Next Article in Journal
The Hypertensive Effect of Amphotericin B-Containing Liposomes (Abelcet) in Mice: Dissecting the Roles of C3a and C5a Anaphylatoxins, Macrophages and Thromboxane
Previous Article in Journal
Hyperreactivity of Salivary Alpha-Amylase to Acute Psychosocial Stress and Norepinephrine Infusion in Essential Hypertension
Previous Article in Special Issue
Metal-Based Nanostructured Therapeutic Strategies for Glioblastoma Treatment—An Update
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Small Molecules and Immunotherapy Agents for Enhancing Radiotherapy in Glioblastoma

1
College of Medicine, The Ohio State University, Columbus, OH 43210, USA
2
Department of Radiation Oncology, The Ohio State University Wexner Medical Center, Columbus, OH 43210, USA
3
Department of Neuro-Oncology, The Ohio State University Wexner Medical Center, Columbus, OH 43210, USA
*
Author to whom correspondence should be addressed.
Biomedicines 2022, 10(7), 1763; https://doi.org/10.3390/biomedicines10071763
Submission received: 29 June 2022 / Revised: 13 July 2022 / Accepted: 19 July 2022 / Published: 21 July 2022
(This article belongs to the Special Issue Potent Agent Research for Glioblastoma Treatment)

Abstract

:
Glioblastoma (GBM) is an aggressive primary brain tumor that is associated with a poor prognosis and quality of life. The standard of care has changed minimally over the past two decades and currently consists of surgery followed by radiotherapy (RT), concomitant and adjuvant temozolomide, and tumor treating fields (TTF). Factors such as tumor hypoxia and the presence of glioma stem cells contribute to the radioresistant nature of GBM. In this review, we discuss the current treatment modalities, mechanisms of radioresistance, and studies that have evaluated promising radiosensitizers. Specifically, we highlight small molecules and immunotherapy agents that have been studied in conjunction with RT in clinical trials. Recent preclinical studies involving GBM radiosensitizers are also discussed.

1. Introduction

Glioblastoma (GBM) is the most common malignant primary brain tumor with a dismal five-year relative survival rate of 6.8% [1]. Despite decades of preclinical and clinical studies, survival and quality of life outcomes have not substantially improved. Currently, the standard treatment approach consists of surgery, radiation treatment (RT), temozolomide (TMZ), and tumor-treating fields (TTF) [2,3]. Numerous clinical trials have attempted to further improve survival outcomes but have generally been disappointing. Although trial results have been underwhelming, there is a greater understanding of the mechanisms driving treatment resistance. Specifically, researchers have uncovered ways the GBM tumor microenvironment (TME) promotes tumorigenesis, disease progression, and radioresistance [4]. More recently, small molecules and immunotherapy agents have been designed to enhance RT efficacy by targeting GBM tumor hypoxia and metabolic reprogramming. In this review article, we aim to highlight studies that have evaluated small molecules and immunotherapy agents as radiosensitizers in GBM patients.

2. Modern Treatment Strategies for Glioblastoma

GBM is among the most challenging cancers to treat because of tumor location, tumor heterogeneity, and the infiltrating growth pattern [4]. For patients with good performance status (Karnofsky performance status ≥60), the median overall survival (OS) rate is approximately 14 months [2]. Standard treatment modalities are discussed below.

2.1. Surgical Resection

Numerous studies have found aggressive, maximal tumor resection is associated with better survival outcomes [5,6]. Unfortunately, complete resection is often not possible due to the diffuse nature of the disease, where the tumor is frequently located in or near the eloquent cortex [7]. Although new surgical techniques have been developed (e.g., neuronavigation, fluorescence, intraoperative imaging), they have failed to significantly improve the prognosis for GBM patients [8].

2.2. Chemotherapy

TMZ has replaced nitrosoureas (e.g., carmustine, lomustine) as the standard chemotherapy for patients with GBM (Figure 1) [9]. TMZ is an alkylating agent that preferentially methylates DNA at the N7 and O6 positions of guanine and the N3 position of adenine [10]. The methyl adducts result in cycles of mismatch repair with eventual strand breaks and ultimately cell cycle arrest at the G2/M phase [11]. Patients with methylation of the MGMT promoter experienced a survival benefit compared to patients who did not have a methylated promoter [12]. This phenomenon is attributed to patients with hypermethylation having decreased expression of the DNA repair enzyme [12]. Lower levels of MGMT therefore prevent mismatch repair and enhance the efficacy of TMZ.

2.3. Radiotherapy

The current standard treatment for newly diagnosed GBM is based on a randomized phase III trial led by the European Organization for Research and Treatment of Cancer (EORTC) 26981-22981/National Cancer Institute of Canada Clinical Trials Group (NCIC CTG). This study found RT plus concomitant and adjuvant TMZ led to improved median progression-free survival (PFS; 6.9 vs. 5.0 months) and median OS (14.6 vs. 12.1 months) compared to RT alone [2]. Based on these findings, GBM patients receive post-operative radiotherapy and concomitant TMZ (75 mg/m2 of body-surface area per day) [13]. The radiation treatment is typically administered over six weeks via 3-D conformal or intensity-modulated radiotherapy (IMRT). A standard treatment schedule is 2 Gy per day given Monday through Friday for a total dose of 60 Gy. Following the radiation course, maintenance TMZ (150–200 mg/m2) is given over 5 days every 28 days for 6 cycles. More recently, Stupp et al. evaluated the safety and efficacy of TTF (low-intensity, intermediate-frequency alternating electric fields) [14] following chemoradiation [3]. Median PFS and OS were significantly prolonged in the TMZ plus TTF vs. TMZ alone [3,15].
Hypofractionation is an alternative strategy where RT is administered in larger doses per fraction over fewer fractions. Recently, radiation oncologists have explored hypofractionation as a shorter, more convenient alternative, particularly in the elderly and/or frail patient population [16,17,18,19,20,21,22]. Preclinical data have suggested high dose per fraction results in a superior immunologic response that creates an abscopal effect [23,24]. Studies have shown hypofractionation has a stimulating effect on the anti-tumor immune response by inducing tumor cell death, normalizing irregular tumor vasculature, and releasing tumor-associated antigens [25,26]. Hypofractionated RT is currently being explored in combination with immunotherapy [27]. Other active areas of RT research include image-guided radiotherapy (IGRT) [28] and particle therapy [29,30,31].

3. Mechanisms of Radiation-Induced Cancer Cell Death

Radiotherapy has been a cornerstone treatment modality for treating GBM. Radiation-induced tumor cell killing can either occur via direct DNA damage or indirectly through the generation of radicals (e.g., peroxyl, hydroxyl radicals) [32]. Radiation can create DNA base lesions (e.g., 8-oxo-guanaine, formamidopyrimidines) and single- or double-strand DNA breaks. By definition, two or more lesions found within two helical turns are defined as a “clustered lesion” [33] while two strands of DNA phosphodiester backbone breaks within 10 base pairs is a double-strand break [34]. A dose of 1 Gy is estimated to produce around 3000 damaged bases, 1000 single-strand breaks, and 40 double-strand breaks [35]. Double-strand breaks are particularly difficult to repair and lead to greater cancer cell death [36].
The Bremsstrahlung process refers to inelastic interactions between an electron and nucleus that releases a photon, and this produces X-rays in linear accelerators used for radiation. Photon interactions with X-rays can be categorized into diagnostic (energy of 20–150 kV and used for imaging), superficial (50–200 kV, used for skin), orthovoltage (200–500 kV, used for skin and ribs), and megavoltage (1–25 MV, used for deep tissues). Irradiation has the ability to inhibit cancer cell proliferation through the stimulation of cell death mechanisms (e.g., apoptosis, necrosis, senescence) or by damaging cell membranes and organelles, impairing important signal transduction pathways [37,38].

4. Radioresistance

Despite advances in the field of radiation oncology (e.g., 2-dimensional WBRT to 3-D conformal RT and IMRT), radioresistance remains a challenging aspect of treating GBM patients. Radiation can induce DNA single- or double-strand breaks, leading to a decrease in radiosensitive tumor cell proliferation. DNA repair mechanisms are then activated, but cell death ultimately results if the damage is irreparable. A small subset of cells can evade apoptosis and instead become overactive. These surviving cells have alterations in tumor suppressor and oncogene expression that lead to radioresistance [39]. Furthermore, the tumor microenvironment (TME), tumor hypoxia, and glioma stem cells (GSCs) are other factors contributing to treatment failure [4]. Understanding these mechanisms has led to the rational development of drug inhibitors.

4.1. Glioma Stem Cells

Cancer stem cells (CSCs) are located within tumor masses and are able to self-renew and differentiate into various tumor cell types [40]. Researchers hypothesize these CSCs have the ability to generate the heterogeneous cell population seen in tumors. CSCs exhibit greater radioresistance due to their DNA-repair mechanisms, ROS scavenging systems, and self-renewal capabilities [41]. In GBM, these cells are known as glioma stem cells (GSCs) and have the ability to propagate as RT-resistant cells [41,42]. GSCs are able to express markers that regulate various pathways, telomerase activity, transporter proteins, cytokine secretion, and pro-angiogenic factors [43]. Glioma initiating cells (GICs) are a subpopulation of GSCs in the tumor microenvironment and play a key role in tumor heterogeneity. The heterogeneity causes challenges in treatment when there are variations in gene status (e.g., IDH, MGMT) [44].

4.2. Hypoxia

In the 1950s, Thomlinson and Gray published a landmark paper that suggested hypoxia may play a key role in tumor radioresistance [45]. Subsequent studies have supported this notion; it is estimated a radiation dose needs to be three times higher for hypoxic regions to induce the same DNA damage as in normal oxygenated regions [46]. This phenomenon is explained through the “oxygen fixation hypothesis,” where radicals are produced through ionizing radiation that interacts with neighboring oxygen, causing the formation of reactive oxygen species (ROSs) [47]. The resulting radical species then irreversibly damages DNA.
Key studies have suggested the TME in GBM plays a key role in the development of tumor hypoxia [48]. The TME is composed of stromal cells, signaling molecules, immune cells, and the surrounding extracellular matrix [49]. This complex matrix of cells creates pockets of hypoxia and acidosis via “microvascular hyperplasia” where rapidly dividing endothelial cells form microaggregates of sprouting vessels [50]. During this rapid growth, there is a complex interplay between cells and the extracellular environment that creates structural abnormalities (e.g., incomplete or absent basement membranes, irregular architecture) [51]. These abnormalities cause irregular blood flow, allowing tumor cells to invade beyond the diffusion distance of oxygen within the tissue. To supply oxygen to the tumor cells, angiogenesis is mediated by hypoxia-inducible factors (HIFs) to create new capillary systems. These inefficient capillary systems maintained by the tumor create an oxygen gradient.

5. Radiosensitizers

The rationale behind combining radiation and chemotherapy originates from the Steel paradigm [52]. Steel et al. proposed that synergy is driven by (1) spatial cooperation, (2) toxicity independence, (3) protection of normal tissues, and (4) enhancement of tumor response. The enhancement effect can be driven by inhibiting radiation-induced damage, reoxygenation following treatment, and/or improved drug access following RT.
Early studies demonstrated some chemotherapeutics such as cisplatin have the ability to sensitize tumor cells to RT, leading to greater radiation efficacy [53]. More recently, radiosensitizers have been developed that work through a variety of mechanisms: (1) Suppression of intracellular thiols or other radioprotective substances, (2) radiation-induced formation of cytotoxic substances via radiolysis of the sensitizer, (3) inhibition of the post-radiation cellular repair processes, (4) structural incorporation of thymine analogues into intracellular DNA, and (5) oxygen mimetic sensitizers [54,55].
Although other disease sites have found success with radiosensitizers, GBM has been particularly challenging due to its anatomic location (e.g., located beyond the blood–brain barrier), cell heterogeneity (e.g., cancer stem cells, tumor microtubes), and increased proliferation rate [56]. To date, TMZ is the most effective and widely used radiosensitizer in the treatment of GBM. TMZ increases the number of RT-induced double-strand DNA breaks as a result of a decrease in DNA repair capacity [57,58]. This review will focus on other small molecule and immunotherapy agents that have shown preclinical promise. Additionally, we will discuss relevant clinical trial findings.

