple antigens [210,211]. Of note, adjuvant choice might be important for patient outcome, as the IMA950/poly-ICLC treatment [211] showed increased OS and PFS rates compared to IMA950/GM-CSF [210]. However, IMA950/poly-ICLC vaccination had no benefit in patients with recurrent GBM [212]. Phase 3 clinical trials are awaited to confirm vaccine efficacy.
Differently from IMA950, HSPPC-96 vaccine targets multiple tumor neoantigens. HSPPC-96 consists of gp96, a 96 kilodalton (kDa) heat shock protein (HSP), and its associated cellular neopeptides. As a chaperone of the ER, HSPPC-96 can be internalized into APCs for efficient antigen presentation [235,236]. Promising phase 1 and 2 results [201,202] sparked numerous clinical trials, some of which still ongoing (i.e., NCT03018288 and NCT01814813). Of note, checkpoint inhibitors may significantly impact vaccine efficacy, as an elevated PD-L1 expression translated into systemic immunosuppression and less response to vaccination [203], warranting further clinical studies on combination therapies of peptide vaccine with ICIs. Recently, researchers utilized whole-exome sequencing data to develop personalized peptide vaccines that consider the patient’s specific neoantigen expression. Phase 1 trials, including the European GAPVAC trial [223] and the American NEOVAX trial [237], have been conducted to assess the efficacy of this approach in newly diagnosed GBM patients. In both trials, the treatments stimulated robust circulating T cell responses against at least one immunizing peptide, involving CD8+ and CD4+ T cells with a memory phenotype. However, a randomized phase 3 trial evaluating personalized peptide vaccines in recurrent GBM patients did not meet the primary nor secondary endpoint for the enrolled participants [225].

4.2.3. Dendritic Cell Vaccines

Another potential immunotherapeutic approach is to exploit the intrinsic antigen presentation ability of DCs to activate adaptive immune responses. Autologous DCs are typically harvested, ex vivo sensitized with antigens and then re-infused into the patient [238,239]. Autologous DCs can be directly isolated from the peripheral blood or differentiated in vitro from monocytes or CD34+ hematopoietic stem cells via IL-4 or GM-CSF [240]. DCs may be “educated” via several forms of antigens ranging from DNA/RNA to peptides and tumor lysates. Peptides loaded on DCs are more efficiently delivered to the target tissue compared to peptide treatments alone. Although the initial clinical results appear promising, there is currently a scarcity of robust evidence regarding the efficacy of DC vaccines in GBM. The outcome of DC vaccines against GBM tumors is variable, reflecting inter-individual heterogeneity and ranging from minimal or no clinical response to significant response. Additionally, without the aid of adjuvants, DCs face challenges in migrating to the lymph nodes, with less than 5% of injected DCs successfully reaching their target destination [241].
CMV proteins are highly expressed in over 90% of GBM tumors but are rarely found in healthy brain tissue [242]. mRNA encoding for the CMV phosphoprotein 65 (pp65) can be loaded into DCs to stimulate CMV-specific T cell immunity able to kill GBM cells [243]. Two phase 1 studies [189,190] demonstrated that, despite the cold microenvironment of GBM, CMV-pp65 RNA-pulsed DCs (also known as CMV-DCs) triggered antigen-specific T cell responses, warranting further follow-up (NCT02771301, NCT02465268). The pre-conditioning of patients with tetanus/diphtheria toxoid actively increased the homing of pp65-specific DCs to the lymph nodes [189].
To date, only two DC vaccines reached randomized phase 3 clinical trials: ICT-107 and DCVax-L. In ICT-107, DCs are pulsed with multiple MHC-I-restricted TAAs highly expressed on GBM: AIM-2, MAGE-1, HER2/neu, TRP-2, gp100, and IL-13Rα2 [244,245]. A phase 2 study demonstrated the safety and immunogenicity of the treatment, as well as an improvement in patients’ PFS compared to the control group [209]. The phase 3 trial (NCT02546102) testing the intradermal administration of ICT-107 in newly diagnosed GBM patients was prematurely suspended because the company was unable to financially support its completion.
For DCVax-L, DCs are pulsed ex vivo with a tumor lysate. In a randomized phase 3 clinical trial, the effectiveness of DCVax-L and standard radiochemotherapy was evaluated in patients with newly diagnosed and recurrent GBM. The addition of DCVax-L to the standard therapy was found to be safe [220]. The multicentric study (NCT00045968) started in 2007 over a period of eight years and included two arms of GBM patients. In addition to standard radiochemotherapy, the first arm was treated with placebo, while the second arm received DCVax-L. The primary endpoint of the trial was PFS. However, in the initial report detailing the trial results, there was no mention of PFS data. Instead, the authors declared an increase in OS [220]. After four years, a second report retrospectively compared the OS of DCVax-L-treated patients with that of an external control population of patients obtained from selected published randomized clinical trials [221]. The data suggested that MGMT-methylated patients show increased survival compared to non-methylated individuals, pointing to a possible cooperative effect of TMZ and DCVax-L. Notably, the treatment led to an extension of median OS for both newly diagnosed GBM (19.3 months vs. 16.5 months) and recurrent GBM (13.2 months vs. 7.8 months) patients compared to external controls receiving standard of care alone [221]. However, concerns were raised regarding the interpretation of the results, emphasizing the necessity to approach the findings with caution. Various design issues, such as a shift in the primary endpoint from PFS to OS based on arguments related to pseudo-progression, an extended duration of the enrollment period, and an inappropriate selection of the control arm, contribute to these concerns [246,247,248,249,250,251].

4.3. Adoptive T Cell Therapy

Adoptive T cell therapy is an immunotherapy technique in which the patient’s T cells are expanded outside the body (ex vivo) and then reinfused back into the patient to target tumors. A few days before T cell reinfusion, patients undergo a lymphodepleting preparative regimen, which involves the use of lymphocyte-directed chemotherapy. This regimen aims to create a favorable environment that prolongs the persistence of infused cells and enhances the effectiveness of the treatment [252]. Currently, adoptive T cell therapy in the context of GBM primarily involves the use of patient-isolated infiltrating T cells (TIL therapy) or patient-isolated T cells genetically engineered ex vivo to regain cancer-fighting properties, such as chimeric antigen receptor T cells (CAR-T cells) (Table 3).

4.3.1. TIL Therapy

The preparation of autologous TILs is a time-consuming process with a low success rate. It involves culturing a resected tumor specimen in a high concentration of recombinant IL-2, along with IL-15 and IL-21 if necessary. The TILs are then selected, expanded, and transferred to the patient. A pilot study demonstrated that the delivery of autologous TILs and IL-2 had limited anti-tumor effects in the context of malignant gliomas [265]. As a potential explanation, patient-isolated TILs are heterogeneous in terms of TCR and level of exhaustion and would therefore react differently against the tumor cells [83,266]. The use of ICIs may therefore promote the anti-tumor efficacy of TIL therapy. Two phase 1 clinical trials (NCT05333588, NCT04943913) are currently recruiting GBM patients to investigate safety of TIL therapy, with results expected for 2024–2025.

4.3.2. CAR-T Cell Therapy

A promising T-cell-based approach involves the genetic engineering of autologous T cells to express a chimeric antigen receptor (CAR) designed to target tumor-specific antigens. CAR is a recombinant receptor that, in its latest generations, consists of four main components: (i) an extracellular antigen-recognition domain, (ii) a spacer region, (iii) a transmembrane domain that anchors CAR to the cell membrane, and (iv) intracellular signaling domains that provide co-stimulation and initiate the signaling cascade [267]. The major advantage of CAR-T cell therapy is that CAR recognizes a tumor antigen independently of MHC-restriction, therefore bypassing antigen presentation. Once bound to a specific antigen, the CAR signaling domains send the signals to the T cell to kill the target cell.
Driven by the success of CAR-T therapies in hematological cancers [268], researchers are currently focusing their efforts on the development of GBM-specific CAR-T therapies. So far, CAR-T cell clinical trials for GBM are still in the early stages, primarily in phase 1/2 trials. While some CAR-T cells have shown promise, they still need to demonstrate clinical benefits conclusively. The interpatient variability in surface antigen expression along with the problem of antigen escape represent major obstacles of this approach. Other barriers to the clinical efficacy of CAR-T cells are T cell engraftment and expansion in vivo and the inhibitory TME, which becomes even more immunosuppressive after CAR-T therapy [269]. Combining lymphodepleting preconditioning and ICIs may address these obstacles. Moreover, the high cost of CAR-T cell manufacturing can affect healthcare expenditures and limit access to this therapy. As reviewed by Luksik et al. (2023), EGFRvIII, IL-13Rα2, and HER2/neu are among the main target antigens of CAR-T cell therapy evaluated in clinics in the last decade [270]. B7-Homolog 3 (B7-H3), the ECM metalloproteinase inducer (EMMPRIN), disialoganglioside (GD2), matrix metalloproteinase 2 (MMP2), and NKGD2 ligands are instead novel targets currently under investigation in ongoing clinical trials [271].
EGFRvIII-directed CAR-T cells were tested in a phase 1 study for the treatment of EGFRvIII+ recurrent GBM, showing safety and feasibility without cross-reactivity to wild-type EGFR. However, the therapy resulted in EGFRvIII antigen escape and adaptive resistance [253]. A subsequent phase 1/2 trial did not yield clinical benefits in recurrent GBM patients [257].
IL-13Rα2 is a potential target found in many human cancers, including GBM (>75%) [272]. Different versions of IL-13Rα2-targeted CARs have been developed so far, with modifications in genetic elements and costimulatory molecules [260,261,273,274]. 13Rα2-targeted CAR-T cells showed promising results in a recurrent GBM patient, with tumor regression, increased cytokine levels, and no therapy-associated toxicity. The clinical response lasted for 7.5 months after treatment [260]. The newest version of IL-13Rα2-targeted CAR-T cells was genetically modified to induce a permanent disruption of the glucocorticoid receptor. In a phase 1 trial, the intracranial administration of the therapy in recurrent GBM patients was well tolerated, with indications of transient tumor reduction and/or tumor necrosis at the site of T cell infusion [274].
HER2/neu, being overexpressed in 80% of GBM, is another common antigenic target used in CAR-T therapies [275]. Despite its expression in both tumor and healthy brain tissue, no off-target toxicity has been observed in GBM patients systemically infused with HER2/neu-specific CAR-T cells [276]. Hedge and colleagues designed and created bivalent HER2/neu- and IL-13Rα2-targeting CAR-T cells that, in preclinical GBM mouse models, reduced antigen escape, enhanced T cell effector functions, and improved animal survival [277]. Trivalent CAR molecules specific for the glioma antigens HER2/neu, IL-13Rα2, and ephrin-A2 (EphA2) have the potential to capture nearly the totality of tumor cells. In preclinical models, these CAR-T cells inhibited tumor growth and extended animal survival compared to monospecific or bispecific CAR-T cells [278]. Clinical trials are still awaited to confirm treatment efficacy in humans.

4.4. Virus-Based Therapy

Virus-based treatments employed for the treatment of GBM can be either gene delivery systems or oncolytic viruses (OVs) (Figure 3, Table 4). Viral vectors are non-lytic and typically deliver pro-inflammatory and anti-angiogenic molecules, tumor suppressor genes, TAAs, ICIs, small interfering RNAs, cancer stroma-degrading enzymes, and cytotoxic convertases [279]. OVs are instead replication-competent viruses that selectively replicate in cancer cells inducing their lysiswhile sparing the heathy counterparts. They can either have inherent oncolytic properties by naturally infecting tumor cells or acquire selective tropism through genetic modifications [280]. Due to their replicative nature, OVs induce cell lysis, which in turn elicits secondary immune responses by releasing viral PAMPs, DAMPs, and TAAs. The infection of tumor cells with OVs has the effect of “warming up” the immunosuppressive TME, resulting in the inhibition of tumor progression and an enhanced suitability of the TME for other therapeutic interventions [280].

4.4.1. Adenovirus (AdV)

In the context of GBM, researchers have primarily focused on the AdV delivery of the herpes simplex virus (HSV) Thymidine kinase (TK) gene, the TP53 tumor suppressor gene, the IL-12-encoding gene, and a transgene encoding for a chimeric death receptor (VB-111).
When administered alongside ganciclovir or valacyclovir, HSV-TK converts them into cytotoxic products that accumulate and selectively eliminate the transduced cancer cells. The various clinical trials testing HSV-TK/ganciclovir gene therapy differed in the promoter used to control TK gene expression: (i) Rous sarcoma virus (RSV) promoter [288,318,319] and (ii) CMV promoter [320,321,322]. AdV-mediated gene therapy was safe and well tolerated [318,319,320]. A phase 2 trial testing the infusion of the suicide gene therapy into the arteries in patients with recurrent GBM revealed an improvement of PFS (29.6 vs. 8.4 weeks) and OS (45.4 vs. 14.3 weeks) compared to standard treatments alone [288]. In a phase 3 randomized, controlled study by Immonen et al. (newly diagnosed GBM and recurrent GBM patients), HSV-TK showed little to moderate improvement in survival rates and moderate tolerability [321,322]. The substitution of ganciclovir with valacyclovir was found to be safe [287] and resulted in improved median OS (17.1 vs. 13.5 months) for newly diagnosed GBM patients compared to standard treatments alone, as observed in a phase 2 study [289].
A second genetic approach used for GBM treatment consists of the upregulation of the TP53 tumor suppressor gene [323]. Restoration of a functionally active p53 protein was achieved via the use of a TP53-armed AdV (INGN 201; ADVEXIN) constructed through cDNA of the wild-type TP53 in place of the AdV E1 region [324]. The virus showed minimal cytotoxicity in vivo but, when intratumorally injected, failed to distribute widely in the tumor tissue, reaching only 5 mm from the injection site. Most notably, one GBM patient enrolled in the clinical study survived nearly 3.5 years after Ad-TP53 treatment without evidence of recurrence [325]. The p53-armed AdV was also investigated in another phase 1 clinical trial (NCT00004080), but the results are not yet available.
Researchers investigated the effects of the proinflammatory cytokine IL-12 on GBM tumors using an engineered AdV-based vector called Ad-RTS-IL-12 [326]. This vector allows for the inducible expression of IL-12, activated via oral administration of veledimex. Preclinical studies showed reduced tumor mass and increased lymphocyte infiltration [326]. In human application, Ad-RTS-IL-12 is injected into the resection cavity of recurrent GBM patients, accompanied by veledimex administration, showing limited toxicity and promising anti-tumor immune responses [290].
Lastly, VB-111 is an AdV-based cancer gene therapy that specifically targets angiogenic endothelial cells with a transgene encoding a chimeric death receptor, linking Fas to human TNF-R. When activated, this receptor induces Fas-mediated apoptosis and vascular disruption, leading to tumor starvation. In a phase 2 study, the combination of VB-111 and bevacizumab doubled the survival of patients with recurrent GBM compared to bevacizumab monotherapy [291]. However, a randomized controlled phase 3 study (GLOBE), testing VB-111 and bevacizumab failed to replicate the phase 2 results in recurrent GBM patients [292].
Alternatively, researchers have tested oncolytic AdVs, also known as conditionally replicative adenoviruses (CRAds), to target GBM tumors. These viruses acquire their tumor specificity via either (i) deletion of genes encoding for cell cycle regulatory proteins, (ii) natural overexpression of virus receptors on the surface of tumor cells, or (iii) use of tumor-specific promoters to control viral replication [327]. In the case of GBM, four main CRAds have reached clinical testing: ONYX-015, DNX-2401, DNX-2440, and CRAd-S-pk7.
ONYX-015 contains a deletion of the E1B gene. The virus preferentially replicates in cancer cells through various, not yet fully characterized mechanisms [328,329]. At the preclinical level, ONYX-015 achieved promising results in terms of tumor cell killing and reduction of tumor mass [330]. In a phase 1 study, ONYX-015 proved to be safe and well tolerated even at the highest dose (1010 viral particles) in all enrolled patients, among which recurrent GBM cases were included [285]. However, no tendency of anti-tumor efficacy could be observed in this study [285].
DNX-2401, previously known as delta-24-RGD (Δ24RGD) or Tasadenoturev, features a 24 bp deletion of the E1A gene that abrogates the binding and inhibition of E1A to the Rb protein and a fiber knob RGD modification to retarget virus entry via cell surface integrins that are typically enriched in glioma cells. These modifications were initially believed to enable selective targeting and replication of the virus to cancer cells with aberrant Rb pathways [331,332]. However, other research groups have been unable to replicate these initial findings [333]. Both as a single agent or in combination with other treatments (i.e., IFN-γ and anti-PD1), DNX-2401 did not raise any safety concerns [283,334,335,336]. Although the 12-month survival objective was achieved, the combination of DNX-2401 with TMZ and pembrolizumab did not meet the primary endpoint of objective response in a phase 2 clinical trial [337]. A new clinical trial (NCT03896568) is actively recruiting recurrent GBM patients to test DNX-2401 oncolytic virus delivered by allogenic bone marrow-derived human mesenchymal stem cells.
DNX-2401 has been recently modified to express the human OX40 co-stimulatory ligand (OX40L), aiming to enhance the antigen presentation in tumor cells. Compared to DNX-2401, this new version exhibited more potent and specific anti-glioma activity, attributed to superior T cell activation and proliferation [338]. Although a phase 1 clinical trial (NCT03714334) was underway to evaluate this modified virus for recurrent GBM treatment, it was terminated due to a stock shortage.
Lastly, Ulasov and colleagues generated a glioma-specific recombinant AdV, called CRAd-S-pk7, by modifying the Ad5 fiber with pk7s and by regulating the expression of the E1A gene via the human survivin promoter [339]. Building on encouraging preclinical results [340,341], CRAd-S-pk7 virus loaded onto neural stem cells was administered during surgery in newly diagnosed GBM patients, along with chemo-radiotherapy [281]. The treatment proved to be safe and well tolerated [281]. Although not the primary objective of the study, the presence of promising survival outcomes provides support for conducting further investigations of CRAd-S-pk7 in phase 2/3 clinical trials.

4.4.2. Retrovirus

In the context of GBM, researchers have primarily focused on the retrovirus delivery of the HSV-TK gene, or of the yeast cytosine deaminase gene (Toca 511). A phase 3 study that tested HSV-TK gene delivery along with intravenous ganciclovir administration demonstrated no significant differences in median OS between treatment and control patients [302]. Toca 511, also known as Vocimagene Amiretrorepvec, is a replication-deficient engineered murine leukemia virus armed with the yeast cytosine deaminase gene [342]. When administered in combination with the prodrug 5-fluorocytosine (Toca FC or 5-FC), the virus-delivered cytosine deaminase converts it into its toxic form 5-Fluorouracil (5-FU) that eventually kills the cancer cells and nearby immunosuppressive cells [343]. Of note, 5-FU can induce so-called “bystander effects”, as it can passively diffuse through cell membranes, therefore not only affecting directly infected cancer cells but also nearby cancer cells [344]. Despite encouraging observations in a phase 1 study [306], similarly to the case of ICIs, Toca 511/5-FC failed to meet the primary endpoint of improve patient survival compared to standard of care when tested in a randomized open label phase 2/3 study [308].

4.4.3. Herpes Simplex Virus (HSV)

The neurotropic HSV-1 belongs to the Herpesviridae family, and it is an enveloped icosahedral virus with a dsDNA genome. To date, three genetically engineered versions of it have been evaluated in completed clinical trials: HSV-1716 [345], G207 [298,346], and G47Δ [347]. First-generation HSV-1716 contains a deletion of γ134.5 genetic loci, which counteracts the normal antiviral response of cells and allows viral protein translation to proceed [345]. Three UK phase 1 clinical trials demonstrated the safety of intratumoral injection of it, either alone or following surgical resection, in glioma patients [293,294,295]. The second-generation G207, which includes an additional insertion of the UL39 gene preventing viral replication in non-dividing cells [298,346], also demonstrated safety [296,297,298]. The third-generation G47Δ (Teserpaturev, DELY-TACT) differs from the G207 backbone for a α47 gene deletion that enhances viral replication and triggers anti-tumor immune-mediated responses via upregulation of MHC-I molecules [347]. Of note, the G47Δ bears the same genetic mutations (γ134.5 and α47) of the first FDA- and EMA-approved oncolytic virus, namely T-VEC (Talimogene Laherparepvec; IMLYGIC®; formerly called OncoVEXGM-CSF) [348]. However, via additional deletion of UL39, G47Δ was more attenuated than T-VEC and therefore safer. At the University of Tokyo, a phase 1/2 study demonstrated the safety of G47Δ when intratumorally injected in recurrent GBM patients [299]. Accordingly, they started the subsequent phase 2 study to test the efficacy of multiple intratumoral G47Δ injections (1 × 109 viral particles; max of six injections) in patients with recurrent GBM [299]. Based on outstanding clinical results, G47Δ received a conditional time-limited approval by the Pharmaceuticals and Medical Devices Agency of Japan (PMDA) for the treatment of brain tumors.

4.4.4. Poliovirus

Polioviruses are positive single-stranded RNA (ssRNA) viruses belonging to the Picornaviridae family. PVSRIPO, or Lerapolturev, is a non-pathogenic poliovirus/rhinovirus chimeric virus with anti-neoplastic activity [349]. PVSRIPO specifically targets tumor cells by utilizing the poliovirus receptor CD155 [350]. In a phase 1 trial, intratumoral treatment with PVSRIPO in recurrent GBM patients demonstrated an improved overall survival compared to historical controls [314]. Ongoing clinical studies include a phase 2 trial (NCT02986178) investigating PVSRIPO as monotherapy, as well as phase 1/2 (NCT03973879) and phase 2 (NCT04479241) trials exploring the combination of PVSRIPO with either anti-PD-L1 atezolizumab or anti-PD1 pembrolizumab, respectively.

4.4.5. Respiratory Enteric Orphan Virus (Reovirus)

Reoviruses are naturally occurring double-stranded RNA viruses that belong to the Reoviridae family. They are non-pathogenic and selectively replicate within cancer cells by taking advantage of the Ras pathway that is commonly upregulated in neoplastic cells [351]. They underwent four phase 1 clinical trials for GBM treatment, with each study exploring a different administration route: intratumoral [315,316] or systemic [317] injection. In all trials, Reolysin proved to be safe. Of note, treatment causes an in vivo upregulation of IFN-regulated genes and PD-1/PDL-1 axis, as well as an increase in T cell infiltration [317]. This makes Reolysin particularly interesting for combination therapies.

4.4.6. Measles Virus (MeV)

MEVs belong to the Paramixoviridae family and contain a negative sense ssRNA genome. They were originally chosen to treat malignancies, as a case report linked their infection to tumor remission [352]. The virus used for GBM treatment is a live attenuated strain called MV-CEA that preferentially enters and replicates within malignant cells, including GBM [353]. MV-CEA demonstrated to be safe in an early phase 1 trial testing the injection of the virus in the tumor resection cavity of recurrent GBM (NCT00390299).

4.4.7. Newcastle Disease Virus (NDV)

NDV is an avian paramyxovirus with intrinsic oncolytic potential [354]. It is a negative-sense ssRNA virus that preferentially replicates within type I IFN-deficient cancer cells [355]. The HUJ [310] and MTH-68/H [311] strains of NDV have been the subject of clinical studies in patients with recurrent GBM. A phase 1/2 study of systemic application of NDV-HUJ revealed minimal toxicity and encouraging anti-tumor responses, with one patient achieving complete tumor remission during maintenance dosing [310]. However, the complete response was not durable.

4.4.8. H-1 Parvovirus (H-1PV)

Another promising strategy in the fight against GBM is the use of the oncolytic H-1PV. It is a rat protoparvovirus of the Parvoviridae family characterized by an ssDNA genome. It is not pathogenic for humans and naturally possesses oncolytic and oncosuppressive properties as demonstrated in various in vitro and in vivo models [356,357]. Wild-type H-1PV treatment was successful in a phase 1/2 clinical trial for recurrent or progressive GBM, where patients received initial H-1PV administration via intravenous or intratumoral injection, followed by surgical resection and virus re-injection into the resection cavity [312]. Results show that the treatment is safe, well tolerated, and associated with surrogate evidence of efficacy, including immune conversion of the TME and extended patient median OS in comparison with historical controls [312,313]. Compassionate use programs explored the combination of H-1PV with different agents, particularly bevacizumab, an anti-angiogenic agent with immunomodulating properties [358], the PD-1 inhibitor Nivolumab, and alongside Valproic acid, owing to encouraging preclinical results [359]. This multimodal therapeutic approach led to partial or complete objective responses in seven out of nine cases [360,361].

4.4.9. Vaccinia Virus (VACV)

Enveloped dsDNA vaccinia viruses belong to the Poxviridae family and, in most of cases, harbor inactivating mutations of the TK-encoding J2R gene (ΔJ2R VACV). ΔJ2R VACV therefore depends on host cells for TK protein, which is overexpressed in tumor cells [362]. Researchers developed the virus TG6002 by engineering a ΔJ2R VACV Copenhagen strain to express the yeast suicide gene FCU1 [363]. When combined with 5-FC, TG6002 activates the prodrug, leading to tumor cell death by inhibiting DNA and protein synthesis. A concluded Phase 1 trial (NCT03294486; ONCOVIRAC) tested the safety and efficacy of TG6002/5-FC in recurrent GBM patients; however, the results are not yet posted.

5. Combination Therapy

It has become increasingly evident that a singular treatment approach is insufficient for effectively addressing tumors, especially when dealing with a complex and heterogeneous entity like GBM. Researchers are now directing their attention toward combination therapies, seeking not only to combine immunotherapeutics with conventional treatments but also to explore synergies among different immune-based approaches (Table 5).
Immune checkpoint inhibitors are currently being tested in combination with CAR-T cells therapies (NCT03726515, NCT04003649), vaccination approaches (NCT03422094, NCT02287428, NCT04013672, NCT03014804, NCT04201873), and with oncolytic viruses such as AdVs (NCT03576612, NCT03636477), HSV (NCT05084430, NCT02798406), and PVSRIPO (NCT04479241, NCT03973879).
In addition to exploring immunotherapeutic strategies, it is crucial to consider the integration of radiation therapy into the treatment landscape for GBM. Being a first-line treatment and integral component of the Stupp protocol, combining radiation with immunotherapy is a logical approach. However, this combination introduces both opportunities and challenges. On the one hand, radiotherapy, with its tumor-targeting ionizing radiations, induces molecular lesions, including DNA breaks (single- and double-stranded) and base modifications triggering immunogenic cell death [369]. As extensively reviewed in De Martino et al. (2021) [370], radiotherapy has the potential to enhance GBM sensitivity to immune-based approaches by actively recruiting effector T cells to the tumor site, an essential requirement for successful immunotherapy. However, the intricate interplay between radiation and immune therapies demands careful consideration, as certain aspects of radiation might counteract immunotherapeutic mechanisms [369]. For instance, B cells, T cells, and NK cells are among the most radiosensitive cells of the TME, while immunosuppressive Tregs and MDSCs are quite resistant to radiation. The combination of radiotherapy with various forms of immunotherapy is an active area of research, with experiments in animal models demonstrating its potential efficacy and benefits. Building on these promising preclinical data, some clinical trials are strategically combining specific types of radiation therapy with immunotherapeutic to harness potential synergies [369]. Understanding the nuances of how radiation influences the immune response is essential for optimizing treatment outcomes and advancing the development of effective combination therapies for GBM.

