CCRG 81045

Elucidating the mechanisms of Temozolomide resistance in gliomas and the strategies to overcome the resistance.
Manendra Singh Tomar a, Ashok Kumar b, Chhitij Srivastava c, Ashutosh Shrivastava a, *
aCenter for Advance Research, Faculty of Medicine, King George’s Medical University, Lucknow 226003, Uttar Pradesh, India
bDepartment of Biochemistry, All India Institute of Medical Sciences (AIIMS) Bhopal, Saket Nagar, Bhopal 462020, Madhya Pradesh, India
cDepartment of Neurosurgery, King George’s Medical University, Lucknow 226003, Uttar Pradesh, India

A R T I C L E I N F O

Keywords:
Glioblastoma multiforme Temozolomide
Drug resistance
O6-methylguanine-DNA methyltransferase (MGMT)
Autophagy
Glioma stem cells (GSCs) miRNAs
Combination therapy
A B S T R A C T

Temozolomide (TMZ) is a first-choice alkylating agent inducted as a gold standard therapy for glioblastoma multiforme (GBM) and astrocytoma. A majority of patients do not respond to TMZ during the course of their treatment. Activation of DNA repair pathways is the principal mechanism for this phenomenon that detaches TMZ-induced O-6-methylguanine adducts and restores genomic integrity. Current understanding in the domain of oncology adds several other novel mechanisms of resistance such as the involvement of miRNAs, drug efflux transporters, gap junction’s activity, the advent of glioma stem cells as well as upregulation of cell survival autophagy. This review describes a multifaceted account of different mechanisms responsible for the intrinsic and acquired TMZ-resistance. Here, we summarize different strategies that intensify the TMZ effect such as MGMT inhibition, development of novel imidazotetrazine analog, and combination therapy; with an aim to incorporate a successful treatment and increased overall survival in GBM patients.

1.Introduction
Glioblastoma Multiforme (GBM) is the primary brain neoplasm of the central nervous system (CNS) with a high degree of lethality and have a low median survival of 15 months after the initial diagnosis [1]. CNS tumors have an incidence of <10 per 100,000 people, and the number has increased steadily in the past decade [2]. The centerpiece of GBM treatment is surgery, followed by radiation and adjuvant chemo- therapy [3]. Delivery of therapeutic agents to the CNS is hampered by the blood-brain barrier (BBB), and accessibility to the desired targets is the main obstacle for the development of new drugs for GBM [4]. Temozolomide (TMZ) is a lipophilic agent, so it can easily penetrate the BBB and thereby serves as an effective agent for the treatment of glioma [5]. It alkylates the genomic DNA at the N7 and O6 sites of guanine and N3 position of adenine, and induces nucleotide mismatch in subsequent replication cycles [6]. Nucleotide mismatch leads to the replacement of cytosine with thymine, opposite of methylated guanine. Mismatch repair identifies small nucleotide insertion/deletion and promotes the cell cycle

arrest in the G2/ M phase, leading to cell death of GBM cells [7].
The main concern in GBM treatment is the development of TMZ resistance. TMZ and other conventional chemotherapeutic drugs exert their anti-tumor action in normal cells by damaging DNA and inciting programmed cell death (PCD) [8]. In contrast, DNA repair pathways are frequently impaired in TMZ-resistant tumor cells; therefore, resistant cancer cells recover from TMZ-induced lesions through alternative repair pathways [8]. DNA repair activity can affect the sensitivity to cytotoxic chemotherapeutic drugs. Accumulating evidence suggest that the hyperactivation of O6-methylguanine-DNA methyltransferase (MGMT) is the most prominent factor that generates TMZ-resistance in GBM via removal of TMZ-induced alkylation from different nucleotides [9]. However, recent studies reveal that MGMT overexpression is not the only factor that causes TMZ-resistance in gliomas, several other path- ways contribute to TMZ-resistance independent of MGMT [10].Towards this, hyperactivation of DNA repair pathways, increased drug efflux, high antiapoptotic potential or survival autophagy, and the advent of glioma stem cells (GSCs) complement to MGMT overexpression in TMZ-

Abbreviations: ABC, ATP binding cassettes; ATGs, autophagy-related genes; BBB, blood-brain barrier; BER, base excision repair; CNS, central nervous system; DSBs, double-strand breaks; FA pathway, fanconi anemia pathway; FANCD2, fanconi anemia complementation group D2; GBM, glioblastoma multiforme; GSCs, glioma stem cells; MGMT, O6-methylguanine-DNA methyltransferase; MMR, mismatch repair; OS, overall survival; PARP1, poly [ADP-ribose] polymerase 1; PARPi, poly [ADP-ribose] polymerase 1 inhibitors; PCD, programmed cell death; SSEA1, stage-specific embryonic antigen-1; TMZ, temozolomide.
* Corresponding author.
E-mail address: [email protected] (A. Shrivastava). https://doi.org/10.1016/j.bbcan.2021.188616
Received 8 May 2021; Received in revised form 25 July 2021; Accepted 15 August 2021 Available online 20 August 2021
0304-419X/© 2021 Elsevier B.V. All rights reserved.

resistance [10,11].
In the first part of the review, we discuss various mechanisms contributing to TMZ-resistance in glioma cells. In the second part, we discuss the cellular alterations which may modulate/increase TMZ ef- ficacy and potential strategies to overcome its resistance.
2.Mechanism of action of TMZ
TMZ (trade name Temodar or Temodal) is an imidazotetrazine class of alkylating agent [12]. At physiological pH, TMZ is converted into 5-
(3-methyl-triazeno) imidazole-4-carboxamide (MTIC) that further breaks down into methyldiazonium cation (Fig. 1). Subsequently, methyldiazonium cation transfers its methyl group to DNA, RNA, and cellular proteins [13]. The conversion of TMZ to its active form depends on pH because TMZ is usually stable at pH ˂ 5.0, but above pH 7.0, it is rapidly converted into MTIC [14]. Methylation on the N7 or O6 position of guanine and N3 of adenine is a cytotoxic lesion, which stimulates the mismatch of nucleotide bases during replication [15]. If MGMT medi- ated repair does not occur, MMR proteins identify mispairing in newly synthesized strand and thymine excision or DNA damage, followed by

Fig. 1. TMZ mechanism of action. At cytosolic pH, TMZ is converted into 5-(3-methyl-triazeno) imidazole-4-carboxamide (MTIC) that generate methyldiazonium cation transfer their methyl group on N7 and O6 sites of guanine and the N3 on adenine. The TMZ adduct imparts mutation in DNA fixed via O6-methylguanine-DNA methyltransferase (MGMT), Mismatch Repair (MMR), and Base excision repair (BER). MMR generates DNA double strand breaks (DSBs) in TMZ sensitive cells and triggers PCD. Whereas overexpression of MGMT and other repair proteins remove TMZ adducts and generate TMZ resistant glioma cells.

cell cycle arrest, leading to PCD (Fig. 1) [10].
3.Mechanisms of TMZ-resistance in GBM
3.1.Role of DNA repair mechanisms in the development of TMZ- resistance phenotype
3.1.1.Role of direct DNA repair through MGMT
MGMT enzyme removes cytotoxic O6-methylguanine adducts from DNA and generates resistance against anti-tumor alkylating agents such

Fig. 2. DNA repair pathways and TMZ resistance: Methyldiazonium cation, the active product of TMZ methylates DNA on N7 or O6 site of guanine and N3 site of adenine. MGMT remove methylation at O6 site of guanine and promote cell survival. If O6-methyl guanine is not repaired through MGMT, it leads to the activation of MMR and promote cell death. Whereas methylation of N7 position of guanine and N3 position of adenine are repaired through BER. During BER, ssDNA breaks are generated that are repaired directly through PARP via recruitment of DNA repair enzymes at damage site. Non-functional BER produces dsDNA break and cell death. In TMZ-resistant GBM, fanconi anemia pathway repair dsDNA break through recombination.

as TMZ and bis-chloroethyl nitrosourea (Fig. 2) [16]. Indeed, MGMT methylation status has a significant impact on histological typing and diagnosis, and is also a prognostic marker for the survival and therapy response in GBM patients [17]. MGMT expression is directly regulated by methylation of the CpG island at the MGMT gene promoter [18]. There are multiple CpG islands in the MGMT promoter that are located at -252 to -155 and -90 to +65 regions, and methylation of these CpG sites shows an inverse relationship with MGMT expression in GBM [19]. Growing evidence have shown a decrease or loss of MGMT methylation in primary and recurrent GBM with different grading [20].
Data from the orthotopic GBM xenograft model has revealed that TMZ responsiveness inversely correlates with cellular MGMT protein level and directly correlates with MGMT promoter methylation. Loss of MGMT promoter methylation is associated with intrinsic resistance to TMZ [21]. Primary GBM patients with MGMT promoter methylation had prolonged median survival (approximately five months) in comparison to MGMT promoter unmethylated patients [22]. Epigenetic silencing of the MGMT through methylation of gene promoter in adult GBM corre- late with progression-free survival (PFS) and overall survival (OS) of patients [23]. Patients with MGMT promoter methylation respond bet- ter to TMZ therapy and have increased survival in both adults and pe- diatric GBM [24,25]. Moreover, Melguizo et al. reported that MGMT promoter methylation status might serve as a better predictive marker in treating patients with GBM [26]. Recently, it has been observed that in a subset of recurrent gliomas, genomic rearrangement leads to MGMT overexpression independent of promoter methylation [27]. However, further studies will be needed to determine the clinical utility of MGMT as a molecular biomarker in GBM patient’s survival and prognosis.

3.1.1.1.Acquired TMZ-resistance in MGMT deficient cells. Approxi- mately 50% GBMs are MGMT-deficient and are still resistant to TMZ; the underlying molecular mechanism of acquired resistance is not fully understood [28,29]. Towards this, a retrospective study reported four times higher non-synonymous mutations in the GBM with MGMT methylation. Furthermore, increased frequencies for mutations in the CDKN2A, PTEN, TP53, and PIK3CA genes were reported in GBM pa- tients with MGMT methylation, compared to non-methylated GBM [29]. Dynein, cytoplasmic 2, heavy chain 1 (DYNC2H1 or DHC2) is overex- pressed in MGMT-deficient recurrent GBMs and cells with MGMT pro- moter methylation, and it generates TMZ-resistance through nuclear localization of DNA repair proteins such as XPC and CBX5. Silencing of XPC, CBX5, or DHC2 enhances TMZ-induced cell death both in vitro and in vivo [30]. Recently Guo et al. identified nuclear proteins that promote acquired TMZ-resistance in MGMT-deficient human glioma cell lines [31]. In that study, in TMZ-treated U87 cells for one week, a total of 455 differentially expressed nuclear proteins were recognized; among them, 327 were highly expressed, and 128 have low expression. Furthermore, in silico analysis revealed that MSH6, MRE11, RPA1, RAD50, MSH2, RBX1, UBR5, CUL4A, CUL4B, DDB1, and DDB2 proteins are enriched in DNA repair pathways and formed protein-protein interaction network [31]. Epigenetic regulation of X-linked inhibitor of apoptosis (XIAP)- associated factor 1 (XAF1), a tumor suppressor protein, acts as the mediator of TMZ-resistance in MGMT methylated GBM. In fact, GBM patients with upregulated XAF1 expression had shorter OS, in compar- ison to patients with downregulated XAF1 expression [32].
Dysregulated expression of mismatch repair (MMR) complex has been reported in MGMT methylated TMZ-resistant GBM cells [33,34]. Inhibition of poly (ADP-ribose) polymerase protein inhibition (PARPi) overcomes TMZ-resistance in MGMT-deficient primary GBM cells, sug- gesting a role of BER in TMZ-resistance independent of MGMT [35]. Moreover, dysregulation of cytochrome P450 17A1 (CYP17A1) in- creases dehydroepiandrosterone (DHEA) secretion that hampers TMZ- induced PCD in MGMT-deficient GBM. DHEA induces specificity pro- tein 1 (Sp1) phosphorylation and localization in TMZ-damaged DNA to attenuate further DNA damage by recruiting proliferating cell nuclear
antigen (PCNA) [36]. Additionally, Sp1 expression is also associated with superoxide dismutase 2 (SOD2) expression that act as a ROS scavenger. Inhibition of Sp1 in TMZ-resistant MGMT-deficient clinical samples and cell lines restores TMZ sensitivity [37].

3.1.2.Role of indirect DNA repair or mismatch repair (MMR)
Mismatch repair plays a decisive part in the correction of nucleotide base mismatch generated during the process of DNA replication [38]. TMZ-induced O6-methyl guanine: thymine mispairing is fixed by the DNA mismatch repair system through excision of the newly synthesized strand [39]. Functional MMR creates a DNA double-strand break leading to PCD via apoptosis (Fig. 2) [40]. The absence of an MMR system is another important mechanism that generates resistance in GBM. Loss of function in MMR genes such as; post-meiotic segregation increased 2 (Saccharomyces cerevisiae) (PMS2), mutL Homology 1 (MLH1), melanocyte-stimulating hormone 2 (MSH2), MSH3 and MSH6 contribute to TMZ-resistance [40–42]. Notably, it has been shown that knockdown of MLH1 or PMS2 confers TMZ-resistance in recurrent GBM [43]. Attenuated expression of MSH2, MSH3 and MSH6 impact TMZ- resistance and MMR activity offers a predictive marker for TMZ ther- apy response [41,42,44]. All these MMR genes are vulnerable to TMZ- induced mutations and alterations in these genes cause resistance to TMZ in GBM.