5.1. Pyrmidine Analogues

Gemcitabine is a difluoro-pyrimidine analog that is phosphorylated and incorporated into the DNA and RNA of cancer cells, leading to chain termination (Figure 2) [59]. The radiosensitizing effects of gemcitabine result from the depletion of phosphorylated deoxynucleotides and cell-cycle redistribution into the S-phase [60,61,62]. To date, gemcitabine has demonstrated activity in breast, ovarian, non-small cell lung, pancreatic, and bladder cancers [63].
In vitro studies have determined the gemcitabine administration schedule is essential for maximal radiosensitzation. Gemcitabine achieved radiosensitization with long exposure (24 h) to low gemcitabine concentrations or brief treatments with increased concentrations [64]. Maraveyas and coworkers conducted a phase I study in brain metastases patients evaluating the maximum tolerated dose of concomitant gemcitabine and RT [65]. A phase I study then evaluated gemcitabine with concomitant RT in newly diagnosed GBM patients [66]. In this study, gemcitabine was delivered at 10 mg/m2/min on a weekly basis for 6 weeks 24 to 72 h prior to concomitant RT (60 Gy in 30 fractions) with the identification of dose-limiting toxicity and maximum tolerated dose as the primary end-points. Based on this study, 175 mg/m2/weekly was recommended for further evaluation in a phase II study. Twenty-three patients were enrolled in their phase II study and found concomitant RT and gemcitabine were well-tolerated with few severe adverse events [67]. Additionally, disease control was observed in both methylated and unmethylated MGMT promoter tumors (91% and 77.5%, respectively).
To date, there is evidence gemcitabine has the ability to cross the blood–brain barrier [68], but some drawbacks include its short plasma half-life, adverse effects related to high drug doses (e.g., myelosuppression, thrombocytopenia, edema), and resistance related to altered expression of nucleoside transporters, kinases, and enzymes [56]. Researchers are currently exploring various delivery strategies for overcoming these limitations (e.g., encapsulation, conjugation, and convention-enhanced delivery) [69,70,71]. For example, Guo et al. surmised gemcitabine coupled to a peripheral benzodiazepine receptor ligand may enhance brain tumor uptake [70]. In their xenograft model, the conjugated agent resulted in a two-fold enhancement in brain tumor selectively compared with gemcitabine alone.

5.2. Kinase Inhibitors

5.2.1. Tyrosine Kinase Inhibitors

Tyrosine kinase inhibitors (TKIs) block receptor signaling, inhibiting cell growth and proliferation. Since the approval of imatinib in 2001 for the treatment of chronic myeloid leukemia, there has been an explosion of TKI utilization in multiple types of cancer [72]. TKIs have incredible potential for treating GBM considering their ability to block cell signaling pathways such as EGFR, PDGFR, and VEGF/VEGFR.
EGFR amplification is seen in approximately 40% of GBM cases, correlating with decreased apoptosis, increased cellular proliferation, tumorigenesis, and radioresistance [73,74,75]. Erlotinib is a TKI that has demonstrated activity against the EGFRvIII mutant receptor in preclinical models [76]. Erlotinib is a quinazoline derivative that reversibly inhibits autophosphorylation of EGFR (Figure 3) [77]. Various phase II studies have evaluated the efficacy of erlotinib with concurrent RT and TMZ, but a range of survival and toxicity outcomes have been reported. The first trial included 97 GBM patients who were given erlotinib alone for 1 week followed by concurrent erlotinib, TMZ (75 mg/m2 daily), and RT (60 Gy total) [78]. Patients had a median survival time of 15.3 months, but there was no significant benefit compared to RT/TMZ arm of the European Organisation for Research and Treatment of Cancer/National Cancer Institute of Canada trial 26981/22981. Furthermore, molecular subset analysis did not reveal that EGFR amplification was predictive of survival. Another phase II trial included 27 newly diagnosed GBM patients [79]. In this trial, erlotinib was determined to be not efficacious with unacceptable toxicity (grade 3 and 4 toxicities including thrombocytopenia, anemia, lymphopenia, fatigue, and febrile neutropenia). Numerous clinical trials have evaluated other EGFR TKIs (e.g., gefitinib, afatinib) in GBM patients [79,80,81]. Unfortunately, all EGFR TKIs to date have failed to show efficacy in GBM. Researchers hypothesize the lack of efficacy may be due to poor blood–brain barrier penetration, altered signaling pathways, and/or genetic heterogeneity [82].
Recently, preclinical studies have tested osimertinib, a third-generation EGFR TKI, in various GBM cell lines and mice [83]. Liu et al. showed osimertinib inhibited GBM cell growth ten-fold higher than first-generation EGFR inhibitors and prolonged survival in GBM-bearing mice.

5.2.2. mTOR Inhibitors

Rapamycin (mTOR) is a protein kinase that is an important regulator of cell survival and proliferation [84]. mTOR is localized in two distinct multi-protein complexes called mTORC1 and mTORC2 [85]. Previous research efforts have uncovered the critical role of mTOR in GBM pathogenesis [86,87]. Recent studies have shown GSCs can activate the mTOR pathway in microglia, creating an immunosuppressive microenvironment that promotes GBM proliferation [88].
Temsirolimus was the first mTORC1 inhibitor investigated in clinical trials (Figure 4). Temsirolimus has been shown to target GICs in preclinical studies, but has failed to demonstrate clinical benefit [89]. Sirolimus, another mTOR inhibitor, also had promising preclinical results, but failed to improve survival, despite being well tolerated [90]. Everolimus, another rapamycin derivative, is a downstream regulator of the EGFR/phosphatidylinositol-3 kinase (PI3K) pathway that has demonstrated radiosensitization in preclinical studies [91]. The North Central Cancer Treatment Group (NCCTG) conducted a phase II trial where weekly everolimus was given concurrently with RT plus TMZ. Ma et al. reported moderate toxicity and survival rates similar to historical phase II trials [92]. The RTOG 0913 trial randomized 171 GBM patients to receive RT with concurrent and adjuvant TMZ with or without daily everolimus (10 mg) [93]. Chinnaiyan and colleagues reported no significant difference in PFS and inferior OS for the patients that received everolimus. There was a significant increase in treatment-related toxicity in patients that received everolimus compared with the control arm; in the experimental arm, there were greater grade 4 and 5 events (30.6% and 11.8%, respectively) than in the control arm (17.9% and 1.3%, respectively).
Researchers surmise that the lack of efficacy may be related to everolimus only selectively inhibiting mTORC1 alone; studies have shown this inhibition can result in increased AKT activation via the activation of mTORC2 [94]. There are ongoing efforts focused on designing a suitable mTORC1/2 inhibitor [95]. AZD2014 is an inhibitor of mTORC1 and mTORC2 (Figure 4) that has shown radiosensitivity in preclinical studies [95] and is being evaluated in a phase I trial (NCT02619864).

5.3. Oxygen Mimetics

Conventional RT induces DNA damage via the formation of free radicals generated from the radiolysis of water. Reductants such as glutathione are able to neutralize the radical-induced damage within the cells, but if oxygen is present, this process is prevented, and the damage becomes irreversible. Hypoxic areas of solid tumors greatly hamper the effects of RT, leading researchers to seek oxygen mimetics [96].
Small molecules have been utilized as oxygen mimetics for decades [97] and have historically contained nitro groups that act as electron acceptors [98]. One of the earlier compounds that demonstrated radiosensitizing effects is misonidazole. Although imidazole showed radiosensitizing effects in murine tumors, its lipophilic properties prevented successful translation into clinical trials [99]. Derivatives of misonidazole were tested, and etanidazole had superior hydrophilicity due to the addition of an amide and hydroxyl group [100]. RRx-001 is a dinitro compound originally used as an ingredient in rocket fuel that has demonstrated radiosensitization properties with low toxicity [101]. Currently, RRx-001 is being evaluated in a phase I trial for patients with newly diagnosed glioblastoma (NCT02871843).
Hydrogen peroxide has been explored as a route for enhancing the efficacy of RT [102] and has been evaluated in a phase I/II trial (NCT02757651). Several studies also explored nicotinamide in combination with carbogen breathing in accelerated RT (ARCON) for various tumor types, including laryngeal, bladder, and head and neck [103,104,105,106]. Nicotinamide is a vasoactive agent that decreases perfusion-limited hypoxia, and carbogen (98% oxygen and 2% CO2) decreases diffusion-limited hypoxia [107]. Transfusion with red blood cells, in theory, should increase the oxygen supply of tumor cells, but this has failed to demonstrate benefit [108].

5.4. Reductive Agents

Bioreductive agents such as quinones and transition metal complexes have garnered attention due to their synergistic effects with RT and their preferential cytotoxicity towards hypoxic cells. Tirapazamine is a pro-drug that can be reduced to a free radical, leading to single- and double-strand DNA breaks under hypoxic environments (Figure 5) [109]. Del Rowe and colleagues conducted a phase II study with RT plus tirapazamine [110]. Although toxicity was acceptable, tirapazamine demonstrated no survival benefit.
An analogue of tirapazamine is SN30000 with more favorable diffusion properties and is currently under development [111,112]. Other analogues such as nimorazole demonstrated efficacy in several trials and are currently used in the treatment of head and neck cancers in Denmark [113].

5.5. Histone Deactylase Inhibitors

Histone deacetylases (HDACs) are enzymes that regulate chromatin structure and gene expression via deacetylation of histones and other cytoplasmic and nuclear proteins [114]. Valproic acid, an HDAC inhibitor, has demonstrated increased RT sensitivity in vitro and in vivo. Although the mechanism is unclear, researchers have proposed radiosensitization may be due to the inhibition of chromatin remodeling [115]. Krauze and colleagues conducted a phase II study evaluating the addition of valproic acid to RT plus TMZ [116]. Median OS was 29.6 months (range, 21–63.8 months), PFS was 10.5 months (range, 6.8–51.2 months), and the addition of valproic acid was generally well tolerated. The utilization of valproic acid remains controversial, though, after a pooled analysis found valproic acid at antiepilepsy doses was not associated with improved PFS or OS [117]. Vorinostat is another HDAC inhibitor that has been explored in one phase I/II trial, but failed to meet its primary efficacy end point [118].

5.6. Targeting DNA Repair Pathways

Ataxia-telangiectasia-mutated (ATM) serine/threonine protein kinase plays a role in the repair of DNA double-strand breaks [119]. ATM activation is induced within minutes of irradiation, and GSCs are particularly resistant following increased activation of ATM [120,121]. Carruthers et al. demonstrated GSCs display a robust intrinsic phosphor-ATM signal that is further enhanced following irradiation [121]. Other studies have found GBM cell lines and GSCs are radiosensitized by ATM inhibition [122].
Recently, medicinal chemists have developed a novel series of ATM inhibitors that demonstrate excellent efficacy and good pharmacokinetic properties [123]. AZD0156 was selected as a suitable candidate for clinical trials (NCT02588105). Further structure–activity relationship lead optimization led to the development of AZD1390, an orally bioavailable inhibitor with greater blood–brain barrier penetrance (Figure 6) [119]. A phase I clinical trial (NCT03423628) is currently recruiting GBM patients for the evaluation of AZD1390 in combination with RT.

5.7. Allosteric Modifiers of Hemoglobin

Phenoxyacetic acid compounds were initially utilized as lipid-lowering drugs but later were found to stabilize the T state of hemoglobin [124]. In a phase III trial, efaproxiral, a phenoxyacetic acid analogue, was found to enhance the effect of RT in patients with advanced lung cancer [125]. Kleinberg et al. then surmised GBM patients may benefit from the radio-enhancing effects of efaproxiral because GBM tumors are known to be hypoxic [126] and radioresistant [127]. Although the results were promising, a large dose was needed to reach a therapeutic effect, and long-term dose-related side effects are a concern [128].

5.8. Immunotherapy

5.8.1. Anti-Angiogenic Therapy

VEGF inhibitors such as bevacizumab have been explored with the hope of targeting angiogenesis [129]. Chinot and colleagues conducted a phase III trial evaluating the addition of bevacizumab to RT (2 Gy per fraction, total of 60 Gy) plus TMZ (75 mg/m2/day for 6 weeks) in patients with newly diagnosed GBM [130]. Although there was increased PFS in the bevacizumab group vs. placebo (10.6 months vs. 6.2 months), there was not a significant difference in OS. Furthermore, there were higher rates of adverse events with bevacizumab than with the placebo. Gilbert et al. also conducted a phase III randomized trial investigating the addition of bevacizumab to RT and TMZ [131]. Their study also demonstrated improved PFS (10.7 months vs. 7.3), although the difference was not significant according to the pre-specified alpha level (p < 0.004). The authors also noted a slight increase in adverse events and, over time, a decreased quality of life and neurocognitive function in the bevacizumab group.