6. Conclusions and Future Directions

GBM patients’ poor prognoses underscore the urgent need for novel treatments to enhance both the quality of life and overall survival for patients. While immunotherapeutic approaches have demonstrated significant efficacy in treating solid tumors, their effectiveness in addressing GBM remains limited. Despite promising results at the preclinical level, anti-GBM immunotherapeutics, whether tested individually or in combination with standard treatments, have so far failed to yield clinically meaningful outcomes when examined in phase 3 clinical trials. This high failure rate highlights the pressing need for more reliable preclinical models and early-stage clinical studies. Moreover, a better understanding of GBM tumor biology, in terms of local TME immunosuppression and systemic T cell dysfunction, is essential in the development of more targeted therapies. Recent advances in patient-derived GBM xenografts in humanized and immunotolerant murine models, as well as in ex vivo 3-D systems and microfluidics, can assist researchers in studying the intricate relationship between GBM and immune cells, leading to the discovery of new ways to efficiently modulate it [371]. Furthermore, these models serve as excellent preclinical settings for the high-throughput screening of therapeutic agents in a time-efficient and cost-effective manner. Artificial intelligence and machine learning can enhance preclinical models, supporting research efforts, and accelerating relevant discoveries. On the clinical side, the majority of phase 2 GBM studies are currently conducted in single-center settings with single-arm designs. A shift towards randomized, controlled, and adequately powered clinical studies can significantly contribute to the development of more robust therapies, preventing the wastage of valuable patient and financial resources and maximizing the reproducibility of results. Clinical trials should also consider including immune-predictive biomarkers and genomic characterization of tumors. This information could provide the key towards more personalized therapies addressing specific tumor signatures and are active areas of intense research.
Standard chemoradiotherapy is well-known for inducing immunosuppression and lymphopenia in GBM patients, posing a significant obstacle to GBM immune-based approaches. Essential changes in current standard treatments are required to increase the success of immunotherapies [372]. Moreover, failed clinical trials have taught us that targeting a single axis, such as a single antigen or immune checkpoint molecule, may not lead to success. Antigens exhibit heterogeneous spatial and temporal expression within tumors, influenced by the tumor microenvironment, treatment, tumor progression, and environmental factors. Consequently, CAR-T therapies are now simultaneously targeting three (trivalent) or more (polyvalent) antigenic targets, and peptide/DC vaccines increasingly utilize the entire tumor lysate rather than a single tumor antigen. Moreover, bispecific T cell engagers (BiTEs), which physically brings T cells in close proximity to tumor cells, have been proposed as a possible solution to overcome antigen escape mechanisms [373]. In addition, various personalized immune-based treatments, customized to individual patient profiles, are currently undergoing clinical evaluation and may hold the key to addressing the challenges posed by GBM. Neoantigen-based personalized vaccines demonstrate significant immunogenicity and safety in GBM, generating robust CD8+ and CD4+ T cell infiltration into the tumor. Alongside the personalization aspect is the consideration of combination therapy; it is crucial to comprehend which therapies synergize effectively and, notably, to determine the optimal timing for their administration to achieve maximum results.
The high costs associated with immunotherapies for GBM, especially in the realm of combination therapies, underscore the pressing need for sustainability in their pricing. To achieve this, stakeholders should focus on increasing research funding, fostering collaborative efforts, implementing regulatory incentives, and promoting value-based pricing. Additionally, encouraging global health partnerships, supporting insurance and health system reforms, and establishing patient assistance programs are crucial steps towards making these treatments more accessible and averting potential healthcare system collapses. By addressing these challenges, we can also work towards mitigating inequalities in access to GBM treatments, ensuring that all patients, regardless of their financial status, have equitable access to life-saving therapies.

Funding

This review was funded by the Luxembourg Cancer Foundation, Luxembourg Institute of Health, FNRS-Télévie (GBModImm no. 7.8513.18, GBModImm2 no. 7651720F, ImmoGBM 7.8505.20, ImmoGBM2 7.6603.22).

Acknowledgments

The authors thank Simone P. Niclou for her contribution in the reviewing of the original text from which this review is based.

Conflicts of Interest

A.M. is an inventor with multiple patents and patent applications related to H-1PV. The authors declare no other conflicts of interest.