3.1.3.Role of base excision repair (BER)
Base excision repair includes correction of DNA damage generated through oxidation, deamination, single-strand breaks, and alkylation [45]. A defective BER is recognized as a mediator of TMZ-resistance in GBM (Fig. 2) [46]. N7-methylguanine and N3-methyladenine adducts are the substrate of DNA glycosylase, an enzyme of BER. In addition to MGMT, the DNA-glycosylase enzyme removes the N7-methylguanine and N3-methyladenine adducts [47]. Furthermore, significantly higher expression of DNA glycosylase has been reported in glioma tissues in comparison to non-neoplastic brains [48]. Alkylpurine-DNA-N- glycosylase or N-methylpurine DNA glycosylase enzymes remove the alkyl group from nucleotide bases that are added by TMZ [49]. Furthermore, alkylpurine-DNA-N-glycosylase confers TMZ-resistance in xenograft models of GBM, and its higher nuclear level is associated with poor survival of GBM patients [50]. The process of healing of the apurinic/apyrimidinic site (AP site) is commonly known as the abasic site after the removal of glycosylase [51]. AP endonuclease 1 (APE1) allows perusing the abasic sites to cut at the 5’-end of the DNA. Subse- quently, the damage repair process is completed with the help of DNA ligase along with DNA polymerase [51].
Usually, enzyme Poly [ADP-ribose] polymerase 1 (PARP 1) repairs ssDNA break generated during BER; however, fragmented DNA accu- mulation in cells leads to apoptosis downstream of PARP1 cleavage [52,53]. Mechanistically, PARP1 activation utilizes NAD+ and modi- fying associated proteins auto as well as via transfer of poly-ADP-ribose (PAR) phenomenon known as PARylation [54]. PARP1 mediated PAR- ylation assigns MGMT protein at TMZ induced O6-methylguanine le- sions in GBM [55]. PARPi-mediated inhibition of PARP-1 catalytic domain prevents its PARylation and then its release from the DNA damage site generating replicative stress and replication-associated dsDNA breaks [56]. Pharmacological inhibition of PARP restores TMZ sensi- tivity in glioma or GSCs by reducing the function of DNA repair proteins [55,57]. The above study shows; BER actively plays a beneficial role in the treatment of TMZ-resistant GBM. Interruption in BER activity is one of the attractive targets to generate TMZ-sensitivity and resistance in GBM cells [58].
3.1.4.Double-strand break repair by Fanconi Anemia (FA) pathway
The FA pathway restores inter-strand crosslinks (ICLs) that cova- lently link both strands of DNA and halt unwinding through DNA heli- cases; thus, hampers the replication [59]. The FA pathway consists of 22 FA proteins (FANCA to FANCW), and mutation in any of the FA genes

generates disorders such as failure of bone marrow and cancer [60]. DNA lesions activate ATR signaling that induces the FA pathway [61]. Homologous recombination assists FA pathways to repair ICLs and various DNA lesions. When DNA lesions-caused by alkylating agents are not repaired by DNA repair pathways such as MGMT, MMR, and BER, then DNA double-strand breaks (DSBs) are produced (Fig. 2) [62]. These
DSBs are the substrate of non-homologous DNA end-joining or can be repaired through homologous recombination repair [63,64]. It has been reported that Brca2/Xrcc2-dependent homologous recombination repair is needed for protection against O6-methylguanine adduct-induced DSBs [65].
The FA gene expression correlates with glioma tumor grade, and

Fig. 3. Schematic process of autophagy in TMZ-resistant glioma cells. The autophagic pathway has several stages namely initiation, nucleation, elongation, and maturation or fusion with the lysosome. (1) TMZ treatment cause DNA damage that leads to activation of ATM/AMPK/ULK1 signaling that can induce acidic ve- sicular organelle formation, LC3 aggregation promotes interaction of autophagosome and lysosome, facilitating survival autophagy and contributes to glioma chemoresistance. (2) Binding of EGF, PDGF, and FGF to receptor tyrosine kinase (RTK) stimulates RAS/RAF/ERK and PI3K/AKT signaling pathways that activate downstream transcription factor and promote autophagy. TMZ treatment induces autophagy through ATGs and promotes phagophore formation. AKT activation inhibits autophagy through mTOR by inhibiting the ULK1 complex. (3) Low O2 condition or hypoxia induces HIF1α that also promotes autophagy. Autophagosome biogenesis starts with the formation of initiation membrane from ER, golgi body, mitochondria, and other membrane sources. Nucleation of vesicle requires Class III PI3K /baclin 1 complex. Follows elongation and maturation of vesicles into autophagosomes requires ATGs (like ATG5-ATG12) and LC3 protein. LC3 protein play crucial role in the binding of autophagosomal membrane and identification of autophagy substrate. Then autophagosome fuses with lysosome and form autolysosome and lyse autophagic substrate. Cytoprotective autophagy promotes protein synthesis, energy generation, and increased cellular invasion of glioma cells.

suppression of the FA pathway sensitizes GBM to alkylating agents [66]. TMZ treatment increases FA complementation group D2 (FANCD2) mono-ubiquitination and FANCD2 nuclear foci, the hallmark of FA pathway, in a time or dose-dependent manner that provokes TMZ- resistance. [66]. Similarly, FANCD2 knockdown through siRNA sensi- tizes glioma cells to TMZ. Recent studies suggest that degradation of FANCD2 via celastrol elicits glioma’s sensitivity towards DNA- crosslinking agent carboplatin [67]. In a similar study, a novel mecha- nism of FA pathway regulation through mTOR signaling in DNA damage and repair has been reported [68,69]. mTOR signaling pathway assists FANDC2 to stimulate DNA DSBs repair and impart chemoresistance [69]. In retrospective studies, it was noted that PARP inhibition stimu- lates FANDC2 mono-ubiquitination to basal level activation of the FA pathway in response to DSBs. Inhibition of DNA damage response pro- cess (FA-pathway, PARP, ATR, or ATM) enhances the sensitivity of GBM cells towards TMZ [70]. FA pathway proteins, FANCG, and FANCD1/
BRCA2 are predominantly involved in the repair of DNA alkylating agents (TMZ or ACNU)-induced lesions [71]. However, siRNA-mediated silencing of FANCD1/BRCA2 increases genomic instability and the sensitivity of human GBM A172 cells to TMZ [71]. The outcome of the above study underscores the significant role of the FA repair pathway in enhancing the complexity of cellular resistance of GBM towards DNA alkylating agents.

3.2.Survival autophagy in GBM cells generate TMZ-resistance
Autophagy is a conserved recycling system that occurs under stress conditions, including nutrient deprivation, hypoxia, damage in the DNA or organelles, and is controlled by series of proteins in which cells maintain homeostasis through self-degradation [72]. Autophagy serves as a double-edged sword in cancer therapy. Although autophagy acts as a tumor suppressor, conversely, it is a mechanism to adapt stress response in tumor cells that helps in the survival of cancer cells during cancer therapy [73]. Upregulated autophagy has a significant impact on motility, cell survival, chemoresistance in glioma cells, and maintenance of GSCs [74].
TMZ-induces autophagy via the ataxia–telangiectasia mutated (ATM) serine-threonine kinase/ AMP-activated protein kinase/ Unc-51 like autophagy activating kinase-1 (ATM/AMPK/ULK1) pathway (Fig. 3). This pathway can induce acidic vesicular organelle formation, LC3 ag- gregation that promotes the interaction of autophagosome with the lysosome, facilitating survival autophagy and thereby contributes to glioma chemoresistance [75]. Besides, receptor tyrosine kinase /rat sarcoma/phosphoinositide 3-kinase (RTK/Ras/PI3K) signaling regulates autophagy (Fig. 3) in the brain and is found to be altered in 89% of GBM patients [76]. Augmentation of survival autophagy further complicates GBM pathogenesis and response to therapy.
Genes linked with the regulation of autophagy are known as autophagy-related genes (ATGs) [77]. To date, 36 different types of ATGs, conserved from yeast to mammals, are reported. Among them, 18 Atg proteins are recruited to phagophore, which is involved in auto- phagosome formation [78,79]. ULK1 complex promotes autophagy initiation, and PI3K/beclin-1 complex (PI3K III C complex) supports autophagosome nucleation (Fig. 3) [80]. The autophagy proteins ATG12–ATG5 are needed for the conjugation of LC3-II or Atg 8 to phosphatidyl-ethanolamine on an autophagic membrane to regulate elon- gation and closure autophagosomal membrane [81]. Sequestosome 1/
p62-like receptors act as adaptor proteins that bind to cargos, induces specific molecular targeting [77]. Membranous components of the en- dosome, endoplasmic reticulum, mitochondria, Golgi apparatus, and plasma membrane elongate autophagosomal membrane through ATG9 [72]. The phagophore membrane elongates gradually and closes to form a double-membrane bounded vesicle autophagosome, which eventually fuses with a lysosome to form an autolysosome [82]. Lysosomes have various acid hydrolases that degrade internalized autophagic cargo [72,82] (Fig. 3).
Recently it has been reported that DOC-2/DAB2-interacting protein (DAB2IP) deprived GBM become resistant to TMZ via ATG9B dysregu- lation [83]. DAB2IP negatively regulates ATG9B through Wnt/ β-catenin pathway and enhance TMZ sensitivity by suppressing TMZ-induced autophagy [83]. Furthermore, in glioma cells, vesicle-associated mem- brane protein 8 (VAMP8) significantly increases cell proliferation and TMZ-resistance by enhancing the expression of autophagy-related pro- teins and autophagosome numbers [84]. Knockdown of VAMP8 sup- presses the growth of glioma cells through cell cycle arrest at G0/G1 phase [84]. Hence, autophagy is a key player of the cytoprotecting mechanism in unfavorable microenvironmental conditions of GBM.

3.3.Glioma stem cells (GSCs)
Subpopulation of cancer stem cells (CSCs) within GBM is another factor that aids the complexity of TMZ- resistance. GSCs have the ability to undergo self-renewal, indefinite proliferation, and multi-lineage dif- ferentiation [85]. GSCs are characterized by overexpression of various transcription factors, including octamer-binding transcription factor 4 (OCT4), sex-determining region Y-box 2 or SOX2, nestin, oligodendro- cyte transcription factor, and inhibitor of DNA binding 1 (ID1). GSCs also overexpress several cell-surface proteins such as the cluster of dif- ferentiation 133 (CD133), CD44, CD15, neural cell adhesion molecule L1 (L1CAM), or A2B5 [86]. Accumulating evidence suggests a linkage between the emergence of GSCs to chemoresistance and tumor recur- rence [87]. Now it is well established that TMZ treatment gradually increases GSCs pool [88].
Consistent exposure of glioma cells with TMZ chemotherapy is capable of interconversion of non-GSCs into new GSCs. Newly emerged GSCs possess several markers associated with normal stem cells like CD133, SSEA1, Oct4, SOX2, and nestin [88]. It has been reported that Wnt signaling pathway impart stemness in GBM [89]. The Wnt signaling inhibitor LGK974 along with TMZ decrease aldehyde dehydrogenase 3A1 (ALDH3A1) expression and markers of stem cells like-CD133, nes- tin, and Sox2 in GBM cells [90]. A phenolic compound 3,4-Dihy- droquinolin-1(2H)-yl)(p-tolyl)methyl)phenol (THTMP) modulates genes of Wnt, Hedgehog, and notch pathways and decrease stemness of GSCs [91]. Additionally, cyclin-dependent kinases (CDKs) inhibitors induce apoptosis in GSCs, resistant to TMZ [92]. Studies have shown that the population of GSCs impart chemoresistance and can be a potential target for tumor regression.

3.3.1.Glioma stem cells markers

3.3.1.1.CD 133. GSCs are characterized by CD133 (Prominin 1) expression, an apical plasma membrane protein, and these GSCs contribute to TMZ-resistance [93]. CD133 expressing cells are known as glioma initiating cells, can form a tumor in vitro as well as in vivo in mice xenografts [93,94]. However, CD133 is a controversial marker because approximately 40% of human-derived GBM specimens did not express CD133+ on the plasma membrane of tumor cells. Several studies reveal that CD133- cells can also generate tumors in immunocompro- mised animal models [85,95]. In addition to CD133, the assessment of several other markers, including nestin, integrin-α6, OCT-4, and SOX-2, is needed to identify GSCs. These studies showed that CD133 expression in glioma is mainly bioenergetic stress-dependent and partially cell cycle-dependent [96,97]. The majority of GBM cells have poor CD133+ expression, but others have no or very high CD133+ expression [98]. A recent study showed the importance of CD133 as a marker of GSCs. This study reported that CD133+ cells exhibit less responsiveness to TMZ, than CD133- cells [99]. Dysregulated expression of notch and sonic- hedgehog pathway-related genes has been reported in CD133+ cells. Inhibition of these pathways augments TMZ sensitivity [100]. Further studies are needed for the characterization of CD133 function and to improve our knowledge about its participation in tumor initiation,

aggressiveness, and maintenance.

3.3.1.2.Stage-specific embryonic antigen-1 (SSEA-1). Read et al. first identified SSEA-1 as a marker of CSCs using the Patched mutant mouse model of medulloblastoma [101]. SSEA-1 is a cluster of differentiation antigen also known as CD15 and expressed in general stem cells as well as in cancer stem cells [102]. The tumor formation ability of SSEA-1+ expressing xenograft encouraged that it may be a potential marker of tumor-initiating cells (TICs) in GBM [103]. Further studies confirmed SSEA1+ as a marker of TICs in GBM since (1) SSEA-1+ expressing GBM cells have significant tumorigenic activity than SSEA-1- cells; (2) It maintains cellular hierarchy; and (3) Cells expressing SSEA1+ have stem cell-like properties such as differentiation capacities, self-renewal properties and ability to proliferate that promote tumor initiation [104]. Charac- terization of SSEA-1 supports the need of a deeper understanding to find other reliable markers for the identification of GSCs/CSCs.
3.3.1.3.Integrin-α6. Receptors of the integrin family are heterodimeric transmembrane glycoprotein that mediates cell-to-cell and cell-to- extracellular matrix adhesion [105]. Alteration in integrin signaling empowers tumor cells to uncontrolled proliferation, invade through tissue boundaries and survive in foreign microenvironments [106]. Integrin-α6 is co-expressed with conventional markers of GSCs, and short hairpin RNA-mediated knockdown of integrin-α6 causes inhibition of proliferation, self-renewal, and tumorigenic potential of GSCs [107]. These findings suggest that elevated expression of integrin-α6 serves as an enrichment marker in GSCs and also acts as a potential target in anti- GBM therapy [107].

3.3.1.4.A2B5. A2B5 epitope belongs to the family of the sialogan- glioside that exists on the plasma membrane of neural and glial cells [108]. A2B5 is synthesized by the ST8 alpha-N-acetyl-neuraminidase α-2,8-sialyltransferase 3 (ST8SIA3) enzyme [109]. GBM cells that ex- press A2B5+ have stem cell-like properties [110]. Overexpression of the ST8SIA3 enzyme induces cellular proliferation, invasion, and xenograft of glioma cells [109]. It has also been shown that GSCs with CD133- also can form a tumor in the presence of glial progenitor marker A2B5
+ [111]. When CD133-/A2B5- and CD133-/A2B5+ cells were propa- gated into immune-compromised mice, these two subpopulations generate a neurosphere with CSCs like characteristics [112]. Whereas A2B5+ cells had a higher invasion ability than that of the A2B5- cells [112]. Human A2B5+ CD133- glioma tissues have lower expression of tumor-suppressing miRNA such as miR-218-5p; highly expressed miR- 218-5p inhibits neurosphere formation by lowering stem cell charac- teristics and invasion of glioma cells [113]. In addition, PAR1 signaling inhibition potently suppressed growth and self-renewal of A2B5
+ derived GBM [114].

3.3.1.5.Oct-4. POU domain, class 5, transcription factor 1 (POU5F1) gene encodes Oct-4, responsible for the self-renewal of embryonic stem cells. Oct-4 overexpression has been reported in glioma, and its expression increases with the advancing grade of gliomas [115,116]. Oct-4 has a remarkable role in the existence of undifferentiated cells in GBM [117]. It is reported that the hypoxic environment of tumor cells aids the expression of OCT-4 and other transcription factors [118]. Further studies suggested that Oct-4 is co-expressed with CD133 and nestin in the perivascular area of astrocytomas; however, its expression was not significantly associated with the survival outcome [119].

3.3.1.6.SOX 2/3. Sox2 is crucial for neural stem cells (NSC); its tran- scription is promoted by the Sox9 transcription factor via notch signaling. In a positive feedforward loop, SOX2 overexpression di- minishes notch1 methylation [120]. This loop exacerbates the poor prognosis in glioma patient’s cohort. Furthermore, the suppression of notch signaling reduces the migration of glioma cells across white-
matter-tract tropism of GSCs extent [120]. Significantly increased expression of SOX2 escapes cell cycle cessation and develops TMZ- resistance. Inhibition of the mTOR signaling pathway abates SOX2/9 expression and revokes chemoresistance [121].
Downregulation of SOX2 through the Wnt/β-catenin pathway en- hances TMZ-sensitivity via reduction of colony-forming potential of tumor cells and provokes cell death [122]. Additionally, SOX3 has a significant role in both general neuronal growth and tumorigenesis. SOX3 overexpresses in a subset of primary GBM samples and patient-derived GSCs. Furthermore, overexpression of SOX3 is accompanied by upre- gulation of the Hedgehog signaling pathway [123].