5.8.2. Immune Checkpoint Inhibitors

Cancer immunotherapy is based on the concept of immunosurveillance where the immune system can actively detect and eliminate cancer cells, but some tumor cells are able to develop the ability to evade the immune system through immunoediting [132]. Immunoediting is a process where the immune system can both constrain and promote tumor progression [133]. Researchers propose this complex dynamic occurs in three phases: Elimination (the immune system can recognize and kill transformed cells), equilibrium (tumor growth is limited), and escape (edited tumors can grow, unrestrained) [134].
Immunotherapy aims to overcome this immunoresistance with immune checkpoint inhibitors (ICIs) [135]. Immune checkpoints are crucial for self-tolerance, and cancer cells exploit this feature via the upregulation of various pathways (e.g., PD-1/PD-L1, CTLA-4) [136]. Over the past decade, ICIs have revolutionized the treatment of solid tumors and have created renewed excitement within the field of cancer immunotherapy [137].
Although radiation is known to create DNA damage, several studies have suggested the immune system may impact the efficacy of radiation [138]. The exact mechanisms dictating how radiation and the immune system interact are still unclear, but data have revealed CD8 T cells play a key role [139,140]. In theory, combining RT and checkpoint blockage immunotherapy should increase radiosensitization.
Immune checkpoint inhibitors were believed to affect the tumor microenvironment by enhancing the expression of cytokine and chemokine release, which increases immune cell infiltration [141,142]. Anti-PD-1 monoclonal antibodies have had success in the setting of hepatocellular carcinoma, non-small cell lung cancer, renal cell carcinoma, melanoma, and a variety of other solid tumors [136]. Anti-CTLA-4 monoclonal antibodies have also demonstrated a survival benefit for metastatic melanoma [143].
Unfortunately, the addition of ICIs to GBM treatment has led to disappointing initial results. The first major clinical trial evaluating ICIs was CheckMate 143 [144]. In this study, patients with a first recurrence of GBM were treated with anti-PD-1 alone or anti-PD-1 and anti-CTLA-4. Adverse events in the anti-PD-1 plus anti-CTLA-4 arm resulted in discontinuation of the trial, but anti-PD-1 monotherapy was better tolerated. The subsequent CheckMate 143 phase III clinical trial with nivolumab unfortunately failed to improve OS [145]. A small study by Cloughesy and colleagues found pembrolizumab prior to salvage surgery may extend survival [146].
The combination of anti-PD-1 and RT with and without TMZ has also been explored and has been found to be well tolerated. These combinations were then studied in two phase III clinical trials: CheckMate 498 and CheckMate 548. CheckMate 498 evaluated anti-PD-1 as an alternative to TMZ in combination with RT while CheckMate 548 evaluated the addition of anti-PD-1 in addition to TMZ plus RT. In both trials, RT in combination with nivolumab was found to not improve survival [136].
To date, GBM ICI phase III clinical trials have yielded disappointing results. Researchers believe the heterogeneity of GBM tumors may contribute to immunotherapy resistance [147,148]. This tumor heterogeneity makes it difficult to find a singular treatment effective for all GBM patients; therefore, combinatorial strategies are being evaluated (NCT02313272, NCT02311582).

6. Recent Preclinical Studies

6.1. Purine Metabolism

There is a growing body of literature suggesting purine synthesis contributes to the aggressive nature of GBM [149]. GICs have high rates of de novo purine and pyrimidine synthesis that may contribute to RT resistance [150]. De novo purine synthesis can generate GTP and ATP. GBM preferentially upregulates GTP synthesis, which promotes nucleolar transformation and GBM proliferation [151]. Mycophenolate mofetil (MMF) has been found to inhibit GTP synthesis by blocking the enzyme inosine monophosphate dehydrogenase (IMPDH; Table 1) [152]. Preclinical studies suggest inhibiting GTP synthesis radiosensitizes GBM cells. Because MMF is already FDA-approved, the barrier to clinical translation is low and should be evaluated in patients with GBM.

6.2. Metabolic Targeting

Tumor cells predominately utilize glycolysis even in the presence of sufficient oxygen, also known as the Warburg effect [153]. As with many malignant solid tumors, GBM is highly glycolytic and produces lactic acid as a byproduct [154]. Studies have shown tumors with high rates of glycolysis are less responsive to RT and chemotherapy [155]; therefore, researchers have been interested in blocking or reducing glycolytic metabolism as a route for overcoming radioresistance. One study by Shen et al. found treating GBM cells (U87, U251) with a PDK inhibitor and radiation reverses the glycolytic shift [154]. The researchers proposed the inhibitor (dichloroacetate) sensitized GBM cells to radiotherapy by causing G2/M phase cell-cycle arrest (Table 1) [154]. This study suggests that altering the glycolytic metabolism may sensitize GBM to RT [32].

6.3. Curcumin

Curcumin has also been explored as a radiosensitizer for GBM [156]. There is evidence that suggests curcumin radiosensitizes tumor cells through various pathways (e.g., modifying activity of RAS-associated proteins, growth factors). Furthermore, curcumin can induce reactive oxygen species generation and inhibit the DNA repair mechanism (Table 1) [157,158]. In a study by Zoi et al., the polyphenol in combination with irradiation (2 to 4 Gy) arrested glioma cells in a synergistic fashion [159].

6.4. Hsp90 Inhibitors

Hsp90 is a molecular chaperone that has been associated with protection against radiation-induced cell death [160]. Inhibitors of Hsp90 (e.g., geldanamycin, 17DMAG, radicicol) have been shown to enhance the radiosensitivity of various cell lines (Table 1) [161,162]. Tani and colleagues found N-vinylpyrrolidone (NVP)-AUY922 enhanced radiosensitivity in CD133-positive GBM cells [163].

6.5. MDM2 Inhibitors

MDM2 has been shown to downregulate p53 activity via ubiquitin-mediated degradation and is amplified or overexpressed in certain GBM patients [164]. MDM2 inhibitors have demonstrated radiosensitizing effects preclinically in other disease sites (e.g., lung cancer, prostate cancer) [165,166], but there are limited data available for GBM. Verreault et al. found MDM2 inhibitor RG7112 reduced tumor growth and increased survival in subcutaneous and orthotopic mouse models (Table 1) [167]. This finding suggests a clinical benefit may be observed in MDM2-amplified GBM patients.

6.6. Chimeric Antigen Receptor (CAR) T Cell Therapy

Following the success of CAR T cell therapy in patients with leukemia and lymphoma, researchers have turned their attention towards developing CAR T cells directed toward solid tumors [168]. O’ Rourke et al. conducted a phase I study of autologous CAR T cells targeted to EGFR variant III in patients with GBM [169]. The authors found CART-EGFRvIII cells trafficked to the brain tumors within the first 2 weeks after infusion, but no significant clinical benefit was observed. RT may increase CAR T cell efficacy via alteration of the TME and increasing expression of tumor antigens [170]. Jin et al. found irradiation led to upregulation of CD70 expression on GBM cells and increased CD70-specific CAR T cell tumor cell elimination (Table 1) [171]. Although CAR T therapy has not demonstrated radiosensitizing effects in GBM, exploring potential synergistic interactions between RT and CAR T is an active area of research.

7. Conclusions

To date, GBM accounts for a disproportionately high percentage of cancer morbidity and mortality. Extensive research efforts over the past two decades have improved our understanding of the mechanisms driving the treatment resistance seen in GBM. Methods for overcoming radioresistance have been of particular interest, and researchers have been exploring various agents (e.g., oxygen mimetics, kinase inhibitors, immunotherapy agents) as radiosensitizers. Although clinical trials have thus far yielded negative results, recent preclinical results have been promising. Developing small molecules to target GBM-specific features such as increased GTP synthesis or amplification of p53 has the potential for selectively radiosensitizing tumor cells. Additionally, RT has the ability to potentiate the efficacy of immunotherapy, suggesting ICIs and CAR T cell therapy in combination with RT may lead to a synergistic effect. Furthermore, the successful implementation of radiosensitizers in other disease sites have gleaned valuable information and may facilitate the rational design of a GBM radiosensitizer.

Author Contributions

Conceptualization by J.K.M., H.K.P. and J.D.P. Writing, review, and editing by J.K.M., H.K.P., A.R.R., R.U., R.R.R., E.M.T., S.J.B., C.P., J.G., S.O., P.G. and J.D.P. Supervision by J.D.P. All authors have read and agreed to the published version of the manuscript.