References

  1. Louis, D.N.; Perry, A.; Wesseling, P.; Brat, D.J.; Cree, I.A.; Figarella-Branger, D.; Hawkins, C.; Ng, H.K.; Pfister, S.M.; Reifenberger, G.; et al. The 2021 WHO Classification of Tumors of the Central Nervous System: A summary. Neuro-Oncology 2021, 23, 1231–1251. [Google Scholar] [CrossRef]
  2. Miller, K.D.; Ostrom, Q.T.; Kruchko, C.; Patil, N.; Tihan, T.; Cioffi, G.; Fuchs, H.E.; Waite, K.A.; Jemal, A.; Siegel, R.L.; et al. Brain and other central nervous system tumor statistics, 2021. CA Cancer J. Clin. 2021, 71, 381–406. [Google Scholar] [CrossRef] [PubMed]
  3. Ohgaki, H.; Kleihues, P. The definition of primary and secondary glioblastoma. Clin. Cancer Res. 2013, 19, 764–772. [Google Scholar] [CrossRef] [PubMed]
  4. Ostrom, Q.T.; Cioffi, G.; Waite, K.; Kruchko, C.; Barnholtz-Sloan, J.S. CBTRUS Statistical Report: Primary Brain and Other Central Nervous System Tumors Diagnosed in the United States in 2014–2018. Neuro-Oncology 2021, 12 (Suppl. 2), iii1–iii105. [Google Scholar] [CrossRef]
  5. Stupp, R.; Mason, W.P.; van den Bent, M.J.; Weller, M.; Fisher, B.; Taphoorn, M.J.B.; Belanger, K.; Brandes, A.A.; Marosi, C.; Bogdahn, U.; et al. Radiotherapy plus Concomitant and Adjuvant Temozolomide for Glioblastoma. N. Engl. J. Med. 2005, 352, 987–996. [Google Scholar] [CrossRef]
  6. Weller, M.; van den Bent, M.; Preusser, M.; Le Rhun, E.; Tonn, J.C.; Minniti, G.; Bendszus, M.; Balana, C.; Chinot, O.; Dirven, L.; et al. EANO guidelines on the diagnosis and treatment of diffuse gliomas of adulthood. Nat. Rev. Clin. Oncol. 2021, 18, 170–186. [Google Scholar] [CrossRef] [PubMed]
  7. Erasimus, H.; Gobin, M.; Niclou, S.; Van Dyck, E. DNA repair mechanisms and their clinical impact in glioblastoma. Mutat. Res. Rev. Mutat. Res. 2016, 769, 19–35. [Google Scholar] [CrossRef]
  8. Hegi, M.E.; Diserens, A.-C.; Gorlia, T.; Hamou, M.-F.; de Tribolet, N.; Weller, M.; Kros, J.M.; Hainfellner, J.A.; Mason, W.; Mariani, L.; et al. MGMT Gene Silencing and Benefit from Temozolomide in Glioblastoma. N. Engl. J. Med. 2005, 352, 997–1003. [Google Scholar] [CrossRef]
  9. Perry, J.R.; Laperriere, N.; O’Callaghan, C.J.; Brandes, A.A.; Menten, J.; Phillips, C.; Fay, M.; Nishikawa, R.; Cairncross, J.G.; Roa, W.; et al. Short-Course Radiation plus Temozolomide in Elderly Patients with Glioblastoma. N. Engl. J. Med. 2017, 376, 1027–1037. [Google Scholar] [CrossRef]
  10. Stupp, R.; Wong, E.T.; Kanner, A.A.; Steinberg, D.; Engelhard, H.; Heidecke, V.; Kirson, E.D.; Taillibert, S.; Liebermann, F.; Dbaly, V.; et al. NovoTTF-100A versus physician’s choice chemotherapy in recurrent glioblastoma: A randomised phase III trial of a novel treatment modality. Eur. J. Cancer 2012, 48, 2192–2202. [Google Scholar] [CrossRef]
  11. Stupp, R.; Taillibert, S.; Kanner, A.; Read, W.; Steinberg, D.; Lhermitte, B.; Toms, S.; Idbaih, A.; Ahluwalia, M.S.; Fink, K.; et al. Effect of Tumor-Treating Fields Plus Maintenance Temozolomide vs. Maintenance Temozolomide Alone on Survival in Patients with Glioblastoma: A Randomized Clinical Trial. JAMA 2017, 318, 2306–2316. [Google Scholar] [CrossRef]
  12. Hottinger, A.F.; Pacheco, P.; Stupp, R. Tumor treating fields: A novel treatment modality and its use in brain tumors. Neuro-Oncology 2016, 18, 1338–1349. [Google Scholar] [CrossRef] [PubMed]
  13. Cloughesy, T.F.; Lassman, A.B. NovoTTF: Where to go from here? Neuro-Oncology 2017, 19, 605–608. [Google Scholar] [CrossRef] [PubMed]
  14. Taphoorn, M.J.B.; Dirven, L.; Kanner, A.A.; Lavy-Shahaf, G.; Weinberg, U.; Taillibert, S.; Toms, S.A.; Honnorat, J.; Chen, T.C.; Sroubek, J.; et al. Influence of Treatment with Tumor-Treating Fields on Health-Related Quality of Life of Patients with Newly Diagnosed Glioblastoma: A Secondary Analysis of a Randomized Clinical Trial. JAMA Oncol. 2018, 4, 495–504. [Google Scholar] [CrossRef]
  15. Weller, M.; Cloughesy, T.; Perry, J.R.; Wick, W. Standards of care for treatment of recurrent glioblastoma--are we there yet? Neuro-Oncology 2013, 15, 4–27. [Google Scholar] [CrossRef] [PubMed]
  16. Audureau, E.; Chivet, A.; Ursu, R.; Corns, R.; Metellus, P.; Noel, G.; Zouaoui, S.; Guyotat, J.; Le Reste, P.J.; Faillot, T.; et al. Prognostic factors for survival in adult patients with recurrent glioblastoma: A decision-tree-based model. J. Neuro-Oncol. 2018, 136, 565–576. [Google Scholar] [CrossRef] [PubMed]
  17. Weller, M.; Le Rhun, E. How did lomustine become standard of care in recurrent glioblastoma? Cancer Treat. Rev. 2020, 87, 102029. [Google Scholar] [CrossRef]
  18. Vaz-Salgado, M.A.; Villamayor, M.; Albarrán, V.; Alía, V.; Sotoca, P.; Chamorro, J.; Rosero, D.; Barrill, A.M.; Martín, M.; Fernandez, E.; et al. Recurrent Glioblastoma: A Review of the Treatment Options. Cancers 2023, 15, 4279. [Google Scholar] [CrossRef] [PubMed]
  19. Taal, W.; Oosterkamp, H.M.; Walenkamp, A.M.; Dubbink, H.J.; Beerepoot, L.V.; Hanse, M.C.; Buter, J.; Honkoop, A.H.; Boerman, D.; de Vos, F.Y.; et al. Single-agent bevacizumab or lomustine versus a combination of bevacizumab plus lomustine in patients with recurrent glioblastoma (BELOB trial): A randomised controlled phase 2 trial. Lancet Oncol. 2014, 15, 943–953. [Google Scholar] [CrossRef] [PubMed]
  20. Wick, W.; Gorlia, T.; Bendszus, M.; Taphoorn, M.; Sahm, F.; Harting, I.; Brandes, A.A.; Taal, W.; Domont, J.; Idbaih, A.; et al. Lomustine and Bevacizumab in Progressive Glioblastoma. N. Engl. J. Med. 2017, 377, 1954–1963. [Google Scholar] [CrossRef]
  21. Batchelor, T.T.; Mulholland, P.; Neyns, B.; Nabors, L.B.; Campone, M.; Wick, A.; Mason, W.; Mikkelsen, T.; Phuphanich, S.; Ashby, L.S.; et al. Phase III randomized trial comparing the efficacy of cediranib as monotherapy, and in combination with lomustine, versus lomustine alone in patients with recurrent glioblastoma. J. Clin. Oncol. 2013, 31, 3212–3218. [Google Scholar] [CrossRef]
  22. Brandes, A.A.; Carpentier, A.F.; Kesari, S.; Sepulveda-Sanchez, J.M.; Wheeler, H.R.; Chinot, O.; Cher, L.; Steinbach, J.P.; Capper, D.; Specenier, P.; et al. A Phase II randomized study of galunisertib monotherapy or galunisertib plus lomustine compared with lomustine monotherapy in patients with recurrent glioblastoma. Neuro-Oncology 2016, 18, 1146–1156. [Google Scholar] [CrossRef] [PubMed]
  23. Lombardi, G.; De Salvo, G.L.; Brandes, A.A.; Eoli, M.; Rudà, R.; Faedi, M.; Lolli, I.; Pace, A.; Daniele, B.; Pasqualetti, F.; et al. Regorafenib compared with lomustine in patients with relapsed glioblastoma (REGOMA): A multicentre, open-label, randomised, controlled, phase 2 trial. Lancet Oncol. 2019, 20, 110–119. [Google Scholar] [CrossRef] [PubMed]
  24. Wick, W.; Puduvalli, V.K.; Chamberlain, M.C.; van den Bent, M.J.; Carpentier, A.F.; Cher, L.M.; Mason, W.; Weller, M.; Hong, S.; Musib, L.; et al. Phase III study of enzastaurin compared with lomustine in the treatment of recurrent intracranial glioblastoma. J. Clin. Oncol. 2010, 28, 1168–1174. [Google Scholar] [CrossRef] [PubMed]
  25. Van Den Bent, M.; Eoli, M.; Sepulveda, J.M.; Smits, M.; Walenkamp, A.; Frenel, J.-S.; Franceschi, E.; Clement, P.M.; Chinot, O.; De Vos, F.; et al. INTELLANCE 2/EORTC 1410 randomized phase II study of Depatux-M alone and with temozolomide vs temozolomide or lomustine in recurrent EGFR amplified glioblastoma. Neuro-Oncology 2019, 22, 684–693. [Google Scholar] [CrossRef] [PubMed]
  26. Duerinck, J.; Du Four, S.; Bouttens, F.; Andre, C.; Verschaeve, V.; Van Fraeyenhove, F.; Chaskis, C.; D’Haene, N.; Le Mercier, M.; Rogiers, A.; et al. Randomized phase II trial comparing axitinib with the combination of axitinib and lomustine in patients with recurrent glioblastoma. J. Neuro-Oncol. 2018, 136, 115–125. [Google Scholar] [CrossRef] [PubMed]
  27. Bonosi, L.; Marrone, S.; Benigno, U.E.; Buscemi, F.; Musso, S.; Porzio, M.; Silven, M.P.; Torregrossa, F.; Grasso, G. Maximal Safe Resection in Glioblastoma Surgery: A Systematic Review of Advanced Intraoperative Image-Guided Techniques. Brain Sci. 2023, 13, 216. [Google Scholar] [CrossRef]
  28. Gerritsen, J.K.W.; Broekman, M.L.D.; De Vleeschouwer, S.; Schucht, P.; Jungk, C.; Krieg, S.M.; Nahed, B.V.; Berger, M.S.; Vincent, A. Decision making and surgical modality selection in glioblastoma patients: An international multicenter survey. J. Neuro-Oncol. 2022, 156, 465–482. [Google Scholar] [CrossRef]
  29. Stummer, W.; Pichlmeier, U.; Meinel, T.; Wiestler, O.D.; Zanella, F.; Reulen, H.-J. Fluorescence-guided surgery with 5-aminolevulinic acid for resection of malignant glioma: A randomised controlled multicentre phase III trial. Lancet Oncol. 2006, 7, 392–401. [Google Scholar] [CrossRef] [PubMed]
  30. Stummer, W.; Tonn, J.C.; Mehdorn, H.M.; Nestler, U.; Franz, K.; Goetz, C.; Bink, A.; Pichlmeier, U.; Group, A.L.-G.S. Counterbalancing risks and gains from extended resections in malignant glioma surgery: A supplemental analysis from the randomized 5-aminolevulinic acid glioma resection study. Clinical article. J. Neurosurg. 2011, 114, 613–623. [Google Scholar] [CrossRef]
  31. Obermeier, B.; Verma, A.; Ransohoff, R.M. The blood-brain barrier. Handb. Clin. Neurol. 2016, 133, 39–59. [Google Scholar] [CrossRef]
  32. Daneman, R.; Prat, A. The blood-brain barrier. Cold Spring Harb. Perspect. Biol. 2015, 7, a020412. [Google Scholar] [CrossRef] [PubMed]
  33. Kadry, H.; Noorani, B.; Cucullo, L. A blood-brain barrier overview on structure, function, impairment, and biomarkers of integrity. Fluids Barriers CNS 2020, 17, 69. [Google Scholar] [CrossRef] [PubMed]
  34. Ehrlich, P. Das sauerstufbudurfnis des organismus. In Eine Farbenanalytische Studies; Hirschwald: Berlin, Germany, 1885; p. 167. [Google Scholar]
  35. Ehrlich, F. Ueber das natürliche Isomere des Leucins. Berichte Dtsch. Chem. Ges. 1904, 37, 1809–1840. [Google Scholar] [CrossRef]
  36. Goldmann, E.E. Die Aussere und Innere Skeretion des Gesunden Organismus im Lichte der “Vitalen Farbung”; Lauppsche: Tübingen, Germany, 1909. [Google Scholar]
  37. Reese, T.S.; Karnovsky, M.J. Fine structural localization of a blood-brain barrier to exogenous peroxidase. J. Cell Biol. 1967, 34, 207–217. [Google Scholar] [CrossRef] [PubMed]
  38. Ghose, A.K.; Viswanadhan, V.N.; Wendoloski, J.J. A Knowledge-Based Approach in Designing Combinatorial or Medicinal Chemistry Libraries for Drug Discovery. 1. A Qualitative and Quantitative Characterization of Known Drug Databases. J. Comb. Chem. 1999, 1, 55–68. [Google Scholar] [CrossRef] [PubMed]
  39. Lipinski, C.A. Drug-like properties and the causes of poor solubility and poor permeability. J. Pharmacol. Toxicol. Methods 2000, 44, 235–249. [Google Scholar] [CrossRef] [PubMed]
  40. Annunziata, P.; Cioni, C.; Toneatto, S.; Paccagnini, E. HIV-1 gp120 increases the permeability of rat brain endothelium cultures by a mechanism involving substance P. Aids 1998, 12, 2377–2385. [Google Scholar] [CrossRef] [PubMed]
  41. Kustova, Y.; Grinberg, A.; Basile, A.S. Increased blood-brain barrier permeability in LP-BM5 infected mice is mediated by neuroexcitatory mechanisms. Brain Res. 1999, 839, 153–163. [Google Scholar] [CrossRef] [PubMed]
  42. St’astný, F.; Skultétyová, I.; Pliss, L.; Jezová, D. Quinolinic acid enhances permeability of rat brain microvessels to plasma albumin. Brain Res. Bull. 2000, 53, 415–420. [Google Scholar] [CrossRef]
  43. Abbott, N.J. Astrocyte-endothelial interactions and blood-brain barrier permeability. J. Anat. 2002, 200, 629–638. [Google Scholar] [CrossRef]
  44. Mo, F.; Pellerino, A.; Soffietti, R.; Ruda, R. Blood-Brain Barrier in Brain Tumors: Biology and Clinical Relevance. Int. J. Mol. Sci. 2021, 22, 12654. [Google Scholar] [CrossRef] [PubMed]
  45. Ahmed, M.H.; Canney, M.; Carpentier, A.; Thanou, M.; Idbaih, A. Unveiling the enigma of the blood-brain barrier in glioblastoma: Current advances from preclinical and clinical studies. Curr. Opin. Oncol. 2023, 35, 522–528. [Google Scholar] [CrossRef]
  46. Brennan, C.W.; Verhaak, R.G.; McKenna, A.; Campos, B.; Noushmehr, H.; Salama, S.R.; Zheng, S.; Chakravarty, D.; Sanborn, J.Z.; Berman, S.H.; et al. The somatic genomic landscape of glioblastoma. Cell 2013, 155, 462–477. [Google Scholar] [CrossRef] [PubMed]
  47. Molinaro, A.M.; Taylor, J.W.; Wiencke, J.K.; Wrensch, M.R. Genetic and molecular epidemiology of adult diffuse glioma. Nat. Rev. Neurol. 2019, 15, 405–417. [Google Scholar] [CrossRef] [PubMed]
  48. Verhaak, R.G.; Hoadley, K.A.; Purdom, E.; Wang, V.; Qi, Y.; Wilkerson, M.D.; Miller, C.R.; Ding, L.; Golub, T.; Mesirov, J.P.; et al. Integrated genomic analysis identifies clinically relevant subtypes of glioblastoma characterized by abnormalities in PDGFRA, IDH1, EGFR, and NF1. Cancer Cell 2010, 17, 98–110. [Google Scholar] [CrossRef] [PubMed]
  49. Wang, Q.; Hu, B.; Hu, X.; Kim, H.; Squatrito, M.; Scarpace, L.; deCarvalho, A.C.; Lyu, S.; Li, P.; Li, Y.; et al. Tumor Evolution of Glioma-Intrinsic Gene Expression Subtypes Associates with Immunological Changes in the Microenvironment. Cancer Cell 2017, 32, 42–56.e6. [Google Scholar] [CrossRef] [PubMed]
  50. Capper, D.; Reifenberger, G.; French, P.J.; Schweizer, L.; Weller, M.; Touat, M.; Niclou, S.P.; Euskirchen, P.; Haberler, C.; Hegi, M.E.; et al. EANO guideline on rational molecular testing of gliomas, glioneuronal, and neuronal tumors in adults for targeted therapy selection. Neuro-Oncology 2023, 25, 813–826. [Google Scholar] [CrossRef] [PubMed]
  51. Yabo, Y.A.; Niclou, S.P.; Golebiewska, A. Cancer cell heterogeneity and plasticity: A paradigm shift in glioblastoma. Neuro-Oncology 2022, 24, 669–682. [Google Scholar] [CrossRef]
  52. White, K.; Connor, K.; Meylan, M.; Bougoüin, A.; Salvucci, M.; Bielle, F.; O’Farrell, A.C.; Sweeney, K.; Weng, L.; Bergers, G.; et al. Identification, validation and biological characterisation of novel glioblastoma tumour microenvironment subtypes: Implications for precision immunotherapy. Ann. Oncol. 2023, 34, 300–314. [Google Scholar] [CrossRef]
  53. Eisenbarth, D.; Wang, Y.A. Glioblastoma heterogeneity at single cell resolution. Oncogene 2023, 42, 2155–2165. [Google Scholar] [CrossRef]
  54. Larsson, I.; Dalmo, E.; Elgendy, R.; Niklasson, M.; Doroszko, M.; Segerman, A.; Jörnsten, R.; Westermark, B.; Nelander, S. Modeling glioblastoma heterogeneity as a dynamic network of cell states. Mol. Syst. Biol. 2021, 17, e10105. [Google Scholar] [CrossRef] [PubMed]
  55. So, J.S.; Kim, H.; Han, K.S. Mechanisms of Invasion in Glioblastoma: Extracellular Matrix, Ca(2+) Signaling, and Glutamate. Front. Cell Neurosci. 2021, 15, 663092. [Google Scholar] [CrossRef] [PubMed]
  56. Xie, Q.; Mittal, S.; Berens, M.E. Targeting adaptive glioblastoma: An overview of proliferation and invasion. Neuro-Oncology 2014, 16, 1575–1584. [Google Scholar] [CrossRef] [PubMed]
  57. Claes, A.; Idema, A.J.; Wesseling, P. Diffuse glioma growth: A guerilla war. Acta Neuropathol. 2007, 114, 443–458. [Google Scholar] [CrossRef] [PubMed]
  58. Friedl, P.; Wolf, K. Tumour-cell invasion and migration: Diversity and escape mechanisms. Nat. Rev. Cancer 2003, 3, 362–374. [Google Scholar] [CrossRef] [PubMed]
  59. Wolf, K.; Friedl, P. Molecular mechanisms of cancer cell invasion and plasticity. Br. J. Dermatol. 2006, 154 (Suppl. 1), 11–15. [Google Scholar] [CrossRef] [PubMed]
  60. Scherer, H.J. Structural Development in Gliomas. Am. J. Cancer 1938, 34, 333–351. [Google Scholar] [CrossRef]
  61. Cuddapah, V.A.; Robel, S.; Watkins, S.; Sontheimer, H. A neurocentric perspective on glioma invasion. Nat. Rev. Neurosci. 2014, 15, 455–465. [Google Scholar] [CrossRef] [PubMed]
  62. Velasquez, C.; Mansouri, S.; Mora, C.; Nassiri, F.; Suppiah, S.; Martino, J.; Zadeh, G.; Fernandez-Luna, J.L. Molecular and Clinical Insights into the Invasive Capacity of Glioblastoma Cells. J. Oncol. 2019, 2019, 1740763. [Google Scholar] [CrossRef]
  63. Fabian, C.; Han, M.; Bjerkvig, R.; Niclou, S.P. Novel facets of glioma invasion. Int. Rev. Cell Mol. Biol. 2021, 360, 33–64. [Google Scholar] [CrossRef]
  64. Beauchesne, P. Extra-neural metastases of malignant gliomas: Myth or reality? Cancers 2011, 3, 461–477. [Google Scholar] [CrossRef] [PubMed]
  65. Lun, M.; Lok, E.; Gautam, S.; Wu, E.; Wong, E.T. The natural history of extracranial metastasis from glioblastoma multiforme. J. Neuro-Oncol. 2011, 105, 261–273. [Google Scholar] [CrossRef] [PubMed]
  66. Hamilton, J.D.; Rapp, M.; Schneiderhan, T.; Sabel, M.; Hayman, A.; Scherer, A.; Kröpil, P.; Budach, W.; Gerber, P.; Kretschmar, U.; et al. Glioblastoma multiforme metastasis outside the CNS: Three case reports and possible mechanisms of escape. J. Clin. Oncol. 2014, 32, e80–e84. [Google Scholar] [CrossRef] [PubMed]
  67. Chen, J.; Li, Y.; Yu, T.S.; McKay, R.M.; Burns, D.K.; Kernie, S.G.; Parada, L.F. A restricted cell population propagates glioblastoma growth after chemotherapy. Nature 2012, 488, 522–526. [Google Scholar] [CrossRef]
  68. Holdhoff, M.; Ye, X.; Supko, J.G.; Nabors, L.B.; Desai, A.S.; Walbert, T.; Lesser, G.J.; Read, W.L.; Lieberman, F.S.; Lodge, M.A.; et al. Timed sequential therapy of the selective T-type calcium channel blocker mibefradil and temozolomide in patients with recurrent high-grade gliomas. Neuro-Oncology 2017, 19, 845–852. [Google Scholar] [CrossRef] [PubMed]
  69. Eisele, G.; Wick, A.; Eisele, A.C.; Clément, P.M.; Tonn, J.; Tabatabai, G.; Ochsenbein, A.; Schlegel, U.; Neyns, B.; Krex, D.; et al. Cilengitide treatment of newly diagnosed glioblastoma patients does not alter patterns of progression. J. Neuro-Oncol. 2014, 117, 141–145. [Google Scholar] [CrossRef] [PubMed]
  70. Tonn, J.C.; Kerkau, S.; Hanke, A.; Bouterfa, H.; Mueller, J.G.; Wagner, S.; Vince, G.H.; Roosen, K. Effect of synthetic matrix-metalloproteinase inhibitors on invasive capacity and proliferation of human malignant gliomas in vitro. Int. J. Cancer 1999, 80, 764–772. [Google Scholar] [CrossRef]
  71. Koutroulis, I.; Zarros, A.; Theocharis, S. The role of matrix metalloproteinases in the pathophysiology and progression of human nervous system malignancies: A chance for the development of targeted therapeutic approaches? Expert. Opin. Ther. Targets 2008, 12, 1577–1586. [Google Scholar] [CrossRef] [PubMed]
  72. Grossman, S.A.; Ye, X.; Chamberlain, M.; Mikkelsen, T.; Batchelor, T.; Desideri, S.; Piantadosi, S.; Fisher, J.; Fine, H.A. Talampanel with standard radiation and temozolomide in patients with newly diagnosed glioblastoma: A multicenter phase II trial. J. Clin. Oncol. 2009, 27, 4155–4161. [Google Scholar] [CrossRef]
  73. Iwamoto, F.M.; Kreisl, T.N.; Kim, L.; Duic, J.P.; Butman, J.A.; Albert, P.S.; Fine, H.A. Phase 2 trial of talampanel, a glutamate receptor inhibitor, for adults with recurrent malignant gliomas. Cancer 2010, 116, 1776–1782. [Google Scholar] [CrossRef]
  74. Drappatz, J.; Norden, A.D.; Wen, P.Y. Therapeutic strategies for inhibiting invasion in glioblastoma. Expert. Rev. Neurother. 2009, 9, 519–534. [Google Scholar] [CrossRef] [PubMed]
  75. Iliff, J.J.; Wang, M.; Liao, Y.; Plogg, B.A.; Peng, W.; Gundersen, G.A.; Benveniste, H.; Vates, G.E.; Deane, R.; Goldman, S.A.; et al. A paravascular pathway facilitates CSF flow through the brain parenchyma and the clearance of interstitial solutes, including amyloid β. Sci. Transl. Med. 2012, 4, 147ra111. [Google Scholar] [CrossRef] [PubMed]
  76. Louveau, A.; Smirnov, I.; Keyes, T.J.; Eccles, J.D.; Rouhani, S.J.; Peske, J.D.; Derecki, N.C.; Castle, D.; Mandell, J.W.; Lee, K.S.; et al. Structural and functional features of central nervous system lymphatic vessels. Nature 2015, 523, 337–341. [Google Scholar] [CrossRef] [PubMed]
  77. Srivastava, S.; Jackson, C.; Kim, T.; Choi, J.; Lim, M. A Characterization of Dendritic Cells and Their Role in Immunotherapy in Glioblastoma: From Preclinical Studies to Clinical Trials. Cancers 2019, 11, 537. [Google Scholar] [CrossRef] [PubMed]
  78. Papadopoulos, Z.; Herz, J.; Kipnis, J. Meningeal Lymphatics: From Anatomy to Central Nervous System Immune Surveillance. J. Immunol. 2020, 204, 286–293. [Google Scholar] [CrossRef] [PubMed]
  79. Kivisäkk, P.; Mahad, D.J.; Callahan, M.K.; Trebst, C.; Tucky, B.; Wei, T.; Wu, L.; Baekkevold, E.S.; Lassmann, H.; Staugaitis, S.M.; et al. Human cerebrospinal fluid central memory CD4+ T cells: Evidence for trafficking through choroid plexus and meninges via P-selectin. Proc. Natl. Acad. Sci. USA 2003, 100, 8389–8394. [Google Scholar] [CrossRef] [PubMed]
  80. Zhang, X.; Zhao, L.; Zhang, H.; Zhang, Y.; Ju, H.; Wang, X.; Ren, H.; Zhu, X.; Dong, Y. The immunosuppressive microenvironment and immunotherapy in human glioblastoma. Front. Immunol. 2022, 13, 1003651. [Google Scholar] [CrossRef]
  81. Gustafson, M.P.; Lin, Y.; New, K.C.; Bulur, P.A.; O’Neill, B.P.; Gastineau, D.A.; Dietz, A.B. Systemic immune suppression in glioblastoma: The interplay between CD14+HLA-DRlo/neg monocytes, tumor factors, and dexamethasone. Neuro-Oncology 2010, 12, 631–644. [Google Scholar] [CrossRef]
  82. Parney, I.F. Basic concepts in glioma immunology. Adv. Exp. Med. Biol. 2012, 746, 42–52. [Google Scholar] [CrossRef] [PubMed]
  83. Chongsathidkiet, P.; Jackson, C.; Koyama, S.; Loebel, F.; Cui, X.; Farber, S.H.; Woroniecka, K.; Elsamadicy, A.A.; Dechant, C.A.; Kemeny, H.R.; et al. Sequestration of T cells in bone marrow in the setting of glioblastoma and other intracranial tumors. Nat. Med. 2018, 24, 1459–1468. [Google Scholar] [CrossRef]
  84. Andaloussi, A.E.; Han, Y.; Lesniak, M.S. Progression of intracranial glioma disrupts thymic homeostasis and induces T-cell apoptosis in vivo. Cancer Immunol. Immunother. 2008, 57, 1807–1816. [Google Scholar] [CrossRef]
  85. Matloubian, M.; Lo, C.G.; Cinamon, G.; Lesneski, M.J.; Xu, Y.; Brinkmann, V.; Allende, M.L.; Proia, R.L.; Cyster, J.G. Lymphocyte egress from thymus and peripheral lymphoid organs is dependent on S1P receptor 1. Nature 2004, 427, 355–360. [Google Scholar] [CrossRef]
  86. Ayasoufi, K.; Pfaller, C.K.; Evgin, L.; Khadka, R.H.; Tritz, Z.P.; Goddery, E.N.; Fain, C.E.; Yokanovich, L.T.; Himes, B.T.; Jin, F.; et al. Brain cancer induces systemic immunosuppression through release of non-steroid soluble mediators. Brain 2020, 143, 3629–3652. [Google Scholar] [CrossRef]
  87. Rodrigues, J.C.; Gonzalez, G.C.; Zhang, L.; Ibrahim, G.; Kelly, J.J.; Gustafson, M.P.; Lin, Y.; Dietz, A.B.; Forsyth, P.A.; Yong, V.W.; et al. Normal human monocytes exposed to glioma cells acquire myeloid-derived suppressor cell-like properties. Neuro-Oncology 2010, 12, 351–365. [Google Scholar] [CrossRef] [PubMed]
  88. Wolburg, H.; Noell, S.; Fallier-Becker, P.; Mack, A.F.; Wolburg-Buchholz, K. The disturbed blood-brain barrier in human glioblastoma. Mol. Asp. Med. 2012, 33, 579–589. [Google Scholar] [CrossRef] [PubMed]
  89. Dubois, L.G.; Campanati, L.; Righy, C.; D’Andrea-Meira, I.; Spohr, T.C.; Porto-Carreiro, I.; Pereira, C.M.; Balca-Silva, J.; Kahn, S.A.; DosSantos, M.F.; et al. Gliomas and the vascular fragility of the blood brain barrier. Front. Cell Neurosci. 2014, 8, 418. [Google Scholar] [CrossRef]
  90. Li, A.; Dubey, S.; Varney, M.L.; Dave, B.J.; Singh, R.K. IL-8 directly enhanced endothelial cell survival, proliferation, and matrix metalloproteinases production and regulated angiogenesis. J. Immunol. 2003, 170, 3369–3376. [Google Scholar] [CrossRef]
  91. Chen, Z.; Mou, L.; Pan, Y.; Feng, C.; Zhang, J.; Li, J. CXCL8 Promotes Glioma Progression By Activating The JAK/STAT1/HIF-1alpha/Snail Signaling Axis. Onco Targets Ther. 2019, 12, 8125–8138. [Google Scholar] [CrossRef] [PubMed]
  92. Groblewska, M.; Litman-Zawadzka, A.; Mroczko, B. The Role of Selected Chemokines and Their Receptors in the Development of Gliomas. Int. J. Mol. Sci. 2020, 21, 3704. [Google Scholar] [CrossRef]
  93. Oushy, S.; Hellwinkel, J.E.; Wang, M.; Nguyen, G.J.; Gunaydin, D.; Harland, T.A.; Anchordoquy, T.J.; Graner, M.W. Glioblastoma multiforme-derived extracellular vesicles drive normal astrocytes towards a tumour-enhancing phenotype. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2018, 373, 20160477. [Google Scholar] [CrossRef]
  94. Simon, T.; Jackson, E.; Giamas, G. Breaking through the glioblastoma micro-environment via extracellular vesicles. Oncogene 2020, 39, 4477–4490. [Google Scholar] [CrossRef] [PubMed]
  95. Constam, D.B.; Philipp, J.; Malipiero, U.V.; ten Dijke, P.; Schachner, M.; Fontana, A. Differential expression of transforming growth factor-beta 1, -beta 2, and -beta 3 by glioblastoma cells, astrocytes, and microglia. J. Immunol. 1992, 148, 1404–1410. [Google Scholar] [CrossRef] [PubMed]