3.4.Role of microRNAs (miRNAs) in TMZ-resistance in GBM
miRNAs belong to the class of small single-stranded non-coding RNAs (~21-25 nucleotides) synthesized through RNA polymerase II/III and regulate posttranscriptional expression of multiple genes [124]. miRNAs interact with the specific sites in the 3’-untranslated region (UTR) of the target mRNA and exert their effects by cleaving mRNA or translational repression of the target protein [124]. Depending on the type, miRNAs act as both tumor suppressors and oncogenic [125,126]. Alteration or overexpression of oncogenic miRNAs and downregulation of tumor-suppressing miRNAs triggers signaling that enhances chemo- therapy resistance in GBM [127]. miRNAs also play a role in cell cycle regulation, proliferation, angiogenesis, invasion, and cancer stem cell behavior [128] (see Fig. 4). miRNAs also act as effective signaling molecules between cancer cells and the surrounding cells that make up the tumor microenvironment (TME) [129].
GBM cells transfer miRNAs via different mechanisms such as circu- lating cell-free miRNAs, gap junction-mediated intercellular transfer of miRNAs, and through extracellular membrane vesicles [125]. These miRNAs generate chemoresistance in GBM mainly by inhibition of PCD, upregulation of DNA damage repair pathways or MGMT upregulation, regulating cellular signaling pathways, increasing drug efflux, medi- ating epithelial interstitial transition, and via generation of drug resis- tance [10,11].
Numerous miRNAs are differentially expressed in the GBM [130]. Srinivasan et al. identified a panel of 10 signature miRNAs that segregate GBM patients in low and high-risk group patients. Among these 10- signature miRNAs; three (miRNA-20a, miRNA-106a and miRNA-17- 5p) were tumor-suppressing whereas the other seven (miRNA-31, miRNA-222, miRNA-148a, miRNA-221, miRNA-146b, miRNA-200b and miRNA-193a) were oncogenic and directly linked to patient prognosis [131].
Higher levels of exosomal miRNA-1238 have been detected in glioma patient’s serum compared to healthy individuals. Furthermore, miRNA- 1238 levels inversely correlate with TMZ-sensitivity to glioma cells. Loss of miRNA-1238 increases TMZ-sensitivity in GBM cells by directly tar- geting the caveolin 1/ epidermal growth factor receptor (CAV1/EGFR) signaling pathway [132]. Circulating miRNA-128 is a potential diag- nostic marker for TMZ sensitivity in GBM [133]. miRNA-128-3p expression is very low in gliomas; and its overexpression in glioma cells downregulates the expression of EMT proteins (c-met, PDGFRα, Notch1, and slug) and enhances the therapeutic benefits of TMZ [134]. Circulating miRNAs may provide support in the differentiation of pri- mary GBM from low-grade gliomas and found a constant drop in the level of miRNA-497 and miRNA-125b in serum as to tumor stages to the lesser extent in GBM than lower grade glioma [135]. To the best of our knowledge, there is no reliable biomarker in GBM for liquid biopsy; circulating miRNAs have great potential to be developed into one.
It has been shown that transfer of miRNA-5096 from glioma cell to astrocytes occurs through gap junction and reprogram the astrocytes for their pro-invasive effect in a Cx43-dependent manner [136]. Exosomal or gap junction-mediated transfer of miRNA-9 confers chemoresistance in GBM. Whereas; gap junction mediated intercellular transfer of anti- miRNA-9 in resistant cells helps to overcome TMZ-resistance by

Fig. 4. A summary diagram represents role of miRNAs involved in TMZ-resistance in GBM. Red arrows represent tumor suppressor miRNAs and green arrows represent oncogenic miRNAs.

reducing the expression of MDR transporter and enhancing PCD [137]. A few studies have explored the role of gap junction-mediated inter- cellular miRNA transfer in chemoresistance in GBM, so further in- vestigations are needed that improve our knowledge and may reveal unique therapeutic opportunities.
Several miRNAs acting as oncogenic miRNAs have been identified; miRNA-21 overexpression inhibits TMZ-induced apoptosis by reducing caspase-3 activity and Bax/Bcl-2 ratio [138]. Inhibition of miRNA-21 sensitizes glioma cells to TMZ by increasing the Bax /Bcl-2 ratio [139]. The overexpression of miRNA-21, miRNA-455-3p, miRNA-195, and miRNA-10a decreases chemosensitivity of GBM cells, whereas the inhibition of miRNA-195 enhances TMZ-induced cell death [140]. Dysfunctional cells are eliminated either via apoptosis or through autophagy; expectedly, oncogenic miRNA-21 and miRNA-221/222 target both the processes, autophagy, and apoptosis. miRNA-21 targets different components of p53, TGF-β, and the genes, responsible for mitochondria-mediated apoptotic pathways [141]. Whereas, down- regulation of miRNA221/222 enhances TMZ efficacy and promote apoptosis independent of p53 status [142]. Oncogenic miRNA-335 tar- gets disheveled-associated activator of morphogenesis 1 (Daam1) post- transcriptionally, Daam1 silencing reverse the oncogenic effects of miRNA-335 [143].
Furthermore, higher expression of miR-132 induces TMZ-resistance by promoting CSCs-like phenotype through downregulation of tumor suppressor candidate 3 (TUSC3) in GBM cells [144]. Higher miRNA-365 expression level has been reported in GSCs and TMZ-resistant glioma cells. In addition, long non-coding RNA (lncRNA) plasmacytoma variant translocation 1 (PVT1) regulates the stemness and TMZ efficacy of gli- oma cells through miR-365/ELF4/ SOX2 axis [145].
Many miRNAs are expressed at a lower rate in chemoresistant glioma and are functionally classified as tumor-suppressing miRNAs. An earlier study reported that downregulation of miRNA-101 significantly en- hances TMZ-resistance and results in poor outcomes in GBM patients. Conversely, upregulation of miRNA-101 overcomes TMZ-resistance in glioma cells through downregulation of GSK3β and repression of MGMT through its promoter methylation [146]. Lower expression of miRNA- 142-3p, miRNA-370-3p, miRNA-603, and miRNA-181d have been demonstrated in TMZ-resistant glioma, and their levels are inversely related to MGMT expression [147–149]. MiR-142-3p and miRNA-603 post-transcriptionally inhibit MGMT expression whereas transfection of miRNA-142-3p sensitizes GBM cells to TMZ [148,149]. Furthermore, delivery or overexpression of miRNA-370-3p enhances TMZ sensitivity of tumor cells by reducing the auto repairability of neoplasm DNA [147]. Recently it has been demonstrated that miRNA-130a-3p inhibits

cellular proliferation, invasion, and TMZ-resistance in GBM cells via targeting specificity protein 1 (SP1) that promotes metastasis in GBM [150].
The above study illustrates the crucial role of dysregulated miRNAs expression in regulating the chemosensitivity of glioma cells but also sensitizes them to targeted therapies. Further studies need to be per- formed on larger cohorts to identify miRNAs exhibiting higher sensi- tivity and specificity as a promising diagnostic biomarker for chemoresistant GBM. Additionally, the combination of conventional chemotherapy along with miRNA-targeted therapy has the potential to intensify therapeutic benefits, especially in TMZ-resistant GBM patients.
Here (Table 1), we illustrate the role of different miRNAs that pro- mote resistance and sensitize the cells to TMZ therapy response.
3.5.Drug efflux transporters
Efflux of cytotoxic drugs or anticancer agents is a prominent way of multidrug resistance generation in cancerous cells. Transporters of ATP- binding cassette (ABC) superfamily mainly participate in this mecha- nism [151]. ABC transporters are ATP-driven efflux pumps that pump multiple molecules across the plasma membrane [151]. Overexpression of ABCB1, ABCE1, or MDR1 transporter protein associated with TMZ- resistance has been reported in gliomas [33]. Modulation of ABCB1 and ABCG2 through leucine-rich repeats and immunoglobulin-like do- mains 1 (LRIG1) improves glioma sensitivity towards TMZ [152]. Repression of ABCE1 increased the efficacy of TMZ on glioma cells both in vitro and in vivo [153]. The above studies suggest that p-glycoprotein and ABC transporters may serve as potential targets that can help in TMZ potency boost up [154].

Table 1
List of different miRNAs that regulates TMZ therapy response.

3.6.Other molecular alterations in TMZ resistant glioma

MicroRNAs (miRNAs) miRNA-21
Expression status Upregulated
Role in TMZ therapy response

Inhibits TMZ-induced apoptosis via reduction of Bax/Bcl-2 ratio and caspase 3 activity. miRNA-21 inhibition

The new WHO update on classification of CNS tumor (2016) is based on molecular alterations such as isocitrate dehydrogenase (IDH) 1 and 2 genes mutation, loss of ATRX, mutant TP53/p53, and 1p/19q co- deletion [155–157]. These all are molecular biomarkers beyond tumor

miRNA-101 miRNA-1238

miRNA-128-3p miRNA195, miR-
455-3p and miRNA-10a
miRNA-9

miRNA-181b miRNA-125b-2 miRNA-145

miRNA-211
sensitizes resistant cells to TMZ [138,193].
Downregulated Reverse TMZ-resistance by GSK3β inhibition [146].
Upregulated Loss of miRNA-1238 increases TMZ sensitivity by targeting CAV1/EGFR signaling pathway [132].
Downregulated Regulate epithelial to mesenchymal transition [134].
Upregulated Acquired TMZ-resistance in MGMT deficient cell lines and knockdown shows cell killing effect [140].
Upregulated Silencing of miRNA-9 downregulated the expression of MDR transpoter and promote apoptosis [194].
Downregulated Reduces neurosphere formation and chemoresistance [195].
Upregulated Inhibition increased expression of proapoptotic protein Bax in GSCs [196].
Downregulated Decreases the level of OCT-4 and SOX-2 and alone suppressed tumorigenesis with stemness [197].
Downregulated MMP-9 mediated regulation of miR-211 lead to the activation of the intrinsic
morphology for the refinement of CNS tumor grading. In addition, tumor cells can have pre-existing genetic alterations that confer complexity in GBM treatment and resistance to TMZ [158]. The emerging molecular targets that aid resistance in GBM and have potential value in the diagnosis and treatment of GBM patients are summarised below (Table 2).

3.6.1.Role of IDH mutation in tumor grading and TMZ response
IDH converts Isocitrate into α-Ketoglutarate (αKG), NADP into NADPH, and releases CO2. IDH1 and 2 are key enzymes linking cellular metabolism to epigenetic regulation and redox states [159]. IDH1 and IDH2 genes are present on chromosomes 2q33 and 15q26.1 respectively [160]. Somatic mutations are the most frequently present in IDH1 gene (75-80%) as compared to the IDH2 gene (20%) in glioma [161]. Ac- cording to WHO reports, IDH mutation was recognized in more than 80% of grade II and III gliomas. IDH mutation is also present in grade IV GBM, but frequently in secondary GBM (73%) but significantly less in primary GBM (3.7%) [162].
Earlier studies had revealed that that the patients with IDH1 or IDH2

mitochondrial/Caspase-9/3 mediated apoptosis in both glioma and GSCs [198].
Table 2
Molecular alterations in TMZ resistant glioma.

miRNA-17
miRNA-17 incorporation modulates autophagy by regulating expression of ATG7 and LC3B gene [199].
Gene Isocitrate
Model of study Role in TMZ-resistance
U-87 GBM cell Overexpression of IDH1 generates

miRNA-9
Upregulated
Involved in overexpression of drug efflux transporter, P-glycoprotein. Delivary of Anti-miRNA-9 reverse chemoresistance in GBM cells [200].
dehydrogenases (IDHs)
TP53/p53
line

PRT-HU2, U-
resistance in glioma cells and mutant IDH1 enhanced cells sensitivity to TMZ [164].
MGMT expression was negatively

miRNA-183/96/182 Upregulated
miRNA-221/222 Upregulated
Overexpression of miRNAs induced ROS mediated apoptosis [201]. Downregulation or knockdown of miRNA-221/222 sensitize glioma cells
87MG and T98G
regulated by p53 in GBM. p53 reactivation may be an effective strategy to overcome TMZ resistance [204].

miRNA-132

Upregulated
to TMZ by regulating apoptosis [142]. Generate GSCs and TMZ-resistance by downregulating TUSC3 in glioma [144].
TERT promoter
mutation EGFR
U373-MG and U87-MG
U87MG and
Lower activity of telomerase is directly linked with TMZ sensitivity of glioma [205].
EGFR/EGFRvIII overexpression

miRNA-142-3p/370-
3p/603/181d miRNA-519a
Downregulated Promote TMZ-resistance by enhancing MGMT mediated direct repair [147–149].
Downregulated miRNA-519a have inverse relation to chemosensitivity of GBM cells and
U251
generates GSCs, tumorigenesis, tumor- recurrence and resistance to chemotherapy. Upregulation of HER2/3 and RTK pathway limiting effect of anti- EGFR therapies [206].

enhance PCD by targeting STAT3/Bcl- 2/Beclin-1 pathway [202].
miRNA-26a Upregulated miRNA-26a reduces the expression of AP-2α that have tumor suppressive role. Inhibitor of miRNA-26a reduces tumor stemness and TMZ resistance [203].
ATRX
LN 229
Upregulation of ATRX expression intensify DNA repair through fortify PARP1 protein. ATRX prompt PARP1 stabilization through FADD suppression via H3K27 methylation and generate TMZ-resistance [207].

mutant glioma had significantly higher OS than the patients with wild- type IDH [163]. However, in vivo and in vitro studies show that over- expression of IDH1 confer TMZ-resistance in glioma cells [164]. Whereas IDH1 mutation induces chemosensitivity, cell cycle arrest at the G1 phase, and inhibits cellular proliferation and invasion [164]. mIDH patients treated with TMZ + RT or RT alone exhibit better OS than wtIDH [165]. Additionally, the introduction of mIDH1/2 or silencing of mtIDH1/2 sensitizes cancer cells to chemotherapy or RT and improves therapy response [159]. A recent study showed that IDH1/2 is an in- dependent prognostic factor, but the combination of IDH1/2 mutation along with MGMT promoter methylation show increased OS in patients with malignant glioma on TMZ treatment [166]. IDH maintains cellular redox status by regulating the balance of NADPH/NAPD+. wtIDH pro- duces NADPH that protects cells from ROS-induced by TMZ, whereas mIDH is associated with an elevated level of ROS [167]. Therefore, IDH mutation is a significant marker of TMZ response in secondary GBM.
4.Strategies to overcome TMZ-resistance in GBM
4.1.Inhibition of MGMT
As discussed earlier MGMT undo alkylation of DNA that has been triggered by TMZ and restore the normal DNA. MGMT-mediated repair of damaged DNA generates resistance to TMZ in gliomas [168]. In previous studies, several pathways have been identified that play essential roles in MGMT expression regulation. The Wnt/β-catenin signaling regulates MGMT expression, and cordycepin-mediated downregulation of Wnt/β catenin signaling subsequently suppresses MGMT expression, thereby augmenting chemosensitivity of glioma cells towards TMZ [169]. BMX, an HDAC8 inhibitor overcomes TMZ- resistance in GBM cells that have wild type p53 by enhancing cell death via downregulation of β catenin/c-Myc/SOX2 cellular signaling and MGMT inhibition [170]. In addition, nuclear factor 1A (NFIA) aids TMZ-resistance in GBM through nuclear factor-κB (NF-κB) signaling and contributes to the poor prognosis and recurrence of GBM patients [171]. Blocking of NF-κB has anti-glioma outcomes and reduces TMZ-resistance by decreasing MGMT gene expression [172]. Resveratrol overcomes TMZ-resistance by suppression of MGMT in glioma cell lines through NF-κB signaling [173]. These results indicate that NFIA-NF-κB axis might be a novel therapeutic approach to TMZ-resistant GBM. Dysre- gulation of IKBKE enhances resistance to TMZ through overexpression of MGMT in GBM [174]. S-nitroso-N-acetyl penicillamine (SNAP), a nitric oxide donor reverses TMZ-resistance by downregulating MGMT expression [175]. Furthermore, the resistance is also overcome by silencing of SATB1 by decreasing MGMT expression and overexpression of solute carrier family 22 member 18 (SLC22A18) in human GBM cells [176].