Funding

J.K.M. received funding from the OSU College of Medicine Research Scholarship.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ostrom, Q.T.; Gittleman, H.; Liao, P.; Rouse, C.; Chen, Y.; Dowling, J.; Wolinsky, Y.; Kruchko, C.; Barnholtz-Sloan, J. CBTRUS Statistical Report: Primary Brain and Central Nervous System Tumors Diagnosed in the United States in 2007–2011. Neuro-Oncology 2014, 16, iv1–iv63. [Google Scholar] [CrossRef] [PubMed]
  2. 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] [PubMed]
  3. Stupp, R.; Taillibert, S.; Kanner, A.A.; Kesari, S.; Steinberg, D.M.; Toms, S.A.; Taylor, L.P.; Lieberman, F.; Silvani, A.; Fink, K.L.; et al. Maintenance Therapy With Tumor-Treating Fields Plus Temozolomide vs Temozolomide Alone for Glioblastoma: A Randomized Clinical Trial. JAMA 2015, 314, 2535–2543. [Google Scholar] [CrossRef] [PubMed]
  4. Ali, M.Y.; Oliva, C.R.; Noman, A.S.M.; Allen, B.G.; Goswami, P.C.; Zakharia, Y.; Monga, V.; Spitz, D.R.; Buatti, J.M.; Griguer, C.E. Radioresistance in Glioblastoma and the Development of Radiosensitizers. Cancers 2020, 12, 2511. [Google Scholar] [CrossRef]
  5. Sanai, N.; Polley, M.-Y.; McDermott, M.W.; Parsa, A.T.; Berger, M.S. An Extent of Resection Threshold for Newly Diagnosed Glioblastomas: Clinical Article. J. Neurosurg. JNS 2011, 115, 3–8. [Google Scholar] [CrossRef] [Green Version]
  6. Li, Y.M.; Suki, D.; Hess, K.; Sawaya, R. The Influence of Maximum Safe Resection of Glioblastoma on Survival in 1229 Patients: Can We Do Better than Gross-Total Resection? J. Neurosurg. JNS 2016, 124, 977–988. [Google Scholar] [CrossRef] [Green Version]
  7. Duffau, H. Long-Term Outcomes after Supratotal Resection of Diffuse Low-Grade Gliomas: A Consecutive Series with 11-Year Follow-Up. Acta Neurochir. 2016, 158, 51–58. [Google Scholar] [CrossRef]
  8. Oronsky, B.; Reid, T.R.; Oronsky, A.; Sandhu, N.; Knox, S.J. A Review of Newly Diagnosed Glioblastoma. Front. Oncol. 2021, 10, 574012. [Google Scholar] [CrossRef]
  9. Rong, L.; Li, N.; Zhang, Z. Emerging Therapies for Glioblastoma: Current State and Future Directions. J. Exp. Clin. Cancer Res. 2022, 41, 142. [Google Scholar] [CrossRef]
  10. Fan, C.-H.; Liu, W.-L.; Cao, H.; Wen, C.; Chen, L.; Jiang, G. O6-Methylguanine DNA Methyltransferase as a Promising Target for the Treatment of Temozolomide-Resistant Gliomas. Cell Death Dis. 2013, 4, e876. [Google Scholar] [CrossRef] [Green Version]
  11. Hirose, Y.; Berger, M.S.; Pieper, R.O. P53 Effects Both the Duration of G2/M Arrest and the Fate of Temozolomide-Treated Human Glioblastoma Cells. Cancer Res. 2001, 61, 1957–1963. [Google Scholar] [PubMed]
  12. 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] [PubMed] [Green Version]
  13. Laperriere, N.; Zuraw, L.; Cairncross, G. Radiotherapy for Newly Diagnosed Malignant Glioma in Adults: A Systematic Review. Radiother. Oncol. 2002, 64, 259–273. [Google Scholar] [CrossRef]
  14. Davies, A.M.; Weinberg, U.; Palti, Y. Tumor Treating Fields: A New Frontier in Cancer Therapy. Ann. N. Y. Acad. Sci. 2013, 1291, 86–95. [Google Scholar] [CrossRef]
  15. Stupp, R.; Taillibert, S.; Kanner, A.; Read, W.; Steinberg, D.M.; 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] [Green Version]
  16. Trone, J.-C.; Vallard, A.; Sotton, S.; Ben Mrad, M.; Jmour, O.; Magné, N.; Pommier, B.; Laporte, S.; Ollier, E. Survival after Hypofractionation in Glioblastoma: A Systematic Review and Meta-Analysis. Radiat. Oncol. 2020, 15, 145. [Google Scholar] [CrossRef]
  17. Liao, G.; Zhao, Z.; Yang, H.; Li, X. Efficacy and Safety of Hypofractionated Radiotherapy for the Treatment of Newly Diagnosed Glioblastoma Multiforme: A Systematic Review and Meta-Analysis. Front. Oncol. 2019, 9, 1017. [Google Scholar] [CrossRef]
  18. Perlow, H.K.; Yaney, A.; Yang, M.; Klamer, B.; Matsui, J.; Raval, R.R.; Blakaj, D.M.; Arnett, A.; Beyer, S.; Elder, J.B.; et al. Dose-Escalated Accelerated Hypofractionation for Elderly or Frail Patients with a Newly Diagnosed Glioblastoma. J. Neurooncol. 2022, 156, 399–406. [Google Scholar] [CrossRef]
  19. Perlow, H.K.; Prasad, R.N.; Yang, M.; Klamer, B.; Matsui, J.; Marrazzo, L.; Detti, B.; Scorsetti, M.; Clerici, E.; Arnett, A.; et al. Accelerated Hypofractionated Radiation for Elderly or Frail Patients with a Newly Diagnosed Glioblastoma: A Pooled Analysis of Patient-Level Data from 4 Prospective Trials. Cancer 2022, 128, 2367–2374. [Google Scholar] [CrossRef]
  20. Cho, K.H.; Hall, W.A.; Gerbi, B.J.; Higgins, P.D.; McGuire, W.A.; Clark, H.B. Single Dose versus Fractionated Stereotactic Radiotherapy for Recurrent High-Grade Gliomas. Int. J. Radiat. Oncol. 1999, 45, 1133–1141. [Google Scholar] [CrossRef]
  21. Combs, S.E.; Thilmann, C.; Edler, L.; Debus, J.; Schulz-Ertner, D. Efficacy of Fractionated Stereotactic Reirradiation in Recurrent Gliomas: Long-Term Results in 172 Patients Treated in a Single Institution. J. Clin. Oncol. 2005, 23, 8863–8869. [Google Scholar] [CrossRef]
  22. Fogh, S.E.; Andrews, D.W.; Glass, J.; Curran, W.; Glass, C.; Champ, C.; Evans, J.J.; Hyslop, T.; Pequignot, E.; Downes, B.; et al. Hypofractionated Stereotactic Radiation Therapy: An Effective Therapy for Recurrent High-Grade Gliomas. J. Clin. Oncol. 2010, 28, 3048–3053. [Google Scholar] [CrossRef]
  23. Brown, J.M.; Koong, A.C. High-Dose Single-Fraction Radiotherapy: Exploiting a New Biology? Int. J. Radiat. Oncol. 2008, 71, 324–325. [Google Scholar] [CrossRef]
  24. Ngwa, W.; Irabor, O.C.; Schoenfeld, J.D.; Hesser, J.; Demaria, S.; Formenti, S.C. Using Immunotherapy to Boost the Abscopal Effect. Nat. Rev. Cancer 2018, 18, 313–322. [Google Scholar] [CrossRef] [PubMed]
  25. Tharmalingam, H.; Hoskin, P.J. The Optimism Surrounding Stereotactic Body Radiation Therapy and Immunomodulation. Chin. Clin. Oncol. 2017, 6, S9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Kotera, Y.; Shimizu, K.; Mulé, J.J. Comparative Analysis of Necrotic and Apoptotic Tumor Cells As a Source of Antigen(s) in Dendritic Cell-Based Immunization. Cancer Res. 2001, 61, 8105–8109. [Google Scholar]
  27. Pouessel, D.; Mervoyer, A.; Larrieu-Ciron, D.; Cabarrou, B.; Attal, J.; Robert, M.; Frenel, J.-S.; Olivier, P.; Poublanc, M.; Mounier, M.; et al. Hypofractionnated Stereotactic Radiotherapy and Anti-PDL1 Durvalumab Combination in Recurrent Glioblastoma: Results of the Phase I Part of the Phase I/II STERIMGLI Trial. J. Clin. Oncol. 2018, 36, 2046. [Google Scholar] [CrossRef]
  28. Nguyen, N.; Nguyen, M.; Vock, J.; Lemanski, C.; Kerr, C.; Vinh-Hung, V.; Chi, A.; Khan, R.; Woods, W.; Altdorfer, G.; et al. Potential Applications of Imaging and Image-Guided Radiotherapy for Brain Metastases and Glioblastoma to Improve Patient Quality of Life. Front. Oncol. 2013, 3, 284. [Google Scholar] [CrossRef] [Green Version]
  29. Mizumoto, M.; Yamamoto, T.; Takano, S.; Ishikawa, E.; Matsumura, A.; Ishikawa, H.; Okumura, T.; Sakurai, H.; Miyatake, S.-I.; Tsuboi, K. Long-Term Survival after Treatment of Glioblastoma Multiforme with Hyperfractionated Concomitant Boost Proton Beam Therapy. Pract. Radiat. Oncol. 2015, 5, e9–e16. [Google Scholar] [CrossRef]
  30. Fitzek, M.M.; Thornton, A.F.; Rabinov, J.D.; Lev, M.H.; Pardo, F.S.; Munzenrider, J.E.; Okunieff, P.; Bussière, M.; Braun, I.; Hochberg, F.H.; et al. Accelerated Fractionated Proton/Photon Irradiation to 90 Cobalt Gray Equivalent for Glioblastoma Multiforme: Results of a Phase II Prospective Trial. J. Neurosurg. 1999, 91, 251–260. [Google Scholar] [CrossRef]
  31. Combs, S.E.; Kieser, M.; Rieken, S.; Habermehl, D.; Jäkel, O.; Haberer, T.; Nikoghosyan, A.; Haselmann, R.; Unterberg, A.; Wick, W.; et al. Randomized Phase II Study Evaluating a Carbon Ion Boost Applied after Combined Radiochemotherapy with Temozolomide versus a Proton Boost after Radiochemotherapy with Temozolomide in Patients with Primary Glioblastoma: The CLEOPATRA Trial. BMC Cancer 2010, 10, 478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Chédeville, A.L.; Madureira, P.A. The Role of Hypoxia in Glioblastoma Radiotherapy Resistance. Cancers 2021, 13, 542. [Google Scholar] [CrossRef]
  33. Riballo, E.; Kühne, M.; Rief, N.; Doherty, A.; Smith, G.C.M.; Recio, M.-J.; Reis, C.; Dahm, K.; Fricke, A.; Krempler, A.; et al. A Pathway of Double-Strand Break Rejoining Dependent upon ATM, Artemis, and Proteins Locating to γ-H2AX Foci. Mol. Cell 2004, 16, 715–724. [Google Scholar] [CrossRef] [PubMed]
  34. van der Schans, G.P. Gamma-Ray Induced Double-Strand Breaks in DNA Resulting from Randomly-Inflicted Single-Strand Breaks: Temporal Local Denaturation, a New Radiation Phenomenon? Int. J. Radiat. Biol. Relat. Stud. Phys. Chem. Med. 1978, 33, 105–120. [Google Scholar] [CrossRef] [PubMed]
  35. Nickoloff, J.A.; Sharma, N.; Taylor, L. Clustered DNA Double-Strand Breaks: Biological Effects and Relevance to Cancer Radiotherapy. Genes 2020, 11, 99. [Google Scholar] [CrossRef] [Green Version]
  36. Hill, R.; Leidal, A.M.; Madureira, P.A.; Gillis, L.D.; Waisman, D.M.; Chiu, A.; Lee, P.W.K. Chromium-Mediated Apoptosis: Involvement of DNA-Dependent Protein Kinase (DNA-PK) and Differential Induction of P53 Target Genes. DNA Repair 2008, 7, 1484–1499. [Google Scholar] [CrossRef]
  37. Corre, I.; Niaudet, C.; Paris, F. Plasma Membrane Signaling Induced by Ionizing Radiation. Mutat. Res. /Rev. Mutat. Res. 2010, 704, 61–67. [Google Scholar] [CrossRef]
  38. Lei, G.; Zhang, Y.; Koppula, P.; Liu, X.; Zhang, J.; Lin, S.H.; Ajani, J.A.; Xiao, Q.; Liao, Z.; Wang, H.; et al. The Role of Ferroptosis in Ionizing Radiation-Induced Cell Death and Tumor Suppression. Cell Res. 2020, 30, 146–162. [Google Scholar] [CrossRef]