  96. Huettner, C.; Paulus, W.; Roggendorf, W. Messenger RNA expression of the immunosuppressive cytokine IL-10 in human gliomas. Am. J. Pathol. 1995, 146, 317–322. [Google Scholar] [PubMed]
  97. Takeshima, H.; Kuratsu, J.; Takeya, M.; Yoshimura, T.; Ushio, Y. Expression and localization of messenger RNA and protein for monocyte chemoattractant protein-1 in human malignant glioma. J. Neurosurg. 1994, 80, 1056–1062. [Google Scholar] [CrossRef] [PubMed]
  98. Wang, C.; Sinha, S.; Jiang, X.; Murphy, L.; Fitch, S.; Wilson, C.; Grant, G.; Yang, F. Matrix Stiffness Modulates Patient-Derived Glioblastoma Cell Fates in Three-Dimensional Hydrogels. Tissue Eng. Part A 2021, 27, 390–401. [Google Scholar] [CrossRef] [PubMed]
  99. Wick, W.; Platten, M.; Weller, M. Glioma cell invasion: Regulation of metalloproteinase activity by TGF-beta. J. Neuro-Oncol. 2001, 53, 177–185. [Google Scholar] [CrossRef] [PubMed]
  100. Park, J.H.; Lee, H.K. Current Understanding of Hypoxia in Glioblastoma Multiforme and Its Response to Immunotherapy. Cancers 2022, 14, 1176. [Google Scholar] [CrossRef]
  101. Hu, M.; Li, Y.; Lu, Y.; Wang, M.; Li, Y.; Wang, C.; Li, Q.; Zhao, H. The regulation of immune checkpoints by the hypoxic tumor microenvironment. PeerJ 2021, 9, e11306. [Google Scholar] [CrossRef] [PubMed]
  102. Sofroniew, M.V. Astrocyte barriers to neurotoxic inflammation. Nat. Rev. Neurosci. 2015, 16, 249–263. [Google Scholar] [CrossRef]
  103. Henrik Heiland, D.; Ravi, V.M.; Behringer, S.P.; Frenking, J.H.; Wurm, J.; Joseph, K.; Garrelfs, N.W.C.; Strahle, J.; Heynckes, S.; Grauvogel, J.; et al. Tumor-associated reactive astrocytes aid the evolution of immunosuppressive environment in glioblastoma. Nat. Commun. 2019, 10, 2541. [Google Scholar] [CrossRef]
  104. Venkatesh, H.S.; Johung, T.B.; Caretti, V.; Noll, A.; Tang, Y.; Nagaraja, S.; Gibson, E.M.; Mount, C.W.; Polepalli, J.; Mitra, S.S.; et al. Neuronal Activity Promotes Glioma Growth through Neuroligin-3 Secretion. Cell 2015, 161, 803–816. [Google Scholar] [CrossRef] [PubMed]
  105. Wu, A.; Wei, J.; Kong, L.Y.; Wang, Y.; Priebe, W.; Qiao, W.; Sawaya, R.; Heimberger, A.B. Glioma cancer stem cells induce immunosuppressive macrophages/microglia. Neuro-Oncology 2010, 12, 1113–1125. [Google Scholar] [CrossRef] [PubMed]
  106. Coniglio, S.J.; Eugenin, E.; Dobrenis, K.; Stanley, E.R.; West, B.L.; Symons, M.H.; Segall, J.E. Microglial stimulation of glioblastoma invasion involves epidermal growth factor receptor (EGFR) and colony stimulating factor 1 receptor (CSF-1R) signaling. Mol. Med. 2012, 18, 519–527. [Google Scholar] [CrossRef] [PubMed]
  107. Pyonteck, S.M.; Akkari, L.; Schuhmacher, A.J.; Bowman, R.L.; Sevenich, L.; Quail, D.F.; Olson, O.C.; Quick, M.L.; Huse, J.T.; Teijeiro, V.; et al. CSF-1R inhibition alters macrophage polarization and blocks glioma progression. Nat. Med. 2013, 19, 1264–1272. [Google Scholar] [CrossRef] [PubMed]
  108. Hambardzumyan, D.; Gutmann, D.H.; Kettenmann, H. The role of microglia and macrophages in glioma maintenance and progression. Nat. Neurosci. 2016, 19, 20–27. [Google Scholar] [CrossRef] [PubMed]
  109. Dumas, A.A.; Pomella, N.; Rosser, G.; Guglielmi, L.; Vinel, C.; Millner, T.O.; Rees, J.; Aley, N.; Sheer, D.; Wei, J.; et al. Microglia promote glioblastoma via mTOR-mediated immunosuppression of the tumour microenvironment. Embo J 2020, 39, e103790. [Google Scholar] [CrossRef] [PubMed]
  110. Suzumura, A.; Sawada, M.; Yamamoto, H.; Marunouchi, T. Transforming growth factor-beta suppresses activation and proliferation of microglia in vitro. J. Immunol. 1993, 151, 2150–2158. [Google Scholar] [CrossRef] [PubMed]
  111. Hussain, S.F.; Yang, D.; Suki, D.; Aldape, K.; Grimm, E.; Heimberger, A.B. The role of human glioma-infiltrating microglia/macrophages in mediating antitumor immune responses. Neuro-Oncology 2006, 8, 261–279. [Google Scholar] [CrossRef] [PubMed]
  112. Kilian, M.; Sheinin, R.; Tan, C.L.; Friedrich, M.; Krämer, C.; Kaminitz, A.; Sanghvi, K.; Lindner, K.; Chih, Y.-C.; Cichon, F.; et al. MHC class II-restricted antigen presentation is required to prevent dysfunction of cytotoxic T cells by blood-borne myeloids in brain tumors. Cancer Cell 2023, 41, 235–251. [Google Scholar] [CrossRef]
  113. Nagaraj, S.; Gabrilovich, D.I. Myeloid-derived suppressor cells. Adv. Exp. Med. Biol. 2007, 601, 213–223. [Google Scholar] [CrossRef]
  114. Dubinski, D.; Wölfer, J.; Hasselblatt, M.; Schneider-Hohendorf, T.; Bogdahn, U.; Stummer, W.; Wiendl, H.; Grauer, O.M. CD4+ T effector memory cell dysfunction is associated with the accumulation of granulocytic myeloid-derived suppressor cells in glioblastoma patients. Neuro-Oncology 2016, 18, 807–818. [Google Scholar] [CrossRef]
  115. Crane, C.A.; Ahn, B.J.; Han, S.J.; Parsa, A.T. Soluble factors secreted by glioblastoma cell lines facilitate recruitment, survival, and expansion of regulatory T cells: Implications for immunotherapy. Neuro-Oncology 2012, 14, 584–595. [Google Scholar] [CrossRef] [PubMed]
  116. Heimberger, A.B.; Abou-Ghazal, M.; Reina-Ortiz, C.; Yang, D.S.; Sun, W.; Qiao, W.; Hiraoka, N.; Fuller, G.N. Incidence and prognostic impact of FoxP3+ regulatory T cells in human gliomas. Clin. Cancer Res. 2008, 14, 5166–5172. [Google Scholar] [CrossRef] [PubMed]
  117. Li, C.; Jiang, P.; Wei, S.; Xu, X.; Wang, J. Regulatory T cells in tumor microenvironment: New mechanisms, potential therapeutic strategies and future prospects. Mol. Cancer 2020, 19, 116. [Google Scholar] [CrossRef] [PubMed]
  118. Crane, C.A.; Han, S.J.; Barry, J.J.; Ahn, B.J.; Lanier, L.L.; Parsa, A.T. TGF-beta downregulates the activating receptor NKG2D on NK cells and CD8+ T cells in glioma patients. Neuro-Oncology 2010, 12, 7–13. [Google Scholar] [CrossRef] [PubMed]
  119. Kelly, W.J.; Giles, A.J.; Gilbert, M. T lymphocyte-targeted immune checkpoint modulation in glioma. J. Immunother. Cancer 2020, 8, e000379. [Google Scholar] [CrossRef] [PubMed]
  120. Shinohara, H.; Yagita, H.; Ikawa, Y.; Oyaizu, N. Fas drives cell cycle progression in glioma cells via extracellular signal-regulated kinase activation. Cancer Res. 2000, 60, 1766–1772. [Google Scholar] [PubMed]
  121. Wischhusen, J.; Jung, G.; Radovanovic, I.; Beier, C.; Steinbach, J.P.; Rimner, A.; Huang, H.; Schulz, J.B.; Ohgaki, H.; Aguzzi, A.; et al. Identification of CD70-mediated apoptosis of immune effector cells as a novel immune escape pathway of human glioblastoma. Cancer Res. 2002, 62, 2592–2599. [Google Scholar] [PubMed]
  122. Chahlavi, A.; Rayman, P.; Richmond, A.L.; Biswas, K.; Zhang, R.; Vogelbaum, M.; Tannenbaum, C.; Barnett, G.; Finke, J.H. Glioblastomas induce T-lymphocyte death by two distinct pathways involving gangliosides and CD70. Cancer Res. 2005, 65, 5428–5438. [Google Scholar] [CrossRef]
  123. Pham, T.; Roth, S.; Kong, J.; Guerra, G.; Narasimhan, V.; Pereira, L.; Desai, J.; Heriot, A.; Ramsay, R. An Update on Immunotherapy for Solid Tumors: A Review. Ann. Surg. Oncol. 2018, 25, 3404–3412. [Google Scholar] [CrossRef]
  124. Philip, M.; Schietinger, A. CD8(+) T cell differentiation and dysfunction in cancer. Nat. Rev. Immunol. 2022, 22, 209–223. [Google Scholar] [CrossRef] [PubMed]
  125. Ansell, S.M.; Lesokhin, A.M.; Borrello, I.; Halwani, A.; Scott, E.C.; Gutierrez, M.; Schuster, S.J.; Millenson, M.M.; Cattry, D.; Freeman, G.J.; et al. PD-1 blockade with nivolumab in relapsed or refractory Hodgkin’s lymphoma. N. Engl. J. Med. 2015, 372, 311–319. [Google Scholar] [CrossRef] [PubMed]
  126. Ferris, R.L.; Blumenschein, G., Jr.; Fayette, J.; Guigay, J.; Colevas, A.D.; Licitra, L.; Harrington, K.; Kasper, S.; Vokes, E.E.; Even, C.; et al. Nivolumab for Recurrent Squamous-Cell Carcinoma of the Head and Neck. N. Engl. J. Med. 2016, 375, 1856–1867. [Google Scholar] [CrossRef] [PubMed]
  127. Bellmunt, J.; de Wit, R.; Vaughn, D.J.; Fradet, Y.; Lee, J.L.; Fong, L.; Vogelzang, N.J.; Climent, M.A.; Petrylak, D.P.; Choueiri, T.K.; et al. Pembrolizumab as Second-Line Therapy for Advanced Urothelial Carcinoma. N. Engl. J. Med. 2017, 376, 1015–1026. [Google Scholar] [CrossRef] [PubMed]
  128. Motzer, R.J.; Tannir, N.M.; McDermott, D.F.; Aren Frontera, O.; Melichar, B.; Choueiri, T.K.; Plimack, E.R.; Barthelemy, P.; Porta, C.; George, S.; et al. Nivolumab plus Ipilimumab versus Sunitinib in Advanced Renal-Cell Carcinoma. N. Engl. J. Med. 2018, 378, 1277–1290. [Google Scholar] [CrossRef] [PubMed]
  129. Overman, M.J.; Lonardi, S.; Wong, K.Y.M.; Lenz, H.J.; Gelsomino, F.; Aglietta, M.; Morse, M.A.; Van Cutsem, E.; McDermott, R.; Hill, A.; et al. Durable Clinical Benefit with Nivolumab Plus Ipilimumab in DNA Mismatch Repair-Deficient/Microsatellite Instability-High Metastatic Colorectal Cancer. J. Clin. Oncol. 2018, 36, 773–779. [Google Scholar] [CrossRef] [PubMed]
  130. Zakharia, Y.; Johnson, T.S.; Colman, H.; Vahanian, N.N.; Link, C.J.; Kennedy, E.; Sadek, R.F.; Kong, F.M.; Vender, J.; Munn, D.; et al. A phase I/II study of the combination of indoximod and temozolomide for adult patients with temozolomide-refractory primary malignant brain tumors. J. Clin. Oncol. 2014, 32 (Suppl. 15), TPS2107. [Google Scholar] [CrossRef]
  131. Schalper, K.A.; Rodriguez-Ruiz, M.E.; Diez-Valle, R.; Lopez-Janeiro, A.; Porciuncula, A.; Idoate, M.A.; Inoges, S.; de Andrea, C.; Lopez-Diaz de Cerio, A.; Tejada, S.; et al. Neoadjuvant nivolumab modifies the tumor immune microenvironment in resectable glioblastoma. Nat. Med. 2019, 25, 470–476. [Google Scholar] [CrossRef] [PubMed]
  132. Sanborn, R.E.; Pishvaian, M.J.; Callahan, M.K.; Weise, A.; Sikic, B.I.; Rahma, O.; Cho, D.C.; Rizvi, N.A.; Sznol, M.; Lutzky, J.; et al. Safety, tolerability and efficacy of agonist anti-CD27 antibody (varlilumab) administered in combination with anti-PD-1 (nivolumab) in advanced solid tumors. J. Immunother. Cancer 2022, 10, e005147. [Google Scholar] [CrossRef]
  133. Jensen, C.; Maarup, S.B.; Poulsen, H.S.; Hasselbalch, B.; Karsdal, M.A.; Svane, I.M.; Lassen, U.N.; Willumsen, N. Indirect assessment of tumor-infiltrating lymphocyte activity in serum for predicting outcome in patients with glioblastoma treated with immunotherapy in the recurrent setting. J. Clin. Oncol. 2022, 40 (Suppl. 16), 2059. [Google Scholar] [CrossRef]
  134. Sim, H.-W.; Lwin, Z.; Barnes, E.; McDonald, K.; Yip, S.; Verhaak, R.; Heimberger, A.; Hall, M.; Wong, M.; Jennens, R.; et al. CTIM-24. Nutmeg: A randomized phase II study of nivolumab and temozolomide versus temozolomide alone in newly diagnosed elderly patients with glioblastoma. Neuro-Oncology 2022, 24 (Suppl. 7), vii65. [Google Scholar] [CrossRef]
  135. Ahluwalia, M.; Peereboom, D.; Schilero, C.; Forst, D.; Wong, E.; Wen, P.; Reardon, D. RBTT-01. randomized phase 2 open label study of nivolumab plus standard dose bevacizumab versus nivolumab plus low dose bevacizumab in recurrent glioblastoma. Neuro-Oncology 2018, 20 (Suppl. 6), vi234. [Google Scholar] [CrossRef]
  136. Ahluwalia, M.S.; Rauf, Y.; Li, H.; Wen, P.Y.; Peereboom, D.M.; Reardon, D.A. Randomized phase 2 study of nivolumab (nivo) plus either standard or reduced dose bevacizumab (bev) in recurrent glioblastoma (rGBM). J. Clin. Oncol. 2021, 39 (Suppl. 15), 2015. [Google Scholar] [CrossRef]
  137. Omuro, A.; Brandes, A.A.; Carpentier, A.F.; Idbaih, A.; Reardon, D.A.; Cloughesy, T.; Sumrall, A.; Baehring, J.; van den Bent, M.; Bähr, O.; et al. Radiotherapy combined with nivolumab or temozolomide for newly diagnosed glioblastoma with unmethylated MGMT promoter: An international randomized phase III trial. Neuro-Oncology 2023, 25, 123–134. [Google Scholar] [CrossRef] [PubMed]
  138. Lim, M.; Weller, M.; Idbaih, A.; Steinbach, J.; Finocchiaro, G.; Raval, R.R.; Ansstas, G.; Baehring, J.; Taylor, J.W.; Honnorat, J.; et al. Phase III trial of chemoradiotherapy with temozolomide plus nivolumab or placebo for newly diagnosed glioblastoma with methylated MGMT promoter. Neuro-Oncology 2022, 24, 1935–1949. [Google Scholar] [CrossRef] [PubMed]
  139. Reardon, D.A.; Kim, T.M.; Frenel, J.S.; Simonelli, M.; Lopez, J.; Subramaniam, D.S.; Siu, L.L.; Wang, H.; Krishnan, S.; Stein, K.; et al. Treatment with pembrolizumab in programmed death ligand 1-positive recurrent glioblastoma: Results from the multicohort phase 1 KEYNOTE-028 trial. Cancer 2021, 127, 1620–1629. [Google Scholar] [CrossRef] [PubMed]
  140. Nowak, A.K.; Cher, L.; Bowyer, S.; Gan, H.K.; Long, A.P.; Balasubramanian, A.; Lee, S.Y.; Lee, W.S.; Yoo, J.-S. Phase Ib study of olinvacimab (O) with pembrolizumab (P) in patients with recurrent glioblastoma (rGBM). J. Clin. Oncol. 2020, 38 (Suppl. 15), e14545. [Google Scholar] [CrossRef]
  141. Sahebjam, S.; Forsyth, P.; Tran, N.; Mokhtari, S.; Arrington, J.; Jaglal, M.; Etame, A.; Liu, J.; Wicklund, M.; Gatewood, T.; et al. ATIM-08. A phase I trial of pembrolizumab and vorinostat combined with temozolomide and radiation therapy for newly diagnosed glioblastoma (NCT03426891). Neuro-Oncology 2018, 20 (Suppl. 6), vi2. [Google Scholar] [CrossRef]
  142. Hwang, H.; Huang, J.; Khaddour, K.; Butt, O.H.; Ansstas, G.; Chen, J.; Katumba, R.G.; Kim, A.H.; Leuthardt, E.C.; Campian, J.L. Prolonged response of recurrent IDH-wild-type glioblastoma to laser interstitial thermal therapy with pembrolizumab. CNS Oncol. 2022, 11, Cns81. [Google Scholar] [CrossRef] [PubMed]
  143. Campian, J.; Butt, O.; Ghinaseddin, A.; Rahman, M.; Chheda, M.; Johanns, T.; Ansstas, G.; Huang, J.; Liu, J.; Talcott, G.; et al. Ctim-26. Phase I/II study of the combination of pembrolizumab (Mk-3475) and laser interstitial thermal therapy (litt) in recurrent glioblastoma. Neuro-Oncology 2021, 23 (Suppl. 6), vi56. [Google Scholar] [CrossRef]
  144. Sloan, A.E.; Rogers, L.R.; Machtay, M. Phase I/II study of laser interstitial thermotherapy (LITT) combined with checkpoint inhibitor for recurrent glioblastoma (rGBM). J. Clin. Oncol. 2018, 36 (Suppl. 15), TPS2074. [Google Scholar] [CrossRef]
  145. Giordano, F.A.; Layer, J.P.; Leonardelli, S.; Friker, L.L.; Seidel, C.; Schaub, C.; Turiello, R.; Sperk, E.; Grau, F.; Paech, D.; et al. Radiotherapy and olaptesed pegol (NOX-A12) in partially resected or biopsy-only MGMT-unmethylated glioblastoma: Interim data from the German multicenter phase 1/2 GLORIA trial. J. Clin. Oncol. 2022, 40 (Suppl. 16), 2050. [Google Scholar] [CrossRef]
  146. Giordano, F.A.; Layer, J.P.; Leonardelli, S.; Friker, L.L.; Turiello, R.; Corvino, D.; Zeyen, T.; Schaub, C.; Mueller, W.; Sperk, E.; et al. Potential predictive biomarker for response to radiotherapy and CXCL12 inhibition in glioblastoma in the phase I/II GLORIA trial. J. Clin. Oncol. 2023, 41 (Suppl. 16), 2048. [Google Scholar] [CrossRef]
  147. Baldini, C.; Cassier, P.A.; Delord, J.-P.; Simonelli, M.; Touat, M.; Yao, L.; Duic, J.P.; Gozman, A.; Marabelle, A. CTIM-03. pembrolizumab monotherapy for microsatellite instability-high (MSI-H) or mismatch repair deficient (DMMR) recurrent gliomas: Results from the multicohort phase 2 keynote-158 study. Neuro-Oncology 2022, 24 (Suppl. 7), vii59–vii60. [Google Scholar] [CrossRef]
  148. Reardon, D.A.; Nayak, L.; Peters, K.B.; Clarke, J.L.; Jordan, J.T.; Groot, J.F.D.; Nghiemphu, P.L.; Kaley, T.J.; Colman, H.; Gaffey, S.C.; et al. Phase II study of pembrolizumab or pembrolizumab plus bevacizumab for recurrent glioblastoma (rGBM) patients. J. Clin. Oncol. 2018, 36, 2006. [Google Scholar] [CrossRef]
  149. Nayak, L.; Molinaro, A.M.; Peters, K.; Clarke, J.L.; Jordan, J.T.; de Groot, J.; Nghiemphu, L.; Kaley, T.; Colman, H.; McCluskey, C.; et al. Randomized Phase II and Biomarker Study of Pembrolizumab plus Bevacizumab versus Pembrolizumab Alone for Patients with Recurrent Glioblastoma. Clin. Cancer Res. 2021, 27, 1048–1057. [Google Scholar] [CrossRef]
  150. Iwamoto, F.; Tanguturi, S.; Desai, A.; Nayak, L.; Uhlmann, E.; Wang, T.; Lustig, R.; Hertan, L.; Bagley, S.; Hayden, J.; et al. CTIM-18. PHASE 2 STUDY OF PD-1 BLOCKADE WITH PEMBROLIZUMAB PLUS RE-IRRADIATION FOR RECURRENT GLIOBLASTOMA (RGBM) (NCT03661723). Neuro-Oncology 2022, 24 (Suppl. 7), vii63–vii64. [Google Scholar] [CrossRef]
  151. Tran, D.D.; Ghiaseddin, A.P.; Chen, D.D.; Le, S.B. Final analysis of 2-THE-TOP: A phase 2 study of TTFields (Optune) plus pembrolizumab plus maintenance temozolomide (TMZ) in patients with newly diagnosed glioblastoma. J. Clin. Oncol. 2023, 41 (Suppl. 16), 2024. [Google Scholar] [CrossRef]
  152. de Groot, J.; Penas-Prado, M.; Alfaro-Munoz, K.; Hunter, K.; Pei, B.L.; O’Brien, B.; Weathers, S.P.; Loghin, M.; Kamiya Matsouka, C.; Yung, W.K.A.; et al. Window-of-opportunity clinical trial of pembrolizumab in patients with recurrent glioblastoma reveals predominance of immune-suppressive macrophages. Neuro-Oncology 2020, 22, 539–549. [Google Scholar] [CrossRef] [PubMed]
  153. Webb, M.; Burns, T.C.; Twohy, E.; Sener, U.; Kizilbash, S.H.; Ruff, M.W.; Uhm, J.H.; Galanis, E.; D’Andre, S.D.; Riviere-Cazaux, C.; et al. Efficacy and safety study of neoadjuvant efineptakin alfa (NT-I7) and pembrolizumab in recurrent glioblastoma. J. Clin. Oncol. 2023, 41 (Suppl. 16), TPS2085. [Google Scholar] [CrossRef]
  154. Lwin, Z.; Gomez-Roca, C.; Saada-Bouzid, E.; Yanez, E.; Muñoz, F.L.; Im, S.A.; Castanon, E.; Senellart, H.; Graham, D.; Voss, M.; et al. LBA41 LEAP-005: Phase II study of lenvatinib (len) plus pembrolizumab (pembro) in patients (pts) with previously treated advanced solid tumours. Ann. Oncol. 2020, 31, S1170. [Google Scholar] [CrossRef]
  155. Jacques, F.H.; Nicholas, G.; Lorimer, I.A.J.; Sikati Foko, V.; Prevost, J.; Dumais, N.; Milne, K.; Nelson, B.H.; Woulfe, J.; Jansen, G.; et al. Avelumab in newly diagnosed glioblastoma. Neuro-Oncol. Adv. 2021, 3, vdab118. [Google Scholar] [CrossRef]
  156. Tiu, C.; Yau, W.H.; Welsh, L.C.; Jones, T.L.; Zachariou, A.; Prout, T.; Parmar, M.; Turner, A.J.; Daly, R.W.; Yap, C.; et al. Abstract CT093: Preliminary evidence of antitumor activity of Ipatasertib (Ipat) and Atezolizumab (A) in glioblastoma (GBM) patients (pts) with PTEN loss in the Phase 1 Ice-CAP trial (NCT03673787). Cancer Res. 2023, 83 (Suppl. 8), CT093. [Google Scholar] [CrossRef]
  157. Weathers, S.-P.S.; Kamiya-Matsuoka, C.; Harrison, R.A.; Liu, D.D.; Dervin, S.; Yun, C.; Loghin, M.E.; Penas-Prado, M.; Majd, N.; Yung, W.K.A.; et al. Phase I/II study to evaluate the safety and clinical efficacy of atezolizumab (atezo; aPDL1) in combination with temozolomide (TMZ) and radiation in patients with newly diagnosed glioblastoma (GBM). J. Clin. Oncol. 2020, 38 (Suppl. 15), 2511. [Google Scholar] [CrossRef]
  158. Reardon, D.A.; Kaley, T.J.; Dietrich, J.; Clarke, J.L.; Dunn, G.; Lim, M.; Cloughesy, T.F.; Gan, H.K.; Park, A.J.; Schwarzenberger, P.; et al. Phase II study to evaluate safety and efficacy of MEDI4736 (durvalumab) + radiotherapy in patients with newly diagnosed unmethylated MGMT glioblastoma (new unmeth GBM). J. Clin. Oncol. 2019, 37 (Suppl. 15), 2032. [Google Scholar] [CrossRef]
  159. Ranjan, S.; Quezado, M.; Garren, N.; Boris, L.; Siegel, C.; Lopes Abath Neto, O.; Theeler, B.J.; Park, D.M.; Nduom, E.; Zaghloul, K.A.; et al. Clinical decision making in the era of immunotherapy for high grade-glioma: Report of four cases. BMC Cancer 2018, 18, 239. [Google Scholar] [CrossRef]
  160. Duerinck, J.; Schwarze, J.K.; Awada, G.; Tijtgat, J.; Vaeyens, F.; Bertels, C.; Geens, W.; Klein, S.; Seynaeve, L.; Cras, L.; et al. Intracerebral administration of CTLA-4 and PD-1 immune checkpoint blocking monoclonal antibodies in patients with recurrent glioblastoma: A phase I clinical trial. J. Immunother. Cancer 2021, 9, e002296. [Google Scholar] [CrossRef] [PubMed]
  161. Reardon, D.A.; Brandes, A.A.; Omuro, A.; Mulholland, P.; Lim, M.; Wick, A.; Baehring, J.; Ahluwalia, M.S.; Roth, P.; Bahr, O.; et al. Effect of Nivolumab vs Bevacizumab in Patients with Recurrent Glioblastoma: The CheckMate 143 Phase 3 Randomized Clinical Trial. JAMA Oncol. 2020, 6, 1003–1010. [Google Scholar] [CrossRef]
  162. Omuro, A.; Reardon, D.A.; Sampson, J.H.; Baehring, J.; Sahebjam, S.; Cloughesy, T.F.; Chalamandaris, A.-G.; Potter, V.; Butowski, N.; Lim, M. Nivolumab plus radiotherapy with or without temozolomide in newly diagnosed glioblastoma: Results from exploratory phase I cohorts of CheckMate 143. Neuro-Oncol. Adv. 2022, 4, vdac025. [Google Scholar] [CrossRef] [PubMed]
  163. Bagley, S.J.; Mathew, D.; Desai, A.S.; Chen, K.; Long, Q.; Shabason, J.E.; Lustig, R.A.; Kurtz, G.; Alonso-Basanta, M.; Maloney, E.; et al. PD1 inhibition and GITR agonism in combination with fractionated stereotactic radiotherapy for the treatment of recurrent glioblastoma: A phase 2, multi-arm study. J. Clin. Oncol. 2023, 41 (Suppl. 16), 2004. [Google Scholar] [CrossRef]
  164. Lukas, R.; Sachdev, S.; Kumthekar, P.; Dixit, K.; Grimm, S.; Gondi, V.; Sharp, L.; Lezon, R.; James, D.; Lesniak, M.; et al. Ctim-12. a phase 1 trial of immunoradiotherapy with the ido enzyme inhibitor (bms-986205) and nivolumab in patients with newly diagnosed mgmt promoter unmethylated idhwt glioblastoma. Neuro-Oncology 2021, 23 (Suppl. 6), vi51–vi52. [Google Scholar] [CrossRef]
  165. Daud, A.; Saleh, M.N.; Hu, J.; Bleeker, J.S.; Riese, M.J.; Meier, R.; Zhou, L.; Serbest, G.; Lewis, K.D. Epacadostat plus nivolumab for advanced melanoma: Updated phase 2 results of the ECHO-204 study. J. Clin. Oncol. 2018, 36, 9511. [Google Scholar] [CrossRef]
  166. Lynes, J.; Jackson, S.; Sanchez, V.; Dominah, G.; Wang, X.; Kuek, A.; Hayes, C.P.; Benzo, S.; Scott, G.C.; Chittiboina, P.; et al. Cytokine Microdialysis for Real-Time Immune Monitoring in Glioblastoma Patients Undergoing Checkpoint Blockade. Neurosurgery 2019, 84, 945–953. [Google Scholar] [CrossRef]
  167. Lim, M.; Ye, X.; Piotrowski, A.F.; Desai, A.S.; Ahluwalia, M.S.; Walbert, T.; Fisher, J.D.; Desideri, S.; Nabors, L.B.; Wen, P.Y.; et al. Updated safety phase I trial of anti-LAG-3 alone and in combination with anti-PD-1 in patients with recurrent GBM. J. Clin. Oncol. 2020, 38 (Suppl. 15), 2512. [Google Scholar] [CrossRef]
  168. Zeng, J.; See, A.P.; Phallen, J.; Jackson, C.M.; Belcaid, Z.; Ruzevick, J.; Durham, N.; Meyer, C.; Harris, T.J.; Albesiano, E.; et al. Anti-PD-1 blockade and stereotactic radiation produce long-term survival in mice with intracranial gliomas. Int. J. Radiat. Oncol. Biol. Phys. 2013, 86, 343–349. [Google Scholar] [CrossRef] [PubMed]
  169. Antonios, J.P.; Soto, H.; Everson, R.G.; Orpilla, J.; Moughon, D.; Shin, N.; Sedighim, S.; Yong, W.H.; Li, G.; Cloughesy, T.F.; et al. PD-1 blockade enhances the vaccination-induced immune response in glioma. JCI Insight 2016, 1, e87059. [Google Scholar] [CrossRef] [PubMed]
  170. Weller, M.; Lim, M.; Idbaih, A.; Steinbach, J.; Finocchiaro, G.; Raval, R.; Ashby, L.; Ansstas, G.; Baehring, J.; Taylor, J.; et al. CTIM-25. A randomized phase 3 study of nivolumab or placebo combined with radiotherapy plus temozolomide in patients with newly diagnosed glioblastoma with methylated mgmt promoter: Checkmate 548. Neuro-Oncology 2021, 23 (Suppl. 6), vi55–vi56. [Google Scholar] [CrossRef]
  171. Lombardi, G.; Barresi, V.; Indraccolo, S.; Simbolo, M.; Fassan, M.; Mandruzzato, S.; Simonelli, M.; Caccese, M.; Pizzi, M.; Fassina, A.; et al. Pembrolizumab Activity in Recurrent High-Grade Gliomas with Partial or Complete Loss of Mismatch Repair Protein Expression: A Monocentric, Observational and Prospective Pilot Study. Cancers 2020, 12, 2283. [Google Scholar] [CrossRef]
  172. Cloughesy, T.F.; Mochizuki, A.Y.; Orpilla, J.R.; Hugo, W.; Lee, A.H.; Davidson, T.B.; Wang, A.C.; Ellingson, B.M.; Rytlewski, J.A.; Sanders, C.M.; et al. Neoadjuvant anti-PD-1 immunotherapy promotes a survival benefit with intratumoral and systemic immune responses in recurrent glioblastoma. Nat. Med. 2019, 25, 477–486. [Google Scholar] [CrossRef] [PubMed]
  173. Berghoff, A.S.; Kiesel, B.; Widhalm, G.; Rajky, O.; Ricken, G.; Wohrer, A.; Dieckmann, K.; Filipits, M.; Brandstetter, A.; Weller, M.; et al. Programmed death ligand 1 expression and tumor-infiltrating lymphocytes in glioblastoma. Neuro-Oncology 2015, 17, 1064–1075. [Google Scholar] [CrossRef] [PubMed]
  174. Koyama, S.; Akbay, E.A.; Li, Y.Y.; Herter-Sprie, G.S.; Buczkowski, K.A.; Richards, W.G.; Gandhi, L.; Redig, A.J.; Rodig, S.J.; Asahina, H.; et al. Adaptive resistance to therapeutic PD-1 blockade is associated with upregulation of alternative immune checkpoints. Nat. Commun. 2016, 7, 10501. [Google Scholar] [CrossRef] [PubMed]