4.2.Modulation of BER
BER modulation potentially enhanced chemosensitivity and patho- logical outcomes in glioma [177].
4.2.1.PARP inhibition
Inhibition of PARP enhances the therapeutic benefits of TMZ in the occurrence of malignant gliomas, especially in tumors with DNA mismatch [178]. The combination of PARPi (PARP inhibitors) and TMZ shows synergistic anti-neoplastic outcomes in 8 out of 10 GSCs tested. Consequently, TMZ dose reduction is associated with PARPi suscepti- bility of cell lines. Notably, a combination of PARPi with TMZ may be considered as a beneficial treatment strategy for the reversal of che- moresistance in glioma [57]. Furthermore, PARP inhibition may be a promising therapy target in MMR-deficient resistant cells and restore TMZ-sensitivity in MSH6-inactivated glioma [179]. Mechanistically, this PARPi-induced synthetic phenotype was unconventional of BER blockage and not recapitulated by loss of PARP1 [179].
4.2.2.Inhibition of APE-1
APE1 or Ref-1 is a crucial mammalian apurinic/apyrimidinic (AP) endonuclease, impart resistance against TMZ in glioma by incising DNA at abasic sites [180]. Overexpression or anomalous subcellular dispersal of APE1 observed in different types of cancer corresponds to drug response, patient survival, and prognosis [181]. The activity of the APE- 1 enzyme significantly contributes to TMZ-resistant gliomas [46,182]. Cut Like homeobox 1 (CUX1) triggers activity of APE1 enzyme and induce resistance to TMZ in GBM cells [183]. Knockdown/silencing of APE1 is strongly linked with decreased growth as well as invasion of resistant cells. APE1 suppression has a limited cytotoxic effect on TMZ- sensitive cell lines as compared to resistant cell lines. APE1 may be the potential target to alter the response of TMZ in resistant GBM cells. Besides chemoresistance, APE1 protein also participates in modifying radiation tolerance and generate radio-resistance in glioma cells [184]. The above studies suggest APE1 act as a potential drug target and have therapeutic benefits in glioma patients.

4.3.Combination therapy to overcome TMZ-resistance
Nowadays, the use of TMZ in combination with other anti-neoplastic agents/ therapeutics has become the primary strategy to treat resistant glioma [185]. The rationale of combination therapy is to use drugs that target key pathways through separate mechanisms, thereby decreasing drug-resistant cancer cells [186]. Importantly, when drugs are used in combination, each drug is used at its optimum dose because it may be toxic to patients with intolerable side effects [187]. Combination treat- ment with TMZ has resulted in prolonging the survival of GBM patients [188]. Here, we enumerate different drugs that are used to improve the OS of GBM patients with TMZ resistant neoplasm (Table 3).

4.4.Imidazotetrazine analogs
As discussed earlier, TMZ is an imidazotetrazine prodrug; however, resistance against TMZ limits efficacy and therapeutic benefits [189]. TMZ-resistance is mainly conferred by MGMT overexpression, which intensifies tolerance to O6-methylguanine lesions and limits the clinical application of TMZ [190]. Therefore, the advancement in TMZ analogs is urgently needed to conquer TMZ-resistance. TMZ analogs C8-imida- zolyl and C8-methylimidazole tetrazines induce anti-neoplastic actions in T98G glioma cell lines that hyperactivate MGMT and promote arrest of cell cycle progression, DNA breakage, and cell death [191]. C8 ana- logs have similar actions like TMZ and methylate DNA adducts, but unlike TMZ; MGMT cannot remove incorporated lesions [191].
Novel TMZ analog DP68 and DP86 are designed to conquer resis- tance in glioma cell lines and primary culture models [192]. The potency of these syntheses is unstrained from MGMT and MMR activities. DP68 cross-links DNA and induces cell cycle arrest with the convergence of cells in the S phase [192]. Besides, DP68 give rise to intense DNA damage response through phosphorylating ATM (a serine/ threonine kinase activated when double-strand DNA damage occurs), KAP1, Chk1, and Chk2 kinase (regulators of the cell cycle), and histone variant H2AX (phosphorylated on serine residue at the time of DNA double-strand break). Blocking expression of FANCD2 (Fanconi anemia group D2 protein) or ATR kinase action increased the anti-neoplastic outcomes of DP68 [192].

5.Conclusion
TMZ induces cell death through DNA damage and have been a mainstay drug to treat glioma patients. Resistance to TMZ in glioma involves multiple molecular pathways and constitute a multifaceted challenge. Elucidating the different mechanisms of TMZ-resistance by using Patient derived GBM xenograft in animal models and the tissue samples of TMZ resistant glioma can provide significant insight into underlying mechanisms of TMZ resistance. Sustained research on how

Table 3
Combination therapy used in TMZ resistant glioma treatment.
authors concur to the final version of the manuscript.

Drug Metformin

Rapamycin Isofuranodiene

Psammaplin C

Chloroquine

Lovastatin Cordycepin

Romidepsin

Cannabinoids

Tamoxifen

Levetiracetam

Difluoromethyl-
ornithine Afatinib

Hypericin BIX01294
Model of Study T98G / Invivo mice
model

Gli36, U87MG Xenograft of GBM6 and GBM10 U87, T98, U251

Primary human GBM cells (CV17, 010627, No3)
U87MG and U373

U87 and U251
LN18, T98G, U87, U251 and rat C6 cells.

T98G, U-138MG, A-172 and U87MG Human glioma cell line
U87MG, A172, SW1783, U373MG, T98G, SW1088, and LN405 U87, U118

Neurospheres of GBM and peritumoral tissue
U87G, U251MG and T98G GBM cell line U87MG and U251

A172 and LA567
LN18 and U251 glioma cell lines
Mechanism of action
Suppress the expression of GSCs marker CD90 and significantly reduce the growth of TMZ resistant tumor in mice model [208].
Rapamycin inhibits mTOR signaling [209].
Induced cell cycle arrest and necrosis through increasing ROS level [210].
Potent inhibitor of carbonic anhydrase XII and helps to overcome P-glycoprotein mediated TMZ-resistance [211]. Promotes TMZ induced autophagy, activation of caspase 3/ p53 dependent apoptosis and induces cell death by inhibiting mitochondrial autophagy via increasing ROS level [212–214]. Impairment of autophagic flux induces apoptosis [215].
Enhance chemosensitivity to TMZ by upregulating AMPK signaling and slowdown the AKT Signaling [216].
Inhibiting PI3K/AKT/mTOR signaling pathways [217].
Enhance Autophagy [218].

Inhibit PKC-pan and promote antiproliferative and apoptotic effects [219].
MGMT suppression, upregulating nuclear transport of HDAC4 and promote apoptosis [220].
Cell cycle arrest at G2/M phase [221].
Down regulates cellular multiplication by repressing EGFRvIII/AKT/EGFRvIII/JAK2/
STAT3 and FAK (Focal adhesion kinase) pathways [206]. Inhibited tumor growth by inducing apoptosis [222]. Inhibitor of histone
methyltransferase G9a, increased the TMZ sensitivity in glioma cell and GSCs [223].
Declaration of Competing Interest
Authors declare no conflict of interest. References
[1]A. Shergalis, A. Bankhead 3rd, U. Luesakul, N. Muangsin, N. Neamati, Current challenges and opportunities in treating glioblastoma, Pharmacol. Rev. 70 (2018) 412–445.
[2]O.G. Taylor, J.S. Brzozowski, K.A. Skelding, Glioblastoma multiforme: an overview of emerging therapeutic targets, Front. Oncol. 9 (2019) 963.
[3]L. Hiddingh, B.A. Tannous, J. Teng, B. Tops, J. Jeuken, E. Hulleman, S.H. Boots- Sprenger, W.P. Vandertop, D.P. Noske, G.J.L. Kaspers, P. Wesseling,
T. Wurdinger,
EFEMP1 induces γ-secretase/Notch-mediated temozolomide resistance in glioblastoma, Oncotarget 5 (2) (2013).
[4]X. Dong, Current strategies for brain drug delivery, Theranostics 8 (2018) 1481–1493.
[5]S.Y. Lee, Temozolomide resistance in glioblastoma multiforme, Genes Dis. 3 (2016) 198–210.
[6]H. Strobel, T. Baisch, R. Fitzel, K. Schilberg, M.D. Siegelin, G. Karpel-Massler, K. M. Debatin, M.A. Westhoff, Temozolomide and other alkylating agents in glioblastoma therapy, Biomedicines 7 (2019).
[7]J. Lips, B. Kaina, Repair of O(6)-methylguanine is not affected by thymine base pairing and the presence of MMR proteins, Mutat. Res. 487 (2001) 59–66.
[8]K. Yoshimoto, M. Mizoguchi, N. Hata, H. Murata, R. Hatae, T. Amano,
A. Nakamizo, T. Sasaki, Complex DNA repair pathways as possible therapeutic targets to overcome temozolomide resistance in glioblastoma, Front. Oncol. 2 (2012) 186.
[9]A.E. Pegg, T.L. Byers, Repair of DNA containing O6-alkylguanine, FASEB J. 6 (1992) 2302–2310.
[10]P. Woo, Y. Li, A. Chan, S. Ng, H. Loong, D. Chan, G. Wong, W.-S. Poon,
A multifaceted review of temozolomide resistance mechanisms in glioblastoma beyond O-6-methylguanine-DNA methyltransferase, Glioma 2 (2019) 68–82.
[11]N. Singh, A. Miner, L. Hennis, S. Mittal, Mechanisms of temozolomide resistance in glioblastoma – a comprehensive review, Cancer Drug Resist. 4 (2021) 17–43.
[12]S.S. Agarwala, J.M. Kirkwood, Temozolomide, a novel alkylating agent with activity in the central nervous system, may improve the treatment of advanced metastatic melanoma, Oncologist 5 (2000) 144–151.
[13]H.S. Friedman, T. Kerby, H. Calvert, Temozolomide and treatment of malignant glioma, clinical cancer research : an official journal of the american association for, Cancer Res. 6 (2000) 2585–2597.
[14]S.D. Baker, M. Wirth, P. Statkevich, P. Reidenberg, K. Alton, S.E. Sartorius,
M. Dugan, D. Cutler, V. Batra, L.B. Grochow, R.C. Donehower, E.K. Rowinsky, Absorption, metabolism, and excretion of 14C-temozolomide following oral administration to patients with advanced cancer, Clin. Cancer Res. 5 (1999) 309–317.
[15]C.H. Fan, W.L. Liu, H. Cao, C. Wen, L. Chen, G. Jiang, O6-methylguanine DNA methyltransferase as a promising target for the treatment of temozolomide- resistant gliomas, Cell Death Dis. 4 (2013), e876.
[16]J.R. Silber, M.S. Bobola, A. Blank, M.C. Chamberlain, O(6)-methylguanine-DNA methyltransferase in glioma therapy: promise and problems, Biochim. Biophys. Acta 2012 (1826) 71–82.
[17]A. Mansouri, L.D. Hachem, S. Mansouri, F. Nassiri, N.J. Laperriere, D. Xia, N. I. Lindeman, P.Y. Wen, A. Chakravarti, M.P. Mehta, M.E. Hegi, R. Stupp, K.
D. Aldape, G. Zadeh, MGMT promoter methylation status testing to guide therapy

these functions are regulated in TMZ-resistance can provide the thera- peutic window to effectively manage the TMZ-resistant gliomas. Therefore, continued research in the area of functional regulation of TMZ resistance can be the basis of custom-tailored therapeutics to treat recalcitrant glioma and enhance patient’s OS.
Funding

Manendra Singh Tomar is the recipient of Junior research fellowship from the University Grants Commission, Government of India. Extra- mural research grant (EMR/2016/005009) from the Science and Engi- neering Research Board to Ashok Kumar is highly acknowledged. Laboratory of Ashutosh Shrivastava is supported by CCRH-Ministry of AYUSH, Government of India and KGMU intramural grant.

Author’s contributions

AS and MST conceptualized the review manuscript. AS, MST and AK wrote and evaluated the manuscript. CS evaluated the manuscript. All
for glioblastoma: refining the approach based on emerging evidence and current challenges, Neuro-Oncology 21 (2019) 167–178.
[18]N. Shah, B. Lin, Z. Sibenaller, T. Ryken, H. Lee, J.G. Yoon, S. Rostad, G. Foltz, Comprehensive analysis of MGMT promoter methylation: correlation with MGMT expression and clinical response in GBM, PLoS One 6 (2011), e16146.
[19]J.F. Costello, B.W. Futscher, R.A. Kroes, R.O. Pieper, Methylation-related chromatin structure is associated with exclusion of transcription factors from and suppressed expression of the O-6-methylguanine DNA methyltransferase gene in human glioma cell lines, Mol. Cell. Biol. 14 (1994) 6515–6521.
[20]K. Storey, K. Leder, A. Hawkins-Daarud, K. Swanson, A.U. Ahmed, R.C. Rockne, J. Foo, Glioblastoma recurrence and the role of O(6)-methylguanine-DNA methyltransferase promoter methylation, JCO Clinical Cancer Informatics 3 (2019) 1–12.
[21]G.J. Kitange, B.L. Carlson, A.C. Mladek, P.A. Decker, M.A. Schroeder, W. Wu, P. T. Grogan, C. Giannini, K.V. Ballman, J.C. Buckner, C.D. James, J.N. Sarkaria, Evaluation of MGMT promoter methylation status and correlation with temozolomide response in orthotopic glioblastoma xenograft model, J. Neuro- Oncol. 92 (2009) 23–31.
[22]A.B. Havik, P. Brandal, H. Honne, H.S. Dahlback, D. Scheie, M. Hektoen, T. R. Meling, E. Helseth, S. Heim, R.A. Lothe, G.E. Lind, MGMT promoter methylation in gliomas-assessment by pyrosequencing and quantitative methylation-specific PCR, J. Transl. Med. 10 (2012) 36.
[23]G. Minniti, M. Salvati, A. Arcella, F. Buttarelli, A. D’Elia, G. Lanzetta, V. Esposito, S. Scarpino, R. Maurizi Enrici, F. Giangaspero, Correlation between O6- methylguanine-DNA methyltransferase and survival in elderly patients with