  39. Carlos-Reyes, A.; Muñiz-Lino, M.A.; Romero-Garcia, S.; López-Camarillo, C.; Hernández-de la Cruz, O.N. Biological Adaptations of Tumor Cells to Radiation Therapy. Front. Oncol. 2021, 11, 718636. [Google Scholar] [CrossRef]
  40. Visvader, J.E.; Lindeman, G.J. Cancer Stem Cells in Solid Tumours: Accumulating Evidence and Unresolved Questions. Nat. Rev. Cancer 2008, 8, 755–768. [Google Scholar] [CrossRef]
  41. Bao, S.; Wu, Q.; McLendon, R.E.; Hao, Y.; Shi, Q.; Hjelmeland, A.B.; Dewhirst, M.W.; Bigner, D.D.; Rich, J.N. Glioma Stem Cells Promote Radioresistance by Preferential Activation of the DNA Damage Response. Nature 2006, 444, 756–760. [Google Scholar] [CrossRef]
  42. 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] [PubMed] [Green Version]
  43. Fidoamore, A.; Cristiano, L.; Antonosante, A.; d’Angelo, M.; Di Giacomo, E.; Astarita, C.; Giordano, A.; Ippoliti, R.; Benedetti, E.; Cimini, A. Glioblastoma Stem Cells Microenvironment: The Paracrine Roles of the Niche in Drug and Radioresistance. Stem Cells Int. 2016, 2016, 6809105. [Google Scholar] [CrossRef] [Green Version]
  44. Parsons, D.W.; Jones, S.; Zhang, X.; Lin, J.C.H.; Leary, R.J.; Angenendt, P.; Mankoo, P.; Carter, H.; Siu, I.M.; Gallia, G.L.; et al. An Integrated Genomic Analysis of Human Glioblastoma Multiforme. Science 2008, 321, 1807–1812. [Google Scholar] [CrossRef] [Green Version]
  45. Thomlinson, R.H.; Gray, L.H. The Histological Structure of Some Human Lung Cancers and the Possible Implications for Radiotherapy. Br. J. Cancer 1955, 9, 539–549. [Google Scholar] [CrossRef] [Green Version]
  46. Gray, L.H.; Conger, A.D.; Ebert, M.; Hornsey, S.; Scott, O.C.A. The Concentration of Oxygen Dissolved in Tissues at the Time of Irradiation as a Factor in Radiotherapy. Br. J. Radiol. 1953, 26, 638–648. [Google Scholar] [CrossRef] [PubMed]
  47. Ewing, D. The Oxygen Fixation Hypothesis: A Reevaluation. Am. J. Clin. Oncol. 1998, 21, 355–361. [Google Scholar] [CrossRef] [PubMed]
  48. Farace, C.; Oliver, J.A.; Melguizo, C.; Alvarez, P.; Bandiera, P.; Rama, A.R.; Malaguarnera, G.; Ortiz, R.; Madeddu, R.; Prados, J. Microenvironmental Modulation of Decorin and Lumican in Temozolomide-Resistant Glioblastoma and Neuroblastoma Cancer Stem-Like Cells. PLOS ONE 2015, 10, e0134111. [Google Scholar] [CrossRef] [Green Version]
  49. Dapash, M.; Hou, D.; Castro, B.; Lee-Chang, C.; Lesniak, M.S. The Interplay between Glioblastoma and Its Microenvironment. Cells 2021, 10, 2257. [Google Scholar] [CrossRef] [PubMed]
  50. Da Ros, M.; De Gregorio, V.; Iorio, A.L.; Giunti, L.; Guidi, M.; De Martino, M.; Genitori, L.; Sardi, I. Glioblastoma Chemoresistance: The Double Play by Microenvironment and Blood-Brain Barrier. Int. J. Mol. Sci. 2018, 19, 2879. [Google Scholar] [CrossRef] [Green Version]
  51. Vaupel, P.; Kallinowski, F.; Okunieff, P. Blood Flow, Oxygen and Nutrient Supply, and Metabolic Microenvironment of Human Tumors: A Review. Cancer Res. 1989, 49, 6449–6465. [Google Scholar] [PubMed]
  52. Gordon Steel, G.; Peckham, M.J. Exploitable Mechanisms in Combined Radiotherapy-Chemotherapy: The Concept of Additivity. Int. J. Radiat. Oncol. 1979, 5, 85–91. [Google Scholar] [CrossRef]
  53. Gill, M.R.; Vallis, K.A. Transition Metal Compounds as Cancer Radiosensitizers. Chem. Soc. Rev. 2019, 48, 540–557. [Google Scholar] [CrossRef]
  54. Fowler, J.F.; Adams, G.E.; Denekamp, J. Radiosensitizers of Hypoxic Cells in Solid Tumours. Cancer Treat. Rev. 1976, 3, 227–256. [Google Scholar] [CrossRef]
  55. Adams, G.E. Chemical radiosensitization of hypoxic cells. Br. Med. Bull. 1973, 29, 48–53. [Google Scholar] [CrossRef] [PubMed]
  56. Bastiancich, C.; Bastiat, G.; Lagarce, F. Gemcitabine and Glioblastoma: Challenges and Current Perspectives. Drug Discov. Today 2018, 23, 416–423. [Google Scholar] [CrossRef] [PubMed]
  57. Mrugala, M.M.; Chamberlain, M.C. Mechanisms of Disease: Temozolomide and Glioblastoma—Look to the Future. Nat. Clin. Pract. Oncol. 2008, 5, 476–486. [Google Scholar] [CrossRef] [PubMed]
  58. Chakravarti, A.; Erkkinen, M.G.; Nestler, U.; Stupp, R.; Mehta, M.; Aldape, K.; Gilbert, M.R.; Black, P.M.L.; Loeffler, J.S. Temozolomide-Mediated Radiation Enhancement in Glioblastoma: A Report on Underlying Mechanisms. Clin. Cancer Res. 2006, 12, 4738–4746. [Google Scholar] [CrossRef] [Green Version]
  59. Lund, B.; Kristjansen, P.E.G.; Hansen, H.H. Clinical and Preclinical Activity of 2′,2′- Difluorodeoxycytidine (Gemcitabine). Cancer Treat. Rev. 1993, 19, 45–55. [Google Scholar] [CrossRef]
  60. Lawrence, T.S.; Blackstock, A.W.; McGinn, C. The Mechanism of Action of Radiosensitization of Conventional Chemotherapeutic Agents. Semin. Radiat. Oncol. 2003, 13, 13–21. [Google Scholar] [CrossRef]
  61. Pauwels, B.; Korst, A.E.C.; Pattyn, G.G.O.; Lambrechts, H.A.J.; Van Bockstaele, D.R.; Vermeulen, K.; Lenjou, M.; de Pooter, C.M.J.; Vermorken, J.B.; Lardon, F. Cell Cycle Effect of Gemcitabine and Its Role in the Radiosensitizing Mechanism in Vitro. Int. J. Radiat. Oncol. 2003, 57, 1075–1083. [Google Scholar] [CrossRef]
  62. Latz, D.; Fleckenstein, K.; Eble, M.; Blatter, J.; Wannenmacher, M.; Weber, K.J. Radiosensitizing Potential of Gemcitabine (2′,2′-Difluoro-2′-Deoxycytidine) within the Cell Cycle in Vitro. Int. J. Radiat. Oncol. 1998, 41, 875–882. [Google Scholar] [CrossRef]
  63. Ciccolini, J.; Serdjebi, C.; Peters, G.J.; Giovannetti, E. Pharmacokinetics and Pharmacogenetics of Gemcitabine as a Mainstay in Adult and Pediatric Oncology: An EORTC-PAMM Perspective. Cancer Chemother. Pharmacol. 2016, 78, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Lawrence, T.; Eisbruch, A.; Shewach, D. Gemcitabine-Mediated Radiosensitization. Semin. Oncol. 1997, 24 (Suppl. 7), 24–28. [Google Scholar]
  65. Maraveyas, A.; Sgouros, J.; Upadhyay, S.; Abdel-Hamid, A.-H.; Holmes, M.; Lind, M. Gemcitabine Twice Weekly as a Radiosensitiser for the Treatment of Brain Metastases in Patients with Carcinoma: A Phase I Study. Br. J. Cancer 2005, 92, 815–819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Fabi, A.; Mirri, A.; Felici, A.; Vidiri, A.; Pace, A.; Occhipinti, E.; Cognetti, F.; Arcangeli, G.; Iandolo, B.; Carosi, M.A.; et al. Fixed Dose-Rate Gemcitabine as Radiosensitizer for Newly Diagnosed Glioblastoma: A Dose-Finding Study. J. Neurooncol. 2008, 87, 79–84. [Google Scholar] [CrossRef]
  67. Metro, G.; Fabi, A.; Mirri, M.A.; Vidiri, A.; Pace, A.; Carosi, M.; Russillo, M.; Maschio, M.; Giannarelli, D.; Pellegrini, D.; et al. Phase II Study of Fixed Dose Rate Gemcitabine as Radiosensitizer for Newly Diagnosed Glioblastoma Multiforme. Cancer Chemother. Pharmacol. 2009, 65, 391. [Google Scholar] [CrossRef] [PubMed]
  68. Sigmond, J.; Honeywell, R.J.; Postma, T.J.; Dirven, C.M.F.; de Lange, S.M.; van der Born, K.; Laan, A.C.; Baayen, J.C.A.; Van Groeningen, C.J.; Bergman, A.M.; et al. Gemcitabine Uptake in Glioblastoma Multiforme: Potential as a Radiosensitizer. Ann. Oncol. 2009, 20, 182–187. [Google Scholar] [CrossRef]
  69. Degen, J.W.; Walbridge, S.; Vortmeyer, A.O.; Oldfield, E.H.; Lonser, R.R. Safety and Efficacy of Convection-Enhanced Delivery of Gemcitabine or Carboplatin in a Malignant Glioma Model in Rats. J. Neurosurg. 2003, 99, 893–898. [Google Scholar] [CrossRef]
  70. Guo, P.; Ma, J.; Li, S.; Guo, Z.; Adams, A.L.; Gallo, J.M. Targeted Delivery of a Peripheral Benzodiazepine Receptor Ligand-Gemcitabine Conjugate to Brain Tumors in a Xenograft Model. Cancer Chemother. Pharmacol. 2001, 48, 169–176. [Google Scholar] [CrossRef]
  71. Wang, C.-X.; Huang, L.-S.; Hou, L.-B.; Jiang, L.; Yan, Z.-T.; Wang, Y.-L.; Chen, Z.-L. Antitumor Effects of Polysorbate-80 Coated Gemcitabine Polybutylcyanoacrylate Nanoparticles in Vitro and Its Pharmacodynamics in Vivo on C6 Glioma Cells of a Brain Tumor Model. Brain Res. 2009, 1261, 91–99. [Google Scholar] [CrossRef] [PubMed]
  72. Kim, G.; Ko, Y.T. Small Molecule Tyrosine Kinase Inhibitors in Glioblastoma. Arch. Pharm. Res. 2020, 43, 385–394. [Google Scholar] [CrossRef] [PubMed]
  73. Libermann, T.A.; Nusbaum, H.R.; Razon, N.; Kris, R.; Lax, I.; Soreq, H.; Whittle, N.; Waterfield, M.D.; Ullrich, A.; Schlessinger, J. Amplification, Enhanced Expression and Possible Rearrangement of EGF Receptor Gene in Primary Human Brain Tumours of Glial Origin. Nature 1985, 313, 144–147. [Google Scholar] [CrossRef]
  74. Chakravarti, A.; Chakladar, A.; Delaney, M.A.; Latham, D.E.; Loeffler, J.S. The Epidermal Growth Factor Receptor Pathway Mediates Resistance to Sequential Administration of Radiation and Chemotherapy in Primary Human Glioblastoma Cells in a RAS-Dependent Manner. Cancer Res. 2002, 62, 4307–4315. [Google Scholar] [PubMed]
  75. Barker, F.G.; Simmons, M.L.; Chang, S.M.; Prados, M.D.; Larson, D.A.; Sneed, P.K.; Wara, W.M.; Berger, M.S.; Chen, P.; Israel, M.A.; et al. EGFR Overexpression and Radiation Response in Glioblastoma Multiforme. Int. J. Radiat. Oncol. 2001, 51, 410–418. [Google Scholar] [CrossRef]
  76. Prados, M.D.; Lamborn, K.R.; Chang, S.; Burton, E.; Butowski, N.; Malec, M.; Kapadia, A.; Rabbitt, J.; Page, M.S.; Fedoroff, A.; et al. Phase 1 Study of Erlotinib HCl Alone and Combined with Temozolomide in Patients with Stable or Recurrent Malignant Glioma. Neuro-Oncology 2006, 8, 67–78. [Google Scholar] [CrossRef] [Green Version]
  77. Schettino, C.; Bareschino, M.A.; Ricci, V.; Ciardiello, F. Erlotinib: An EGF Receptor Tyrosine Kinase Inhibitor in Non-Small-Cell Lung Cancer Treatment. Expert Rev. Respir. Med. 2008, 2, 167–178. [Google Scholar] [CrossRef]
  78. Brown, P.D.; Krishnan, S.; Sarkaria, J.N.; Wu, W.; Jaeckle, K.A.; Uhm, J.H.; Geoffroy, F.J.; Arusell, R.; Kitange, G.; Jenkins, R.B.; et al. Phase I/II Trial of Erlotinib and Temozolomide With Radiation Therapy in the Treatment of Newly Diagnosed Glioblastoma Multiforme: North Central Cancer Treatment Group Study N0177. J. Clin. Oncol. 2008, 26, 5603–5609. [Google Scholar] [CrossRef]