  175. Carter, T.; Shaw, H.; Cohn-Brown, D.; Chester, K.; Mulholland, P. Ipilimumab and Bevacizumab in Glioblastoma. Clin. Oncol. 2016, 28, 622–626. [Google Scholar] [CrossRef] [PubMed]
  176. Wolchok, J.D.; Kluger, H.; Callahan, M.K.; Postow, M.A.; Rizvi, N.A.; Lesokhin, A.M.; Segal, N.H.; Ariyan, C.E.; Gordon, R.A.; Reed, K.; et al. Nivolumab plus ipilimumab in advanced melanoma. N. Engl. J. Med. 2013, 369, 122–133. [Google Scholar] [CrossRef]
  177. Larkin, J.; Chiarion-Sileni, V.; Gonzalez, R.; Grob, J.J.; Cowey, C.L.; Lao, C.D.; Schadendorf, D.; Dummer, R.; Smylie, M.; Rutkowski, P.; et al. Combined Nivolumab and Ipilimumab or Monotherapy in Untreated Melanoma. N. Engl. J. Med. 2015, 373, 23–34. [Google Scholar] [CrossRef] [PubMed]
  178. Omuro, A. Immune-checkpoint inhibitors for glioblastoma: What have we learned? Arq. Neuro-Psiquiatr. 2022, 80 (Suppl. 1), 266–269. [Google Scholar] [CrossRef]
  179. Lim, M.; Ye, X.; Piotrowski, A.F.; Desai, A.S.; Ahluwalia, M.S.; Walbert, T.; Fisher, J.D.; Desideri, S.; Belcaid, Z.; Jackson, C.; et al. Updated phase I trial of anti-LAG-3 or anti-CD137 alone and in combination with anti-PD-1 in patients with recurrent GBM. J. Clin. Oncol. 2019, 37, 2017. [Google Scholar] [CrossRef]
  180. Arrieta, V.A.; Dmello, C.; McGrail, D.J.; Brat, D.J.; Lee-Chang, C.; Heimberger, A.B.; Chand, D.; Stupp, R.; Sonabend, A.M. Immune checkpoint blockade in glioblastoma: From tumor heterogeneity to personalized treatment. J. Clin. Investig. 2023, 133, e163447. [Google Scholar] [CrossRef] [PubMed]
  181. Hodges, T.R.; Ott, M.; Xiu, J.; Gatalica, Z.; Swensen, J.; Zhou, S.; Huse, J.T.; de Groot, J.; Li, S.; Overwijk, W.W.; et al. Mutational burden, immune checkpoint expression, and mismatch repair in glioma: Implications for immune checkpoint immunotherapy. Neuro-Oncology 2017, 19, 1047–1057. [Google Scholar] [CrossRef]
  182. Leuthardt, E.C.; Duan, C.; Kim, M.J.; Campian, J.L.; Kim, A.H.; Miller-Thomas, M.M.; Shimony, J.S.; Tran, D.D. Hyperthermic Laser Ablation of Recurrent Glioblastoma Leads to Temporary Disruption of the Peritumoral Blood Brain Barrier. PLoS ONE 2016, 11, e0148613. [Google Scholar] [CrossRef]
  183. Salehi, A.; Paturu, M.R.; Patel, B.; Cain, M.D.; Mahlokozera, T.; Yang, A.B.; Lin, T.H.; Leuthardt, E.C.; Yano, H.; Song, S.K.; et al. Therapeutic enhancement of blood-brain and blood-tumor barriers permeability by laser interstitial thermal therapy. Neuro-Oncol. Adv. 2020, 2, vdaa071. [Google Scholar] [CrossRef]
  184. Frederico, S.C.; Hancock, J.C.; Brettschneider, E.E.S.; Ratnam, N.M.; Gilbert, M.R.; Terabe, M. Making a Cold Tumor Hot: The Role of Vaccines in the Treatment of Glioblastoma. Front. Oncol. 2021, 11, 672508. [Google Scholar] [CrossRef]
  185. Khong, H.; Overwijk, W.W. Adjuvants for peptide-based cancer vaccines. J. Immunother. Cancer 2016, 4, 56. [Google Scholar] [CrossRef]
  186. Rudnick, J.D.; Fink, K.L.; Landolfi, J.C.; Markert, J.; Piccioni, D.E.; Glantz, M.J.; Swanson, S.J.; Gringeri, A.; Yu, J. Immunological targeting of CD133 in recurrent glioblastoma: A multi-center phase I translational and clinical study of autologous CD133 dendritic cell immunotherapy. J. Clin. Oncol. 2017, 35, 2059. [Google Scholar] [CrossRef]
  187. Reap, E.A.; Suryadevara, C.M.; Batich, K.A.; Sanchez-Perez, L.; Archer, G.E.; Schmittling, R.J.; Norberg, P.K.; Herndon, J.E., 2nd; Healy, P.; Congdon, K.L.; et al. Dendritic Cells Enhance Polyfunctionality of Adoptively Transferred T Cells That Target Cytomegalovirus in Glioblastoma. Cancer Res. 2018, 78, 256–264. [Google Scholar] [CrossRef]
  188. Vlahovic, G.; Archer, G.E.; Reap, E.; Desjardins, A.; Peters, K.B.; Randazzo, D.; Healy, P.; Herndon, J.E.; Friedman, A.H.; Friedman, H.S.; et al. Phase I trial of combination of antitumor immunotherapy targeted against cytomegalovirus (CMV) plus regulatory T-cell inhibition in patients with newly-diagnosed glioblastoma multiforme (GBM). J. Clin. Oncol. 2016, 34 (Suppl. 15), e13518. [Google Scholar] [CrossRef]
  189. Mitchell, D.A.; Batich, K.A.; Gunn, M.D.; Huang, M.N.; Sanchez-Perez, L.; Nair, S.K.; Congdon, K.L.; Reap, E.A.; Archer, G.E.; Desjardins, A.; et al. Tetanus toxoid and CCL3 improve dendritic cell vaccines in mice and glioblastoma patients. Nature 2015, 519, 366–369. [Google Scholar] [CrossRef] [PubMed]
  190. Batich, K.A.; Reap, E.A.; Archer, G.E.; Sanchez-Perez, L.; Nair, S.K.; Schmittling, R.J.; Norberg, P.; Xie, W.; Herndon, J.E., 2nd; Healy, P.; et al. Long-term Survival in Glioblastoma with Cytomegalovirus pp65-Targeted Vaccination. Clin. Cancer Res. 2017, 23, 1898–1909. [Google Scholar] [CrossRef] [PubMed]
  191. Rahman, M.; Ghiaseddin, A.; Deleyrolle, P.; Peters, K.B.; Archer, G.; Sampson, J.; Mitchell, D. CTIM-07 – Phase II randomized, blinded, placebo-controlled trial testing pp65 CMV mRNA dendritic cell vaccine and tetanus-diphtheria toxoid for newly diagnosed GBM (ATTAC II, NCT02465268). Neuro-Oncology 2022, 24 (Suppl. 7), vii60–vii61. [Google Scholar] [CrossRef]
  192. Batich, K.A.; Mitchell, D.A.; Healy, P.; Herndon, J.E., 2nd; Sampson, J.H. Once, Twice, Three Times a Finding: Reproducibility of Dendritic Cell Vaccine Trials Targeting Cytomegalovirus in Glioblastoma. Clin. Cancer Res. 2020, 26, 5297–5303. [Google Scholar] [CrossRef]
  193. Sampson, J.H.; Batich, K.A.; Mitchell, D.A.; Herndon, J.E.; Broadwater, G.; Healy, P.; Sanchez-Perez, L.; Nair, S.; Congdon, K.; Norberg, P.; et al. Reproducibility of outcomes in sequential trials using CMV-targeted dendritic cell vaccination for glioblastoma. J. Clin. Oncol. 2022, 40 (Suppl. 16), 2005. [Google Scholar] [CrossRef]
  194. Sampson, J.H.; Schmittling, R.J.; Archer, G.E.; Congdon, K.L.; Nair, S.K.; Reap, E.A.; Desjardins, A.; Friedman, A.H.; Friedman, H.S.; Herndon, J.E., 2nd; et al. A pilot study of IL-2Rα blockade during lymphopenia depletes regulatory T-cells and correlates with enhanced immunity in patients with glioblastoma. PLoS ONE 2012, 7, e31046. [Google Scholar] [CrossRef]
  195. Reardon, D.A.; Desjardins, A.; Vredenburgh, J.J.; O’Rourke, D.M.; Tran, D.D.; Fink, K.L.; Nabors, L.B.; Li, G.; Bota, D.A.; Lukas, R.V.; et al. Rindopepimut with Bevacizumab for Patients with Relapsed EGFRvIII-Expressing Glioblastoma (ReACT): Results of a Double-Blind Randomized Phase II Trial. Clin. Cancer Res. 2020, 26, 1586–1594. [Google Scholar] [CrossRef]
  196. Schuster, J.; Lai, R.K.; Recht, L.D.; Reardon, D.A.; Paleologos, N.A.; Groves, M.D.; Mrugala, M.M.; Jensen, R.; Baehring, J.M.; Sloan, A.; et al. A phase II, multicenter trial of rindopepimut (CDX-110) in newly diagnosed glioblastoma: The ACT III study. Neuro-Oncology 2015, 17, 854–861. [Google Scholar] [CrossRef]
  197. Sampson, J.H.; Heimberger, A.B.; Archer, G.E.; Aldape, K.D.; Friedman, A.H.; Friedman, H.S.; Gilbert, M.R.; Herndon, J.E., 2nd; McLendon, R.E.; Mitchell, D.A.; 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. [Google Scholar] [CrossRef]
  198. Sampson, J.H.; Aldape, K.D.; Archer, G.E.; Coan, A.; Desjardins, A.; Friedman, A.H.; Friedman, H.S.; Gilbert, M.R.; Herndon, J.E.; McLendon, R.E.; et al. Greater chemotherapy-induced lymphopenia enhances tumor-specific immune responses that eliminate EGFRvIII-expressing tumor cells in patients with glioblastoma. Neuro-Oncology 2011, 13, 324–333. [Google Scholar] [CrossRef]
  199. Schmittling, R.J.; Archer, G.E.; Mitchell, D.A.; Heimberger, A.; Pegram, C.; Herndon, J.E., 2nd; Friedman, H.S.; Bigner, D.D.; Sampson, J.H. Detection of humoral response in patients with glioblastoma receiving EGFRvIII-KLH vaccines. J. Immunol. Methods 2008, 339, 74–81. [Google Scholar] [CrossRef]
  200. Weller, M.; Butowski, N.; Tran, D.D.; Recht, L.D.; Lim, M.; Hirte, H.; Ashby, L.; Mechtler, L.; Goldlust, S.A.; Iwamoto, F.; et al. Rindopepimut with temozolomide for patients with newly diagnosed, EGFRvIII-expressing glioblastoma (ACT IV): A randomised, double-blind, international phase 3 trial. Lancet Oncol. 2017, 18, 1373–1385. [Google Scholar] [CrossRef]
  201. Crane, C.A.; Han, S.J.; Ahn, B.; Oehlke, J.; Kivett, V.; Fedoroff, A.; Butowski, N.; Chang, S.M.; Clarke, J.; Berger, M.S.; et al. Individual patient-specific immunity against high-grade glioma after vaccination with autologous tumor derived peptides bound to the 96 KD chaperone protein. Clin. Cancer Res. 2013, 19, 205–214. [Google Scholar] [CrossRef]
  202. Bloch, O.; Crane, C.A.; Fuks, Y.; Kaur, R.; Aghi, M.K.; Berger, M.S.; Butowski, N.A.; Chang, S.M.; Clarke, J.L.; McDermott, M.W.; et al. Heat-shock protein peptide complex-96 vaccination for recurrent glioblastoma: A phase II, single-arm trial. Neuro-Oncology 2014, 16, 274–279. [Google Scholar] [CrossRef]
  203. Bloch, O.; Lim, M.; Sughrue, M.E.; Komotar, R.J.; Abrahams, J.M.; O’Rourke, D.M.; D’Ambrosio, A.; Bruce, J.N.; Parsa, A.T. Autologous Heat Shock Protein Peptide Vaccination for Newly Diagnosed Glioblastoma: Impact of Peripheral PD-L1 Expression on Response to Therapy. Clin. Cancer Res. 2017, 23, 3575–3584. [Google Scholar] [CrossRef]
  204. Bloch, O.; Shi, Q.; Anderson, S.K.; Knopp, M.; Raizer, J.; Clarke, J.; Waziri, A.; Colman, H.; Bruce, J.; Olson, J.J.; et al. ATIM-14. Alliance A071101: A phase II randomized trial comparing the efficacy of heat shock protein peptide complex-96 (HSPPC-96) vaccine given with bevacizumab versus bevacizumab alone in the treatment of surgically resectable recurrent glioblastoma. Neuro-Oncology 2017, 19 (Suppl. 6), vi29. [Google Scholar] [CrossRef]
  205. Carpentier, A.F.; Verlut, C.; Ghiringhelli, F.; Bronnimann, C.; Ursu, R.; Fumet, J.D.; Gherga, E.; Lefort, F.; Belin, C.; Vernerey, D.; et al. Anti-telomerase vaccine in patients with newly diagnosed, unmethylated MGMT glioblastoma: A phase II study. J. Clin. Oncol. 2023, 41 (Suppl. 16), 2005. [Google Scholar] [CrossRef]
  206. Fenstermaker, R.A.; Ciesielski, M.J.; Qiu, J.; Yang, N.; Frank, C.L.; Lee, K.P.; Mechtler, L.R.; Belal, A.; Ahluwalia, M.S.; Hutson, A.D. Clinical study of a survivin long peptide vaccine (SurVaxM) in patients with recurrent malignant glioma. Cancer Immunol. Immunother. 2016, 65, 1339–1352. [Google Scholar] [CrossRef]
  207. Ahluwalia, M.S.; Ciesielski, M.; Abad, A.; Reardon, D.; Aiken, R.; Barbaro, M.; Sinicrope, K.; Peereboom, D.M.; Odia, Y.; Brenner, A.; et al. P07.09.B A randomized phase 2B study of survivin vaccine survaxm plus adjuvant temozolomide for newly-diagnosed glioblastoma (survive). Neuro-Oncology 2023, 25 (Suppl. 2), ii52–ii53. [Google Scholar] [CrossRef]
  208. Ahluwalia, M.S.; Reardon, D.A.; Abad, A.P.; Curry, W.T.; Wong, E.T.; Figel, S.A.; Mechtler, L.L.; Peereboom, D.M.; Hutson, A.D.; Withers, H.G.; et al. Phase IIa Study of SurVaxM Plus Adjuvant Temozolomide for Newly Diagnosed Glioblastoma. J. Clin. Oncol. 2023, 41, 1453–1465. [Google Scholar] [CrossRef]
  209. Wen, P.Y.; Reardon, D.A.; Armstrong, T.S.; Phuphanich, S.; Aiken, R.D.; Landolfi, J.C.; Curry, W.T.; Zhu, J.J.; Glantz, M.; Peereboom, D.M.; et al. A Randomized Double-Blind Placebo-Controlled Phase II Trial of Dendritic Cell Vaccine ICT-107 in Newly Diagnosed Patients with Glioblastoma. Clin. Cancer Res. 2019, 25, 5799–5807. [Google Scholar] [CrossRef]
  210. Rampling, R.; Peoples, S.; Mulholland, P.J.; James, A.; Al-Salihi, O.; Twelves, C.J.; McBain, C.; Jefferies, S.; Jackson, A.; Stewart, W.; et al. A Cancer Research UK First Time in Human Phase I Trial of IMA950 (Novel Multipeptide Therapeutic Vaccine) in Patients with Newly Diagnosed Glioblastoma. Clin. Cancer Res. 2016, 22, 4776–4785. [Google Scholar] [CrossRef]
  211. Migliorini, D.; Dutoit, V.; Allard, M.; Grandjean Hallez, N.; Marinari, E.; Widmer, V.; Philippin, G.; Corlazzoli, F.; Gustave, R.; Kreutzfeldt, M.; et al. Phase I/II trial testing safety and immunogenicity of the multipeptide IMA950/poly-ICLC vaccine in newly diagnosed adult malignant astrocytoma patients. Neuro-Oncology 2019, 21, 923–933. [Google Scholar] [CrossRef]
  212. Boydell, E.; Marinari, E.; Migliorini, D.; Dietrich, P.Y.; Patrikidou, A.; Dutoit, V. Exploratory Study of the Effect of IMA950/Poly-ICLC Vaccination on Response to Bevacizumab in Relapsing High-Grade Glioma Patients. Cancers 2019, 11, 464. [Google Scholar] [CrossRef]
  213. Reardon, D.A.; Brem, S.; Desai, A.S.; Bagley, S.J.; Kurz, S.C.; Fuente, M.I.D.L.; Nagpal, S.; Welch, M.R.; Hormigo, A.; Forsyth, P.A.J.; et al. Intramuscular (IM) INO-5401 + INO-9012 with electroporation (EP) in combination with cemiplimab (REGN2810) in newly diagnosed glioblastoma. J. Clin. Oncol. 2022, 40 (Suppl. 16), 2004. [Google Scholar] [CrossRef]
  214. Olin, M.R.; Low, W.; McKenna, D.H.; Haines, S.J.; Dahlheimer, T.; Nascene, D.; Gustafson, M.P.; Dietz, A.B.; Clark, H.B.; Chen, W.; et al. Vaccination with dendritic cells loaded with allogeneic brain tumor cells for recurrent malignant brain tumors induces a CD4+IL17+ response. J. ImmunoTherapy Cancer 2014, 2, 4. [Google Scholar] [CrossRef]
  215. Prins, R.M.; Wang, X.; Soto, H.; Young, E.; Lisiero, D.N.; Fong, B.; Everson, R.; Yong, W.H.; Lai, A.; Li, G.; et al. Comparison of glioma-associated antigen peptide-loaded versus autologous tumor lysate-loaded dendritic cell vaccination in malignant glioma patients. J. Immunother. 2013, 36, 152–157. [Google Scholar] [CrossRef]
  216. Hu, J.L.; Omofoye, O.A.; Rudnick, J.D.; Kim, S.; Tighiouart, M.; Phuphanich, S.; Wang, H.; Mazer, M.; Ganaway, T.; Chu, R.M.; et al. A Phase I Study of Autologous Dendritic Cell Vaccine Pulsed with Allogeneic Stem-like Cell Line Lysate in Patients with Newly Diagnosed or Recurrent Glioblastoma. Clin. Cancer Res. 2022, 28, 689–696. [Google Scholar] [CrossRef]
  217. Erhart, F.; Buchroithner, J.; Reitermaier, R.; Fischhuber, K.; Klingenbrunner, S.; Sloma, I.; Hibsh, D.; Kozol, R.; Efroni, S.; Ricken, G.; et al. Immunological analysis of phase II glioblastoma dendritic cell vaccine (Audencel) trial: Immune system characteristics influence outcome and Audencel up-regulates Th1-related immunovariables. Acta Neuropathol. Commun. 2018, 6, 135. [Google Scholar] [CrossRef]
  218. Inogés, S.; Tejada, S.; de Cerio, A.L.; Gállego Pérez-Larraya, J.; Espinós, J.; Idoate, M.A.; Domínguez, P.D.; de Eulate, R.G.; Aristu, J.; Bendandi, M.; et al. A phase II trial of autologous dendritic cell vaccination and radiochemotherapy following fluorescence-guided surgery in newly diagnosed glioblastoma patients. J. Transl. Med. 2017, 15, 104. [Google Scholar] [CrossRef]
  219. Fadul, C.E.; Fisher, J.L.; Hampton, T.H.; Lallana, E.C.; Li, Z.; Gui, J.; Szczepiorkowski, Z.M.; Tosteson, T.D.; Rhodes, C.H.; Wishart, H.A.; et al. Immune response in patients with newly diagnosed glioblastoma multiforme treated with intranodal autologous tumor lysate-dendritic cell vaccination after radiation chemotherapy. J. Immunother. 2011, 34, 382–389. [Google Scholar] [CrossRef]
  220. Liau, L.M.; Ashkan, K.; Tran, D.D.; Campian, J.L.; Trusheim, J.E.; Cobbs, C.S.; Heth, J.A.; Salacz, M.; Taylor, S.; D’Andre, S.D.; et al. First results on survival from a large Phase 3 clinical trial of an autologous dendritic cell vaccine in newly diagnosed glioblastoma. J. Transl. Med. 2018, 16, 142. [Google Scholar] [CrossRef]
  221. Liau, L.M.; Ashkan, K.; Brem, S.; Campian, J.L.; Trusheim, J.E.; Iwamoto, F.M.; Tran, D.D.; Ansstas, G.; Cobbs, C.S.; Heth, J.A.; et al. Association of Autologous Tumor Lysate-Loaded Dendritic Cell Vaccination with Extension of Survival Among Patients with Newly Diagnosed and Recurrent Glioblastoma: A Phase 3 Prospective Externally Controlled Cohort Trial. JAMA Oncol. 2023, 9, 112–121. [Google Scholar] [CrossRef]
  222. Wick, W.; Dietrich, P.-Y.; Kuttruff, S.; Hilf, N.; Frenzel, K.; Admon, A.; Burg, S.H.v.d.; Deimling, A.v.; Gouttefangeas, C.; Kroep, J.R.; et al. GAPVAC-101: First-in-human trial of a highly personalized peptide vaccination approach for patients with newly diagnosed glioblastoma. J. Clin. Oncol. 2018, 36 (Suppl. 15), 2000. [Google Scholar] [CrossRef]
  223. Hilf, N.; Kuttruff-Coqui, S.; Frenzel, K.; Bukur, V.; Stevanović, S.; Gouttefangeas, C.; Platten, M.; Tabatabai, G.; Dutoit, V.; van der Burg, S.H.; et al. Actively personalized vaccination trial for newly diagnosed glioblastoma. Nature 2019, 565, 240–245. [Google Scholar] [CrossRef]
  224. Kodysh, J.; Rubinsteyn, A.; Blazquez, A.; Mandeli, J.; Bhardwaj, N.; Hormigo, A. CTIM-17. phase I study of the safety and immunogenicity of personalized neoantigen vaccines and tumor treating fields in patients with newly diagnosed glioblastoma. Neuro-Oncology 2020, 22 (Suppl. 2), ii36. [Google Scholar]
  225. Narita, Y.; Arakawa, Y.; Yamasaki, F.; Nishikawa, R.; Aoki, T.; Kanamori, M.; Nagane, M.; Kumabe, T.; Hirose, Y.; Ichikawa, T.; et al. A randomized, double-blind, phase III trial of personalized peptide vaccination for recurrent glioblastoma. Neuro-Oncology 2019, 21, 348–359. [Google Scholar] [CrossRef]
  226. Dunn-Pirio, A.; Peters, K.; DesJardins, A.; Randazzo, D.; Friedman, H.; Healy, P.; II, J.H.; Reap, E.; Archer, G.; Li, Q.-J.; et al. Tumor stem cell RNA-loaded dendritic cell vaccine for recurrent glioblastoma: A phase 1 trial (S41.004). Neurology 2017, 88, S41.004. [Google Scholar] [CrossRef]
  227. Vik-Mo, E.O.; Nyakas, M.; Mikkelsen, B.V.; Moe, M.C.; Due-Tønnesen, P.; Suso, E.M.; Sæbøe-Larssen, S.; Sandberg, C.; Brinchmann, J.E.; Helseth, E.; et al. Therapeutic vaccination against autologous cancer stem cells with mRNA-transfected dendritic cells in patients with glioblastoma. Cancer Immunol. Immunother. 2013, 62, 1499–1509. [Google Scholar] [CrossRef]
  228. Jouanneau, E.; Black, K.L.; Veiga, L.; Cordner, R.; Goverdhana, S.; Zhai, Y.; Zhang, X.X.; Panwar, A.; Mardiros, A.; Wang, H.; et al. Intrinsically de-sialylated CD103(+) CD8 T cells mediate beneficial anti-glioma immune responses. Cancer Immunol. Immunother. 2014, 63, 911–924. [Google Scholar] [CrossRef] [PubMed]
  229. Lopes, A.; Vandermeulen, G.; Préat, V. Cancer DNA vaccines: Current preclinical and clinical developments and future perspectives. J. Exp. Clin. Cancer Res. 2019, 38, 146. [Google Scholar] [CrossRef]
  230. Melnick, K.; Dastmalchi, F.; Mitchell, D.; Rahman, M.; Sayour, E.J. Contemporary RNA Therapeutics for Glioblastoma. Neuromolecular Med. 2022, 24, 8–12. [Google Scholar] [CrossRef]
  231. Herrada, A.A.; Rojas-Colonelli, N.; González-Figueroa, P.; Roco, J.; Oyarce, C.; Ligtenberg, M.A.; Lladser, A. Harnessing DNA-induced immune responses for improving cancer vaccines. Hum. Vaccin. Immunother. 2012, 8, 1682–1693. [Google Scholar] [CrossRef]
  232. Reardon, D.A.; Brem, S.; Desai, A.S.; Bagley, S.J.; Kurz, S.C.; Fuente, M.I.D.L.; Nagpal, S.; Welch, M.R.; Hormigo, A.; Carroll, N.; et al. INO-5401 and INO-9012 delivered intramuscularly (IM) with electroporation (EP) in combination with cemiplimab (REGN2810) in newly diagnosed glioblastoma (GBM): Interim results. J. Clin. Oncol. 2020, 38 (Suppl. 15), 2514. [Google Scholar] [CrossRef]
  233. Huang, B.; Li, X.; Li, Y.; Zhang, J.; Zong, Z.; Zhang, H. Current Immunotherapies for Glioblastoma Multiforme. Front. Immunol. 2020, 11, 603911. [Google Scholar] [CrossRef]
  234. Uematsu, M.; Ohsawa, I.; Aokage, T.; Nishimaki, K.; Matsumoto, K.; Takahashi, H.; Asoh, S.; Teramoto, A.; Ohta, S. Prognostic significance of the immunohistochemical index of survivin in glioma: A comparative study with the MIB-1 index. J. Neuro-Oncol. 2005, 72, 231–238. [Google Scholar] [CrossRef]
  235. Caudill, M.M.; Li, Z. HSPPC-96: A personalised cancer vaccine. Expert. Opin. Biol. Ther. 2001, 1, 539–547. [Google Scholar] [CrossRef]
  236. Amato, R.J. Heat shock protein-peptide complex-96 (Vitespen) for the treatment of cancer. Oncol. Rev. 2008, 2, 29–35. [Google Scholar] [CrossRef]
  237. Keskin, D.B.; Anandappa, A.J.; Sun, J.; Tirosh, I.; Mathewson, N.D.; Li, S.; Oliveira, G.; Giobbie-Hurder, A.; Felt, K.; Gjini, E.; et al. Neoantigen vaccine generates intratumoral T cell responses in phase Ib glioblastoma trial. Nature 2019, 565, 234–239. [Google Scholar] [CrossRef] [PubMed]
  238. Filley, A.C.; Dey, M. Dendritic cell based vaccination strategy: An evolving paradigm. J. Neuro-Oncol. 2017, 133, 223–235. [Google Scholar] [CrossRef] [PubMed]
  239. Lynes, J.; Sanchez, V.; Dominah, G.; Nwankwo, A.; Nduom, E. Current Options and Future Directions in Immune Therapy for Glioblastoma. Front. Oncol. 2018, 8, 578. [Google Scholar] [CrossRef] [PubMed]
  240. Reinhard, G.; Märten, A.; Kiske, S.M.; Feil, F.; Bieber, T.; Schmidt-Wolf, I.G. Generation of dendritic cell-based vaccines for cancer therapy. Br. J. Cancer 2002, 86, 1529–1533. [Google Scholar] [CrossRef] [PubMed]
  241. De Vries, I.J.; Krooshoop, D.J.; Scharenborg, N.M.; Lesterhuis, W.J.; Diepstra, J.H.; Van Muijen, G.N.; Strijk, S.P.; Ruers, T.J.; Boerman, O.C.; Oyen, W.J.; et al. Effective migration of antigen-pulsed dendritic cells to lymph nodes in melanoma patients is determined by their maturation state. Cancer Res. 2003, 63, 12–17. [Google Scholar] [PubMed]
  242. Mitchell, D.A.; Xie, W.; Schmittling, R.; Learn, C.; Friedman, A.; McLendon, R.E.; Sampson, J.H. Sensitive detection of human cytomegalovirus in tumors and peripheral blood of patients diagnosed with glioblastoma. Neuro-Oncology 2008, 10, 10–18. [Google Scholar] [CrossRef]
  243. Nair, S.K.; De Leon, G.; Boczkowski, D.; Schmittling, R.; Xie, W.; Staats, J.; Liu, R.; Johnson, L.A.; Weinhold, K.; Archer, G.E.; et al. Recognition and killing of autologous, primary glioblastoma tumor cells by human cytomegalovirus pp65-specific cytotoxic T cells. Clin. Cancer Res. 2014, 20, 2684–2694. [Google Scholar] [CrossRef] [PubMed]
  244. Liu, G.; Yuan, X.; Zeng, Z.; Tunici, P.; Ng, H.; Abdulkadir, I.R.; Lu, L.; Irvin, D.; Black, K.L.; Yu, J.S. Analysis of gene expression and chemoresistance of CD133+ cancer stem cells in glioblastoma. Mol. Cancer 2006, 5, 67. [Google Scholar] [CrossRef]
  245. Saikali, S.; Avril, T.; Collet, B.; Hamlat, A.; Bansard, J.Y.; Drenou, B.; Guegan, Y.; Quillien, V. Expression of nine tumour antigens in a series of human glioblastoma multiforme: Interest of EGFRvIII, IL-13Ralpha2, gp100 and TRP-2 for immunotherapy. J. Neuro-Oncol. 2007, 81, 139–148. [Google Scholar] [CrossRef]
  246. Pasqualetti, F.; Zanotti, S. Nonrandomised controlled trial in recurrent glioblastoma patients: The promise of autologous tumour lysate-loaded dendritic cell vaccination. Br. J. Cancer 2023, 129, 895–896. [Google Scholar] [CrossRef]
  247. Preusser, M.; van den Bent, M.J. Autologous tumor lysate-loaded dendritic cell vaccination (DCVax-L) in glioblastoma: Breakthrough or fata morgana? Neuro-Oncology 2023, 25, 631–634. [Google Scholar] [CrossRef] [PubMed]
  248. Gatto, L.; Di Nunno, V.; Tosoni, A.; Bartolini, S.; Ranieri, L.; Franceschi, E. DCVax-L Vaccination in Patients with Glioblastoma: Real Promise or Negative Trial? The Debate Is Open. Cancers 2023, 15, 3251. [Google Scholar] [CrossRef] [PubMed]
  249. Rahman, R.; Ventz, S.; Trippa, L. External Control Arms and Data Analysis Methods in Nonrandomized Trial of Patients with Glioblastoma. JAMA Oncol. 2023, 9, 1006–1007. [Google Scholar] [CrossRef] [PubMed]
  250. Mandel, J.J.; de Groot, J.F. External Control Arms and Data Analysis Methods in Nonrandomized Trial of Patients with Glioblastoma. JAMA Oncol. 2023, 9, 1006. [Google Scholar] [CrossRef] [PubMed]
  251. Van Gool, S.W.; Makalowski, J.; Kampers, L.F.C.; Van de Vliet, P.; Sprenger, T.; Schirrmacher, V.; Stucker, W. Dendritic cell vaccination for glioblastoma multiforme patients: Has a new milestone been reached? Transl. Cancer Res. 2023, 12, 2224–2228. [Google Scholar] [CrossRef] [PubMed]
  252. Wang, J.; Shen, F.; Yao, Y.; Wang, L.L.; Zhu, Y.; Hu, J. Adoptive Cell Therapy: A Novel and Potential Immunotherapy for Glioblastoma. Front. Oncol. 2020, 10, 59. [Google Scholar] [CrossRef] [PubMed]
  253. O’Rourke, D.M.; Nasrallah, M.P.; Desai, A.; Melenhorst, J.J.; Mansfield, K.; Morrissette, J.J.D.; Martinez-Lage, M.; Brem, S.; Maloney, E.; Shen, A.; et al. A single dose of peripherally infused EGFRvIII-directed CAR T cells mediates antigen loss and induces adaptive resistance in patients with recurrent glioblastoma. Sci. Transl. Med. 2017, 9, eaaa0984. [Google Scholar] [CrossRef] [PubMed]