glioblastoma treated with radiotherapy plus concomitant and adjuvant temozolomide, J. Neuro-Oncol. 102 (2011) 311–316.
[24]A.M. Donson, S.O. Addo-Yobo, M.H. Handler, L. Gore, N.K. Foreman, MGMT promoter methylation correlates with survival benefit and sensitivity to temozolomide in pediatric glioblastoma, Pediatr. Blood Cancer 48 (2007) 403–407.
[25]Q. Li, J. Guo, W. Wang, D. Wang, Relationship between MGMT gene expression and treatment effectiveness and prognosis in glioma, Oncol. Lett. 14 (2017) 229–233.
[26]C. Melguizo, J. Prados, B. Gonz´alez, R. Ortiz, A. Concha, P.J. Alvarez, R. Madeddu, G. Perazzoli, J.A. Oliver, R. L´opez, F. Rodríguez-Serrano,
A. Ar´anega, MGMT promoter methylation status and MGMT and CD133 immunohistochemical expression as prognostic markers in glioblastoma patients treated with temozolomide plus radiotherapy, J. Transl. Med. 10 (2012) 250.
[27]B. Oldrini, N. Vaquero-Siguero, Q. Mu, P. Kroon, Y. Zhang, M. Gal´an-Ganga,
Z. Bao, Z. Wang, H. Liu, J.K. Sa, J. Zhao, H. Kim, S. Rodriguez-Perales, D.-H. Nam, R.G.W. Verhaak, R. Rabadan, T. Jiang, J. Wang, M. Squatrito, MGMT genomic rearrangements contribute to chemotherapy resistance in gliomas, Nat. Commun. 11 (2020) 3883.
[28]M. Esteller, S.R. Hamilton, P.C. Burger, S.B. Baylin, J.G. Herman, Inactivation of the DNA repair gene O6-methylguanine-DNA methyltransferase by promoter hypermethylation is a common event in primary human neoplasia, Cancer Res. 59 (1999) 793–797.
[29]K.L. McDonald, T. Tabone, A.K. Nowak, W.N. Erber, Somatic mutations in glioblastoma are associated with methylguanine-DNA methyltransferase methylation, Oncol. Lett. 9 (2015) 2063–2067.
[30]G.Z. Yi, G. Huang, M. Guo, X. Zhang, H. Wang, S. Deng, Y. Li, W. Xiang, Z. Chen, J. Pan, Z. Li, L. Yu, B. Lei, Y. Liu, S. Qi, Acquired temozolomide resistance in MGMT-deficient glioblastoma cells is associated with regulation of DNA repair by DHC2, Brain 142 (2019) 2352–2366.
[31]J. Guo, G.Z. Yi, Z. Liu, X. Sun, R. Yang, M. Guo, Y. Li, K. Li, K. Li, X. Wang,
H. Song, S. Qi, G. Huang, Y. Liu, Quantitative proteomics analysis reveals nuclear perturbation in human glioma U87 cells treated with temozolomide, Cell Biochem. Funct. 38 (2020) 185–194.
[32]Q. Wu, A.E. Berglund, D. Wang, R.J. MacAulay, J.J. Mul´e, A.B. Etame, Paradoxical epigenetic regulation of XAF1 mediates plasticity towards adaptive resistance evolution in MGMT-methylated glioblastoma, Sci. Rep. 9 (2019) 14072.
[33]G. Perazzoli, J. Prados, R. Ortiz, O. Caba, L. Cabeza, M. Berdasco, B. G´onzalez, C. Melguizo, Temozolomide resistance in glioblastoma cell lines: implication of MGMT, MMR, P-glycoprotein and CD133 expression, PloS one 10 (2015), e0140131.
[34]J. Stritzelberger, L. Distel, R. Buslei, R. Fietkau, F. Putz, Acquired temozolomide resistance in human glioblastoma cell line U251 is caused by mismatch repair deficiency and can be overcome by lomustine, Clin. Transla. Oncol. 20 (2018) 508–516.
[35]R.K. Balvers, M.L. Lamfers, J.J. Kloezeman, A. Kleijn, L.M. Berghauser Pont, C. M. Dirven, S. Leenstra, ABT-888 enhances cytotoxic effects of temozolomide independent of MGMT status in serum free cultured glioma cells, J. Transl. Med. 13 (2015) 74.
[36]W.B. Yang, J.Y. Chuang, C.Y. Ko, W.C. Chang, T.I. Hsu, Dehydroepiandrosterone induces temozolomide resistance through modulating phosphorylation and acetylation of Sp1 in glioblastoma, Mol. Neurobiol. 56 (2019) 2301–2313.
[37]K.Y. Chang, T.I. Hsu, C.C. Hsu, S.Y. Tsai, J.J. Liu, S.W. Chou, M.S. Liu, J.P. Liou, C.Y. Ko, K.Y. Chen, J.J. Hung, W.C. Chang, C.K. Chuang, T.J. Kao, J.Y. Chuang, Specificity protein 1-modulated superoxide dismutase 2 enhances temozolomide resistance in glioblastoma, which is independent of O(6)-methylguanine-DNA methyltransferase, Redox Biol. 13 (2017) 655–664.
[38]G.M. Li, Mechanisms and functions of DNA mismatch repair, Cell Res. 18 (2008) 85–98.
[39]B. Verbeek, T.D. Southgate, D.E. Gilham, G.P. Margison, O6-methylguanine-DNA methyltransferase inactivation and chemotherapy, Br. Med. Bull. 85 (2008) 17–33.
[40]Q. Sun, C. Pei, Q. Li, T. Dong, Y. Dong, W. Xing, P. Zhou, Y. Gong, Z. Zhen, Y. Gao, Y. Xiao, J. Su, H. Ren, Up-regulation of MSH6 is associated with temozolomide resistance in human glioblastoma, Biochem. Biophys. Res. Commun. 496 (2018) 1040–1046.
[41]S. Choi, Y. Yu, M.R. Grimmer, M. Wahl, S.M. Chang, J.F. Costello, Temozolomide- associated hypermutation in gliomas, Neuro-Oncology 20 (2018) 1300–1309.
[42]J.L. McFaline-Figueroa, C.J. Braun, M. Stanciu, Z.D. Nagel, P. Mazzucato,
D. Sangaraju, E. Cerniauskas, K. Barford, A. Vargas, Y. Chen, N. Tretyakova, J. A. Lees, M.T. Hemann, F.M. White, L.D. Samson, Minor changes in expression of the mismatch repair protein MSH2 exert a major impact on glioblastoma response to temozolomide, Cancer Res. 75 (2015) 3127–3138.
[43]Y. Shinsato, T. Furukawa, S. Yunoue, H. Yonezawa, K. Minami, Y. Nishizawa, R. Ikeda, K. Kawahara, M. Yamamoto, H. Hirano, H. Tokimura, K. Arita,
Reduction of MLH1 and PMS2 confers temozolomide resistance and is associated with recurrence of glioblastoma, Oncotarget 4 (2013) 2261–2270.
[44]S. Yip, J. Miao, D.P. Cahill, A.J. Iafrate, K. Aldape, C.L. Nutt, D.N. Louis, MSH6 mutations arise in glioblastomas during temozolomide therapy and mediate temozolomide resistance, Clinical Cancer Res. 15 (2009) 4622–4629.
[45]H.E. Krokan, M. Bjørås, Base excision repair, Cold Spring Harb. Perspect. Biol. 5 (2013) a012583.
[46]A.P. Montaldi, P.R. Godoy, E.T. Sakamoto-Hojo, APE1/REF-1 down-regulation enhances the cytotoxic effects of temozolomide in a resistant glioblastoma cell line, mutation research, Genet. Toxicol. Environ. Mutagen. 793 (2015) 19–29.

[47]J.-B. Tang, D. Svilar, R.N. Trivedi, X.-H. Wang, E.M. Goellner, B. Moore, R. L. Hamilton, L.A. Banze, A.R. Brown, R.W. Sobol, N-methylpurine DNA
glycosylase and DNA polymerase beta modulate BER inhibitor potentiation of glioma cells to temozolomide, Neuro-Oncology 13 (2011) 471–486.
[48]G. Serrano-Heras, B. Castro-Robles, C.M. Romero-S´anchez, B. Carri´on,
R. Barbella-Aponte, H. Sandoval, T. Segura, Involvement of N-methylpurine DNA glycosylase in resistance to temozolomide in patient-derived glioma cells, Sci. Rep. 10 (2020) 22185.
[49]A.L. Jacobs, P. Sch¨ar, DNA glycosylases: in DNA repair and beyond, Chromosoma 121 (2012) 1–20.
[50]S. Agnihotri, A.S. Gajadhar, C. Ternamian, T. Gorlia, K.L. Diefes, P.S. Mischel,
J.Kelly, G. McGown, M. Thorncroft, B.L. Carlson, J.N. Sarkaria, G.P. Margison,
K.Aldape, C. Hawkins, M. Hegi, A. Guha, Alkylpurine-DNA-N-glycosylase confers resistance to temozolomide in xenograft models of glioblastoma multiforme and is associated with poor survival in patients, J. Clin. Invest. 122 (2012) 253–266.
[51]K. Yoshimoto, M. Mizoguchi, N. Hata, H. Murata, R. Hatae, T. Amano,
A. Nakamizo, T. Sasaki, Complex DNA repair pathways as possible therapeutic targets to overcome temozolomide resistance in glioblastoma, Front. Oncol. 2 (2012).
[52]S. Jiapaer, T. Furuta, S. Tanaka, T. Kitabayashi, M. Nakada, Potential strategies overcoming the temozolomide resistance for glioblastoma, Neurol. Med. Chir. (Tokyo) 58 (2018) 405–421.
[53]G.E. Ronson, A.L. Piberger, M.R. Higgs, A.L. Olsen, G.S. Stewart, P.J. McHugh, E. Petermann, N.D. Lakin, PARP1 and PARP2 stabilise replication forks at base excision repair intermediates through Fbh1-dependent Rad51 regulation, Nat. Commun. 9 (2018) 746.
[54]S. Taipakova, A. Kuanbay, C. Saint-Pierre, D. Gasparutto, Y. Baiken, R. Groisman, A.A. Ishchenko, M. Saparbaev, A.K. Bissenbaev, The arabidopsis thaliana poly (ADP-ribose) polymerases 1 and 2 modify DNA by ADP-ribosylating terminal phosphate residues, Front. Cell Dev. Biol. 8 (2020).
[55]S. Wu, X. Li, F. Gao, J.F. de Groot, D. Koul, W.K.A. Yung, PARP-mediated PARylation of MGMT is critical to promote repair of temozolomide-induced O6- methylguanine DNA damage in glioblastoma, Neuro-Oncology 23 (2021) 920–931.
[56]D. Criscuolo, F. Morra, R. Giannella, R. Visconti, A. Cerrato, A. Celetti, New combinatorial strategies to improve the PARP inhibitors efficacy in the urothelial bladder cancer treatment, J. Experimental Clin. Cancer Res. 38 (2019) 91.
[57]L. Tentori, L. Ricci-Vitiani, A. Muzi, F. Ciccarone, F. Pelacchi, R. Calabrese,
D. Runci, R. Pallini, P. Caiafa, G. Graziani, Pharmacological inhibition of poly (ADP-ribose) polymerase-1 modulates resistance of human glioblastoma stem cells to temozolomide, BMC Cancer 14 (2014) 151.
[58]E.M. Goellner, B. Grimme, A.R. Brown, Y.-C. Lin, X.-H. Wang, K.F. Sugrue, L. Mitchell, R.N. Trivedi, J.-B. Tang, R.W. Sobol, Overcoming temozolomide resistance in glioblastoma via dual inhibition of NAD biosynthesis and base excision repair, Cancer Res. 71 (2011) 2308–2317.
[59]H. Kim, A. D’Andrea, Regulation of DNA cross-link repair by the Fanconi anemia/
BRCA pathway, Genes Dev. 26 (2012) 1393–1408.
[60]C.-B. Fang, H.-T. Wu, M.-L. Zhang, J. Liu, G.-J. Zhang, Fanconi anemia pathway: mechanisms of breast cancer predisposition development and potential therapeutic targets, Front. Cell Dev. Biol. 8 (2020).
[61]P.R. Andreassen, A.D. D’Andrea, T. Taniguchi, ATR couples FANCD2 monoubiquitination to the DNA-damage response, Genes Dev. 18 (2004) 1958–1963.
[62]N. Kondo, A. Takahashi, K. Ono, T. Ohnishi, DNA damage induced by alkylating agents and repair pathways, J. Nucleic Acids 2010 (2010), 543531.
[63]A.J. Davis, D.J. Chen, DNA double strand break repair via non-homologous end- joining, Transl. Cancer Res. 2 (2013) 130–143.
[64]A. Dud´as, M. Chovanec, DNA double-strand break repair by homologous recombination, Mutat. Res. 566 (2004) 131–167.
[65]W.P. Roos, T. Nikolova, S. Quiros, S.C. Naumann, O. Kiedron, M.Z. Zdzienicka, B. Kaina, Brca2/Xrcc2 dependent HR, but not NHEJ, is required for protection against O(6)-methylguanine triggered apoptosis, DSBs and chromosomal aberrations by a process leading to SCEs, DNA repair 8 (2009) 72–86.
[66]A.A. Patil, P. Sayal, M.L. Depondt, R.D. Beveridge, A. Roylance, D.H. Kriplani, K. N. Myers, A. Cox, D. Jellinek, M. Fernando, T.A. Carroll, S.J. Collis, FANCD2 re- expression is associated with glioma grade and chemical inhibition of the Fanconi Anaemia pathway sensitises gliomas to chemotherapeutic agents, Oncotarget 5 (2014) 6414–6424.
[67]D.S. Metselaar, M.H. Meel, B. Benedict, P. Waranecki, J. Koster, G.J.L. Kaspers, E. Hulleman, Celastrol-induced degradation of FANCD2 sensitizes pediatric high- grade gliomas to the DNA-crosslinking agent carboplatin, EBioMedicine 50 (2019) 81–92.
[68]F. Guo, Mtor-Fanconi anemia DNA damage repair pathway in cancer, J. Oncobiomarkers 2 (2014).
[69]C. Shen, D. Oswald, D. Phelps, H. Cam, C.E. Pelloski, Q. Pang, P.J. Houghton, Regulation of FANCD2 by the mTOR pathway contributes to the resistance of cancer cells to DNA double-strand breaks, Cancer Res. 73 (2013) 3393–3401.
[70]vi209-vi209 O. Rominiyi, K. Myers, N. Gomez-Roman, N. Lad, D. Dar, D. Jellinek, A. Chalmers, T. Carroll, B. Chen, Y. Al-Tamimi, S. Collis, RDNA-12., The fanconi anaemia (FA) pathway and glioblastoma: a new foundation for dna damage response targeted combinations, Neuro Oncol 21 (2019).
[71]N. Kondo, A. Takahashi, E. Mori, T. Noda, M.Z. Zdzienicka, L.H. Thompson, T. Helleday, M. Suzuki, Y. Kinashi, S. Masunaga, K. Ono, M. Hasegawa,
T. Ohnishi, FANCD1/BRCA2 plays predominant role in the repair of DNA damage induced by ACNU or TMZ, PloS one 6 (2011), e19659.