  79. Peereboom, D.M.; Shepard, D.R.; Ahluwalia, M.S.; Brewer, C.J.; Agarwal, N.; Stevens, G.H.J.; Suh, J.H.; Toms, S.A.; Vogelbaum, M.A.; Weil, R.J.; et al. Phase II Trial of Erlotinib with Temozolomide and Radiation in Patients with Newly Diagnosed Glioblastoma Multiforme. J. Neurooncol. 2010, 98, 93–99. [Google Scholar] [CrossRef]
  80. Uhm, J.H.; Ballman, K.V.; Wu, W.; Giannini, C.; Krauss, J.C.; Buckner, J.C.; James, C.D.; Scheithauer, B.W.; Behrens, R.J.; Flynn, P.J.; et al. Phase II Evaluation of Gefitinib in Patients With Newly Diagnosed Grade 4 Astrocytoma: Mayo/North Central Cancer Treatment Group Study N0074. Int. J. Radiat. Oncol. 2011, 80, 347–353. [Google Scholar] [CrossRef] [Green Version]
  81. Reardon, D.A.; Nabors, L.B.; Mason, W.P.; Perry, J.R.; Shapiro, W.; Kavan, P.; Mathieu, D.; Phuphanich, S.; Cseh, A.; Fu, Y.; et al. Phase I/Randomized Phase II Study of Afatinib, an Irreversible ErbB Family Blocker, with or without Protracted Temozolomide in Adults with Recurrent Glioblastoma. Neuro-Oncology 2015, 17, 430–439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Brown, N.; McBain, C.; Nash, S.; Hopkins, K.; Sanghera, P.; Saran, F.; Phillips, M.; Dungey, F.; Clifton-Hadley, L.; Wanek, K.; et al. Multi-Center Randomized Phase II Study Comparing Cediranib plus Gefitinib with Cediranib plus Placebo in Subjects with Recurrent/Progressive Glioblastoma. PLoS ONE 2016, 11, e0156369. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, X.; Chen, X.; Shi, L.; Shan, Q.; Cao, Q.; Yue, C.; Li, H.; Li, S.; Wang, J.; Gao, S.; et al. The Third-Generation EGFR Inhibitor AZD9291 Overcomes Primary Resistance by Continuously Blocking ERK Signaling in Glioblastoma. J. Exp. Clin. Cancer Res. 2019, 38, 219. [Google Scholar] [CrossRef]
  84. Li, X.; Wu, C.; Chen, N.; Gu, H.; Yen, A.; Cao, L.; Wang, E.; Wang, L. PI3K/Akt/MTOR Signaling Pathway and Targeted Therapy for Glioblastoma. Oncotarget 2016, 7, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Laplante, M.; Sabatini, D.M. MTOR Signaling in Growth Control and Disease. Cell 2012, 149, 274–293. [Google Scholar] [CrossRef] [Green Version]
  86. Jhanwar-Uniyal, M.; Gillick, J.L.; Neil, J.; Tobias, M.; Thwing, Z.E.; Murali, R. Distinct Signaling Mechanisms of MTORC1 and MTORC2 in Glioblastoma Multiforme: A Tale of Two Complexes. Adv. Biol. Regul. 2015, 57, 64–74. [Google Scholar] [CrossRef]
  87. Jhanwar-Uniyal, M.; Wainwright, J.V.; Mohan, A.L.; Tobias, M.E.; Murali, R.; Gandhi, C.D.; Schmidt, M.H. Diverse Signaling Mechanisms of MTOR Complexes: MTORC1 and MTORC2 in Forming a Formidable Relationship. Adv. Biol. Regul. 2019, 72, 51–62. [Google Scholar] [CrossRef]
  88. 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]
  89. Wick, W.; Gorlia, T.; Bady, P.; Platten, M.; van den Bent, M.J.; Taphoorn, M.J.B.; Steuve, J.; Brandes, A.A.; Hamou, M.-F.; Wick, A.; et al. Phase II Study of Radiotherapy and Temsirolimus versus Radiochemotherapy with Temozolomide in Patients with Newly Diagnosed Glioblastoma without MGMT Promoter Hypermethylation (EORTC 26082). Clin. Cancer Res. 2016, 22, 4797–4806. [Google Scholar] [CrossRef] [Green Version]
  90. Reardon, D.A.; Desjardins, A.; Vredenburgh, J.J.; Gururangan, S.; Friedman, A.H.; Herndon, J.E.; Marcello, J.; Norfleet, J.A.; McLendon, R.E.; Sampson, J.H.; et al. Phase 2 Trial of Erlotinib plus Sirolimus in Adults with Recurrent Glioblastoma. J. Neurooncol. 2010, 96, 219–230. [Google Scholar] [CrossRef] [Green Version]
  91. Rao, R.D.; Mladek, A.C.; Lamont, J.D.; Goble, J.M.; Erlichman, C.; James, C.D.; Sarkaria, J.N. Disruption of Parallel and Converging Signaling Pathways Contributes to the Synergistic Antitumor Effects of Simultaneous MTOR and EGFR Inhibition in GBM Cells. Neoplasia 2005, 7, 921–929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Ma, D.J.; Galanis, E.; Anderson, S.K.; Schiff, D.; Kaufmann, T.J.; Peller, P.J.; Giannini, C.; Brown, P.D.; Uhm, J.H.; McGraw, S.; et al. A Phase II Trial of Everolimus, Temozolomide, and Radiotherapy in Patients with Newly Diagnosed Glioblastoma: NCCTG N057K. Neuro-Oncology 2015, 17, 1261–1269. [Google Scholar] [CrossRef] [PubMed]
  93. Chinnaiyan, P.; Won, M.; Wen, P.Y.; Rojiani, A.M.; Werner-Wasik, M.; Shih, H.A.; Ashby, L.S.; Michael Yu, H.-H.; Stieber, V.W.; Malone, S.C.; et al. A Randomized Phase II Study of Everolimus in Combination with Chemoradiation in Newly Diagnosed Glioblastoma: Results of NRG Oncology RTOG 0913. Neuro-Oncology 2018, 20, 666–673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. O’Reilly, K.E.; Rojo, F.; She, Q.-B.; Solit, D.; Mills, G.B.; Smith, D.; Lane, H.; Hofmann, F.; Hicklin, D.J.; Ludwig, D.L.; et al. MTOR Inhibition Induces Upstream Receptor Tyrosine Kinase Signaling and Activates Akt. Cancer Res. 2006, 66, 1500–1508. [Google Scholar] [CrossRef] [Green Version]
  95. Kahn, J.; Hayman, T.J.; Jamal, M.; Rath, B.H.; Kramp, T.; Camphausen, K.; Tofilon, P.J. The MTORC1/MTORC2 Inhibitor AZD2014 Enhances the Radiosensitivity of Glioblastoma Stem-like Cells. Neuro-Oncology 2014, 16, 29–37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Wang, H.; Mu, X.; He, H.; Zhang, X.-D. Cancer Radiosensitizers. Trends Pharmacol. Sci. 2018, 39, 24–48. [Google Scholar] [CrossRef]
  97. Adams, G.E.; Asquith, J.C.; Dewey, D.L.; Foster, J.L.; Michael, B.D.; Willson, R.L. Electron Affinic Sensitization. Int. J. Radiat. Biol. Relat. Stud. Phys. Chem. Med. 1971, 19, 575–585. [Google Scholar] [CrossRef]
  98. De, P.; Roy, K. Nitroaromatics as Hypoxic Cell Radiosensitizers: A 2D-QSAR Approach to Explore Structural Features Contributing to Radiosensitization Effectiveness. Eur. J. Med. Chem. Rep. 2022, 4, 100035. [Google Scholar] [CrossRef]
  99. Overgaard, J.; Sand Hansen, H.; Andersen, A.P.; Hjelm-Hansen, M.; Jørgensen, K.; Sandberg, E.; Berthelsen, A.; Hammer, R.; Pedersen, M. Misonidazole Combined with Split-Course Radiotherapy in the Treatment of Invasive Carcinoma of Larynx and Pharynx: Report from the DAHANCA 2 Study. Int. J. Radiat. Oncol. 1989, 16, 1065–1068. [Google Scholar] [CrossRef]
  100. Coleman, C.N.; Wasserman, T.H.; Urtasun, R.C.; Halsey, J.; Noll, L.; Hancock, S.; Phillips, T.L. Final Report of the Phase i Trial of the Hypoxic Cell Radiosensitizer SR 2508 (Etanidazole) Radiation Therapy Oncology Group 83-03. Int. J. Radiat. Oncol. 1990, 18, 389–393. [Google Scholar] [CrossRef]
  101. Oronsky, B.; Scicinski, J.; Ning, S.; Peehl, D.; Oronsky, A.; Cabrales, P.; Bednarski, M.; Knox, S. RRx-001, A Novel Dinitroazetidine Radiosensitizer. Investig. New Drugs 2016, 34, 371–377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Takaoka, T.; Shibamoto, Y.; Matsuo, M.; Sugie, C.; Murai, T.; Ogawa, Y.; Miyakawa, A.; Manabe, Y.; Kondo, T.; Nakajima, K.; et al. Biological Effects of Hydrogen Peroxide Administered Intratumorally with or without Irradiation in Murine Tumors. Cancer Sci. 2017, 108, 1787–1792. [Google Scholar] [CrossRef] [PubMed]
  103. Janssens, G.O.; Rademakers, S.E.; Terhaard, C.H.; Doornaert, P.A.; Bijl, H.P.; van den Ende, P.; Chin, A.; Marres, H.A.; de Bree, R.; van der Kogel, A.J.; et al. Accelerated Radiotherapy With Carbogen and Nicotinamide for Laryngeal Cancer: Results of a Phase III Randomized Trial. J. Clin. Oncol. 2012, 30, 1777–1783. [Google Scholar] [CrossRef] [PubMed]
  104. Yang, L.; Taylor, J.; Eustace, A.; Irlam, J.J.; Denley, H.; Hoskin, P.J.; Alsner, J.; Buffa, F.M.; Harris, A.L.; Choudhury, A.; et al. A Gene Signature for Selecting Benefit from Hypoxia Modification of Radiotherapy for High-Risk Bladder Cancer Patients. Clin. Cancer Res. 2017, 23, 4761–4768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Rademakers, S.E.; Hoogsteen, I.J.; Rijken, P.F.; Terhaard, C.H.; Doornaert, P.A.; Langendijk, J.A.; van den Ende, P.; van der Kogel, A.J.; Bussink, J.; Kaanders, J.H. Prognostic Value of the Proliferation Marker Ki-67 in Laryngeal Carcinoma: Results of the Accelerated Radiotherapy with Carbogen Breathing and Nicotinamide Phase III Randomized Trial. Head Neck 2015, 37, 171–176. [Google Scholar] [CrossRef]
  106. Hoskin, P.J.; Rojas, A.M.; Saunders, M.I.; Bentzen, S.M.; Motohashi, K.J. Carbogen and Nicotinamide in Locally Advanced Bladder Cancer: Early Results of a Phase-III Randomized Trial. Radiother. Oncol. 2009, 91, 120–125. [Google Scholar] [CrossRef]
  107. Horsman, M.R.; Overgaard, J.; Chaplin, D.J. Combination of Nicotinamide and Hyperthermia to Eliminate Radioresistant Chronically and Acutely Hypoxic Tumor Cells. Cancer Res. 1990, 50, 7430–7436. [Google Scholar]
  108. Welsh, L.; Panek, R.; Riddell, A.; Wong, K.; Leach, M.O.; Tavassoli, M.; Rahman, D.; Schmidt, M.; Hurley, T.; Grove, L.; et al. Blood Transfusion during Radical Chemo-Radiotherapy Does Not Reduce Tumour Hypoxia in Squamous Cell Cancer of the Head and Neck. Br. J. Cancer 2017, 116, 28–35. [Google Scholar] [CrossRef] [Green Version]
  109. Daniels, J.S.; Gates, K.S. DNA Cleavage by the Antitumor Agent 3-Amino-1,2,4-Benzotriazine 1,4-Dioxide (SR4233):  Evidence for Involvement of Hydroxyl Radical. J. Am. Chem. Soc. 1996, 118, 3380–3385. [Google Scholar] [CrossRef]
  110. Del Rowe, J.; Scott, C.; Werner-Wasik, M.; Bahary, J.P.; Curran, W.J.; Urtasun, R.C.; Fisher, B. Single-Arm, Open-Label Phase II Study of Intravenously Administered Tirapazamine and Radiation Therapy for Glioblastoma Multiforme. J. Clin. Oncol. 2000, 18, 1254–1259. [Google Scholar] [CrossRef]
  111. Wang, J.; Guise, C.P.; Dachs, G.U.; Phung, Y.; Hsu, A.H.-L.; Lambie, N.K.; Patterson, A.V.; Wilson, W.R. Identification of One-Electron Reductases That Activate Both the Hypoxia Prodrug SN30000 and Diagnostic Probe EF5. Biochem. Pharmacol. 2014, 91, 436–446. [Google Scholar] [CrossRef] [PubMed]