  254. Durgin, J.S.; Henderson, F., Jr.; Nasrallah, M.P.; Mohan, S.; Wang, S.; Lacey, S.F.; Melenhorst, J.J.; Desai, A.S.; Lee, J.Y.K.; Maus, M.V.; et al. Case Report: Prolonged Survival Following EGFRvIII CAR T Cell Treatment for Recurrent Glioblastoma. Front. Oncol. 2021, 11, 669071. [Google Scholar] [CrossRef]
  255. Tang, O.Y.; Tian, L.; Yoder, T.; Xu, R.; Kulikovskaya, I.; Gupta, M.; Melenhorst, J.J.; Lacey, S.F.; O’Rourke, D.M.; Binder, Z.A. PD1 Expression in EGFRvIII-Directed CAR T Cell Infusion Product for Glioblastoma Is Associated with Clinical Response. Front. Immunol. 2022, 13, 872756. [Google Scholar] [CrossRef]
  256. Suryadevara, C.M.; Desai, R.; Abel, M.L.; Riccione, K.A.; Batich, K.A.; Shen, S.H.; Chongsathidkiet, P.; Gedeon, P.C.; Elsamadicy, A.A.; Snyder, D.J.; et al. Temozolomide lymphodepletion enhances CAR abundance and correlates with antitumor efficacy against established glioblastoma. Oncoimmunology 2018, 7, e1434464. [Google Scholar] [CrossRef]
  257. Goff, S.L.; Morgan, R.A.; Yang, J.C.; Sherry, R.M.; Robbins, P.F.; Restifo, N.P.; Feldman, S.A.; Lu, Y.C.; Lu, L.; Zheng, Z.; et al. Pilot Trial of Adoptive Transfer of Chimeric Antigen Receptor-transduced T Cells Targeting EGFRvIII in Patients with Glioblastoma. J. Immunother. 2019, 42, 126–135. [Google Scholar] [CrossRef] [PubMed]
  258. Liu, Z.; Zhou, J.; Yang, X.; Liu, Y.; Zou, C.; Lv, W.; Chen, C.; Cheng, K.K.-y.; Chen, T.; Chang, L.-J.; et al. Safety and antitumor activity of GD2-Specific 4SCAR-T cells in patients with glioblastoma. Mol. Cancer 2023, 22, 3. [Google Scholar] [CrossRef] [PubMed]
  259. Badhiwala, J.; Decker, W.K.; Berens, M.E.; Bhardwaj, R.D. Clinical trials in cellular immunotherapy for brain/CNS tumors. Expert. Rev. Neurother. 2013, 13, 405–424. [Google Scholar] [CrossRef] [PubMed]
  260. Brown, C.E.; Alizadeh, D.; Starr, R.; Weng, L.; Wagner, J.R.; Naranjo, A.; Ostberg, J.R.; Blanchard, M.S.; Kilpatrick, J.; Simpson, J.; et al. Regression of Glioblastoma after Chimeric Antigen Receptor T-Cell Therapy. N. Engl. J. Med. 2016, 375, 2561–2569. [Google Scholar] [CrossRef] [PubMed]
  261. Brown, C.E.; Badie, B.; Barish, M.E.; Weng, L.; Ostberg, J.R.; Chang, W.C.; Naranjo, A.; Starr, R.; Wagner, J.; Wright, C.; et al. Bioactivity and Safety of IL13Ralpha2-Redirected Chimeric Antigen Receptor CD8+ T Cells in Patients with Recurrent Glioblastoma. Clin. Cancer Res. 2015, 21, 4062–4072. [Google Scholar] [CrossRef] [PubMed]
  262. Litten, J.B.; Ramakrishnan, A.; Astrow, S.H.; Harrison, C.; Aliki, A.; Badie, B. Phase 1b multicenter study to evaluate CHM 1101 in patients with recurrent or progressive glioblastoma. J. Clin. Oncol. 2023, 41 (Suppl. 16), TPS2086. [Google Scholar] [CrossRef]
  263. Lin, Q.; Ba, T.; Ho, J.; Chen, D.; Cheng, Y.; Wang, L.; Xu, G.; Xu, L.; Zhou, Y.; Wei, Y.; et al. First-in-Human Trial of EphA2-Redirected CAR T-Cells in Patients with Recurrent Glioblastoma: A Preliminary Report of Three Cases at the Starting Dose. Front. Oncol. 2021, 11, 694941. [Google Scholar] [CrossRef] [PubMed]
  264. Yao, Y.; Chen, D.; Tang, C.; Ji, C.; Li, Z.; Qian, Q. Safety, efficacy, and biomarker analysis of response to engineered tumor-infiltrating lymphocytes secreting anti-PD-1 antibody in recurrent glioblastoma: An open-label, two-arms, phase 1 study. J. Clin. Oncol. 2023, 41 (Suppl. 16), 2042. [Google Scholar] [CrossRef]
  265. Quattrocchi, K.B.; Miller, C.H.; Cush, S.; Bernard, S.A.; Dull, S.T.; Smith, M.; Gudeman, S.; Varia, M.A. Pilot study of local autologous tumor infiltrating lymphocytes for the treatment of recurrent malignant gliomas. J. Neuro-Oncol. 1999, 45, 141–157. [Google Scholar] [CrossRef]
  266. Sims, J.S.; Grinshpun, B.; Feng, Y.; Ung, T.H.; Neira, J.A.; Samanamud, J.L.; Canoll, P.; Shen, Y.; Sims, P.A.; Bruce, J.N. Diversity and divergence of the glioma-infiltrating T-cell receptor repertoire. Proc. Natl. Acad. Sci. USA 2016, 113, E3529–E3537. [Google Scholar] [CrossRef]
  267. Feins, S.; Kong, W.; Williams, E.F.; Milone, M.C.; Fraietta, J.A. An introduction to chimeric antigen receptor (CAR) T-cell immunotherapy for human cancer. Am. J. Hematol. 2019, 94, S3–S9. [Google Scholar] [CrossRef] [PubMed]
  268. Zhang, X.; Zhu, L.; Zhang, H.; Chen, S.; Xiao, Y. CAR-T Cell Therapy in Hematological Malignancies: Current Opportunities and Challenges. Front. Immunol. 2022, 13, 927153. [Google Scholar] [CrossRef] [PubMed]
  269. Karachi, A.; Dastmalchi, F.; Nazarian, S.; Huang, J.; Sayour, E.J.; Jin, L.; Yang, C.; Mitchell, D.A.; Rahman, M. Optimizing T Cell-Based Therapy for Glioblastoma. Front. Immunol. 2021, 12, 705580. [Google Scholar] [CrossRef]
  270. Luksik, A.S.; Yazigi, E.; Shah, P.; Jackson, C.M. CAR T Cell Therapy in Glioblastoma: Overcoming Challenges Related to Antigen Expression. Cancers 2023, 15, 1414. [Google Scholar] [CrossRef]
  271. Maggs, L.; Cattaneo, G.; Dal, A.E.; Moghaddam, A.S.; Ferrone, S. CAR T Cell-Based Immunotherapy for the Treatment of Glioblastoma. Front. Neurosci. 2021, 15, 662064. [Google Scholar] [CrossRef]
  272. Brown, C.E.; Warden, C.D.; Starr, R.; Deng, X.; Badie, B.; Yuan, Y.C.; Forman, S.J.; Barish, M.E. Glioma IL13Ralpha2 is associated with mesenchymal signature gene expression and poor patient prognosis. PLoS ONE 2013, 8, e77769. [Google Scholar] [CrossRef] [PubMed]
  273. Brown, C.E.; Aguilar, B.; Starr, R.; Yang, X.; Chang, W.C.; Weng, L.; Chang, B.; Sarkissian, A.; Brito, A.; Sanchez, J.F.; et al. Optimization of IL13Ralpha2-Targeted Chimeric Antigen Receptor T Cells for Improved Anti-tumor Efficacy against Glioblastoma. Mol. Ther. 2018, 26, 31–44. [Google Scholar] [CrossRef]
  274. Brown, C.E.; Rodriguez, A.; Palmer, J.; Ostberg, J.R.; Naranjo, A.; Wagner, J.R.; Aguilar, B.; Starr, R.; Weng, L.; Synold, T.W.; et al. Off-the-shelf, steroid-resistant, IL13Ralpha2-specific CAR T cells for treatment of glioblastoma. Neuro-Oncology 2022, 24, 1318–1330. [Google Scholar] [CrossRef]
  275. Mineo, J.F.; Bordron, A.; Baroncini, M.; Maurage, C.A.; Ramirez, C.; Siminski, R.M.; Berthou, C.; Dam Hieu, P. Low HER2-expressing glioblastomas are more often secondary to anaplastic transformation of low-grade glioma. J. Neuro-Oncol. 2007, 85, 281–287. [Google Scholar] [CrossRef]
  276. Ahmed, N.; Brawley, V.; Hegde, M.; Bielamowicz, K.; Kalra, M.; Landi, D.; Robertson, C.; Gray, T.L.; Diouf, O.; Wakefield, A.; et al. HER2-Specific Chimeric Antigen Receptor-Modified Virus-Specific T Cells for Progressive Glioblastoma: A Phase 1 Dose-Escalation Trial. JAMA Oncol. 2017, 3, 1094–1101. [Google Scholar] [CrossRef]
  277. Hegde, M.; Mukherjee, M.; Grada, Z.; Pignata, A.; Landi, D.; Navai, S.A.; Wakefield, A.; Fousek, K.; Bielamowicz, K.; Chow, K.K.; et al. Tandem CAR T cells targeting HER2 and IL13Ralpha2 mitigate tumor antigen escape. J. Clin. Investig. 2016, 126, 3036–3052. [Google Scholar] [CrossRef] [PubMed]
  278. Bielamowicz, K.; Fousek, K.; Byrd, T.T.; Samaha, H.; Mukherjee, M.; Aware, N.; Wu, M.F.; Orange, J.S.; Sumazin, P.; Man, T.K.; et al. Trivalent CAR T cells overcome interpatient antigenic variability in glioblastoma. Neuro-Oncology 2018, 20, 506–518. [Google Scholar] [CrossRef] [PubMed]
  279. Wan, P.K.; Ryan, A.J.; Seymour, L.W. Beyond cancer cells: Targeting the tumor microenvironment with gene therapy and armed oncolytic virus. Mol. Ther. 2021, 29, 1668–1682. [Google Scholar] [CrossRef]
  280. Chiocca, E.A.; Rabkin, S.D. Oncolytic viruses and their application to cancer immunotherapy. Cancer Immunol. Res. 2014, 2, 295–300. [Google Scholar] [CrossRef] [PubMed]
  281. Fares, J.; Ahmed, A.U.; Ulasov, I.V.; Sonabend, A.M.; Miska, J.; Lee-Chang, C.; Balyasnikova, I.V.; Chandler, J.P.; Portnow, J.; Tate, M.C.; et al. Neural stem cell delivery of an oncolytic adenovirus in newly diagnosed malignant glioma: A first-in-human, phase 1, dose-escalation trial. Lancet Oncol. 2021, 22, 1103–1114. [Google Scholar] [CrossRef] [PubMed]
  282. Chen, S.R.; Chen, M.M.; Ene, C.; Lang, F.F.; Kan, P. Perfusion-guided endovascular super-selective intra-arterial infusion for treatment of malignant brain tumors. J. Neurointerv Surg. 2022, 14, 533–538. [Google Scholar] [CrossRef] [PubMed]
  283. Lang, F.F.; Conrad, C.; Gomez-Manzano, C.; Yung, W.K.A.; Sawaya, R.; Weinberg, J.S.; Prabhu, S.S.; Rao, G.; Fuller, G.N.; Aldape, K.D.; et al. Phase I Study of DNX-2401 (Delta-24-RGD) Oncolytic Adenovirus: Replication and Immunotherapeutic Effects in Recurrent Malignant Glioma. J. Clin. Oncol. 2018, 36, 1419–1427. [Google Scholar] [CrossRef]
  284. Rivera-Molina, Y.; Fueyo, J.; Jiang, H.; Nguyen, T.; Ho Shin, D.; Youssef, G.; Fan, X.; Gumin, J.; Alonso, M.M.; Phadnis, S.; et al. EXTH-27. Activating the immunity within the tumor using viroimmunotherapy: Delta-24-RGD oncolytic adenovirus armed with the immunopositive regulator gitrl. Neuro-Oncology 2019, 21 (Suppl. 6), vi87. [Google Scholar] [CrossRef]
  285. Chiocca, E.A.; Abbed, K.M.; Tatter, S.; Louis, D.N.; Hochberg, F.H.; Barker, F.; Kracher, J.; Grossman, S.A.; Fisher, J.D.; Carson, K.; et al. A phase I open-label, dose-escalation, multi-institutional trial of injection with an E1B-Attenuated adenovirus, ONYX-015, into the peritumoral region of recurrent malignant gliomas, in the adjuvant setting. Mol. Ther. 2004, 10, 958–966. [Google Scholar] [CrossRef]
  286. Umemura, Y.; Orringer, D.; Junck, L.; Varela, M.L.; West, M.E.J.; Faisal, S.M.; Comba, A.; Heth, J.; Sagher, O.; Leung, D.; et al. Combined cytotoxic and immune-stimulatory gene therapy for primary adult high-grade glioma: A phase 1, first-in-human trial. Lancet Oncol. 2023, 24, 1042–1052. [Google Scholar] [CrossRef]
  287. Chiocca, E.A.; Aguilar, L.K.; Bell, S.D.; Kaur, B.; Hardcastle, J.; Cavaliere, R.; McGregor, J.; Lo, S.; Ray-Chaudhuri, A.; Chakravarti, A.; et al. Phase IB study of gene-mediated cytotoxic immunotherapy adjuvant to up-front surgery and intensive timing radiation for malignant glioma. J. Clin. Oncol. 2011, 29, 3611–3619. [Google Scholar] [CrossRef]
  288. Ji, N.; Weng, D.; Liu, C.; Gu, Z.; Chen, S.; Guo, Y.; Fan, Z.; Wang, X.; Chen, J.; Zhao, Y.; et al. Adenovirus-mediated delivery of herpes simplex virus thymidine kinase administration improves outcome of recurrent high-grade glioma. Oncotarget 2016, 7, 4369–4378. [Google Scholar] [CrossRef] [PubMed]
  289. Wheeler, L.A.; Manzanera, A.G.; Bell, S.D.; Cavaliere, R.; McGregor, J.M.; Grecula, J.C.; Newton, H.B.; Lo, S.S.; Badie, B.; Portnow, J.; et al. Phase II multicenter study of gene-mediated cytotoxic immunotherapy as adjuvant to surgical resection for newly diagnosed malignant glioma. Neuro-Oncology 2016, 18, 1137–1145. [Google Scholar] [CrossRef] [PubMed]
  290. Chiocca, E.A.; Yu, J.S.; Lukas, R.V.; Solomon, I.H.; Ligon, K.L.; Nakashima, H.; Triggs, D.A.; Reardon, D.A.; Wen, P.; Stopa, B.M.; et al. Regulatable interleukin-12 gene therapy in patients with recurrent high-grade glioma: Results of a phase 1 trial. Sci. Transl. Med. 2019, 11, eaaw5680. [Google Scholar] [CrossRef] [PubMed]
  291. Brenner, A.J.; Peters, K.B.; Vredenburgh, J.; Bokstein, F.; Blumenthal, D.T.; Yust-Katz, S.; Peretz, I.; Oberman, B.; Freedman, L.S.; Ellingson, B.M.; et al. Safety and efficacy of VB-111, an anticancer gene therapy, in patients with recurrent glioblastoma: Results of a phase I/II study. Neuro-Oncology 2020, 22, 694–704. [Google Scholar] [CrossRef] [PubMed]
  292. Cloughesy, T.F.; Brenner, A.; de Groot, J.F.; Butowski, N.A.; Zach, L.; Campian, J.L.; Ellingson, B.M.; Freedman, L.S.; Cohen, Y.C.; Lowenton-Spier, N.; et al. A randomized controlled phase III study of VB-111 combined with bevacizumab vs bevacizumab monotherapy in patients with recurrent glioblastoma (GLOBE). Neuro-Oncology 2020, 22, 705–717. [Google Scholar] [CrossRef] [PubMed]
  293. Rampling, R.; Cruickshank, G.; Papanastassiou, V.; Nicoll, J.; Hadley, D.; Brennan, D.; Petty, R.; MacLean, A.; Harland, J.; McKie, E.; et al. Toxicity evaluation of replication-competent herpes simplex virus (ICP 34.5 null mutant 1716) in patients with recurrent malignant glioma. Gene Ther. 2000, 7, 859–866. [Google Scholar] [CrossRef] [PubMed]
  294. Papanastassiou, V.; Rampling, R.; Fraser, M.; Petty, R.; Hadley, D.; Nicoll, J.; Harland, J.; Mabbs, R.; Brown, M. The potential for efficacy of the modified (ICP 34.5(-)) herpes simplex virus HSV1716 following intratumoural injection into human malignant glioma: A proof of principle study. Gene Ther. 2002, 9, 398–406. [Google Scholar] [CrossRef]
  295. Harrow, S.; Papanastassiou, V.; Harland, J.; Mabbs, R.; Petty, R.; Fraser, M.; Hadley, D.; Patterson, J.; Brown, S.M.; Rampling, R. HSV1716 injection into the brain adjacent to tumour following surgical resection of high-grade glioma: Safety data and long-term survival. Gene Ther. 2004, 11, 1648–1658. [Google Scholar] [CrossRef] [PubMed]
  296. Markert, J.M.; Razdan, S.N.; Kuo, H.C.; Cantor, A.; Knoll, A.; Karrasch, M.; Nabors, L.B.; Markiewicz, M.; Agee, B.S.; Coleman, J.M.; et al. A phase 1 trial of oncolytic HSV-1, G207, given in combination with radiation for recurrent GBM demonstrates safety and radiographic responses. Mol. Ther. 2014, 22, 1048–1055. [Google Scholar] [CrossRef]
  297. Markert, J.M.; Liechty, P.G.; Wang, W.; Gaston, S.; Braz, E.; Karrasch, M.; Nabors, L.B.; Markiewicz, M.; Lakeman, A.D.; Palmer, C.A.; et al. Phase Ib trial of mutant herpes simplex virus G207 inoculated pre-and post-tumor resection for recurrent GBM. Mol. Ther. 2009, 17, 199–207. [Google Scholar] [CrossRef]
  298. Markert, J.M.; Medlock, M.D.; Rabkin, S.D.; Gillespie, G.Y.; Todo, T.; Hunter, W.D.; Palmer, C.A.; Feigenbaum, F.; Tornatore, C.; Tufaro, F.; et al. Conditionally replicating herpes simplex virus mutant, G207 for the treatment of malignant glioma: Results of a phase I trial. Gene Ther. 2000, 7, 867–874. [Google Scholar] [CrossRef] [PubMed]
  299. Todo, T.; Ito, H.; Ino, Y.; Ohtsu, H.; Ota, Y.; Shibahara, J.; Tanaka, M. Intratumoral oncolytic herpes virus G47∆ for residual or recurrent glioblastoma: A phase 2 trial. Nat. Med. 2022, 28, 1630–1639. [Google Scholar] [CrossRef] [PubMed]
  300. Chiocca, E.A.; Nakashima, H.; Kasai, K.; Fernandez, S.A.; Oglesbee, M. Preclinical Toxicology of rQNestin34.5v.2: An Oncolytic Herpes Virus with Transcriptional Regulation of the ICP34.5 Neurovirulence Gene. Mol. Ther. Methods Clin. Dev. 2020, 17, 871–893. [Google Scholar] [CrossRef] [PubMed]
  301. Chiocca, E.A.; Solomon, I.; Nakashima, H.; Lawler, S.E.; Triggs, D.; Zhang, A.; Grant, J.; Reardon, D.A.; Wen, P.Y.; Lee, E.Q.; et al. First-in-human CAN-3110 (ICP-34.5 expressing HSV-1 oncolytic virus) in patients with recurrent high-grade glioma. J. Clin. Oncol. 2021, 39 (Suppl. 15), 2009. [Google Scholar] [CrossRef]
  302. Rainov, N.G. A phase III clinical evaluation of herpes simplex virus type 1 thymidine kinase and ganciclovir gene therapy as an adjuvant to surgical resection and radiation in adults with previously untreated glioblastoma multiforme. Hum. Gene Ther. 2000, 11, 2389–2401. [Google Scholar] [CrossRef] [PubMed]
  303. Kalkanis, S.; Jolly, D.J.; Pertschuk, D.; Ostertag, D.; Robbins, J.M.; Huang, T.T.; Gruber, H.; Mikkelsen, T. AT-29: INTRAVENOUS ADMINISTRATION OF TOCA 511 IN PATIENTS WITH RECURRENT GLIOBLASTOMA. Neuro-Oncology 2014, 16 (Suppl. 5), v15. [Google Scholar] [CrossRef]
  304. Merchan, J.R.; Ahnert, J.R.; Falchook, G.; Ostertag, D.; Tejera, D.; Gruber, H.E.; Jolly, D.J.; Shorr, J. Toca 6: A phase 1b study of Toca 511 and Toca FC in patients with advanced solid tumors or lymphoma. J. Clin. Oncol. 2018, 36 (Suppl. 15), TPS2613. [Google Scholar] [CrossRef]
  305. Ostertag, D.; Amundson, K.K.; Lopez Espinoza, F.; Martin, B.; Buckley, T.; Galvão da Silva, A.P.; Lin, A.H.; Valenta, D.T.; Perez, O.D.; Ibañez, C.E.; et al. Brain tumor eradication and prolonged survival from intratumoral conversion of 5-fluorocytosine to 5-fluorouracil using a nonlytic retroviral replicating vector. Neuro-Oncology 2012, 14, 145–159. [Google Scholar] [CrossRef] [PubMed]
  306. Cloughesy, T.F.; Landolfi, J.; Vogelbaum, M.A.; Ostertag, D.; Elder, J.B.; Bloomfield, S.; Carter, B.; Chen, C.C.; Kalkanis, S.N.; Kesari, S.; et al. Durable complete responses in some recurrent high-grade glioma patients treated with Toca 511 + Toca FC. Neuro-Oncology 2018, 20, 1383–1392. [Google Scholar] [CrossRef] [PubMed]
  307. Aghi, M.; Vogelbaum, M.A.; Kesari, S.; Chen, C.C.; Liau, L.M.; Piccioni, D.; Portnow, J.; Chang, S.; Robbins, J.M.; Boyce, T.; et al. AT-02 Intratumoral delivery of the retroviral replicating vector (RRV) TOCA 511 in subjects with recurrent high grade glioma: Interim report of phase 1 study (NCT 01156584). Neuro-Oncology 2014, 16 (Suppl. 5), v8. [Google Scholar] [CrossRef]
  308. Cloughesy, T.F.; Petrecca, K.; Walbert, T.; Butowski, N.; Salacz, M.; Perry, J.; Damek, D.; Bota, D.; Bettegowda, C.; Zhu, J.-J.; et al. Effect of Vocimagene Amiretrorepvec in Combination with Flucytosine vs Standard of Care on Survival Following Tumor Resection in Patients with Recurrent High-Grade Glioma: A Randomized Clinical Trial. JAMA Oncol. 2020, 6, 1939–1946. [Google Scholar] [CrossRef]
  309. Ahluwalia, M.; Pugh, S.; Ellingson, B.; Kotecha, R.; Cloughesy, T.; Vogelbaum, M.; Aldape, K.; Cui, Y.; Armstrong, T.; Mehta, M. RBTT-11. NRG Oncology NRG-BN006: A Phase II/III randomized, open-label study of Toca 511 and Toca FC with standard of care compared to standard of care in patients with newly diagnosed glioblastoma. Neuro-Oncology 2019, 21 (Suppl. 6), vi220–vi221. [Google Scholar] [CrossRef]
  310. Freeman, A.I.; Zakay-Rones, Z.; Gomori, J.M.; Linetsky, E.; Rasooly, L.; Greenbaum, E.; Rozenman-Yair, S.; Panet, A.; Libson, E.; Irving, C.S.; et al. Phase I/II trial of intravenous NDV-HUJ oncolytic virus in recurrent glioblastoma multiforme. Mol. Ther. 2006, 13, 221–228. [Google Scholar] [CrossRef] [PubMed]
  311. Csatary, L.K.; Gosztonyi, G.; Szeberenyi, J.; Fabian, Z.; Liszka, V.; Bodey, B.; Csatary, C.M. MTH-68/H oncolytic viral treatment in human high-grade gliomas. J. Neuro-Oncol. 2004, 67, 83–93. [Google Scholar] [CrossRef] [PubMed]
  312. Geletneky, K.; Hajda, J.; Angelova, A.L.; Leuchs, B.; Capper, D.; Bartsch, A.J.; Neumann, J.O.; Schöning, T.; Hüsing, J.; Beelte, B.; et al. Oncolytic H-1 Parvovirus Shows Safety and Signs of Immunogenic Activity in a First Phase I/IIa Glioblastoma Trial. Mol. Ther. 2017, 25, 2620–2634. [Google Scholar] [CrossRef] [PubMed]
  313. Angelova, A.; Rommelaere, J. Immune System Stimulation by Oncolytic Rodent Protoparvoviruses. Viruses 2019, 11, 415. [Google Scholar] [CrossRef] [PubMed]
  314. Desjardins, A.; Gromeier, M.; Herndon, J.E., 2nd; Beaubier, N.; Bolognesi, D.P.; Friedman, A.H.; Friedman, H.S.; McSherry, F.; Muscat, A.M.; Nair, S.; et al. Recurrent Glioblastoma Treated with Recombinant Poliovirus. N. Engl. J. Med. 2018, 379, 150–161. [Google Scholar] [CrossRef] [PubMed]
  315. Forsyth, P.; Roldán, G.; George, D.; Wallace, C.; Palmer, C.A.; Morris, D.; Cairncross, G.; Matthews, M.V.; Markert, J.; Gillespie, Y.; et al. A Phase I Trial of Intratumoral Administration of Reovirus in Patients with Histologically Confirmed Recurrent Malignant Gliomas. Mol. Ther. 2008, 16, 627–632. [Google Scholar] [CrossRef] [PubMed]
  316. Kicielinski, K.P.; Chiocca, E.A.; Yu, J.S.; Gill, G.M.; Coffey, M.; Markert, J.M. Phase 1 clinical trial of intratumoral reovirus infusion for the treatment of recurrent malignant gliomas in adults. Mol. Ther. 2014, 22, 1056–1062. [Google Scholar] [CrossRef]
  317. Samson, A.; Scott, K.J.; Taggart, D.; West, E.J.; Wilson, E.; Nuovo, G.J.; Thomson, S.; Corns, R.; Mathew, R.K.; Fuller, M.J.; et al. Intravenous delivery of oncolytic reovirus to brain tumor patients immunologically primes for subsequent checkpoint blockade. Sci. Transl. Med. 2018, 10, eaam7577. [Google Scholar] [CrossRef] [PubMed]
  318. Trask, T.W.; Trask, R.P.; Aguilar-Cordova, E.; Shine, H.D.; Wyde, P.R.; Goodman, J.C.; Hamilton, W.J.; Rojas-Martinez, A.; Chen, S.H.; Woo, S.L.; et al. Phase I study of adenoviral delivery of the HSV-tk gene and ganciclovir administration in patients with current malignant brain tumors. Mol. Ther. 2000, 1, 195–203. [Google Scholar] [CrossRef]
  319. Smitt, P.S.; Driesse, M.; Wolbers, J.; Kros, M.; Avezaat, C. Treatment of relapsed malignant glioma with an adenoviral vector containing the herpes simplex thymidine kinase gene followed by ganciclovir. Mol. Ther. 2003, 7, 851–858. [Google Scholar] [CrossRef]
  320. Sandmair, A.M.; Loimas, S.; Puranen, P.; Immonen, A.; Kossila, M.; Puranen, M.; Hurskainen, H.; Tyynelä, K.; Turunen, M.; Vanninen, R.; et al. Thymidine kinase gene therapy for human malignant glioma, using replication-deficient retroviruses or adenoviruses. Hum. Gene Ther. 2000, 11, 2197–2205. [Google Scholar] [CrossRef] [PubMed]
  321. Germano, I.M.; Fable, J.; Gultekin, S.H.; Silvers, A. Adenovirus/herpes simplex-thymidine kinase/ganciclovir complex: Preliminary results of a phase I trial in patients with recurrent malignant gliomas. J. Neuro-Oncol. 2003, 65, 279–289. [Google Scholar] [CrossRef]
  322. Immonen, A.; Vapalahti, M.; Tyynelä, K.; Hurskainen, H.; Sandmair, A.; Vanninen, R.; Langford, G.; Murray, N.; Ylä-Herttuala, S. AdvHSV-tk gene therapy with intravenous ganciclovir improves survival in human malignant glioma: A randomised, controlled study. Mol. Ther. 2004, 10, 967–972. [Google Scholar] [CrossRef] [PubMed]
  323. Bögler, O.; Su Huang, H.-J.; Kleihues, P.; Cavenee, W.K. The p53 gene and its role in human brain tumors. Glia 1995, 15, 308–327. [Google Scholar] [CrossRef] [PubMed]
  324. Zhang, W.W.; Alemany, R.; Wang, J.; Koch, P.E.; Ordonez, N.G.; Roth, J.A. Safety evaluation of Ad5CMV-p53 in vitro and in vivo. Hum. Gene Ther. 1995, 6, 155–164. [Google Scholar] [CrossRef] [PubMed]
  325. Lang, F.F.; Bruner, J.M.; Fuller, G.N.; Aldape, K.; Prados, M.D.; Chang, S.; Berger, M.S.; McDermott, M.W.; Kunwar, S.M.; Junck, L.R.; et al. Phase I trial of adenovirus-mediated p53 gene therapy for recurrent glioma: Biological and clinical results. J. Clin. Oncol. 2003, 21, 2508–2518. [Google Scholar] [CrossRef]
  326. Barrett, J.A.; Cai, H.; Miao, J.; Khare, P.D.; Gonzalez, P.; Dalsing-Hernandez, J.; Sharma, G.; Chan, T.; Cooper, L.J.N.; Lebel, F. Regulated intratumoral expression of IL-12 using a RheoSwitch Therapeutic System((R)) (RTS((R))) gene switch as gene therapy for the treatment of glioma. Cancer Gene Ther. 2018, 25, 106–116. [Google Scholar] [CrossRef] [PubMed]
  327. Nandi, S.; Lesniak, M.S. Adenoviral virotherapy for malignant brain tumors. Expert. Opin. Biol. Ther. 2009, 9, 737–747. [Google Scholar] [CrossRef]
  328. Cheng, P.-H.; Wechman, S.L.; McMasters, K.M.; Zhou, H.S. Oncolytic Replication of E1b-Deleted Adenoviruses. Viruses 2015, 7, 5767–5779. [Google Scholar] [CrossRef]
  329. Bischoff, J.R.; Kirn, D.H.; Williams, A.; Heise, C.; Horn, S.; Muna, M.; Ng, L.; Nye, J.A.; Sampson-Johannes, A.; Fattaey, A.; et al. An adenovirus mutant that replicates selectively in p53-deficient human tumor cells. Science 1996, 274, 373–376. [Google Scholar] [CrossRef]
  330. Heise, C.C.; Williams, A.M.; Xue, S.; Propst, M.; Kirn, D.H. Intravenous administration of ONYX-015, a selectively replicating adenovirus, induces antitumoral efficacy. Cancer Res. 1999, 59, 2623–2628. [Google Scholar] [PubMed]
  331. Fueyo, J.; Gomez-Manzano, C.; Alemany, R.; Lee, P.S.; McDonnell, T.J.; Mitlianga, P.; Shi, Y.X.; Levin, V.A.; Yung, W.K.; Kyritsis, A.P. A mutant oncolytic adenovirus targeting the Rb pathway produces anti-glioma effect in vivo. Oncogene 2000, 19, 2–12. [Google Scholar] [CrossRef]
  332. Suzuki, K.; Fueyo, J.; Krasnykh, V.; Reynolds, P.N.; Curiel, D.T.; Alemany, R. A conditionally replicative adenovirus with enhanced infectivity shows improved oncolytic potency. Clin. Cancer Res. 2001, 7, 120–126. [Google Scholar] [PubMed]
  333. Stepanenko, A.A.; Sosnovtseva, A.O.; Valikhov, M.P.; Chernysheva, A.A.; Cherepanov, S.A.; Yusubalieva, G.M.; Ruzsics, Z.; Lipatova, A.V.; Chekhonin, V.P. Superior infectivity of the fiber chimeric oncolytic adenoviruses Ad5/35 and Ad5/3 over Ad5-delta-24-RGD in primary glioma cultures. Mol. Ther. Oncolytics 2022, 24, 230–248. [Google Scholar] [CrossRef]
  334. Alonso, M.M.; García-Moure, M.; Gonzalez-Huarriz, M.; Marigil, M.; Hernandez-Alcoceba, R.; Buñales, M.; Hervás, S.; Gallego, J.; Gomez-Manzano, C.; Fueyo, J.; et al. Abstract CT027: Oncolytic virus DNX-2401 with a short course of temozolomide for glioblastoma at first recurrence: Clinical data and prognostic biomarkers. Cancer Res. 2017, 77 (Suppl. 13), CT027. [Google Scholar] [CrossRef]