[72]Y. Chun, J. Kim, Autophagy: an essential degradation program for cellular homeostasis and life, Cells 7 (2018) 278.
[73]E. White, R.S. DiPaola, The double-edged sword of autophagy modulation in cancer, Clin. Cancer Res. 15 (2009) 5308–5316.
[74]J.E. Simpson, N. Gammoh, The impact of autophagy during the development and survival of glioblastoma, Open Biol. 10 (2020), 200184.
[75]Y. Zou, Q. Wang, B. Li, B. Xie, W. Wang, Temozolomide induces autophagy via ATM-AMPK-ULK1 pathways in glioma, Mol. Med. Rep. 10 (2014) 411–416.
[76]J.R.D. Pearson, T. Regad, Targeting cellular pathways in glioblastoma multiforme, Signal Transduct. Target. Therapy 2 (2017) 17040.
[77]M. Buccarelli, M. Marconi, S. Pacioni, I. De Pascalis, Q.G. D’Alessandris,
M. Martini, B. Ascione, W. Malorni, L.M. Larocca, R. Pallini, L. Ricci-Vitiani, P. Matarrese, Inhibition of autophagy increases susceptibility of glioblastoma
stem cells to temozolomide by igniting ferroptosis, Cell Death Dis. 9 (2018) 841.
[78]M. Mehrpour, A. Esclatine, I. Beau, P. Codogno, Overview of macroautophagy regulation in mammalian cells, Cell Res. 20 (2010) 748–762.
[79]S.A. Tooze, Current views on the source of the autophagosome membrane, Essays Biochem. 55 (2013) 29–38.
[80]H. Xiang, J. Zhang, C. Lin, L. Zhang, B. Liu, L. Ouyang, Targeting autophagy- related protein kinases for potential therapeutic purpose, Acta Pharm. Sin. B 10 (2020) 569–581.
[81]A.H. Lystad, A. Simonsen, Mechanisms and pathophysiological roles of the ATG8 conjugation machinery, Cells 8 (2019).
[82]K.R. Parzych, D.J. Klionsky, An overview of autophagy: morphology, mechanism, and regulation, Antioxid. Redox Signal. 20 (2014) 460–473.
[83]E.-J. Yun, S. Kim, J.-T. Hsieh, S.T. Baek, Wnt/ß-catenin signaling pathway induces autophagy-mediated temozolomide-resistance in human glioblastoma, Cell Death Dis. 11 (2020) 771.
[84]Y. Chen, D. Meng, H. Wang, R. Sun, D. Wang, S. Wang, J. Fan, Y. Zhao, J. Wang, S. Yang, C. Huai, X. Song, R. Qin, T. Xu, D. Yun, L. Hu, J. Yang, X. Zhang, H. Chen, J. Chen, H. Chen, D. Lu, VAMP8 facilitates cellular proliferation and temozolomide resistance in human glioma cells, Neuro-Oncology 17 (2015) 407–418.
[85]A.U. Ahmed, B. Auffinger, M.S. Lesniak, Understanding glioma stem cells: rationale, clinical relevance and therapeutic strategies, Expert. Rev. Neurother. 13 (2013) 545–555.
[86]M. Guerra-Rebollo, C. Garrido, L. S´anchez-Cid, C. Soler-Botija, O. Meca-Cort´es, N. Rubio, J. Blanco, Targeting of replicating CD133 and OCT4/SOX2 expressing glioma stem cells selects a cell population that reinitiates tumors upon release of therapeutic pressure, Sci. Rep. 9 (2019) 9549.
[87]J. Chen, Y. Li, T.S. Yu, R.M. McKay, D.K. Burns, S.G. Kernie, L.F. Parada,
A restricted cell population propagates glioblastoma growth after chemotherapy, Nature 488 (2012) 522–526.
[88]B. Auffinger, A.L. Tobias, Y. Han, G. Lee, D. Guo, M. Dey, M.S. Lesniak, A.
U. Ahmed, Conversion of differentiated cancer cells into cancer stem-like cells in a glioblastoma model after primary chemotherapy, Cell Death Differ. 21 (2014) 1119–1131.
[89]R. Guan, X. Zhang, M. Guo, Glioblastoma stem cells and Wnt signaling pathway: molecular mechanisms and therapeutic targets, Chin. Neurosurg. J. 6 (2020) 25.
[90]A.K. Suwala, K. Koch, D.H. Rios, P. Aretz, C. Uhlmann, I. Ogorek, J. Felsberg, G. Reifenberger, K. K¨ohrer, R. Deenen, H.J. Steiger, U.D. Kahlert, J. Maciaczyk, Inhibition of Wnt/beta-catenin signaling downregulates expression of aldehyde dehydrogenase isoform 3A1 (ALDH3A1) to reduce resistance against temozolomide in glioblastoma in vitro, Oncotarget 9 (2018) 22703–22716.
[91]P. Doan, A. Musa, A. Murugesan, V. Sipil¨a, N.R. Candeias, F. Emmert-Streib, P. Ruusuvuori, K. Granberg, O. Yli-Harja, M. Kandhavelu, Glioblastoma
multiforme stem cell cycle arrest by alkylaminophenol through the modulation of EGFR and CSC signaling pathways, Cells 9 (2020).
[92]V. Juric, H. Düssmann, M.L.M. Lamfers, J.H.M. Prehn, M. Rehm, B.M. Murphy, Transcriptional CDK inhibitors CYC065 and THZ1 induce apoptosis in glioma stem cells derived from recurrent GBM, Cells 10 (2021).
[93]S.K. Singh, C. Hawkins, I.D. Clarke, J.A. Squire, J. Bayani, T. Hide, R.
M. Henkelman, M.D. Cusimano, P.B. Dirks, Identification of human brain tumour initiating cells, Nature 432 (2004) 396–401.
[94]S. Hombach-Klonisch, M. Mehrpour, S. Shojaei, C. Harlos, M. Pitz, A. Hamai, K. Siemianowicz, W. Likus, E. Wiechec, B.D. Toyota, R. Hoshyar, A. Seyfoori, Z. Sepehri, S.R. Ande, F. Khadem, M. Akbari, A.M. Gorman, A. Samali,
T. Klonisch, S. Ghavami, Glioblastoma and chemoresistance to alkylating agents: involvement of apoptosis, autophagy, and unfolded protein response, Pharmacol. Ther. 184 (2018) 13–41.
[95]D. Beier, P. Hau, M. Proescholdt, A. Lohmeier, J. Wischhusen, P.J. Oefner, L. Aigner, A. Brawanski, U. Bogdahn, C.P. Beier, CD133(+) and CD133(-)
glioblastoma-derived cancer stem cells show differential growth characteristics and molecular profiles, Cancer Res. 67 (2007) 4010–4015.
[96]C.E. Griguer, C.R. Oliva, E. Gobin, P. Marcorelles, D.J. Benos, J.R. Lancaster Jr., G.Y. Gillespie, CD133 is a marker of bioenergetic stress in human glioma, PloS one 3 (2008), e3655.
[97]M. Jaksch, J. Múnera, R. Bajpai, A. Terskikh, R.G. Oshima, Cell cycle-dependent variation of a CD133 epitope in human embryonic stem cell, colon cancer, and melanoma cell lines, Cancer Res. 68 (2008) 7882–7886.
[98]J. Wang, P. Sakariassen, O. Tsinkalovsky, H. Immervoll, S.O. Bøe, A. Svendsen, L. Prestegarden, G. Røsland, F. Thorsen, L. Stuhr, A. Molven, R. Bjerkvig,
P. Enger, CD133 negative glioma cells form tumors in nude rats and give rise to CD133 positive cells, Int. J. Cancer 122 (2008) 761–768.

[99]K. Ohnishi, T. Tani, N. Tojo, J.I. Sagara, Glioblastoma cell line shows phenotypes of cancer stem cells in hypoxic microenvironment of spheroids, Biochem. Biophys. Res. Commun. 546 (2021) 150–154.
[100]I.V. Ulasov, S. Nandi, M. Dey, A.M. Sonabend, M.S. Lesniak, Inhibition of sonic hedgehog and notch pathways enhances sensitivity of CD133(+) glioma stem cells to temozolomide therapy, Mol. Med. (Cambridge, Mass.) 17 (2011) 103–112.
[101]T.A. Read, M.P. Fogarty, S.L. Markant, R.E. McLendon, Z. Wei, D.W. Ellison, P. G. Febbo, R.J. Wechsler-Reya, Identification of CD15 as a marker for tumor- propagating cells in a mouse model of medulloblastoma, Cancer cell 15 (2009) 135–147.
[102]J. Xu, H. Hardin, R. Zhang, K. Sundling, D. Buehler, R.V. Lloyd, Stage-specific embryonic Antigen-1 (SSEA-1) expression in thyroid tissues, Endocr. Pathol. 27 (2016) 271–275.
[103]W. Lin, J.F. Modiano, D. Ito, Stage-specific embryonic antigen: determining expression in canine glioblastoma, melanoma, and mammary cancer cells, J. Vet. Sci. 18 (2017) 101–104.
[104]M.J. Son, K. Woolard, D.H. Nam, J. Lee, H.A. Fine, SSEA-1 is an enrichment marker for tumor-initiating cells in human glioblastoma, Cell Stem Cell 4 (2009) 440–452.
[105]J. Kwon, T.S. Lee, H.W. Lee, M.C. Kang, H.J. Yoon, J.H. Kim, J.H. Park, Integrin alpha 6: a novel therapeutic target in esophageal squamous cell carcinoma, Int. J. Oncol. 43 (2013) 1523–1530.
[106]J. Cooper, F.G. Giancotti, Integrin signaling in cancer: mechanotransduction, stemness, epithelial plasticity, and therapeutic resistance, Cancer Cell 35 (2019) 347–367.
[107]J.D. Lathia, J. Gallagher, J.M. Heddleston, J. Wang, C.E. Eyler, J. Macswords, Q. Wu, A. Vasanji, R.E. McLendon, A.B. Hjelmeland, J.N. Rich, Integrin alpha 6 regulates glioblastoma stem cells, Cell Stem Cell 6 (2010) 421–432.
[108]I. Tooyama, T. Yamada, S.U. Kim, P.L. McGeer, Immunohistochemical study of A2B5-positive ganglioside in postmortem human brain tissue of alzheimer disease, amyotrophic lateral sclerosis, progressive supranuclear palsy and control cases, Neurosci. Lett. 136 (1992) 91–94.
[109]N. Baeza-Kallee, R. Berg`es, A. Soub´eran, C. Colin, E. Denicolaï, R. Appay,
A. Tchoghandjian, D. Figarella-Branger, Glycolipids recognized by A2B5 antibody promote proliferation, migration, and clonogenicity in glioblastoma cells, Cancers 11 (2019).
[110]A. Tchoghandjian, N. Baeza, C. Colin, M. Cayre, P. Metellus, C. Beclin, L. Ouafik, D. Figarella-Branger, A2B5 cells from human glioblastoma have cancer stem cell properties, Brain Pathol. (Zurich, Switzerland) 20 (2010) 211–221.
[111]A.T. Ogden, A.E. Waziri, R.A. Lochhead, D. Fusco, K. Lopez, J.A. Ellis, J. Kang, M. Assanah, G.M. McKhann, M.B. Sisti, P.C. McCormick, P. Canoll, J.N. Bruce, Identification of A2B5 CD133- tumor-initiating cells in adult human gliomas, Neurosurgery 62 (2008) 505–514, discussion 514–505.
[112]Y. Han, H. Wang, Y. Huang, Z. Cheng, T. Sun, G. Chen, X. Xie, Y. Zhou, Z. Du, Isolation and characteristics of CD133-/A2B5 and CD133-/A2B5- cells from the SHG139s cell line, Mol. Med. Rep. 12 (2015) 7949–7956.
[113]Z. Wu, Y. Han, Y. Li, X. Li, T. Sun, G. Chen, Y. Huang, Y. Zhou, Z. Du, MiR-218-5p inhibits the stem cell properties and invasive ability of the A2B5 CD133- subgroup of human glioma stem cells, Oncol. Rep. 35 (2016) 869–877.
[114]R. Auvergne, C. Wu, A. Connell, S. Au, A. Cornwell, M. Osipovitch, A. Benraiss, S. Dangelmajer, H. Guerrero-Cazares, A. Quinones-Hinojosa, S.A. Goldman, PAR1 inhibition suppresses the self-renewal and growth of A2B5-defined glioma progenitor cells and their derived gliomas in vivo, Oncogene 35 (2016) 3817–3828.
[115]Z. Du, D. Jia, S. Liu, F. Wang, G. Li, Y. Zhang, X. Cao, E.A. Ling, A. Hao, Oct4 is expressed in human gliomas and promotes colony formation in glioma cells, Glia 57 (2009) 724–733.
[116]C709-718 Y.J. Wang, M. Herlyn, The emerging roles of Oct4 in tumor-initiating cells, Am. J. Physiol. Cell Physiol. 309 (2015).
[117]A.K. Rooj, A. Bronisz, J. Godlewski, The role of octamer binding transcription factors in glioblastoma multiforme, Biochim. Biophys. Acta 2016 (1859) 805–811.
[118]P. Wang, W.W. Wan, S.L. Xiong, H. Feng, N. Wu, Cancer stem-like cells can be induced through dedifferentiation under hypoxic conditions in glioma, hepatoma and lung cancer, Cell Death Discovery 3 (2017) 16105.
[119]J. Krogh Petersen, P. Jensen, M. Dahl Sørensen, B. Winther Kristensen, Expression and prognostic value of Oct-4 in astrocytic brain tumors, PLoS One 11 (2016), e0169129.
[120]J. Wang, S.L. Xu, J.J. Duan, L. Yi, Y.F. Guo, Y. Shi, L. Li, Z.Y. Yang, X.M. Liao, J. Cai, Y.Q. Zhang, H.L. Xiao, L. Yin, H. Wu, J.N. Zhang, S.Q. Lv, Q.K. Yang, X. J. Yang, T. Jiang, X. Zhang, X.W. Bian, S.C. Yu, Invasion of white matter tracts by glioma stem cells is regulated by a NOTCH1-SOX2 positive-feedback loop, Nat. Neurosci. 22 (2019) 91–105.
[121]L. Garros-Regulez, P. Aldaz, O. Arrizabalaga, V. Moncho-Amor, E. Carrasco- Garcia, L. Manterola, L. Moreno-Cugnon, C. Barrena, J. Villanua, I. Ruiz,
S. Pollard, R. Lovell-Badge, N. Sampron, I. Garcia, A. Matheu, mTOR inhibition decreases SOX2-SOX9 mediated glioma stem cell activity and temozolomide resistance, Expert Opin. Ther. Targets 20 (2016) 393–405.
[122]W. Luo, D. Yan, Z. Song, X. Zhu, X. Liu, X. Li, S. Zhao, miR-126-3p sensitizes glioblastoma cells to temozolomide by inactivating Wnt/ß-catenin signaling via targeting SOX2, Life Sci. 226 (2019) 98–106.
[123]J. Marjanovic Vicentic, D. Drakulic, I. Garcia, V. Vukovic, P. Aldaz, N. Puskas, I. Nikolic, G. Tasic, S. Raicevic, L. Garros-Regulez, N. Sampron, M.J. Atkinson, N. Anastasov, A. Matheu, M. Stevanovic, SOX3 can promote the malignant behavior of glioblastoma cells, Cell. Oncol. 42 (2019) 41–54.