  112. Mistry, I.N.; Thomas, M.; Calder, E.D.D.; Conway, S.J.; Hammond, E.M. Clinical Advances of Hypoxia-Activated Prodrugs in Combination with Radiation Therapy. Int. J. Radiat. Oncol. 2017, 98, 1183–1196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Overgaard, J.; Sand Hansen, H.; Overgaard, M.; Bastholt, L.; Berthelsen, A.; Specht, L.; Lindeløv, B.; Jørgensen, K. A Randomized Double-Blind Phase III Study of Nimorazole as a Hypoxic Radiosensitizer of Primary Radiotherapy in Supraglottic Larynx and Pharynx Carcinoma. Results of the Danish Head and Neck Cancer Study (DAHANCA) Protocol 5-85. Radiother. Oncol. 1998, 46, 135–146. [Google Scholar] [CrossRef]
  114. Cerna, D.; Camphausen, K.; Tofilon, P.J. Histone Deacetylation as a Target for Radiosensitization. In Current Topics in Developmental Biology; Academic Press: Cambridge, MA, USA, 2006; Volume 73, pp. 173–204. ISBN 0070-2153. [Google Scholar]
  115. Chinnaiyan, P.; Cerna, D.; Burgan, W.E.; Beam, K.; Williams, E.S.; Camphausen, K.; Tofilon, P.J. Postradiation Sensitization of the Histone Deacetylase Inhibitor Valproic Acid. Clin. Cancer Res. 2008, 14, 5410–5415. [Google Scholar] [CrossRef] [Green Version]
  116. Krauze, A.V.; Myrehaug, S.D.; Chang, M.G.; Holdford, D.J.; Smith, S.; Shih, J.; Tofilon, P.J.; Fine, H.A.; Camphausen, K. A Phase 2 Study of Concurrent Radiation Therapy, Temozolomide, and the Histone Deacetylase Inhibitor Valproic Acid for Patients With Glioblastoma. Int. J. Radiat. Oncol. 2015, 92, 986–992. [Google Scholar] [CrossRef] [Green Version]
  117. Happold, C.; Gorlia, T.; Chinot, O.; Gilbert, M.R.; Nabors, L.B.; Wick, W.; Pugh, S.L.; Hegi, M.; Cloughesy, T.; Roth, P.; et al. Does Valproic Acid or Levetiracetam Improve Survival in Glioblastoma? A Pooled Analysis of Prospective Clinical Trials in Newly Diagnosed Glioblastoma. J. Clin. Oncol. 2016, 34, 731–739. [Google Scholar] [CrossRef]
  118. Galanis, E.; Anderson, S.K.; Miller, C.R.; Sarkaria, J.N.; Jaeckle, K.; Buckner, J.C.; Ligon, K.L.; Ballman, K.V.; Moore, D.F., Jr.; Nebozhyn, M.; et al. Phase I/II Trial of Vorinostat Combined with Temozolomide and Radiation Therapy for Newly Diagnosed Glioblastoma: Results of Alliance N0874/ABTC 02. Neuro-Oncology 2018, 20, 546–556. [Google Scholar] [CrossRef] [Green Version]
  119. Durant, S.T.; Zheng, L.; Wang, Y.; Chen, K.; Zhang, L.; Zhang, T.; Yang, Z.; Riches, L.; Trinidad, A.G.; Fok, J.H.; et al. The Brain-Penetrant Clinical ATM Inhibitor AZD1390 Radiosensitizes and Improves Survival of Preclinical Brain Tumor Models. Sci. Adv. 2018, 4, eaat1719. [Google Scholar] [CrossRef] [Green Version]
  120. Bakkenist, C.J.; Kastan, M.B. DNA Damage Activates ATM through Intermolecular Autophosphorylation and Dimer Dissociation. Nature 2003, 421, 499–506. [Google Scholar] [CrossRef]
  121. Carruthers, R.; Ahmed, S.U.; Strathdee, K.; Gomez-Roman, N.; Amoah-Buahin, E.; Watts, C.; Chalmers, A.J. Abrogation of Radioresistance in Glioblastoma Stem-like Cells by Inhibition of ATM Kinase. Mol. Oncol. 2015, 9, 192–203. [Google Scholar] [CrossRef]
  122. Golding, S.E.; Rosenberg, E.; Valerie, N.; Hussaini, I.; Frigerio, M.; Cockcroft, X.F.; Chong, W.Y.; Hummersone, M.; Rigoreau, L.; Menear, K.A.; et al. Improved ATM Kinase Inhibitor KU-60019 Radiosensitizes Glioma Cells, Compromises Insulin, AKT and ERK Prosurvival Signaling, and Inhibits Migration and Invasion. Mol. Cancer Ther. 2009, 8, 2894–2902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Pike, K.G.; Barlaam, B.; Cadogan, E.; Campbell, A.; Chen, Y.; Colclough, N.; Davies, N.L.; de-Almeida, C.; Degorce, S.L.; Didelot, M.; et al. The Identification of Potent, Selective, and Orally Available Inhibitors of Ataxia Telangiectasia Mutated (ATM) Kinase: The Discovery of AZD0156 (8-{6-[3-(Dimethylamino)Propoxy]Pyridin-3-Yl}-3-Methyl-1-(Tetrahydro-2H-Pyran-4-Yl)-1,3-Dihydro-2H-Imidazo[4,5-c]Quinolin-2-One). J. Med. Chem. 2018, 61, 3823–3841. [Google Scholar] [CrossRef]
  124. Abraham, D.J.; Kennedy, P.E.; Mehanna, A.S.; Patwa, D.C.; Williams, F.L. Design, Synthesis, and Testing of Potential Antisickling Agents. 4. Structure-Activity Relationships of Benzyloxy and Phenoxy Acids. J. Med. Chem. 1984, 27, 967–978. [Google Scholar] [CrossRef] [PubMed]
  125. Suh, J.H.; Stea, B.; Nabid, A.; Kresl, J.J.; Fortin, A.; Mercier, J.-P.; Senzer, N.; Chang, E.L.; Boyd, A.P.; Cagnoni, P.J.; et al. Phase III Study of Efaproxiral As an Adjunct to Whole-Brain Radiation Therapy for Brain Metastases. J. Clin. Oncol. 2006, 24, 106–114. [Google Scholar] [CrossRef] [PubMed]
  126. Rampling, R.; Cruickshank, G.; Lewis, A.D.; Fitzsimmons, S.A.; Workman, P. Direct Measurement of PO2 Distribution and Bioreductive Enzymes in Human Malignant Brain Tumors. Int. J. Radiat. Oncol. 1994, 29, 427–431. [Google Scholar] [CrossRef]
  127. Taghian, A.; duBois, W.; Budach, W.; Baumann, M.; Freeman, J.; Suit, H. In Vivo Radiation Sensitivity of Glioblastoma Multiforme. Int. J. Radiat. Oncol. 1995, 32, 99–104. [Google Scholar] [CrossRef]
  128. Kleinberg, L.; Grossman, S.A.; Carson, K.; Lesser, G.; O’Neill, A.; Pearlman, J.; Phillips, P.; Herman, T.; Gerber, M. Survival of Patients With Newly Diagnosed Glioblastoma Multiforme Treated With RSR13 and Radiotherapy: Results of a Phase II New Approaches to Brain Tumor Therapy CNS Consortium Safety and Efficacy Study. J. Clin. Oncol. 2002, 20, 3149–3155. [Google Scholar] [CrossRef]
  129. Jain, R.K.; di Tomaso, E.; Duda, D.G.; Loeffler, J.S.; Sorensen, A.G.; Batchelor, T.T. Angiogenesis in Brain Tumours. Nat. Rev. Neurosci. 2007, 8, 610–622. [Google Scholar] [CrossRef]
  130. Chinot, O.L.; Wick, W.; Mason, W.; Henriksson, R.; Saran, F.; Nishikawa, R.; Carpentier, A.F.; Hoang-Xuan, K.; Kavan, P.; Cernea, D.; et al. Bevacizumab plus Radiotherapy–Temozolomide for Newly Diagnosed Glioblastoma. N. Engl. J. Med. 2014, 370, 709–722. [Google Scholar] [CrossRef] [Green Version]
  131. Gilbert, M.R.; Dignam, J.J.; Armstrong, T.S.; Wefel, J.S.; Blumenthal, D.T.; Vogelbaum, M.A.; Colman, H.; Chakravarti, A.; Pugh, S.; Won, M.; et al. A Randomized Trial of Bevacizumab for Newly Diagnosed Glioblastoma. N. Engl. J. Med. 2014, 370, 699–708. [Google Scholar] [CrossRef] [Green Version]
  132. Dunn, G.P.; Bruce, A.T.; Ikeda, H.; Old, L.J.; Schreiber, R.D. Cancer Immunoediting: From Immunosurveillance to Tumor Escape. Nat. Immunol. 2002, 3, 991–998. [Google Scholar] [CrossRef] [PubMed]
  133. Vesely, M.D.; Schreiber, R.D. Cancer Immunoediting: Antigens, Mechanisms, and Implications to Cancer Immunotherapy. Ann. N. Y. Acad. Sci. 2013, 1284, 1–5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. O’Donnell, J.S.; Teng, M.W.L.; Smyth, M.J. Cancer Immunoediting and Resistance to T Cell-Based Immunotherapy. Nat. Rev. Clin. Oncol. 2019, 16, 151–167. [Google Scholar] [CrossRef]
  135. Ribas, A.; Wolchok, J.D. Cancer Immunotherapy Using Checkpoint Blockade. Science 2018, 359, 1350–1355. [Google Scholar] [CrossRef] [Green Version]
  136. Medikonda, R.; Dunn, G.; Rahman, M.; Fecci, P.; Lim, M. A Review of Glioblastoma Immunotherapy. J. Neurooncol. 2021, 151, 41–53. [Google Scholar] [CrossRef]
  137. Robert, C. A Decade of Immune-Checkpoint Inhibitors in Cancer Therapy. Nat. Commun. 2020, 11, 3801. [Google Scholar] [CrossRef]
  138. Sharabi, A.B.; Lim, M.; DeWeese, T.L.; Drake, C.G. Radiation and Checkpoint Blockade Immunotherapy: Radiosensitisation and Potential Mechanisms of Synergy. Lancet Oncol. 2015, 16, e498–e509. [Google Scholar] [CrossRef]
  139. Lee, Y.; Auh, S.L.; Wang, Y.; Burnette, B.; Wang, Y.; Meng, Y.; Beckett, M.; Sharma, R.; Chin, R.; Tu, T.; et al. Therapeutic Effects of Ablative Radiation on Local Tumor Require CD8+ T Cells: Changing Strategies for Cancer Treatment. Blood 2009, 114, 589–595. [Google Scholar] [CrossRef] [PubMed]
  140. Lugade, A.A.; Moran, J.P.; Gerber, S.A.; Rose, R.C.; Frelinger, J.G.; Lord, E.M. Local Radiation Therapy of B16 Melanoma Tumors Increases the Generation of Tumor Antigen-Specific Effector Cells That Traffic to the Tumor. J. Immunol. 2005, 174, 7516. [Google Scholar] [CrossRef] [Green Version]
  141. Schalper, K.A.; Rodriguez-Ruiz, M.E.; Diez-Valle, R.; López-Janeiro, A.; Porciuncula, A.; Idoate, M.A.; Inogés, S.; de Andrea, C.; López-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]
  142. Zhao, J.; Chen, A.X.; Gartrell, R.D.; Silverman, A.M.; Aparicio, L.; Chu, T.; Bordbar, D.; Shan, D.; Samanamud, J.; Mahajan, A.; et al. Immune and Genomic Correlates of Response to Anti-PD-1 Immunotherapy in Glioblastoma. Nat. Med. 2019, 25, 462–469. [Google Scholar] [CrossRef] [PubMed]
  143. Hodi, F.S.; O’Day, S.J.; McDermott, D.F.; Weber, R.W.; Sosman, J.A.; Haanen, J.B.; Gonzalez, R.; Robert, C.; Schadendorf, D.; Hassel, J.C.; et al. Improved Survival with Ipilimumab in Patients with Metastatic Melanoma. N. Engl. J. Med. 2010, 363, 711–723. [Google Scholar] [CrossRef] [PubMed]
  144. Sampson, J.H.; Vlahovic, G.; Sahebjam, S.; Omuro, A.M.P.; Baehring, J.M.; Hafler, D.A.; Voloschin, A.D.; Paliwal, P.; Grosso, J.; Coric, V.; et al. Preliminary Safety and Activity of Nivolumab and Its Combination with Ipilimumab in Recurrent Glioblastoma (GBM): CHECKMATE-143. J. Clin. Oncol. 2015, 33, 3010. [Google Scholar] [CrossRef]
  145. Reardon, D.A.; Brandes, A.A.; Omuro, A.; Mulholland, P.; Lim, M.; Wick, A.; Baehring, J.; Ahluwalia, M.S.; Roth, P.; Bähr, 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]
  146. 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]
  147. Jackson, C.M.; Choi, J.; Lim, M. Mechanisms of Immunotherapy Resistance: Lessons from Glioblastoma. Nat. Immunol. 2019, 20, 1100–1109. [Google Scholar] [CrossRef]