  335. Lang, F.F.; Tran, N.D.; Puduvalli, V.K.; Elder, J.B.; Fink, K.L.; Conrad, C.A.; Yung, W.K.A.; Penas-Prado, M.; Gomez-Manzano, C.; Peterkin, J.; et al. Phase 1b open-label randomized study of the oncolytic adenovirus DNX-2401 administered with or without interferon gamma for recurrent glioblastoma. J. Clin. Oncol. 2017, 35 (Suppl. 15), 2002. [Google Scholar] [CrossRef]
  336. van Putten, E.H.P.; Kleijn, A.; van Beusechem, V.W.; Noske, D.; Lamers, C.H.J.; de Goede, A.L.; Idema, S.; Hoefnagel, D.; Kloezeman, J.J.; Fueyo, J.; et al. Convection Enhanced Delivery of the Oncolytic Adenovirus Delta24-RGD in Patients with Recurrent GBM: A Phase I Clinical Trial Including Correlative Studies. Clin. Cancer Res. 2022, 28, 1572–1585. [Google Scholar] [CrossRef]
  337. Nassiri, F.; Patil, V.; Yefet, L.S.; Singh, O.; Liu, J.; Dang, R.M.A.; Yamaguchi, T.N.; Daras, M.; Cloughesy, T.F.; Colman, H.; et al. Oncolytic DNX-2401 virotherapy plus pembrolizumab in recurrent glioblastoma: A phase 1/2 trial. Nat. Med. 2023, 29, 1370–1378. [Google Scholar] [CrossRef]
  338. Jiang, H.; Rivera-Molina, Y.; Gomez-Manzano, C.; Clise-Dwyer, K.; Bover, L.; Vence, L.M.; Yuan, Y.; Lang, F.F.; Toniatti, C.; Hossain, M.B.; et al. Oncolytic Adenovirus and Tumor-Targeting Immune Modulatory Therapy Improve Autologous Cancer Vaccination. Cancer Res. 2017, 77, 3894–3907. [Google Scholar] [CrossRef]
  339. Ulasov, I.V.; Rivera, A.A.; Sonabend, A.M.; Rivera, L.B.; Wang, M.; Zhu, Z.B.; Lesniak, M.S. Comparative evaluation of survivin, midkine and CXCR4 promoters for transcriptional targeting of glioma gene therapy. Cancer Biol. Ther. 2007, 6, 679–685. [Google Scholar] [CrossRef]
  340. Ulasov, I.V.; Zhu, Z.B.; Tyler, M.A.; Han, Y.; Rivera, A.A.; Khramtsov, A.; Curiel, D.T.; Lesniak, M.S. Survivin-driven and fiber-modified oncolytic adenovirus exhibits potent antitumor activity in established intracranial glioma. Hum. Gene Ther. 2007, 18, 589–602. [Google Scholar] [CrossRef] [PubMed]
  341. Kim, J.W.; Auffinger, B.; Spencer, D.A.; Miska, J.; Chang, A.L.; Kane, J.R.; Young, J.S.; Kanojia, D.; Qiao, J.; Mann, J.F.; et al. Single dose GLP toxicity and biodistribution study of a conditionally replicative adenovirus vector, CRAd-S-pk7, administered by intracerebral injection to Syrian hamsters. J. Transl. Med. 2016, 14, 134. [Google Scholar] [CrossRef]
  342. Perez, O.D.; Logg, C.R.; Hiraoka, K.; Diago, O.; Burnett, R.; Inagaki, A.; Jolson, D.; Amundson, K.; Buckley, T.; Lohse, D.; et al. Design and selection of Toca 511 for clinical use: Modified retroviral replicating vector with improved stability and gene expression. Mol. Ther. 2012, 20, 1689–1698. [Google Scholar] [CrossRef] [PubMed]
  343. Mitchell, L.A.; Lopez Espinoza, F.; Mendoza, D.; Kato, Y.; Inagaki, A.; Hiraoka, K.; Kasahara, N.; Gruber, H.E.; Jolly, D.J.; Robbins, J.M. Toca 511 gene transfer and treatment with the prodrug, 5-fluorocytosine, promotes durable antitumor immunity in a mouse glioma model. Neuro-Oncology 2017, 19, 930–939. [Google Scholar] [CrossRef] [PubMed]
  344. Kuriyama, S.; Masui, K.; Sakamoto, T.; Nakatani, T.; Kikukawa, M.; Tsujinoue, H.; Mitoro, A.; Yamazaki, M.; Yoshiji, H.; Fukui, H.; et al. Bystander effect caused by cytosine deaminase gene and 5-fluorocytosine in vitro is substantially mediated by generated 5-fluorouracil. Anticancer. Res. 1998, 18, 3399–3406. [Google Scholar] [PubMed]
  345. McKie, E.A.; MacLean, A.R.; Lewis, A.D.; Cruickshank, G.; Rampling, R.; Barnett, S.C.; Kennedy, P.G.; Brown, S.M. Selective in vitro replication of herpes simplex virus type 1 (HSV-1) ICP34.5 null mutants in primary human CNS tumours--evaluation of a potentially effective clinical therapy. Br. J. Cancer 1996, 74, 745–752. [Google Scholar] [CrossRef] [PubMed]
  346. Mineta, T.; Rabkin, S.D.; Yazaki, T.; Hunter, W.D.; Martuza, R.L. Attenuated multi-mutated herpes simplex virus-1 for the treatment of malignant gliomas. Nat. Med. 1995, 1, 938–943. [Google Scholar] [CrossRef]
  347. Todo, T. Oncolytic virus therapy using genetically engineered herpes simplex viruses. Front. Biosci. 2008, 13, 2060–2064. [Google Scholar] [CrossRef]
  348. Coffin, R. Interview with Robert Coffin, inventor of T-VEC: The first oncolytic immunotherapy approved for the treatment of cancer. Immunotherapy 2016, 8, 103–106. [Google Scholar] [CrossRef]
  349. Gromeier, M.; Nair, S.K. Recombinant Poliovirus for Cancer Immunotherapy. Annu. Rev. Med. 2018, 69, 289–299. [Google Scholar] [CrossRef]
  350. Merrill, M.K.; Bernhardt, G.; Sampson, J.H.; Wikstrand, C.J.; Bigner, D.D.; Gromeier, M. Poliovirus receptor CD155-targeted oncolysis of glioma. Neuro-Oncology 2004, 6, 208–217. [Google Scholar] [CrossRef] [PubMed]
  351. Prior, I.A.; Hood, F.E.; Hartley, J.L. The Frequency of Ras Mutations in Cancer. Cancer Res. 2020, 80, 2969–2974. [Google Scholar] [CrossRef] [PubMed]
  352. Kelly, E.; Russell, S.J. History of Oncolytic Viruses: Genesis to Genetic Engineering. Mol. Ther. 2007, 15, 651–659. [Google Scholar] [CrossRef] [PubMed]
  353. Engeland, C.E.; Ungerechts, G. Measles Virus as an Oncolytic Immunotherapy. Cancers 2021, 13, 544. [Google Scholar] [CrossRef] [PubMed]
  354. Bai, Y.; Chen, Y.; Hong, X.; Liu, X.; Su, X.; Li, S.; Dong, X.; Zhao, G.; Li, Y. Newcastle disease virus enhances the growth-inhibiting and proapoptotic effects of temozolomide on glioblastoma cells in vitro and in vivo. Sci. Rep. 2018, 8, 11470. [Google Scholar] [CrossRef] [PubMed]
  355. Reichard, K.W.; Lorence, R.M.; Cascino, C.J.; Peeples, M.E.; Walter, R.J.; Fernando, M.B.; Reyes, H.M.; Greager, J.A. Newcastle disease virus selectively kills human tumor cells. J. Surg. Res. 1992, 52, 448–453. [Google Scholar] [CrossRef] [PubMed]
  356. Marchini, A.; Daeffler, L.; Pozdeev, V.I.; Angelova, A.; Rommelaere, J. Immune Conversion of Tumor Microenvironment by Oncolytic Viruses: The Protoparvovirus H-1PV Case Study. Front. Immunol. 2019, 10, 1848. [Google Scholar] [CrossRef] [PubMed]
  357. Hartley, A.; Kavishwar, G.; Salvato, I.; Marchini, A. A Roadmap for the Success of Oncolytic Parvovirus-Based Anticancer Therapies. Annu. Rev. Virol. 2020, 7, 537–557. [Google Scholar] [CrossRef]
  358. Geletneky, K.; Angelova, A.; Leuchs, B.; Bartsch, A.; Capper, D.; Hajda, J.; Rommelaere, J. Atnt-07favorable Response of Patients with Glioblastoma at Second or Third Recurrence to Repeated Injection of Oncolytic Parvovirus H-1 in Combination with Bevacicumab. Neuro-Oncology 2015, 17 (Suppl. 5), v11. [Google Scholar] [CrossRef]
  359. Li, J.; Bonifati, S.; Hristov, G.; Marttila, T.; Valmary-Degano, S.; Stanzel, S.; Schnölzer, M.; Mougin, C.; Aprahamian, M.; Grekova, S.P.; et al. Synergistic combination of valproic acid and oncolytic parvovirus H-1PV as a potential therapy against cervical and pancreatic carcinomas. EMBO Mol. Med. 2013, 5, 1537–1555. [Google Scholar] [CrossRef]
  360. Geletneky, K.; Bartsch, A.; Weiss, C.; Bernhard, H.; Marchini, A.; Rommelaere, J. ATIM-40. High rate of objective anti-tumor response in 9 patients with glioblastoma after viro-immunotherapy with oncolytic parvovirus H-1 in combination with bevacicumab and PD-1 checkpoint blockade. Neuro-Oncology 2018, 20 (Suppl. 6), vi10. [Google Scholar] [CrossRef]
  361. Geletneky, K.; Weiss, C.; Bernhard, H.; Capper, D.; Leuchs, B.; Marchini, A.; Rommelaere, J. ATIM-29. First clinical observation of improved anti-tumor effects of viro-immunotherapy with oncolytic parvovirus H-1 in combination with PD-1 checkpoint blockade and bevacicumab in patients with recurrent glioblastoma. Neuro-Oncology 2016, 18 (Suppl. 6), vi24. [Google Scholar] [CrossRef]
  362. Hengstschläger, M.; Knöfler, M.; Müllner, E.W.; Ogris, E.; Wintersberger, E.; Wawra, E. Different regulation of thymidine kinase during the cell cycle of normal versus DNA tumor virus-transformed cells. J. Biol. Chem. 1994, 269, 13836–13842. [Google Scholar] [CrossRef] [PubMed]
  363. Foloppe, J.; Kintz, J.; Futin, N.; Findeli, A.; Cordier, P.; Schlesinger, Y.; Hoffmann, C.; Tosch, C.; Balloul, J.M.; Erbs, P. Targeted delivery of a suicide gene to human colorectal tumors by a conditionally replicating vaccinia virus. Gene Ther. 2008, 15, 1361–1371. [Google Scholar] [CrossRef] [PubMed]
  364. Dutoit, V.; Marinari, E.; Dietrich, P.-Y.; Migliorini, D. Combination of the Ima950/Poly-Iclc Multipeptide Vaccine with Pembrolizumab in Relapsing Glioblastoma Patients. Neuro-Oncology 2020, 22 (Suppl. 2), ii34. [Google Scholar] [CrossRef]
  365. Miller, A.; Kosaloglu-Yalcin, Z.; Westernberg, L.; Montero, L.; Bahmanof, M.; Frentzen, A.; Premlal, A.L.R.; Greenbaum, J.; Seumois, G.; Habbaba, R.; et al. A phase 1b study of personalized neoantigen vaccine plus pembrolizumab in adults with advanced cancer. J. Clin. Oncol. 2021, 39 (Suppl. 15), 2615. [Google Scholar] [CrossRef]
  366. Ahluwalia, M.S.; Peereboom, D.M.; Ciolfi, M.; Schilero, C.; Hobbs, B.; Ciesielski, M.J.; Fenstermaker, R.A. Phase II study of pembrolizumab plus SurVaxM for glioblastoma at first recurrence. J. Clin. Oncol. 2020, 38 (Suppl. 15), TPS2581. [Google Scholar] [CrossRef]
  367. Chiocca, E.A.; Gelb, A.B.; Chen, C.C.; Rao, G.; Reardon, D.A.; Wen, P.Y.; Bi, W.L.; Peruzzi, P.; Amidei, C.; Triggs, D.; et al. Combined immunotherapy with controlled interleukin-12 gene therapy and immune checkpoint blockade in recurrent glioblastoma: An open-label, multi-institutional phase I trial. Neuro-Oncology 2022, 24, 951–963. [Google Scholar] [CrossRef]
  368. Sloan, A.E.; Buerki, R.A.; Murphy, C.; Kelly, A.T.; Ambady, P.; Brown, M.; Butowski, N.A.; Cavaliere, R.; Curry, W.T.; Desjardins, A.; et al. LUMINOS-101: Phase 2 study of PVSRIPO with pembrolizumab in recurrent glioblastoma. J. Clin. Oncol. 2021, 39 (Suppl. 15), TPS2065. [Google Scholar] [CrossRef]
  369. Awada, H.; Paris, F.; Pecqueur, C. Exploiting radiation immunostimulatory effects to improve glioblastoma outcome. Neuro-Oncology 2023, 25, 433–446. [Google Scholar] [CrossRef]
  370. De Martino, M.; Padilla, O.; Daviaud, C.; Wu, C.C.; Gartrell, R.D.; Vanpouille-Box, C. Exploiting Radiation Therapy to Restore Immune Reactivity of Glioblastoma. Front. Oncol. 2021, 11, 671044. [Google Scholar] [CrossRef] [PubMed]
  371. Slika, H.; Karimov, Z.; Alimonti, P.; Abou-Mrad, T.; De Fazio, E.; Alomari, S.; Tyler, B. Preclinical Models and Technologies in Glioblastoma Research: Evolution, Current State, and Future Avenues. Int. J. Mol. Sci. 2023, 24, 16316. [Google Scholar] [CrossRef] [PubMed]
  372. Stepanenko, A.A.; Sosnovtseva, A.O.; Valikhov, M.P.; Chernysheva, A.A.; Abramova, O.V.; Naumenko, V.A.; Chekhonin, V.P. The need for paradigm shift: Prognostic significance and implications of standard therapy-related systemic immunosuppression in glioblastoma for immunotherapy and oncolytic virotherapy. Front. Immunol. 2024, 15, 1326757. [Google Scholar] [CrossRef]
  373. Choi, B.D.; Yu, X.; Castano, A.P.; Bouffard, A.A.; Schmidts, A.; Larson, R.C.; Bailey, S.R.; Boroughs, A.C.; Frigault, M.J.; Leick, M.B.; et al. CAR-T cells secreting BiTEs circumvent antigen escape without detectable toxicity. Nat. Biotechnol. 2019, 37, 1049–1058. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Therapeutic challenges for the cure of GBM. Abbreviations: APC, antigen-presenting cell; bFGF, basic fibroblast growth factor; ECM, extracellular matrix; GBM, glioblastoma; IL, interleukin; LAG-3, lymphocyte-activation gene 3; MDSC, myeloid-derived suppressor cells; PD-1, programmed cell death protein 1; PGE2, prostaglandin E2; TAM, tumor-associated microglia and macrophages; TGF-β, transforming growth factor-β; TIM-3, T-cell immunoglobulin and mucin domain; TME, tumor microenvironment; Treg, regulatory T cell; WHO, the World Health Organization. The figure illustrates the distinctive characteristics of GBM (WHO grade 4) that are understood to hinder the development of effective anti-tumor therapies. These include (1) an anatomical location shielded by the blood–brain barrier, (2) intra- and inter-patient tumor heterogeneity, (3) infiltrative behavior, and (4) a highly immunosuppressive TME. The latter showcases the presence of GBM-driven cytokines with immunosuppressive and tumor-promoting properties, along with immunosuppressive cell populations such as Tregs and M1-like TAMs, accompanied by upregulated exhaustion markers. Additionally, GBMs strategically downregulate antigen-processing and presentation molecules to effectively evade T cell activation. The image was created using BioRender (https://www.biorender.com/, accessed on 18 December 2023).
Figure 1. Therapeutic challenges for the cure of GBM. Abbreviations: APC, antigen-presenting cell; bFGF, basic fibroblast growth factor; ECM, extracellular matrix; GBM, glioblastoma; IL, interleukin; LAG-3, lymphocyte-activation gene 3; MDSC, myeloid-derived suppressor cells; PD-1, programmed cell death protein 1; PGE2, prostaglandin E2; TAM, tumor-associated microglia and macrophages; TGF-β, transforming growth factor-β; TIM-3, T-cell immunoglobulin and mucin domain; TME, tumor microenvironment; Treg, regulatory T cell; WHO, the World Health Organization. The figure illustrates the distinctive characteristics of GBM (WHO grade 4) that are understood to hinder the development of effective anti-tumor therapies. These include (1) an anatomical location shielded by the blood–brain barrier, (2) intra- and inter-patient tumor heterogeneity, (3) infiltrative behavior, and (4) a highly immunosuppressive TME. The latter showcases the presence of GBM-driven cytokines with immunosuppressive and tumor-promoting properties, along with immunosuppressive cell populations such as Tregs and M1-like TAMs, accompanied by upregulated exhaustion markers. Additionally, GBMs strategically downregulate antigen-processing and presentation molecules to effectively evade T cell activation. The image was created using BioRender (https://www.biorender.com/, accessed on 18 December 2023).
Cancers 16 01276 g001
Figure 2. Overview of the main immunotherapeutic modalities against GBM.
Figure 2. Overview of the main immunotherapeutic modalities against GBM.
Cancers 16 01276 g002
Figure 3. Past and ongoing clinical trials in virus-based therapies for GBM. This figure provides a comprehensive overview of the clinical studies investigating oncolytic virus (OV) or non-lytic viral vectors for the treatment of GBM. A check mark under the “OV” section signifies the virus is oncolytic, while a cross mark indicates its use as a non-lytic viral vector. The image was created with BioRender (https://www.biorender.com/, accessed on 18 December 2023).
Figure 3. Past and ongoing clinical trials in virus-based therapies for GBM. This figure provides a comprehensive overview of the clinical studies investigating oncolytic virus (OV) or non-lytic viral vectors for the treatment of GBM. A check mark under the “OV” section signifies the virus is oncolytic, while a cross mark indicates its use as a non-lytic viral vector. The image was created with BioRender (https://www.biorender.com/, accessed on 18 December 2023).
Cancers 16 01276 g003
Table 1. List of clinical trials involving ICIs in adult GBM patients. The table includes concluded or terminated studies, as well as those currently underway or preparing to enroll participants. Data were sourced from Clinicaltrials.Gov, retrieved on 13 December 2023.
Table 1. List of clinical trials involving ICIs in adult GBM patients. The table includes concluded or terminated studies, as well as those currently underway or preparing to enroll participants. Data were sourced from Clinicaltrials.Gov, retrieved on 13 December 2023.
InhibitorNCT
Number
PhaseStudy
Status
Tumor TargetInterventionOutcome
Anti-CTLA-4
(Ipilimumab)
NCT050749922TerminatedndGBMIpi
Anti-IDO1
(Indoximod)
NCT02052648 [130]1/2CompletedMalignant Brain TumorsIND + TMZ
IND + TMZ + Bev
IND + TMZ + Stereotactic RT
Anti-PD-1
(Nivolumab)
NCT026486331TerminatedrGBMValproate + Stereotactic RT + Nivo
NCT02550249 [131]2CompletedGBMNeoadjuvant NivomOS: 7.3 months (95% CI, 5.4–7.9), mPFS: 4.1 months (95% CI, 2.8–5.5)
NCT02335918 [132]2CompletedAdvanced Solid TumorsNivo + VarlilumabOS-12: 40.9%
NCT03890952 [133]2Active Not RecruitingrGBMNivo + Bev + Surgery
Nivo + Bev
NCT04195139 [134]2Active Not RecruitingndGBMRT + TMZ + NivomOS: 11.8 months, PFS-6: 64%
RT + TMZmOS: 12.0 months, PFS-6: 49%
NCT037436622Active Not RecruitingrGBM (MGMT-M)RT + Bev + Nivo
RT + Bev + Nivo + Surgery
NCT03452579 [135,136]2Active Not RecruitingrGBMNivo + Bev (10 mg/Kg)OS-12: 41.1%, OS-12 (age > 60 year): 46.2%, OS-12 (age ≤ 60 years): 35.6%.
Nivo + Bev (3 mg/Kg)OS-12: 37.7%, OS-12 (age > 60 year): 23.8%, OS-12 (age ≤ 60 years): 56.4%.
NCT047041542Active Not RecruitingRecurrent or Metastatic TumorsNivo + Regorafenib
NCT059096182Not Yet RecruitingGBM and Brain Metastases (MGMT-UN)Crizanlizumab
Crizanlizumab + Nivo
NCT02617589 [137]3CompletedndGBM (MGMT-UN)Nivo + RTmPFS: 6.0 months (95% CI, 5.7–6.2), mOS: 13.4 months (95% CI, 12.6–14.3)
TMZ + RTmPFS: 6.2 months (95% CI, 5.9–6.7), mOS: 14.9 months (95% CI, 13.3–16.1)
NCT02667587 [138]3Active Not RecruitingndGBM (MGMT-M)RT + TMZ + NivomPFS: 10.64 months (95% CI, 8.90–11.79), mOS: 28.91 months (95% CI, 24.38–31.57),
RT + TMZ + PlacebomPFS: 10.32 months (95% CI, 9.69–12.45), mOS: 32.07 months (95% CI, 29.37–33.77),
Anti-PD-1
(Pembrolizumab)
NCT028526551CompletedrGBMPembro
NCT02054806 [139]1CompletedAdvanced Solid TumorsPembrorGBM = mPFS: 2.8 months (95% CI, 1.9–8.1), mOS: 13.1 months (95% CI, 8.0–26.6)
NCT057009551RecruitingrGBMPembro + TMZ
NCT025305021TerminatedndGBMPembro + TMZ + RT
NCT03722342 [140]1Active Not RecruitingrGBMPembro + Olinvacimab
NCT03426891 [141]1CompletedndGBMPembro + Vorinostat + TMZ + RT
NCT02311582 [142,143]1/2Active Not RecruitingRecurrent Malignant GliomasPembro + LITTmPFS: 10.5 months, mOS: 11.4 months
PembromPFS: 2.1 months, mOS: 5.2 months
NCT03277638 [144]1/2RecruitingrGBMPembro (7 days before LITT)
Pembro (14 days after LITT)
Pembro (35 days after LITT)
NCT049773751/2RecruitingrGBMPembro + Stereotactic RT + Surgery
NCT024303631/2UnknownGBM or
Gliosarcoma
Pembro
Pictilisib
NCT050538801/2UnknownrGBMPembro
Pembro + ACT001
NCT04121455 [145,146]1/2Active Not RecruitingndGBM (MGMT-UN)NOX-A12 (200 mg) + RT
NOX-A12 (400 mg) + RT
NOX-A12 (600 mg) + RT
NOX-A12 (600 mg) + RT + Bev
NOX-A12 (600 mg) + RT
NOX-A12 (600 mg) + RT + Pembro
NCT059739031/2Not Yet RecruitingrGBMLenvatinib + Pembro + TTF
NCT02628067 [147]2RecruitingAdvanced Solid TumorsPembroGlioma = mPFS: 1.4 (95% CI, 1.0–2.1), mOS: 5.6 months (95% CI, 2.6–16.2)
NCT02337491 [148,149]2CompletedrGBMPembro + BevPFS-6: 26% (95% CI, 16.3–41.5), mOS: 8.8 months (95% CI, 7.7–14.2)
PembroPFS-6: 6.7% (95% CI, 1.7–25.4), mOS: 10.3 months (95% CI, 8.5–12.5)
NCT03661723 [150]2Active Not RecruitingrGBMPembro + RT (lead-in)ORR: 3.3%, OS-6: 83.3 (95% CI, 6.5–92.7), OS-12: 40.0 (95% CI, 22.8–56.6)
Pembro + Bev + RT (lead-in)ORR: 10.0%, OS-6: 56.7 (95% CI, 37.3–72.1), OS-12: 16.6 (95% CI, 6.0–31.7)
Pembro + RT
Pembro + Bev + RT
NCT054638482RecruitingrGBMPembro + Olaparib + TMZ (Safety Lead In)
Pembro + Olaparib + TMZ (Surgical Cohort)
Pembro (Surgical Cohort)
NCT033476172Active Not RecruitingndGBMFerumoxytol MRI + Pembro
NCT031975062SuspendedndGBMPembro + Surgery + TMZ + RT
Pembro + TMZ + RT
NCT058791202Not Yet RecruitingrGBMMRgFUS + Neoadjuvant Pembro + Adjuvant Pembro
Neoadjuvant Pembro + Adjuvant Pembro
NCT03405792 [151]2Active Not RecruitingndGBMTTF + TMZ + PembromPFS: 12.0 months, PFS-12: 50.0%, mOS: 24.8 months, OS-24: 52.4%
TTF + TMZmPFS: 5.8 months, PFS-12: 28.2%, mOS: 14.7 months, OS-24: 12%
NCT02337686 [152]2Active Not RecruitingrGBMPembro + SurgerymPFS: 4.5 months (95% CI, 2.27–6.83), PFS-6: 40%, mOS: 20 months, estimated OS-12: 63%
NCT05465954 [153]2RecruitingrGBMPembro + Efineptakin alfa
NCT03797326 [154]2Active Not RecruitingSolid Tumors Pembro + Lenvatinib
Lenvatinib
NCT052357374RecruitingndGBMNeoadjuvant Pembro + Adjuvant Pembro + SOC
Neoadjuvant Pembro + SOC
SOC
Anti-PD-L1
(Avelumab)
NCT03047473 [155]2CompletedndGBMAvelumabORR: 23.3%, mPFS: 9.7 months (95% CI, 8.2–15.5), mOS: 15.3 months (95% CI, 10.7–21.5)
Anti-PD-L1 (Atezolizumab)NCT054232101Active Not RecruitingndGBMAtezo + Fractionated Stereotactic RT
NCT041604941Active Not RecruitingRecurrent GliomasD2C7-IT (6.92 μg/mL) + Atezo
D2C7-IT (4.61 μg/mL) + Atezo
NCT031583891/2CompletedndGBM (MGMT-UN)APG101 + RT
Alectinib + RT
Idasanutlin + RT
Atezo + RT
Vismodegib + RT
Temsirolimus + RT
Palbociclib + RT
NCT03673787 [156]1/2RecruitingAdvanced Solid TumorsAtezo + Ipatasertib
NCT03174197 [157]1/2Active Not RecruitingndGBMAtezo + TMZ + RTmOS: 17.1 months (95% CI, 13.9-N/A), mPFS: 9.7 months (95% CI, 7.6–15), mPFS (MGMT-M): 16.7 months (95% CI, 7.85-N/A), mPFS (MGMT-UN): 7.9 months (95% CI, 6.70–12.4)
NCT050392811/2RecruitingrGBMAtezo + Cabozantinib
NCT060697262Not Yet RecruitingrGBMPre-Surgery Atezo
NCT047299592SuspendedrGBMAtezo + Tocilizumab + Stereotactic RT
Atezo + Tocilizumab + Stereotactic RT + Surgery
Anti-PD-L1
(Durvalumab)
NCT02336165 [158]2CompletedGBMndGBM = Durva + RTOS-12: 60% (90% CI, 46.1–71.4)
Bev-Naïve rGBM = DurvaPFS-6: 19.4% (90% CI, 9.3–32.1)
Bev-Naïve rGBM = Durva + Bev (10 mg/Kg)PFS-6: 15.2% (90% CI, 6.7–26.8)
Bev-Naïve rGBM = Durva + Bev (3 mg/Kg)PFS-6: 17.2% (90% CI, 7.7–29.7)
Bev-Resistant rGBM = Durva + BevOS-6: 36.4% (80% CI, 23.5–49.3)
Anti-PD-1 +
Anti-CTLA-4
NCT02311920 [159]1CompletedndGBM
or
Gliosarcoma
TMZ + Ipi
TMZ + Nivo
TMZ + Ipi + Nivo
NCT046063161RecruitingrGBMNivo + Ipi
Nivo + Placebo
Placebo
NCT03233152 [160]1Active Not RecruitingrGBMNivo + IpimPFS: 11.7 weeks (2–152), mOS: 38 weeks (95% CI, 27–49),
NCT060979751Not Yet RecruitingrGBMNivo + Ipi
NCT033677152CompletedndGBM (MGMT-UN)Nivo + Ipi + Short-Course RTOS-12: 80%, mOS: 16.85 months (4.5–32.9), mPFS: 5.92 months (1.5–13.9)
NCT034307912TerminatedrGBMTTF + Nivo
TTF + Nivo + Ipi
NCT048172542RecruitingndGBMNivo + Ipi (1 mg/Kg) + TMZ
Nivo + Ipi (3 mg/Kg) + TMZ
NCT041451152RecruitingrGBMNivo + Ipi
NCT043968602/3Active, not recruitingndGBM (MGMT-UN)RT + TMZ
RT + Nivo + Ipi
NCT02017717 [161,162]3Active, not recruitingrGBMNivoOS-12: 41.8% (95% CI, 34.7–48.8), mOS: 9.8 months (95% CI, 8.2–11.8), mPFS: 1.51 months (95% CI, 1.48–1.61)
Nivo + Ipi
BevOS-12: 42.4% (95% CI, 34.9–49.6), mOS: 10.05 months (95% CI, 9–11.99), mPFS: 3.61 months (95% CI, 2.99–4.6)
Anti-PD-1 +
Anti-GITR
NCT04225039 [163]2Active, not recruitingrGBMRetifanlimab + INCAGN01876 + Stereotactic RTmPFS: 3.9 months (95% CI, 2.1–6.2), mOS: 9.4 months (95% CI, 8.2–10.6)
Retifanlimab + INCAGN01876 + Stereotactic RT prior to SurgerymPFS: 11.7 months, mOS: 20.1 months
Retifanlimab + INCAGN01876 prior to SurgerymPFS: 2.0 months, mOS: 9.4 months
Anti-PD-1 +
Anti-IDO1
NCT04047706 [164]1Active, not recruitingndGBMRT + TMZ + Nivo + BMS-986205
RT + Nivo + BMS-986205
NCT02327078 [165]1/2CompletedAdvanced TumorsNivo + Epacadostat
Anti-PD-1 +
Anti-LAG-3
NCT03493932 [166]1CompletedGBMNivo + Relatlimab
NCT02658981 [167]1CompletedrGBMBMS-986016
BMS-986016 + Nivo
Anti-PD-1 +
Anti-TIGIT
NCT046565350/1RecruitingGBMDomvanalimab + Placebo
Zimberelimab + Placebo
Domvanalimab + Zimberelimab
Placebo
NCT048263931Active Not RecruitingRecurrent GliomasCemiplimab + ASP8374
Anti-PD-1 +
Anti-TIM-3
NCT039619711Active Not RecruitingrGBMSpartalizumab + Sabatolimab + Stereotactic RT
Anti-PD-1 +
Anti-GITR or
Anti-IDO1 or Anti-CTLA-4
NCT037074571TerminatedrGBMNivo + MK-4166
Nivo + Epacadostat
Nivo + Ipi
Anti-PD-L1 +
Anti-CTLA-4
NCT027948832CompletedRecurrent Malignant GliomasSurgery + DurvamOS: 11.71 (95% CI, 8.332–32.71), mPFS: 4.356 (95% CI, 2.941–32.74)
Surgery + Tremelimumab mOS: 7.246 (95% CI, 2.746–16.32), mPFS: 2.746 (95% CI, 2.68–8.727)
Surgery + Durva + TremelimumabmOS: 7.703 (95% CI, 7.41–40.14), mPFS: 4.913 (95% CI, 2.905–120.4)
VariousNCT060473791/2Not Yet RecruitingMalignant Gliomas
or
Brain Metastases
NEO212 + Ipi
NEO212 + Pembro
NEO212 + Nivo
NEO212 + Regorafenib
NEO212 + CarbolaUn + Paclitaxel
NEO212 + FOLFIRI + Bev
NEO212
NEO212 + SOC