[124]J. O’Brien, H. Hayder, Y. Zayed, C. Peng, Overview of MicroRNA biogenesis, mechanisms of actions, and circulation, Front. Endocrinol. 9 (2018) 402.
[125]A. Buruiana, A.I.Florian I. Florian ? T.L. Timi?, The Roles of miRNA in Glioblastoma Tumor Cell Communication: Diplomatic and Aggressive Negotiations, International journal of molecular sciences 21 (2020).
[126]B. Zhang, X. Pan, G.P. Cobb, T.A. Anderson, microRNAs as oncogenes and tumor suppressors, Dev. Biol. 302 (2007) 1–12.
[127]B. Banelli, A. Forlani, G. Allemanni, A. Morabito, M.P. Pistillo, M. Romani, MicroRNA in glioblastoma: an overview, Int. J. Genom 2017 (2017) 7639084.
[128]J. Godlewski, H.B. Newton, E.A. Chiocca, S.E. Lawler, MicroRNAs and glioblastoma; the stem cell connection, Cell Death Differ. 17 (2010) 221–228.
[129]K. Wu, F. Xing, S.Y. Wu, K. Watabe, Extracellular vesicles as emerging targets in cancer: recent development from bench to bedside, biochimica et biophysica acta, Rev. Cancer 2017 (1868) 538–563.
[130]M. Mizoguchi, Y. Guan, K. Yoshimoto, N. Hata, T. Amano, A. Nakamizo, T. Sasaki, MicroRNAs in human malignant gliomas, J. Oncol. 2012 (2012), 732874.
[131]S. Srinivasan, I.R. Patric, K. Somasundaram, A ten-microRNA expression signature predicts survival in glioblastoma, PloS one 6 (2011), e17438.
[132]J. Yin, A. Zeng, Z. Zhang, Z. Shi, W. Yan, Y. You, Exosomal transfer of miR-1238 contributes to temozolomide-resistance in glioblastoma, EBioMedicine 42 (2019) 238–251.
[133]R.F. Liang, M. Li, Y. Yang, X. Wang, Q. Mao, Y.H. Liu, Circulating miR-128 as a potential diagnostic biomarker for glioma, Clin. Neurol. Neurosurg. 160 (2017) 88–91.
[134]C. Zhao, R. Guo, F. Guan, S. Ma, M. Li, J. Wu, X. Liu, H. Li, B. Yang, MicroRNA- 128-3p enhances the chemosensitivity of temozolomide in glioblastoma by targeting c-met and EMT, Sci. Rep. 10 (2020) 9471.
[135]G. Regazzo, I. Terrenato, M. Spagnuolo, M. Carosi, G. Cognetti, L. Cicchillitti,
F. Sperati, V. Villani, C. Carapella, G. Piaggio, A. Pelosi, M.G. Rizzo, A restricted signature of serum miRNAs distinguishes glioblastoma from lower grade gliomas, J. Experiment. Clin. Cancer Res. 35 (2016) 124.
[136]X. Hong, W.C. Sin, A.L. Harris, C.C. Naus, Gap junctions modulate glioma invasion by direct transfer of microRNA, Oncotarget 6 (2015) 15566–15577.
[137]J.L. Munoz, S.A. Bliss, S.J. Greco, S.H. Ramkissoon, K.L. Ligon, P. Rameshwar, Delivery of functional anti-miR-9 by mesenchymal stem cell-derived exosomes to glioblastoma multiforme cells conferred chemosensitivity, Mol. Therapy. Nucleic Acids 2 (2013), e126.
[138]L. Shi, J. Chen, J. Yang, T. Pan, S. Zhang, Z. Wang, MiR-21 protected human glioblastoma U87MG cells from chemotherapeutic drug temozolomide induced apoptosis by decreasing Bax/Bcl-2 ratio and caspase-3 activity, Brain Res. 1352 (2010) 255–264.
[139]S. Zhang, Y. Wan, T. Pan, X. Gu, C. Qian, G. Sun, L. Sun, Y. Xiang, Z. Wang, L. Shi, MicroRNA-21 inhibitor sensitizes human glioblastoma U251 stem cells to chemotherapeutic drug temozolomide, J. Mol. Neurosci. 47 (2012) 346–356.
[140]K. Ujifuku, N. Mitsutake, S. Takakura, M. Matsuse, V. Saenko, K. Suzuki,
K. Hayashi, T. Matsuo, K. Kamada, I. Nagata, S. Yamashita, miR-195, miR-455-3p and miR-10a(*) are implicated in acquired temozolomide resistance in glioblastoma multiforme cells, Cancer Lett. 296 (2010) 241–248.
[141]T. Papagiannakopoulos, A. Shapiro, K.S. Kosik, MicroRNA-21 targets a network of key tumor-suppressive pathways in glioblastoma cells, Cancer Res. 68 (2008) 8164–8172.
[142]L. Chen, J. Zhang, L. Han, A. Zhang, C. Zhang, Y. Zheng, T. Jiang, P. Pu, C. Jiang, C. Kang, Downregulation of miR-221/222 sensitizes glioma cells to temozolomide by regulating apoptosis independently of p53 status, Oncol. Rep. 27 (2012) 854–860.
[143]M. Shu, X. Zheng, S. Wu, H. Lu, T. Leng, W. Zhu, Y. Zhou, Y. Ou, X. Lin, Y. Lin, D. Xu, Y. Zhou, G. Yan, Targeting oncogenic miR-335 inhibits growth and invasion of malignant astrocytoma cells, Mol. Cancer 10 (2011) 59.
[144]Z.X. Cheng, W.B. Yin, Z.Y. Wang, MicroRNA-132 induces temozolomide resistance and promotes the formation of cancer stem cell phenotypes by targeting tumor suppressor candidate 3 in glioblastoma, Int. J. Mol. Med. 40 (2017) 1307–1314.
[145]R. Gong, Z.Q. Li, K. Fu, C. Ma, W. Wang, J.C. Chen, Long noncoding RNA PVT1 promotes stemness and temozolomide resistance through miR-365/ELF4/SOX2 Axis in glioma, Experiment. Neurobiol. 30 (2021) 244–255.
[146]T. Tian, M. Mingyi, X. Qiu, Y. Qiu, MicroRNA-101 reverses temozolomide resistance by inhibition of GSK3beta in glioblastoma, Oncotarget 7 (2016) 79584–79595.
[147]Y.T. Gao, X.B. Chen, H.L. Liu, Up-regulation of miR-370-3p restores glioblastoma multiforme sensitivity to temozolomide by influencing MGMT expression, Sci. Rep. 6 (2016) 32972.
[148]D. Kushwaha, V. Ramakrishnan, K. Ng, T. Steed, T. Nguyen, D. Futalan, J.
C. Akers, J. Sarkaria, T. Jiang, D. Chowdhury, B.S. Carter, C.C. Chen, A genome- wide miRNA screen revealed miR-603 as a MGMT-regulating miRNA in glioblastomas, Oncotarget 5 (2014) 4026–4039.
[149]Y.Y. Lee, A.A. Yarmishyn, M.L. Wang, H.Y. Chen, S.H. Chiou, Y.P. Yang, C.F. Lin, P.I. Huang, Y.W. Chen, H.I. Ma, M.T. Chen, MicroRNA-142-3p is involved in regulation of MGMT expression in glioblastoma cells, Cancer Manag. Res. 10 (2018) 775–785.
[150]Z. Wang, Z. Li, Y. Fu, L. Han, Y. Tian, MiRNA-130a-3p inhibits cell proliferation, migration, and TMZ resistance in glioblastoma by targeting Sp1, Am. J. Transl. Res. 11 (2019) 7272–7285.
[151]S.A. Mirzaei, S. Reiisi, P. Ghiasi Tabari, A. Shekari, F. Aliakbari, E. Azadfallah, F. Elahian, Broad blocking of MDR efflux pumps by acetylshikonin and acetoxyisovalerylshikonin to generate hypersensitive phenotype of malignant carcinoma cells, Sci. Rep. 8 (2018) 3446.

[152]B. Liu, Z. Guo, H. Dong, T. Daofeng, Q. Cai, B. Ji, S. Zhang, L. Wu, J. Wang, L. Wang, X. Zhu, Y. Liu, Q. Chen, LRIG1, human EGFR inhibitor, reverses
multidrug resistance through modulation of ABCB1 and ABCG2, Brain Res. 1611 (2015) 93–100.
[153]P. Zhang, X.B. Chen, B.Q. Ding, H.L. Liu, T. He, Down-regulation of ABCE1 inhibits temozolomide resistance in glioma through the PI3K/Akt/NF-?B signaling pathway, Biosci. Rep. 38 (2018).
[154]X. Zhang, X. Liu, W. Zhou, M. Yang, Y. Ding, Q. Wang, R. Hu, Fasudil increases temozolomide sensitivity and suppresses temozolomide-resistant glioma growth via inhibiting ROCK2/ABCG2, Cell Death Dis. 9 (2018) 190.
[155]D.N. Louis, A. Perry, G. Reifenberger, A. von Deimling, D. Figarella-Branger, W. K. Cavenee, H. Ohgaki, O.D. Wiestler, P. Kleihues, D.W. Ellison, The 2016 World Health Organization classification of tumors of the central nervous system: a summary, Acta Neuropathol. 131 (2016) 803–820.
[156]M.S. Tomar, A. Shrivastava, TERT promoter mutations correlate with IDHs, MGMT and EGFR in glioblastoma multiforme, Neurol. India 69 (2021) 135–136.
[157]M. Barange, S. Epari, M. Gurav, O. Shetty, A. Sahay, P. Shetty, J. Goda,
A. Moyiadi, T. Gupta, R. Jalali, TERT promoter mutation in adult glioblastomas: it’s correlation with other relevant molecular markers, Neurol. India 69 (2021) 126–134.
[158]A. Olar, K.D. Aldape, Using the molecular classification of glioblastoma to inform personalized treatment, J. Pathol. 232 (2014) 165–177.
[159]R.J. Molenaar, J.P. Maciejewski, J.W. Wilmink, C.J.F. van Noorden, Wild-type and mutated IDH1/2 enzymes and therapy responses, Oncogene 37 (2018) 1949–1960.
[160]L. Deng, P. Xiong, Y. Luo, X. Bu, S. Qian, W. Zhong, S. Lv, Association between IDH1/2 mutations and brain glioma grade, Oncol. Lett. 16 (2018) 5405–5409.
[161]J. Mondesir, C. Willekens, M. Touat, S. de Botton, IDH1 and IDH2 mutations as novel therapeutic targets: current perspectives, J. Blood Med. 7 (2016) 171–180.
[162]S. Han, Y. Liu, S.J. Cai, M. Qian, J. Ding, M. Larion, M.R. Gilbert, C. Yang, IDH mutation in glioma: molecular mechanisms and potential therapeutic targets, Br. J. Cancer 122 (2020) 1580–1589.
[163]A.L. Cohen, S.L. Holmen, H. Colman, IDH1 and IDH2 mutations in gliomas, Curr. Neurol. Neurosci. Rep. 13 (2013) 345.
[164]J.B. Wang, D.F. Dong, M.D. Wang, K. Gao, IDH1 overexpression induced chemotherapy resistance and IDH1 mutation enhanced chemotherapy sensitivity in glioma cells in vitro and in vivo, Asian Pacific J. Cancer Prevent 15 (2014) 427–432.
[165]P. Yang, W. Zhang, Y. Wang, X. Peng, B. Chen, X. Qiu, G. Li, S. Li, C. Wu, K. Yao, W. Li, W. Yan, J. Li, Y. You, C.C. Chen, T. Jiang, IDH mutation and MGMT promoter methylation in glioblastoma: results of a prospective registry, Oncotarget 6 (2015) 40896–40906.
[166]A.A. Pandith, I. Qasim, S.M. Baba, A. Koul, W. Zahoor, D. Afroze, A. Lateef, U. Manzoor, I.A. Bhat, D. Sanadhya, A.R. Bhat, A.U. Ramzan, F. Mohammad,
I. Anwar, Favorable role of IDH1/2 mutations aided with MGMT promoter gene methylation in the outcome of patients with malignant glioma, Future Sci. OA 7 (2020) FSO663.
[167]Y. Liu, F. Lang, F.J. Chou, K.A. Zaghloul, C. Yang, Isocitrate dehydrogenase mutations in glioma: genetics, biochemistry, and clinical indications, Biomedicines 8 (2020).
[168]Z.K. Qiu, D. Shen, Y.S. Chen, Q.Y. Yang, C.C. Guo, B.H. Feng, Z.P. Chen, Enhanced MGMT expression contributes to temozolomide resistance in glioma stem-like cells, Chin. J. Cancer 33 (2014) 115–122.
[169]Y. Bi, H. Li, D. Yi, Y. Bai, S. Zhong, Q. Liu, Y. Chen, G. Zhao, ß-catenin contributes to cordycepin-induced MGMT inhibition and reduction of temozolomide resistance in glioma cells by increasing intracellular reactive oxygen species, Cancer Lett. 435 (2018) 66–79.
[170]C.Y. Tsai, H.J. Ko, S.J. Chiou, Y.L. Lai, C.C. Hou, T. Javaria, Z.Y. Huang, T.
S. Cheng, T.I. Hsu, J.Y. Chuang, A.L. Kwan, T.H. Chuang, C.F. Huang, J.K. Loh, Y. R. Hong, NBM-BMX, an HDAC8 inhibitor, overcomes temozolomide resistance in glioblastoma multiforme by downregulating the ß-catenin/c-Myc/SOX2 pathway and upregulating p53-mediated MGMT inhibition, Int. J. Mol. Sci. 22 (2021).
[171]X. Yu, M. Wang, J. Zuo, A. Wahafu, P. Mao, R. Li, W. Wu, W. Xie, J. Wang, Nuclear factor I a promotes temozolomide resistance in glioblastoma via activation of nuclear factor κB pathway, Life Sci. 236 (2019), 116917.
[172]Z. Yu, Y. Chen, S. Wang, P. Li, G. Zhou, Y. Yuan,
Inhibition of NF- κB results in anti-glioma activity and reduces temozolomide-induced chemoresistance by down-regulating MGMT gene expression, Cancer Lett. 428 (2018) 77–89.
[173]H. Huang, H. Lin, X. Zhang, J. Li,
Resveratrol reverses temozolomide resistance by downregulation of MGMT in T98G glioblastoma cells by the NF-κB-dependent pathway, Oncol. Rep. 27 (2012) 2050–2056.
[174]G. Guo, Y. Sun, R. Hong, J. Xiong, Y. Lu, Y. Liu, J. Lu, Z. Zhang, C. Guo, Y. Nan, Q. Huang, IKBKE enhances TMZ-chemoresistance through upregulation of MGMT expression in glioblastoma, Clin. Translat. Oncol. 22 (2020) 1252–1262.
[175]C.K. Tsai, L.C. Huang, Y.P. Wu, I.Y. Kan, D.Y. Hueng, SNAP reverses temozolomide resistance in human glioblastoma multiforme cells through down- regulation of MGMT, FASEB J. 33 (2019) 14171–14184.
[176]B. Yang, Y.B. Ma, S.H. Chu, Silencing SATB1 overcomes temozolomide resistance by downregulating MGMT expression and upregulating SLC22A18 expression in human glioblastoma cells, Cancer Gene Ther. 25 (2018) 309–316.
[177]M.Z. Mohammed, V.N. Vyjayanti, C.A. Laughton, L.V. Dekker, P.M. Fischer, D. M. Wilson 3rd, R. Abbotts, S. Shah, P.M. Patel, I.D. Hickson, S. Madhusudan, Development and evaluation of human AP endonuclease inhibitors in melanoma and glioma cell lines, Br. J. Cancer 104 (2011) 653–663.

[178]C.L. Cheng, S.P. Johnson, S.T. Keir, J.A. Quinn, F. Ali-Osman, C. Szabo, H. Li, A. L. Salzman, M.E. Dolan, P. Modrich, D.D. Bigner, H.S. Friedman, Poly(ADP- ribose) polymerase-1 inhibition reverses temozolomide resistance in a DNA mismatch repair-deficient malignant glioma xenograft, Mol. Cancer Ther. 4 (2005) 1364–1368.
[179]F. Higuchi, H. Nagashima, J. Ning, M.V.A. Koerner, H. Wakimoto, D.P. Cahill, Restoration of temozolomide sensitivity by PARP inhibitors in mismatch repair deficient glioblastoma is independent of base excision repair, Clin. Cancer Res. 26 (2020) 1690–1699.
[180]J.R. Silber, M.S. Bobola, A. Blank, K.D. Schoeler, P.D. Haroldson, M.B. Huynh, D. D. Kolstoe, The apurinic/apyrimidinic endonuclease activity of Ape1/Ref-1 contributes to human glioma cell resistance to alkylating agents and is elevated by oxidative stress, Clin. Cancer Res. 8 (2002) 3008–3018.
[181]T. Str¨obel, S. Madlener, S. Tuna, S. Vose, T. Lagerweij, T. Wurdinger,
K. Vierlinger, A. W¨ohrer, B.D. Price, B. Demple, O. Saydam, N. Saydam, Ape1 guides DNA repair pathway choice that is associated with drug tolerance in glioblastoma, Sci. Rep. 7 (2017) 9674.
[182]A.L. Hudson, N.R. Parker, P. Khong, J.F. Parkinson, T. Dwight, R.J. Ikin, Y. Zhu, J. Chen, H.R. Wheeler, V.M. Howell, Glioblastoma recurrence correlates with increased APE1 and polarization toward an immuno-suppressive microenvironment, Front. Oncol. 8 (2018) 314.
[183]S. Kaur, Z.M. Ramdzan, M.C. Guiot, L. Li, L. Leduy, D. Ramotar, S. Sabri, B. Abdulkarim, A. Nepveu, CUX1 stimulates APE1 enzymatic activity and increases the resistance of glioblastoma cells to the mono-alkylating agent temozolomide, Neuro-Oncology 20 (2018) 484–493.
[184]M.D. Naidu, J.M. Mason, R.V. Pica, H. Fung, L.A. Pe˜na, Radiation resistance in glioma cells determined by DNA damage repair activity of Ape1/Ref-1, J. Radiat. Res. 51 (2010) 393–404.
[185]D. Ghosh, S. Nandi, S. Bhattacharjee, Combination therapy to checkmate glioblastoma: clinical challenges and advances, Clin. Translat. Med. 7 (2018) 33.
[186]R. Bayat Mokhtari, T.S. Homayouni, N. Baluch, E. Morgatskaya, S. Kumar, B. Das, H. Yeger, Combination therapy in combating cancer, Oncotarget 8 (2017) 38022–38043.
[187]J. Li, Y. Wang, S. Xue, J. Sun, W. Zhang, P. Hu, L. Ji, Z. Mao, Effective combination treatment of lung cancer cells by single vehicular delivery of siRNA and different anticancer drugs, Int. J. Nanomed. 11 (2016) 4609–4624.
[188]6204676 T.C. Carter, R. Medina-Flores, B.E. Lawler, Glioblastoma treatment with temozolomide and bevacizumab and overall survival in a rural tertiary healthcare practice, Biomed. Res. Int. 2018 (2018).
[189]M.C. Chamberlain, Temozolomide: therapeutic limitations in the treatment of adult high-grade gliomas, Expert. Rev. Neurother. 10 (2010) 1537–1544.
[190]G.J. Kitange, B.L. Carlson, M.A. Schroeder, P.T. Grogan, J.D. Lamont, P.A. Decker, W. Wu, C.D. James, J.N. Sarkaria, Induction of MGMT expression is associated with temozolomide resistance in glioblastoma xenografts, Neuro-Oncology 11 (2009) 281–291.
[191]Z. Yang, D. Wei, X. Dai, M.F.G. Stevens, T.D. Bradshaw, Y. Luo, J. Zhang, C8- substituted imidazotetrazine analogs overcome temozolomide resistance by inducing DNA adducts and DNA damage, Front. Oncol. 9 (2019) 485.
[192]Y.P. Ramirez, A.C. Mladek, R.M. Phillips, M. Gynther, J. Rautio, A.H. Ross, R. T. Wheelhouse, J.N. Sakaria, Evaluation of novel imidazotetrazine analogues
designed to overcome temozolomide resistance and glioblastoma regrowth, Mol. Cancer Ther. 14 (2015) 111–119.
[193]S.T. Wong, X.Q. Zhang, J.T. Zhuang, H.L. Chan, C.H. Li, G.K. Leung, MicroRNA-21 inhibition enhances in vitro chemosensitivity of temozolomide-resistant glioblastoma cells, Anticancer Res. 32 (2012) 2835–2841.
[194]J.L. Munoz, V. Rodriguez-Cruz, S.H. Ramkissoon, K.L. Ligon, S.J. Greco,
P. Rameshwar, Temozolomide resistance in glioblastoma occurs by miRNA-9- targeted PTCH1, independent of sonic hedgehog level, Oncotarget 6 (2015) 1190–1201.
[195]P. Li, X. Lu, Y. Wang, L. Sun, C. Qian, W. Yan, N. Liu, Y. You, Z. Fu, MiR-181b suppresses proliferation of and reduces chemoresistance to temozolomide in U87 glioma stem cells, J. Biomed. Res. 24 (2010) 436–443.
[196]L. Shi, S. Zhang, K. Feng, F. Wu, Y. Wan, Z. Wang, J. Zhang, Y. Wang, W. Yan, Z. Fu, Y. You, MicroRNA-125b-2 confers human glioblastoma stem cells resistance to temozolomide through the mitochondrial pathway of apoptosis, Int. J. Oncol. 40 (2012) 119–129.
[197]Y.P. Yang, Y. Chien, G.Y. Chiou, J.Y. Cherng, M.L. Wang, W.L. Lo, Y.L. Chang, P. I. Huang, Y.W. Chen, Y.H. Shih, M.T. Chen, S.H. Chiou, Inhibition of cancer stem cell-like properties and reduced chemoradioresistance of glioblastoma using microRNA145 with cationic polyurethane-short branch PEI, Biomaterials 33 (2012) 1462–1476.
[198]S. Asuthkar, K.K. Velpula, C. Chetty, B. Gorantla, J.S. Rao, Epigenetic regulation of miRNA-211 by MMP-9 governs glioma cell apoptosis, chemosensitivity and radiosensitivity, Oncotarget 3 (2012) 1439–1454.
[199]S. Comincini, G. Allavena, S. Palumbo, M. Morini, F. Durando, F. Angeletti, L. Pirtoli, C. Miracco, microRNA-17 regulates the expression of ATG7 and
modulates the autophagy process, improving the sensitivity to temozolomide and low-dose ionizing radiation treatments in human glioblastoma cells, Cancer Biol. Therapy 14 (2013) 574–586.
[200]J.L. Munoz, S.A. Bliss, S.J. Greco, S.H. Ramkissoon, K.L. Ligon, P. Rameshwar, Delivery of functional anti-miR-9 by mesenchymal stem cell-derived exosomes to glioblastoma multiforme cells conferred chemosensitivity, Mol. Therapy Nucleic Acids 2 (2013), e126.
[201]H. Tang, Y. Bian, C. Tu, Z. Wang, Z. Yu, Q. Liu, G. Xu, M. Wu, G. Li, The miR-183/
96/182 cluster regulates oxidative apoptosis and sensitizes cells to chemotherapy in gliomas, Curr. Cancer Drug Targets 13 (2013) 221–231.

[202]H. Li, L. Chen, J.J. Li, Q. Zhou, A. Huang, W.W. Liu, K. Wang, L. Gao, S.T. Qi, Y. T. Lu, miR-519a enhances chemosensitivity and promotes autophagy in glioblastoma by targeting STAT3/Bcl2 signaling pathway, J. Hematol. Oncol. 11 (2018) 70.
[203]W. Huang, Z. Zhong, C. Luo, Y. Xiao, L. Li, X. Zhang, L. Yang, K. Xiao, Y. Ning, L. Chen, Q. Liu, X. Hu, J. Zhang, X. Ding, S. Xiang, The miR-26a/AP-2a/Nanog signaling axis mediates stem cell self-renewal and temozolomide resistance in glioma, Theranostics 9 (2019) 5497–5516.
[204]I.M. Forte, P. Indovina, C.A. Iannuzzi, D. Cirillo, D. Di Marzo, D. Barone,
F. Capone, F. Pentimalli, A. Giordano, Targeted therapy based on p53 reactivation reduces both glioblastoma cell growth and resistance to temozolomide, Int. J. Oncol. 54 (2019) 2189–2199.
[205]T. Kanzawa, I.M. Germano, Y. Kondo, H. Ito, S. Kyo, S. Kondo, Inhibition of telomerase activity in malignant glioma cells correlates with their sensitivity to temozolomide, Br. J. Cancer 89 (2003) 922–929.
[206]R. Vengoji, M.A. Macha, R.K. Nimmakayala, S. Rachagani, J.A. Siddiqui,
K. Mallya, S. Gorantla, M. Jain, M.P. Ponnusamy, S.K. Batra, N. Shonka, Afatinib and temozolomide combination inhibits tumorigenesis by targeting EGFRvIII- cMet signaling in glioblastoma cells, J. Experiment. Clinical Cancer Res. 38 (2019) 266.
[207]B. Han, X. Meng, P. Wu, Z. Li, S. Li, Y. Zhang, C. Zha, Q. Ye, C. Jiang, J. Cai, T. Jiang, ATRX/EZH2 complex epigenetically regulates FADD/PARP1 axis, contributing to TMZ resistance in glioma, Theranostics 10 (2020) 3351–3365.
[208]S. Valtorta, A. Lo Dico, I. Raccagni, D. Gaglio, S. Belloli, L.S. Politi, C. Martelli, C. Diceglie, M. Bonanomi, G. Ercoli, V. Vaira, L. Ottobrini, R.M. Moresco, Metformin and temozolomide, a synergic option to overcome resistance in glioblastoma multiforme models, Oncotarget 8 (2017) 113090–113104.
[209]D.Q. Chong, X.Y. Toh, I.A. Ho, K.C. Sia, J.P. Newman, Y. Yulyana, W.H. Ng, S. H. Lai, M.M. Ho, N. Dinesh, C.K. Tham, P.Y. Lam, Combined treatment of nimotuzumab and rapamycin is effective against temozolomide-resistant human gliomas regardless of the EGFR mutation status, BMC Cancer 15 (2015) 255.
[210]A. Brunetti, O. Marinelli, M.B. Morelli, R. Iannarelli, C. Amantini, D. Russotti,
G. Santoni, F. Maggi, M. Nabissi, Isofuranodiene synergizes with temozolomide in inducing glioma cells death, Phytomedicine 52 (2019) 51–59.
[211]I.C. Salaroglio, P. Mujumdar, L. Annovazzi, J. Kopecka, M. Mellai, D. Schiffer, S. A. Poulsen, C. Riganti, Carbonic anhydrase XII inhibitors overcome P- glycoprotein-mediated resistance to temozolomide in glioblastoma, Mol. Cancer Ther. 17 (2018) 2598–2609.
[212]S.W. Lee, H.K. Kim, N.H. Lee, H.Y. Yi, H.S. Kim, S.H. Hong, Y.K. Hong, Y.A. Joe, The synergistic effect of combination temozolomide and chloroquine treatment is dependent on autophagy formation and p53 status in glioma cells, Cancer Lett. 360 (2015) 195–204.
[213]E.B. Golden, H.Y. Cho, A. Jahanian, F.M. Hofman, S.G. Louie, A.H. Sch¨onthal, T. C. Chen, Chloroquine enhances temozolomide cytotoxicity in malignant gliomas by blocking autophagy, Neurosurg. Focus. 37 (2014) E12.
[214]Y.S. Hori, R. Hosoda, Y. Akiyama, R. Sebori, M. Wanibuchi, T. Mikami, T. Sugino, K. Suzuki, M. Maruyama, M. Tsukamoto, N. Mikuni, Y. Horio, A. Kuno, Chloroquine potentiates temozolomide cytotoxicity by inhibiting mitochondrial autophagy in glioma cells, J. Neuro-Oncol. 122 (2015) 11–20.
[215]2710693 Z. Zhu, P. Zhang, N. Li, K.M.Y. Kiang, S.Y. Cheng, V.K. Wong, G. K. Leung, Lovastatin enhances cytotoxicity of temozolomide via impairing autophagic flux in glioblastoma cells, Biomed. Res. Int. 2019 (2019).
[216]Y. Bi, H. Li, D. Yi, Y. Sun, Y. Bai, S. Zhong, Y. Song, G. Zhao, Y. Chen, Cordycepin augments the chemosensitivity of human glioma cells to temozolomide by activating AMPK and inhibiting the AKT signaling pathway, Mol. Pharm. 15 (2018) 4912–4925.
[217]Y. Wu, L. Dong, S. Bao, M. Wang, Y. Yun, R. Zhu, FK228 augmented temozolomide sensitivity in human glioma cells by blocking PI3K/AKT/mTOR signal pathways, Biomed. Pharmacotherapy 84 (2016) 462–469.
[218]S. Torres, M. Lorente, F. Rodríguez-Forn´es, S. Hern´andez-Tiedra, M. Salazar, E. García-Taboada, J. Barcia, M. Guzm´an, G. Velasco, A combined preclinical
therapy of cannabinoids and temozolomide against glioma, Mol. Cancer Ther. 10 (2011) 90–103.
[219]J. Balça-Silva, D. Matias, A. do Carmo, H. Gir˜ao, V. Moura-Neto, A.B. Sarmento- Ribeiro, M.C. Lopes, Tamoxifen in combination with temozolomide induce a synergistic inhibition of PKC-pan in GBM cell lines, Biochim Biophys Acta 1850 (2015) 722–732.
[220]B.M. Scicchitano, S. Sorrentino, G. Proietti, G. Lama, G. Dobrowolny, A. Catizone, E. Binda, L.M. Larocca, G. Sica, Levetiracetam enhances the temozolomide effect on glioblastoma stem cell proliferation and apoptosis, Cancer Cell Int. 18 (2018) 136.
[221]G.A. Alexiou, E. Vartholomatos, I.T.K.E. Peponi, G. Markopoulos, A.P.V.I. Tasiou, V. Ragos, P. Tsekeris, A.P. Kyritsis, V. Galani, Combination treatment for glioblastoma with temozolomide, DFMO and radiation, J. B.U.ON. 24 (2019) 397–404.
[222]V. Gupta, Y.S. Su, W. Wang, A. Kardosh, L.F. Liebes, F.M. Hofman, A.
H. Sch¨onthal, T.C. Chen, Enhancement of glioblastoma cell killing by combination treatment with temozolomide and tamoxifen or hypericin, Neurosurg. Focus. 20 (2006) E20.CCRG 81045
[223]I.A. Ciechomska, M.P. Marciniak, J. Jackl, B. Kaminska, Pre-treatment or post- treatment of human glioma cells with BIX01294, the inhibitor of histone methyltransferase G9a, sensitizes cells to temozolomide, Front. Pharmacol. 9 (2018) 1271.