  148. Lim, M.; Xia, Y.; Bettegowda, C.; Weller, M. Current State of Immunotherapy for Glioblastoma. Nat. Rev. Clin. Oncol. 2018, 15, 422–442. [Google Scholar] [CrossRef]
  149. Hu, J.; Locasale, J.W.; Bielas, J.H.; O’Sullivan, J.; Sheahan, K.; Cantley, L.C.; Heiden, M.G.V.; Vitkup, D. Heterogeneity of Tumor-Induced Gene Expression Changes in the Human Metabolic Network. Nat. Biotechnol. 2013, 31, 522–529. [Google Scholar] [CrossRef]
  150. Wang, X.; Yang, K.; Xie, Q.; Wu, Q.; Mack, S.C.; Shi, Y.; Kim, L.J.Y.; Prager, B.C.; Flavahan, W.A.; Liu, X.; et al. Purine Synthesis Promotes Maintenance of Brain Tumor Initiating Cells in Glioma. Nat. Neurosci. 2017, 20, 661–673. [Google Scholar] [CrossRef] [Green Version]
  151. Kofuji, S.; Hirayama, A.; Eberhardt, A.O.; Kawaguchi, R.; Sugiura, Y.; Sampetrean, O.; Ikeda, Y.; Warren, M.; Sakamoto, N.; Kitahara, S.; et al. IMP Dehydrogenase-2 Drives Aberrant Nucleolar Activity and Promotes Tumorigenesis in Glioblastoma. Nat. Cell Biol. 2019, 21, 1003–1014. [Google Scholar] [CrossRef]
  152. Zhou, W.; Yao, Y.; Scott, A.J.; Wilder-Romans, K.; Dresser, J.J.; Werner, C.K.; Sun, H.; Pratt, D.; Sajjakulnukit, P.; Zhao, S.G.; et al. Purine Metabolism Regulates DNA Repair and Therapy Resistance in Glioblastoma. Nat. Commun. 2020, 11, 3811. [Google Scholar] [CrossRef] [PubMed]
  153. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg Effect: The Metabolic Requirements of Cell Proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [Green Version]
  154. Shen, H.; Hau, E.; Joshi, S.; Dilda, P.J.; McDonald, K.L. Sensitization of Glioblastoma Cells to Irradiation by Modulating the Glucose Metabolism. Mol. Cancer Ther. 2015, 14, 1794–1804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Pitroda, S.P.; Wakim, B.T.; Sood, R.F.; Beveridge, M.G.; Beckett, M.A.; MacDermed, D.M.; Weichselbaum, R.R.; Khodarev, N.N. STAT1-Dependent Expression of Energy Metabolic Pathways Links Tumour Growth and Radioresistance to the Warburg Effect. BMC Med. 2009, 7, 68. [Google Scholar] [CrossRef] [Green Version]
  156. Zoi, V.; Galani, V.; Tsekeris, P.; Kyritsis, A.P.; Alexiou, G.A. Radiosensitization and Radioprotection by Curcumin in Glioblastoma and Other Cancers. Biomedicines 2022, 10, 312. [Google Scholar] [CrossRef] [PubMed]
  157. Nakamae, I.; Morimoto, T.; Shima, H.; Shionyu, M.; Fujiki, H.; Yoneda-Kato, N.; Yokoyama, T.; Kanaya, S.; Kakiuchi, K.; Shirai, T.; et al. Curcumin Derivatives Verify the Essentiality of ROS Upregulation in Tumor Suppression. Molecules 2019, 24, 4067. [Google Scholar] [CrossRef] [Green Version]
  158. Zeng, Y.; Du, Q.; Zhang, Z.; Ma, J.; Han, L.; Wang, Y.; Yang, L.; Tao, N.; Qin, Z. Curcumin Promotes Cancer-Associated Fibroblasts Apoptosis via ROS-Mediated Endoplasmic Reticulum Stress. Arch. Biochem. Biophys. 2020, 694, 108613. [Google Scholar] [CrossRef]
  159. Zoi, V.; Galani, V.; Vartholomatos, E.; Zacharopoulou, N.; Tsoumeleka, E.; Gkizas, G.; Bozios, G.; Tsekeris, P.; Chousidis, I.; Leonardos, I.; et al. Curcumin and Radiotherapy Exert Synergistic Anti-Glioma Effect In Vitro. Biomedicines 2021, 9, 1562. [Google Scholar] [CrossRef]
  160. Dote, H.; Burgan, W.E.; Camphausen, K.; Tofilon, P.J. Inhibition of Hsp90 Compromises the DNA Damage Response to Radiation. Cancer Res. 2006, 66, 9211–9220. [Google Scholar] [CrossRef] [Green Version]
  161. Bisht, K.S.; Bradbury, C.M.; Mattson, D.; Kaushal, A.; Sowers, A.; Markovina, S.; Ortiz, K.L.; Sieck, L.K.; Isaacs, J.S.; Brechbiel, M.W.; et al. Geldanamycin and 17-Allylamino-17-Demethoxygeldanamycin Potentiate the in Vitro and in Vivo Radiation Response of Cervical Tumor Cells via the Heat Shock Protein 90-Mediated Intracellular Signaling and Cytotoxicity. Cancer Res. 2003, 63, 8984–8995. [Google Scholar]
  162. Machida, H.; Matsumoto, Y.; Shirai, M.; Kubota, N. Geldanamycin, an Inhibitor of Hsp90, Sensitizes Human Tumour Cells to Radiation. Int. J. Radiat. Biol. 2003, 79, 973–980. [Google Scholar] [CrossRef] [PubMed]
  163. Tani, T.; Tojo, N.; Ohnishi, K. Preferential Radiosensitization to Glioblastoma Cancer Stem Cell-like Cells by a Hsp90 Inhibitor, N-vinylpyrrolidone-AUY922. Oncol. Lett. 2022, 23, 102. [Google Scholar] [CrossRef] [PubMed]
  164. Miles, X.; Vandevoorde, C.; Hunter, A.; Bolcaen, J. MDM2/X Inhibitors as Radiosensitizers for Glioblastoma Targeted Therapy. Front. Oncol. 2021, 11, 703442. [Google Scholar] [CrossRef] [PubMed]
  165. Stewart-Ornstein, J.; Iwamoto, Y.; Miller, M.A.; Prytyskach, M.A.; Ferretti, S.; Holzer, P.; Kallen, J.; Furet, P.; Jambhekar, A.; Forrester, W.C.; et al. P53 Dynamics Vary between Tissues and Are Linked with Radiation Sensitivity. Nat. Commun. 2021, 12, 898. [Google Scholar] [CrossRef]
  166. Spiegelberg, D.; Mortensen, A.C.; Lundsten, S.; Brown, C.J.; Lane, D.P.; Nestor, M. The MDM2/MDMX-P53 Antagonist PM2 Radiosensitizes Wild-Type P53 Tumors. Cancer Res. 2018, 78, 5084–5093. [Google Scholar] [CrossRef] [Green Version]
  167. Verreault, M.; Schmitt, C.; Goldwirt, L.; Pelton, K.; Haidar, S.; Levasseur, C.; Guehennec, J.; Knoff, D.; Labussière, M.; Marie, Y.; et al. Preclinical Efficacy of the MDM2 Inhibitor RG7112 in MDM2-Amplified and TP53 Wild-Type Glioblastomas. Clin. Cancer Res. 2016, 22, 1185–1196. [Google Scholar] [CrossRef] [Green Version]
  168. Kershaw, M.H.; Westwood, J.A.; Parker, L.L.; Wang, G.; Eshhar, Z.; Mavroukakis, S.A.; White, D.E.; Wunderlich, J.R.; Canevari, S.; Rogers-Freezer, L.; et al. A Phase I Study on Adoptive Immunotherapy Using Gene-Modified T Cells for Ovarian Cancer. Clin. Cancer Res. 2006, 12, 6106–6115. [Google Scholar] [CrossRef] [Green Version]
  169. O’Rourke, D.M.; Nasrallah, M.P.; Desai, A.; Melenhorst, J.J.; Mansfield, K.; Morrissette, J.J.; 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] [Green Version]
  170. Flynn, J.P.; O’Hara, M.H.; Gandhi, S.J. Preclinical Rationale for Combining Radiation Therapy and Immunotherapy beyond Checkpoint Inhibitors (i.e., CART). Transl. Lung Cancer Res. 2007, 6, 159–168. [Google Scholar] [CrossRef] [Green Version]
  171. Jin, L.; Ge, H.; Long, Y.; Yang, C.; Chang, Y.; Mu, L.; Sayour, E.J.; De Leon, G.; Wang, Q.J.; Yang, J.C.; et al. CD70, a Novel Target of CAR T-Cell Therapy for Gliomas. Neuro-Oncology 2018, 20, 55–65. [Google Scholar] [CrossRef]
Figure 1. TMZ replacing nitrosoureas as the standard chemotherapy agent for GBM.
Figure 1. TMZ replacing nitrosoureas as the standard chemotherapy agent for GBM.
Biomedicines 10 01763 g001
Figure 2. The conversion of gemcitabine to gemcitabine-5′-triphosphate before being incorporated into DNA and RNA, eventually leading to strand termination.
Figure 2. The conversion of gemcitabine to gemcitabine-5′-triphosphate before being incorporated into DNA and RNA, eventually leading to strand termination.
Biomedicines 10 01763 g002
Figure 3. Erlotinib reversibly inhibits EGFR tyrosine kinase activity, which prevents cell growth and proliferation of cancer cells.
Figure 3. Erlotinib reversibly inhibits EGFR tyrosine kinase activity, which prevents cell growth and proliferation of cancer cells.
Biomedicines 10 01763 g003
Figure 4. Small molecule inhibitors of mTOR.
Figure 4. Small molecule inhibitors of mTOR.
Biomedicines 10 01763 g004
Figure 5. One proposed mechanism for tirapazamine-mediated DNA cleavage under hypoxic conditions.
Figure 5. One proposed mechanism for tirapazamine-mediated DNA cleavage under hypoxic conditions.
Biomedicines 10 01763 g005
Figure 6. ATM inhibitors: AZD0156 was modified to AZD1390, an orally available compound with greater blood–brain barrier penetrance. The preserved core is highlighted in red.
Figure 6. ATM inhibitors: AZD0156 was modified to AZD1390, an orally available compound with greater blood–brain barrier penetrance. The preserved core is highlighted in red.
Biomedicines 10 01763 g006
Table 1. Potential avenues for glioblastoma (GBM) treatment.
Table 1. Potential avenues for glioblastoma (GBM) treatment.
CategoryAgent(s)Proposed Mechanism
Purine synthesis inhibitorMycophenolate mofetilGBM upregulates GTP synthesis and mycophenolate mofetil inhibits GTP synthesis
PDK inhibitorDichloroacetatePDK inhibitor that sensitizes GBM cells to RT via G2/M phase cell-cycle arrest.
DNA repair inhibitorCurcuminCurcumin radiosensitizes tumor cells and leads to greater G2/M cell-cycle arrest.
Hsp90 inhibitorGeldanamycin, 17DMAG, radicicol, NVP-AUY922Targets Hsp90, a chaperone involved in protecting cells against radiation-induced death.
MDM2 inhibitorRG7112MDM2 inhibitors increase expression of p53 and may be beneficial in patients with TP53 wildtype and MDM2 amplification.
CAR T cell therapyCD70 CAR T cellsTargets CD70-expressing GBM tumors and may offset the immunosuppressive effects.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Matsui, J.K.; Perlow, H.K.; Ritter, A.R.; Upadhyay, R.; Raval, R.R.; Thomas, E.M.; Beyer, S.J.; Pillainayagam, C.; Goranovich, J.; Ong, S.; et al. Small Molecules and Immunotherapy Agents for Enhancing Radiotherapy in Glioblastoma. Biomedicines 2022, 10, 1763. https://doi.org/10.3390/biomedicines10071763

AMA Style

Matsui JK, Perlow HK, Ritter AR, Upadhyay R, Raval RR, Thomas EM, Beyer SJ, Pillainayagam C, Goranovich J, Ong S, et al. Small Molecules and Immunotherapy Agents for Enhancing Radiotherapy in Glioblastoma. Biomedicines. 2022; 10(7):1763. https://doi.org/10.3390/biomedicines10071763

Chicago/Turabian Style

Matsui, Jennifer K., Haley K. Perlow, Alex R. Ritter, Rituraj Upadhyay, Raju R. Raval, Evan M. Thomas, Sasha J. Beyer, Clement Pillainayagam, Justin Goranovich, Shirley Ong, and et al. 2022. "Small Molecules and Immunotherapy Agents for Enhancing Radiotherapy in Glioblastoma" Biomedicines 10, no. 7: 1763. https://doi.org/10.3390/biomedicines10071763

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