Atezo, Atezolizumab; Bev, Bevacizumab; CI, confidence interval; CTLA-4, cytotoxic T-lymphocyte-associated protein 4; Durva, Durvalumab; GITR, glucocorticoid-induced TNFR-related protein; IDO1, Indoleamine 2,3-dioxygenase 1; IND, Indoximod; Ipi, Ipilimumab; LAG-3, Lymphocyte-Activation Gene 3; mOS, median overall survival; mPFS, median progression-free survival; MRgFUS, MRI-guided focused ultrasound; ndGBM, newly diagnosed GBM; Nivo, Nivolumab; ORR, objective response rate; OS-12, overall survival at 12 months; OS-24, overall survival at 24 months; Pembro, Pembrolizumab; PFS-6, progression-free survival at 6 months; PD-1, Programmed Cell Death-Protein 1; PD-L1, programmed Death-Ligand 1; rGBM, recurrent GBM; RT, radiotherapy; SOC, standard of care; TIGIT, T Cell immunoreceptor with Ig and ITIM domains; TIM-3, T cell immunoglobulin and mucin domain-containing protein 3; TMZ, Temozolomide; TTF, tumor-treating fields.
Table 2. List of clinical trials involving vaccination strategies in adult GBM patients. The table includes concluded or terminated studies, as well as those currently underway or preparing to enroll participants. Data were sourced from ClinicalTrials.gov, retrieved on 13 December 2023.
Table 2. List of clinical trials involving vaccination strategies in adult GBM patients. The table includes concluded or terminated studies, as well as those currently underway or preparing to enroll participants. Data were sourced from ClinicalTrials.gov, retrieved on 13 December 2023.
AntigenVaccine/
Delivery
NCT
Number
PhaseStudy StatusTumor TargetInterventionOutcome
CD133DC vaccine NCT02049489
[186]
1 CompletedrGBM ICT-121
CMV-pp65 Peptide VaccineNCT018540991 WithdrawnndGBMTMZ (5 Day) + PEP-CMV (Day 6–8)
TMZ (5 Day) + PEP-CMV (day 22–24)
TMZ (21 Day) + PEP-CMV (day 22–24)
Peptide VaccineNCT028643681 TerminatedndGBMTd + TMZ (5 Day) + PEP-CMV (Component A + Component B) + Td
Td + TMZ (21 Day) + PEP-CMV (Component A + Component B) + Td
Td + TMZ (5 Day) + PEP-CMV (Safety Cohort) + Td
Td + TMZ (5 Day) + PEP-CMV (Component A) + Td
Td + TMZ (21 Day) + PEP-CMV (Component A) + Td
DC vaccine NCT049634131 Active, not recruitingndGBMCMV-DC + GM-CSF
DC vaccine NCT00693095
[187]
1 CompletedndGBMCMV-ALT + CMV-DC
CMV-ALT
DC vaccine NCT00626483
[188]
1 CompletedndGBMCMV-DC + GM-CSF + BasiliximabmOS: 5.6 months (95% CI, 3.6–9.9), mPFS: 7.7 months (95% CI, 3.4–13.8)
DC vaccine NCT047419841 WithdrawnndGBM (MGMT-UN)Monocyte loaded with mRNA encoding for CMV-pp65 (MT-201)
DC vaccine NCT00639639
[189,190]
1CompletedndGBMCMV-ALT + CMV-DC + Unpulsed DCs (or Td)
CMV-DC + Unpulsed DCs (or Td)
CMV-DC + GM-CSF + Unpulsed DCs (or Td)
DC vaccine NCT02465268
[191]
2Active, not recruitingndGBMTd + TMZ + Short-Length CMV-DC + GM-CSF
Td + TMZ + Full-Length CMV-DC + GM-CSF
Unpulsed PBMCs
DC vaccine NCT02366728
[192,193]
2CompletedndGBMCMV-DCmOS: 16 months (95% CI, 12.8–25.5), mPFS: 6.5 months (95% CI, 4.4–12.1)
CMV-DC + Td mOS: 20 months (95% CI, 16.7–25.6), mPFS: 6.7 months (95% CI, 4.6–15.2)
CMV-DC + Td + Basiliximab mOS: 19 months (95% CI, 10.2-N/A), mPFS: 7.1 months (95% CI, 6-N/A)
Liposome NCT045731401 RecruitingndGBM (MGMT-UN)Liposome loaded with mRNA encoding for CMV-pp65 (RNA-LP)
EGFRvIIIPeptide VaccineNCT00626015 [194]1 CompletedndGBM (EGFRvIII+)Rindopepimut + TMZ + Daclizumab
Rindopepimut + TMZ + Placebo
Rindopepimut + Basiliximab
Peptide VaccineNCT01498328
[195]
2CompletedrGBM (EGFRvIII+)Bev-Naïve = Bev + Rindopepimut + GM-CSFPFS-6: 28%, ORR: 30%, mDOR: 7.8 months (95% CI, 3.5–22.2)
Bev-Naïve = Bev + KLHPFS-6: 16%, ORR: 18%, mDOR: 5.6 months (95% CI, 3.7–7.4)
Bev-Resistant = Bev + Rindopepimut + GM-CSF
Peptide VaccineNCT00458601 [196]2CompletedndGBM (EGFRvIII+)Rindopepimut + GM-CSF + TMZmOS: 21.8 months, OS-36: 26%
Peptide VaccineNCT00643097 [197,198,199]2CompletedndGBM (EGFRvIII+)Rindopepimut + GM-CSF mPFS: 14.2 (95% CI, 9.9–17.6)
Rindopepimut + GM-CSF + TMZ (5 Day, 200 mg/m2)mPFS: 12.1 (95% CI, 10.5–23.7)
Rindopepimut + GM-CSF + TMZ (21 Day, 100 mg/m2)mPFS: 11.6 (95% CI, 8.1–12.7)
Peptide VaccineNCT01480479
[200]
3 CompletedndGBM (EGFRvIII+)Rindopepimut + GM-CSF + TMZmOS: 20.1 months (95% CI, 18.5–22.1)
KLH + TMZmOS: 20.0 months (95% CI, 18.1–21.9)
HSPPC-96Peptide VaccineNCT00293423 [201,202]1/2CompletedRecurrent GliomasHSPPC-96 VaccineOS-12: 29.3% (95% CI, 16.6–45.7), mOS: 42.6 weeks (95% CI, 34.7–50.5)
Peptide VaccineNCT00905060
[203]
2CompletedndGBMHSPPC-96 Vaccine + TMZmOS: 23.8 months (95% CI, 9.8–30.2), mPFS: 18 (95% CI, 12.4–21.8)
Peptide VaccineNCT01814813
[204]
2TerminatedrGBM HSPPC-96 Vaccine + Concomitant BevmOS: 6.6 months (95% CI, 5.4–10.4), mPFS: 3.7 months (95% CI, 2.9–5.4)
HSPPC-96 Vaccine + Bev At ProgressionmOS: 9.2 months (95% CI, 5.7–11.6), mPFS: 2.5 months (95% CI, 2.0–3.5)
BevmOS: 10.7 months (95% CI, 8.8–17.2), mPFS: 5.3 months (95% CI, 3.7–8.0)
hTERTPeptide VaccineNCT000699401CompletedSarcoma and Brain Tumors (HLA-A2+) 540–548 hTERT Vaccine + GM-CSF
Peptide VaccineNCT04280848
[205]
2Active, not recruitingndGBM (MGMT-UN)MGMT-UN = UCPVaxmPFS: 8.9 months (95% CI, 7.6–10.6), mOS: 17.9 months (95% CI, 16–23), OS-24: 26%
MGMT-UN or MGMT m = UCPVax + TMZ
SurvivinPeptide VaccineNCT01250470
[206]
1 CompletedRecurrent Malignant Gliomas SurVaxM/Montanide ISA-51 + GM-CSFmPFS: 17.6 weeks, mOS: 86.6 weeks
Peptide VaccineNCT05163080 [207]2RecruitingndGBMSurVaxM/Montanide ISA-51 + GM-CSF + TMZ
Placebo/Montanide ISA-51 + GM-CSF + TMZ
Peptide VaccineNCT02455557 [208]2Active, not recruitingndGBMSurVaxM/Montanide ISA-51 + GM-CSF + TMZPFS-6: 95% (95% CI, 86–98), mPFS: 11.4 months, mOS: 25.8 months (95% CI, 19.5–43.5)
AIM-2, MAGE-1, HER2/neu, TRP-2, gp100, and IL-13Rα2DC vaccine NCT01280552
[209]
2CompletedndGBMICT-107mOS: 18.3 months (95% CI, 14.9–21.2), mPFS: 11.2 months (95% CI, 8.2–13.0)
Unpulsed DCsmOS: 16.7 months (95% CI, 12.3–23.0), mPFS: 9.0 months (95% CI, 5.5–10.3)
NCT025461023SuspendedndGBMICT-107 + TMZ
Placebo + TMZ
EGFRvIII, IL-13Rα2, EphA2, HER2/neu, YKL-40Peptide VaccineNCT027543622WithdrawnrGBM Bev + Multipeptide Vaccine + Poly-ICLC
EphA2, CMV-pp65, and SurvivinPeptide VaccineNCT052831091RecruitingndGBM (MGMT-UN)P30-EPS + Poly-ICLC
BCAN, CSPG4, FABP7, IGF2BP3, NLGN4X, NRCAM, PTPRZ1 (2 peptides), and TNCPeptide VaccineNCT014032851 TerminatedGBM (HLA-A2+)IMA950 + GM-CSF + Imiquimod + Cyclophosphamide
Peptide VaccineNCT01222221
[210]
1CompletedndGBM (HLA-A2+)IMA950 + GM-CSF + Chemoradiotherapy (Vaccine before TMZ)mOS: 14.4 months
IMA950 + GM-CSF + Chemoradiotherapy (Vaccine after TMZ)mOS: 15.7 months
Peptide VaccineNCT01920191
[211,212]
1/2CompletedndGBM (HLA-A2+)IMA950 + Poly-ICLCmOS: 19 months (95% CI: 17.25–27.87), PFS-6: 81%, mPFS: 9.5 months
WT-1, PSMA, hTERT, IL-2ElectroporationNCT03491683
[213]
1/2Active, not recruitingndGBMMGMT-UN = INO-5401 + INO-9012 + Cemiplimab + RT + TMZ mOS: 17.9 months (95% CI, 14.5–19.8)
MGMT m = INO-5401 + INO-9012 + Cemiplimab + RT + TMZ mOS: 32.5 months (95% CI, 18.4-N/A)
Tumor LysateDC vaccine NCT01171469
[214]
1 CompletedRecurrent or Progressive Malignant GliomasDCs pulsed with Tumor Lysate (from BTSCs) + Imiquimod
DC vaccine NCT00068510
[215]
1 CompletedMalignant Gliomas DCs pulsed with Tumor Lysate
DC vaccine NCT018088201 CompletedMalignant GliomasDCs pulsed with Tumor Lysate + Imiquimod
DC vaccine NCT02010606
[216]
1 CompletedGBM ndGBM = DCs pulsed with Tumor Lysate (from Allogeneic Stem-like Cells) + RT + TMZmPFS: 8.75 months, mOS: 20.36 months
rGBM = DCs pulsed with Tumor Lysate (from Allogeneic Stem-like Cells) + Bev (optional) mPFS: 3.23 months, PFS-6: 24%, mOS: 11.97 months
DC vaccineNCT01213407
[217]
2CompletedMalignant GliomasSOC + DCs pulsed with Tumor Lysate (Trivax)
SOC
DC vaccine NCT01006044
[218]
2CompletedGBM DCs pulsed with Tumor Lysate mPFS: 12.7 months (95% CI, 7–16), mOS: 23.4 months (95% CI, 16–33.1)
DC vaccine NCT00323115
[219]
2CompletedndGBMDCs pulsed with Tumor Lysate + RT + TMZPFS-6: 90%, mPFS: 9.5 months, mOS: 28 months
DC vaccineNCT00045968
[220,221]
3Active, not recruitingGBM DCs pulsed with Tumor Lysate (DCVax-L)ndGBM = mOS: 19.3 months (95% CI, 17.5–21.3)
rGBM = mOS: 13.2 months (95% CI, 9.7–16.8)
Unpulsed PBMCsndGBM = mOS: 16.5 months (95% CI, 16.0–17.5)
rGBM = mOS: 7.8 months (95% CI, 7.2–8.2)
PersonalizedPeptide VaccineNCT02149225
[222,223]
1 CompletedndGBMAPVAC1/APVAC2 + Poly-ICLC + GM-CSF + TMZmPFS: 14.2 months, mOS: 29 months
Peptide VaccineNCT025109501 TerminatedndGBMPersonalized Peptide Vaccine + Poly-ICLC + TMZ
Peptide VaccineNCT03223103
[224]
1 Active, not recruiting ndGBMMutation-derived Tumor Antigen Vaccine + Poly-ICLC + TTFEstimated PFS-12: 62.5%, estimated OS-12: 83.3%
Peptide VaccineNCT055572401 RecruitingndGBMNPVAC1 + Poly-ICLC + TMZ
NPVAC2 + Poly-ICLC + TMZ
ElectroporationNCT040157001 Active, not recruitingndGBM (MGMT-UN)Personalized DNA Vaccine (GNOS-PV01) + INO-9012
Peptide Vaccine[225]3 ConcludedrGBM (HLA-A24+)Personalized Peptide Vaccine mOS: 8.4 months (95% CI, 6.6–10.6)
PlacebomOS: 8.0 months (95% CI, 4.8–12.9)
N/APeptide VaccineNCT048425131 RecruitingndGBM (HLA-A2+, MGMT-M)Multipeptide Vaccine + XS15 + Montanide ISA-51
DC vaccine NCT049683661 RecruitingndGBMDCs pulsed with Multiple Neopeptides + TMZ
DC vaccine NCT00612001
[215]
1 CompletedMalignant Gliomas DCs pulsed with Multiple Glioma-associated Peptides
DC vaccine NCT00890032
[226]
1 CompletedrGBM DCs pulsed with mRNA from BTSCsmPFS: 3.2 months (95.0% CI, 1.8–7.2), mOS: 11 months (95.0% CI, 8.2–14.8)
DC vaccine NCT028205841 CompletedrGBM DCs pulsed with mRNA from Glioma Stem Cells
DC vaccine NCT00846456
[227]
1/2CompletedGBM DCs pulsed with mRNA from Glioma Stem CellsmOS (treated group): 759 days, mOS (control group): 585 days
DC vaccine NCT00576641
[228]
1 CompletedBrain Stem Glioma and GBMDCs pulsed with Tumor Peptides
Bev, Bevacizumab; BTSC, brain tumor stem cell; CAR-T, chimeric antigen receptor T cell; CI, confidence interval; CMV-ALT, CMV-autologous lymphocyte transfer; DC, dendritic cell; GM-CSF, granulocyte-macrophage colony-stimulating factor; KLH, Keyhole Limpet Haemocyanin; mDOR, median duration of response; mOS, median overall survival; mPFS, median progression-free survival; ndGBM, newly diagnosed GBM; NPVAC, NeoPep vaccine; ORR, objective response rate; OS-12, overall survival at 12 months; OS-24, overall survival at 24 months; PBMC, peripheral blood mononuclear cell; Poly-ICLC, polyinosinic–polycytidylic acid stabilized with polylysine and carboxymethylcellulose; PFS-12, progression-free survival at 12 months; PFS-6, progression-free survival at 6 months; rGBM, recurrent GBM; RT, radiotherapy; SOC, standard of care; Td, tetanus and diphtheria toxoid; TMZ, Temozolomide.
Table 3. List of clinical trials involving adoptive T cell therapies in adult GBM patients. The table includes concluded or terminated studies, as well as those currently underway or preparing to enroll participants. Data were sourced from ClinicalTrials.gov, retrieved on 13 December 2023.
Table 3. List of clinical trials involving adoptive T cell therapies in adult GBM patients. The table includes concluded or terminated studies, as well as those currently underway or preparing to enroll participants. Data were sourced from ClinicalTrials.gov, retrieved on 13 December 2023.
AntigenNCT
Number
PhaseStudy
Status
Tumor
Target
InterventionOutcome
Monovalent CAR-TB7-H3NCT052413921RecruitingrGBM B7-H3 CAR-T
NCT053661791RecruitingrGBM B7-H3 CAR-T
NCT054743781RecruitingrGBM B7-H3 CAR-T
NCT043851731RecruitingrGBM or Refractory GBMB7-H3 CAR-T + TMZ
NCT040778661/2RecruitingrGBM or Refractory GBMTMZ
TMZ + B7-H3 CAR-T
CD70NCT053535301RecruitingndGBM (MGMT-UN, CD70+)CD70 CAR-T
EGFRvIIINCT058026931Not yet recruitingrGBM (EGFRvIII+)EGFRvIII CAR-T
NCT02209376 [253,254,255]1TerminatedrGBM EGFRvIII CAR-TmOS: 251 days
NCT02664363 [256]1TerminatedndGBM (EGFRvIII+)EGFRvIII CAR-T
NCT028440621Unknown rGBM (EGFRvIII+)EGFRvIII CAR-T
NCT032836311TerminatedrGBM EGFRvIII CAR-T
NCT050636821Unknown Leptomeningeal GBM (EGFRvIII+) EGFRvIII CAR-T
NCT056603691RecruitingGBMEGFRvIII BiTE-secreting CAR-T
NCT05024175ObservationalNot yet recruitingGBM/
NCT01454596 [257]1/2CompletedMalignant Gliomas (EGFRvIII+)EGFRvIII CAR-TmOS: 6.9 months (2.8–10)
NCT039416261/2Unknown Solid Tumors (EGFRvIII+)EGFRvIII CAR-T
NCT036382061/2Unknown Solid Tumors (EGFRvIII+)EGFRvIII CAR-T
EMMPRINNCT040458471Unknown Recurrent Malignant Gliomas (CD147+)EMMPRIN CAR-T
GD2NCT03170141 [258]1Enrolling by invitationrGBM (GD2+)GD2 CAR-TmOS = 10 months (3–24)
HER2/neuNCT01109095 [259]1CompletedGBMHER2 CAR-T
NCT033892301Active, not recruitingRecurrent or Refractory Malignant GliomasHER2 CAR-T
IL-13Rα2NCT02208362 [260]1Active, not RecruitingRecurrent Malignant GliomasIL-13Rα2 CAR-T (intratumoral)
IL-13Rα2 CAR-T (intracavitary)
IL-13Rα2 CAR-T (intraventricular)
IL-13Rα2 CAR-T (intratumoral/intraventricular)
NCT046613841RecruitingLeptomeningeal GBM, Ependymoma, or MedulloblastomaIL-13Rα2 CAR-T
NCT055408731RecruitingRecurrent or Refractory Malignant GliomasIL-13Rα2 CAR-T
NCT00730613 [261]1CompletedRecurrent or Refractory Malignant GliomasIL-13Rα2 CTLs
MMP2 (Chlorotoxin)NCT042143921RecruitingrGBM (MMP2+)MMP2 CAR-T (intratumoral)
MMP2 CAR-T (intratumoral/intraventricular)
NCT05627323 [262]1RecruitingrGBM (MMP2+)MMP2 CAR-T
NKG2DNCT042704611WithdrawnRecurrent Solid Tumors (NKG2DL+)NKG2D CAR-T
NCT051317631RecruitingRecurrent Solid Tumors (NKG2DL+)NKG2D CAR-T
NCT04717999N/ANot yet recruitingrGBM NKG2D CAR-T
NCT045506631Unknown Relapsed or Refractory Solid Tumors (NKG2DL+)NKG2D CAR-T
PD-L1NCT029378441Unknown rGBM PD-L1 CAR-T
SNC-109NCT058680831RecruitingrGBM SNC-109 CAR-T
Polyvalent CAR-TIL-7Ra, CD44 and CD133NCT055770911Not yet recruitingrGBM Tris-CAR-T
EGFRvIII, IL-13Rα2, HER2/neu, EphA2, CD133, GD2NCT03423992 [263]1Unknown Recurrent Malignant GliomasPersonalized CAR-TmOS (EphA2-specific CAR-T) = 86–181 days
TILs NCT053335881RecruitingGBMTILs
NCT03347097 [264]1Unknown rGBMTILsmOS: 16.1 months
PD-1-TILsmOS: 11.2 months
NCT049439131RecruitingGliomasTILs
BiTE, bispecific T-cell engager; CAR-T, chimeric antigen receptor T cell; MGMT-unmethylated, MGMT-UN; mOS, median overall survival; ndGBM, newly diagnosed GBM; rGBM, recurrent GBM; TIL, tumor-infiltrating lymphocyte.
Table 4. List of clinical trials involving virus-based therapies in adult GBM patients. The table includes concluded or terminated studies, as well as those currently underway or preparing to enroll participants. Data were sourced from ClinicalTrials.gov, retrieved on 13 December 2023.
Table 4. List of clinical trials involving virus-based therapies in adult GBM patients. The table includes concluded or terminated studies, as well as those currently underway or preparing to enroll participants. Data were sourced from ClinicalTrials.gov, retrieved on 13 December 2023.
Virus Name Virus Type NCT
Number
Phase Study
Status
Tumor
Target
Intervention Outcome
AdenovirusOVCRAd-S-pk7NCT051390561RecruitingRecurrent Malignant GliomasNSC-expressing CRAd-S-pk7
NCT03072134 [281]1CompletedNewly Diagnosed Malignant GliomasNSC-expressing CRAd-S-pk7mPFS: 9.1 months (95% CI, 8.5–36), mOS: 18.4 months (95% CI, 6.5–36)
DNX-2401NCT03896568 [282]1RecruitingRecurrent Malignant GliomasBM-hMSCs loaded with DNX-2401
NCT02197169 [283]1CompletedrGBM or GliosarcomaDNX-2401
DNX-2401 + IFN-γ
NCT01956734 [284]1CompletedrGBMDNX-2401 + TMZ
NCT015825161/2CompletedrGBMDNX-2401
NCT00805376 [283]1CompletedRecurrent Malignant GliomasDNX-2401mOS: 9.5 months
DNX-2401 + SurgerymOS: 13 months
DNX-2440NCT037143341TerminatedrGBMDNX-2440
ONYX-015[285]1CompletedRecurrent Malignant GliomasONYX-015mOS (all patients): 6.2 months (1.3–28.0), mOS (GBM patients): 4.9 months
Non-LyticAdV-ECRT-122T NCT061025251/2Not yet recruitingGBM (hTERT+)AdV-ECRT-122T + Valganciclovir
AdV-HSV-TKNCT000028241CompletedPrimary Brain TumorsAdV-HSV-TK + Ganciclovir
NCT01811992 [286]1CompletedMalignant GliomasAdV-HSV-TK + AdV-Flt3L + ValacyclovirmOS: 21.3 months (95% CI, 11.1–26.1)
NCT00751270 [287]1CompletedMalignant GliomasResectable Gliomas = AdV-HSV-TK + Valacyclovir + RT
Unresectable Gliomas = AdV-HSV-TK + Valacyclovir + RT
NCT035960861/2RecruitingrGBMAdV-HSV-TK + Valacyclovir + Radiochemotherapy
NCT036034051/2RecruitingndGBMAdV-HSV-TK + Valacyclovir + Radiochemotherapy
NCT00870181 [288]2CompletedRecurrent Malignant GliomasAdV-HSV-TK + GanciclovirPFS-6: 71.4%, mPFS: 34.9 weeks (9.0–238.4), mOS: 45.7 weeks (9.0–238.4)
SOCPFS-6: 5.6%, mPFS: 7.4 weeks (1.1–35.3), mOS: 8.6 weeks (1.1–45.0)
NCT00589875 [289]2CompletedMalignant GliomasAdV-HSV-TK + Valacyclovir + RTmOS: 17.1 months
SOCmOS: 13.5 months
Ad-RTS-IL-12 NCT02026271
[290]
1CompletedMalignant GliomasAd-RTS-IL-12 + Veledimex
AdV-IFN-βNCT059149351 RecruitingrGBMAdV-IFN-β
NCT000310831CompletedMalignant GliomasAdV-IFN-β
AdV-p53NCT000040411CompletedRecurrent Malignant GliomasAdV-p53
NCT000040801CompletedRecurrent or Progressive Brain Tumors AdV-p53
VB-111 NCT01260506 [291]1/2CompletedrGBMVB-111 until progressionmOS: 223 days, OS-12: 18%
VB-111 upon progression + Bev (primed combination)mOS: 414 days, OS-12: 57%
VB-111 + Bev (unprimed combination)mOS: 141.5 days, OS-12: 10%
NCT02511405 [292]3CompletedrGBMVB-111 + BevmOS: 6.8 months, ORR: 27.3%
BevmOS: 7.9 months, ORR: 21.9%
Herpes Simplex VirusOVC134NCT036575761RecruitingrGBMC134
C5252NCT050954411Not yet recruitingrGBM or Progressive GBMC5252
HSV-1716NCT020319651TerminatedRecurrent Malignant GliomasHSV-1716
[293]1CompletedRecurrent Malignant GliomasHSV-1716
[294]1CompletedMalignant GliomasHSV-1716
[295]1CompletedMalignant GliomasHSV-1716
G207NCT00157703 [296]1CompletedMalignant GliomasG207 + RTmOS: 7.5 months (95% CI, 3.0–12.7)
NCT00028158 [297]1/2CompletedRecurrent Brain Tumors G207
NCT00036699 [298]1/2CompletedRecurrent Brain Tumors G207
G47ΔUMIN000002661
[299]
1/2CompletedrGBM or Progressive GBMG47ΔmOS: 30.5 (95% CI, 19.2–52.7)
M032NCT020628271Active, not recruitingRecurrent Malignant GliomasM032 (NSC 733972)
rQnestin34.5v.2NCT03152318 [300,301]1RecruitingRecurrent Malignant GliomasrQNestin34.5v.2
rQNestin34.5v.2 + Cyclophosphamide
rQNestin34.5v.2 (Multiple doses)
RetrovirusNon-LyticRV-HSV-TK[302]3CompletedndGBMSOC mOS: 354 days (95% CI, 315–372), OS-12: 55%
SOC + RV-HSV-TK + GanciclovirmOS: 365 days (95% CI, 334–416), OS-12: 50%
Toca 511NCT01985256 [303]1CompletedRecurrent Brain Tumors Toca 511 + 5-FC
NCT02576665 [304]1TerminatedSolid Tumors or LymphomasToca 511 + 5-FC
NCT01470794 [305,306]1CompletedRecurrent Malignant Brain Tumors Toca 511 + 5-FC
NCT01156584 [307]1CompletedRecurrent Malignant GliomasToca 511 + 5-FC
NCT043270111Terminated/Toca 511 + 5-FC (Long term safety follow-up)
NCT02414165 [308]2/3TerminatedRecurrent Malignant GliomasToca 511 + 5-FCmOS: 11.10 months
Lomustine, TMZ or BevmOS: 12.22 months
NCT04105374 [309]2/3WithdrawnndGBMSOC
SOC + Toca 511 + 5-FC
Measles Virus OVMV-CEANCT003902991CompletedrGBMMV-CEA (Intracavitary)PFS-6: 22.2% (95% CI, 6.6–75.4), mOS: 11.8 months (95% CI, 4.4-N/A)
MV CEA (Intratumoral/Intracavitary)PFS-6: 23.1% (95% CI, 8.6–62.3), mOS: 11.4 months (95% CI, 4.3-N/A)
Newcastle Disease Virus OVNDV-HUJNCT01174537 [310]1/2WithdrawnrGBM, Sarcoma or NeuroblastomaNDV (HUJ strain)
NDV-MTH-68/H[311]//Malignant GliomasNDV (MTH-68/H strain)
ParvovirusOVH-1PVNCT01301430 [312,313]1/2CompletedrGBM or Progressive GBMH-1PV
PoliovirusOVPVSRIPONCT01491893 [314]1CompletedrGBMPVSRIPOmOS (PVSRIPO): 12.5 months (95% CI, 9.9–15.2), mOS (historical controls): 11.3 months (95% CI, 9.8–12.5)
NCT029861782Active, not recruitingRecurrent Malignant GliomasPVSRIPO
Reovirus OVReolysinNCT005286841CompletedMalignant GliomasReolysin
[315]1CompletedRecurrent Malignant GliomasReolysinmOS: 21 weeks (6 to 234)
[316]1CompletedRecurrent Malignant GliomasReolysinmOS: 140 days (97 to 989)
[317]1CompletedMalignant Gliomas and Brain MetastasesReolysinmOS: 469 days (118 to 1079)
Vaccinia Virus OVTG6002 NCT032944861/2CompletedrGBMTG6002 + 5-FC
5-FC, 5-FluoroCytosine; AdV, Adenovirus; Bev, Bevacizumab; BM-hMSCs, allogeneic bone marrow-derived human mesenchymal stem cells; CI, confidence interval; HSV, herpes simplex virus; MGMT-methylated, MGMT-M; MGMT-unmethylated, MGMT-UN; mOS, median overall survival; OV, oncolytic virus; mPFS, median progression-free survival; MV, measles virus; ndGBM, newly diagnosed GBM; NDV, Newcastle disease virus; NSC, neural stem cells; ORR, objective response rate; OS-12, overall survival at 12 months; PFS-6, progression-free survival at 6 months; rGBM, recurrent GBM; RT, radiotherapy; RV, retrovirus; SOC, standard of care; TMZ, Temozolomide.
Table 5. List of clinical trials combining immunotherapeutic strategies in adult GBM patients. The table includes concluded or terminated studies, as well as those currently underway or preparing to enroll participants. Data were sourced from ClinicalTrials.gov, retrieved on 13 December 2023.
Table 5. List of clinical trials combining immunotherapeutic strategies in adult GBM patients. The table includes concluded or terminated studies, as well as those currently underway or preparing to enroll participants. Data were sourced from ClinicalTrials.gov, retrieved on 13 December 2023.
CombinationNCT
Number
PhaseStudy
Status
Tumor
Target
InterventionOutcome
ICT
+
ACT
Anti-PD-1 + CAR-TNCT037265151CompletedndGBM (MGMT-UN)EGFRvIII CAR-T + Pembro
NCT040036491RecruitingrGBM or Refractory GBMNivo + IL-13Rα2 CAR-T + Ipi
Nivo + IL-13Rα2 CAR-T
IL-13Rα2 CAR-T
ICT
+
Vaccine
Anti-PD-1 + CMV-DCNCT025290721CompletedRecurrent Brain TumorsNivo + Surgery + Nivo&CMV-DC
Nivo&CMV-DC + Surgery + Nivo&CMV-DC
Anti-PD-1 + HSPPC-96NCT030182882CompletedndGBM (MGMT-UN)RT + TMZ
RT + TMZ + Pembro
RT + TMZ + Pembro + HSPPC-96 Vaccine
RT + TMZ + Pembro + Placebo
Anti-PD-1 + IMA950NCT03665545 [364]1/2Active, not recruitingrGBMIMA950 + Poly-ICLC
IMA950 + Poly-ICLC + Pembro
Anti-PD-1 or Anti-CTLA-4
+ NeoVax
NCT034220941TerminatedndGBM (MGMT-UN)NeoVax + Nivo (start at time of progression)
NeoVax + Nivo (start with Cycle 1)
NeoVax + Nivo (start with Cycle 2)
NeoVax + Ipi + Nivo (start with Cycle 3)
NeoVax + Ipi + Nivo (day 1&15 each cycle)
NCT02287428 [237,365]1RecruitingndGBMRT + NeoVaxmPFS: 7.6 months (90% CI, 6.2–9.5), mOS: 16.8 months (90% CI, 9.6–21.3)
RT + Pembro followed by NeoVax + Pembro
RT followed by NeoVax + Pembro
RT + 1 dose Pembro followed by NeoVax + Pembro
MGMT m = RT + TMZ Followed by TMZ + NeoVax + Pembro
Anti-PD-1 + SurVaxMNCT04013672 [366]Phase 2Active, not recruitingrGBMPembro + SurVaxM/Montanide ISA-51 + GM-CSF (no prior immunotherapy)
Pembro + SurVaxM/Montanide ISA-51 + GM-CSF (prior failed immunotherapy)
Anti-PD-1 + DC-Tumor LysateNCT030148042WithdrawnrGBMDCVax-L
DCVax-L + Nivo
NCT042018731RecruitingrGBMPembro + ATL-DC + Poly-ICLC
Placebo + ATL-DC + Poly-ICLC
ICT
+
Virus
Anti-PD-1 + AdVNCT035766121Active, not recruitingNewly Diagnosed Malignant GliomasMGMT-UN = AdV-HSV-TK/Valacyclovir + RT + TMZ + Nivo
MGMT m and undetermined = AdV-HSV-TK/Valacyclovir + RT + TMZ + Nivo
NCT03636477 [367]1CompletedrGBM or Progressive GBMAd-RTS-IL-12 + Veledimex + NivomOS: 16.9 months
NivomOS: 9.8 months
Anti-PD-1 + HSVNCT050844301/2RecruitingRecurrent Malignant GliomasrGBM = Pembro + M032
ndGBM = Pembro + M032
NCT02798406 [337]2CompletedrGBM or GliosarcomaDNX-2401 + PembroORR: 10.4% (90% CI, 4.2–20.7), OS-12: 52.7% (95% CI, 40.1–69.2), mOS: 12.5 months (10.7–13.5)
Anti-PD-1 + PoliovirusNCT04479241 [368]2Active, not recruitingrGBMPVSRIPO + Pembro
Anti-PD-L1 + PoliovirusNCT039738791/2WithdrawnRecurrent Malignant GliomasPVSRIPO + Atezo
ACT, adoptive cell therapy; AdV, Adenovirus; Atezo, Atezolizumab; CAR-T, chimeric antigen receptor T cells; CI, confidence interval; DC, dendritic cell; HSV, herpes simplex virus; ICT, immune checkpoint therapy; Ipi, Ipilimumab; MGMT-methylated, MGMT-M; MGMT-unmethylated, MGMT-UN; mOS, median overall survival; mPFS, median progression-free survival; ndGBM, newly diagnosed GBM; ORR, objective response rate; Nivo, Nivolumab; OS-12, overall survival at 12 months; Pembro, Pembrolizumab; rGBM, recurrent GBM; RT, radiotherapy; TMZ, temozolomide.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Salvato, I.; Marchini, A. Immunotherapeutic Strategies for the Treatment of Glioblastoma: Current Challenges and Future Perspectives. Cancers 2024, 16, 1276. https://doi.org/10.3390/cancers16071276

AMA Style

Salvato I, Marchini A. Immunotherapeutic Strategies for the Treatment of Glioblastoma: Current Challenges and Future Perspectives. Cancers. 2024; 16(7):1276. https://doi.org/10.3390/cancers16071276

Chicago/Turabian Style

Salvato, Ilaria, and Antonio Marchini. 2024. "Immunotherapeutic Strategies for the Treatment of Glioblastoma: Current Challenges and Future Perspectives" Cancers 16, no. 7: 1276. https://doi.org/10.3390/cancers16071276

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop