Endoplasmic reticulum stress‑induced cell death as a potential mechanism for targeted therapy in glioblastoma (Review)

  • Authors:
    • Pengfei Shi
    • Zhuohang Zhang
    • Jie Xu
    • Li Zhang
    • Hongjuan Cui
  • View Affiliations

  • Published online on: July 6, 2021     https://doi.org/10.3892/ijo.2021.5240
  • Article Number: 60
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

The endoplasmic reticulum (ER) is an essential organelle for protein synthesis, folding and modification, lipid synthesis, and calcium storage. When endogenous or exogenous stimuli lead to ER‑synthesized protein folding dysfunction, numerous unfolded or misfolded proteins accumulate in the ER cavity and cause a series of subsequent responses, referred to as ER stress. If ER stress is continuous, the unfolded protein response (UPR) is not enough to remove the accumulated unfolded and misfolded proteins, and thus, UPR signaling pathways will drive cell apoptosis. Glioblastoma (GBM) is currently the most aggressive and common malignant tumor of the nervous system. Since ER stress may increase the sensitivity of GBM to temozolomide, this article reviews the possible mechanisms of ER stress‑induced apoptosis and the factors affecting ER stress, and evaluates the potential of ER stress as a therapeutic target.

1. Introduction

Glioblastoma (GBM) is a common primary malignant brain tumor in the cranial cavity accounting for 45.2% of malignant primary brain and central nervous system tumors (1). Tumors formed by brain glial cells, including astrocytes, oligodendrocytes and ependymal cells, can be referred to as GBM (2). In 2018, the International and European Society for Pediatric Oncology found that GBM grows fast, and 70-80% of patients have a course of 3-6 months, and only 10% of patients have a course of >1 year based on clinical patient data from 10 countries (3,4). Those with a longer course may evolve from low-malignant astrocytomas (5). Due to the rapid growth of GBM, cerebral edema and intracranial pressure are markedly increased, and all patients have symptoms, including headache and vomiting (6). Although excision, radiation and chemotherapy are standard treatments, the prognosis remains poor for patients with GBM (7,8).

The main function of the endoplasmic reticulum (ER) is to synthesize proteins and lipids. In addition to meeting its own needs, the lipids are also provided to the Golgi apparatus, lysosomes, plasma membranes, mitochondria and other membranous cell structures (9). The accumulation of unfolded or misfolded proteins in the ER will cause ER stress, and in turn triggers the unfolded protein response (UPR) to ensure that the protein is folded correctly (10). ER stress can induce the expression of glucose regulatory proteins [78-kDa glucose-regulated protein (GRP78) and GRP94] and other ER molecular chaperones to protect cell proliferation (10). However, ER stress can also independently induce endogenous cell apoptosis and ultimately affect cell fate, such as adaptation, injury or apoptosis (11). Generally, if the ER stress is continuous, the protein kinase R-like endoplasmic reticulum kinase (PERK), inositol-requiring enzyme 1 (IRE1) and activating transcription factor (ATF)6 signaling pathways will be used to transduce the downstream apoptotic pathways (12,13). Therefore, researchers are gradually turning their attention to how to induce GBM cell apoptosis by ER stress. This review discusses the factors that influence the occurrence of ER stress, and the possibility of ER stress as a treatment for GBM.

2. Role of ER stress in cell survival and death

The perception and response to exogenous stress is an important component of cell physiology. Certain studies have demonstrated that the ER can initiate the cell response to exogenous stress (14,15). ER synthetic proteins should be properly folded, glycosylated and disulfide-bonded to form functional proteins (16). Therefore, a quality control mechanism is important for detecting misfolded or unfolded proteins and performing cell functions, such as cell division and cell-to-cell interactions (16,17). Due to the proliferation of cancer cells, the probability of misfolded or unfolded proteins is higher than that of normal cells (18). Therefore, the UPR can prevent the continuous synthesis of misfolded or unfolded proteins to a certain extent, and has a protective effect on GBM cells (19). The UPR has three classic transmembrane ER-resident UPR sensors: IRE1α (20), PERK (21) and ATF6 (22). These sensors can detect misfolded and unfolded proteins, and accelerate the recovery and maintenance of ER homeostasis (17).

Furthermore, the signaling network within the cells is complex. When a signal changes, it must trigger a series of signal changes, suppression or assistance (23). Additionally, sustained ER stress can induce other signaling pathways, such as DNA damage signaling and death receptor-mediated signaling, to promote cell apoptosis (24). The PERK-elF2α-ATF4-DNA damage inducible transcript 3 (GADD153) signaling pathway is one of the classic signaling pathways of ER stress (25). Therefore, the expression levels of members of this pathway will also cause changes in other genes, which will synergistically promote cell apoptosis (25). p53 is a key player in the DNA damage response (DDR), and its expression is specifically induced by the PERK kinase during the UPR following ER stress (25). p53 activation during the DDR has been well studied (26,27). Once activated, p53 will stimulate and suppress different gene products, which aim to either prevent abnormal proliferation by a reversible arrest of the cell cycle to facilitate the repair processes, or to induce irreversible outcomes, including apoptosis or senescence (25). In addition, PERK expression decreases RNA component of mitochondrial RNA processing endoribonuclease expression, and increases microRNA-206 to inhibit Bcl-2, and consequently induces cleaved caspase3 (28). Cannabidiol (CBD) has the ability to inhibit the proliferation of GBM cells, and CBD can promote the expression of GADD153 to trigger ER stress, and induces mitochondrial dysfunction and lethal mitophagy arrest via the GADD153-tribbles pseudokinase 3-AKT-mTOR axis (29). Salinomycin and its ester derivatives 5-7 increase the levels of phosphorylated (p-)eukaryotic initiation factor (eIF)2α (Ser51) and IRE1α proteins, and also increase the levels of DNA damage indicators, such as γ-H2A histone family member X (γH2AX) protein and modified guanine (8-oxoG), by upregulating the expression levels of GADD153 (30). eIF5B has been demonstrated to serve a critical role in canonical translation, and eIF5B depletion results in the upregulation of DNA damage-inducible protein 34 (GADD34) transcription, and leads to activation of JNK to promote cell death (31). X-box binding protein 1 (XBP1) expression increases DNA damage, protein ATM phosphorylation, and the expression levels of MRE11-RAD50-NBS1 complex and γH2AX (32). In addition, IRE1 can interact with adaptor protein TNF receptor-associated factor 2 (TRAF2) and then initiates JNK, which has been demonstrated to be involved in cell death (33). IRE1 may contribute to apoptosis by activating the RIDD signaling pathway (34).

3. Apoptosis mechanism of ER stress

GRP78 is a key regulator of the UPR (35). As a Ca2+-binding molecular chaperone in the ER, GRP78 maintains ER homeostasis, suppresses stress-induced apoptosis and controls UPR signaling (35). In the case of protein folding, GRP78 will inhibit the activation of these sensors (PERK, IRE1α and ATF6) (36). When misfolded and unfolded proteins exceed critical thresholds, GRP78 separates from the three sensors, leading to the activation of three distinct but partially functionally overlapping signaling pathways (37-39) (Fig. 1).

PERK, located at the ER membrane, is a type I transmembrane Ser/Thr kinase (40). PERK has a luminal stress-sensing domain and a cytosolic kinase domain (41). PERK is activated through a homo-oligomerization process leading to PERK trans-autophosphorylation and phosphorylation of its main substrate (42). During the early phase of ER stress, PERK activity can promote survival by inducing translational arrest, upregulating chaperones and enhancing the expression of the antioxidant gene, such as quinone oxidoreductase 2 and heme oxygenase 1 (43). Activated PERK induces the phosphorylation of serine 51 of the eukaryotic translation initiation factor 2α (eIF2α) and inhibits the synthesis of misfolded proteins (40). eIF2α can activate activating transcription factor 4 (ATF4), a transcription factor controlling the expression of certain genes involved in folding, autophagy, amino acid metabolism, antioxidant response, apoptosis and DNA damage, such as GADD153 (42). Under normal physiological conditions, GADD153 is present in the cytoplasm; however, continuous ER stress can promote GADD153 activation and transfer to the nucleus (44,45). Additionally, GADD153 has been reported to result in the decrease of Bcl-2 protein and the translocation of the pro-apoptotic molecule Bax from the cytosol to the mitochondria, which in turn induces the mitochondrial apoptotic pathway (46). Cytochrome'C can cause the activation of the caspase adapter apoptotic peptidase activating factor 1 and pro-caspase-9 by forming an enzyme complex with them. This complex is referred to as 'apoptotic bodies' (46). Additionally, caspase-9 further activates caspase-3/7 (46). The inhibitor of apoptosis protein inhibits activation of caspase-3; however, the release of second mitochondria-derived activator of caspase can remove the inhibitory effect (47).

Although a few studies have demonstrated that the existence of IRE1α is beneficial to the neovascularization of GBM, IRE1α also induces cell death and serves a unique role in ER stress (48-50). GRP78 can bind the luminal domain of IRE1α (49). Under ER stress, IRE1α dissociates from GRP78 and promotes the separation of TRAF2, and also induces cleavage of XBP1 into a splicing variant of XBP1 and transcription of GADD153 (51,52). TRAF2 forms a complex with TNF receptor superfamily member 1A-associated death domain protein, transforming growth factor β-activated kinase 1 and receptor-interacting protein 1 (53). This complex can activate apoptosis signal-regulating kinase 1 (ASK1), and then activate mitogen-activated protein kinase 4/7 (54-56) and the JNK apoptosis signaling pathway (57).

ATF6 is a type II transmembrane protein. ATF6 has a cytosolic bZIP transcription factor domain (58). During ER stress, ATF6 translocates to the Golgi where it is cleaved by proteases (59). The cleaved ATF6 cytosolic fragment can then act as a transcription factor (59). Prior to that, ATF6 needs to be modified, such as by reduction and glycosylation (60). ATF6 also induces GADD153 expression (61,62).

4. Factors inducing ER stress in GBM

Abnormal gene expression and ER stress

Normal tissue cells need to undergo a series of changes before they can become cancer cells, such as genetic changes, or activation or inactivation of certain signaling pathways (8). For example, stearoyl CoA desaturase (SCD1) and proto-oncogene SEC61 translocon subunit γ can promote ER homeostasis, avoid the production of misfolded proteins and inhibit the occurrence of tumors (63). Paraoxonase 2 (PON2), a paraoxonase protein, consists of lactone hydrolases with different substrate specificities (64). When PON2 is located on the nuclear membrane and ER, it can increase the stability of the ER and protect cancer cells from adverse environmental conditions and chemotherapy (64). Hypoxia-stimulated galectin-1 (Gal1) is an effective regulator of GBM cell migration and an angiogenic molecule (65). Additionally, the reduction of Gal1 weakens the expression levels of seven genes related to chemical resistance: ORP150, BNIP3L, HERP, TRA1, GADD45B, GRP78 and CYR61 (65). The absence of cytochrome P450 17A1 can induce the occurrence of ER stress and reactive oxygen species (ROS) generation by regulating secretion-associated Ras-related GTPase 1 (66). Furthermore, heat-shock protein 27 and 90 can decrease the levels of cytochrome C, caspase3, caspase9 and caspase12 in GBM cells (67,68). Additionally, cyclophilin B is a prolyl isomerase residing in the ER, and its absence can damage the ER structure (69).

In addition, knockdown of cAMP-responsive element-binding protein 3 induces cell apoptosis by increasing p-PERK, p-eIF2α, ATF4, Bax and caspase3 (70). Reversion inducing cysteine rich protein with kazal motifs (RECK) is a key suppressor gene in regulating cancer cell invasion and metastasis (71). Highly expressive RECK can modulate ER stress by binding to or sequestering GRP78 to activate p-eIF2α (71). Tumor necrosis factor receptor-associated protein 1 can markedly induce the occurrence of ER stress by activating ATF4 (72). In addition, neural precursor cells can migrate to advanced astrocytoma via the release of the vanillin receptor [transient receptor potential vanillin subfamily member 1 (TRPV1)] (73). TRPV1 induces GBM cell death via ER stress (73). Therefore, drugs that target these proteins located on the ER membrane or stabilizing the structure of the ER can be developed.

ROS and ER stress

The redox environment of the ER determines the fate of proteins entering the ER, and the level of redox signaling mediators also regulates the level of ROS (74). ROS can induce the occurrence of ER stress through redox signaling mediators, such as NADPH-P450 reductase, protein disulfide isomerase-endoplasmic reticulum oxidoreductase 1, NADPH oxidase 4, glutathione/glutathione disulfide and calcium (74,75).

In addition, ROS-mediated hypoxia-inducible factor 1α serves an important role in promoting tumor microenvironment, anti-apoptosis and drug resistance in GBM (76). Mitochondrial PTEN-induced kinase 1 (PINK1), a regulator of the Warburg effect, is a negative regulator of proliferation in GBM cells (77). PINK1 can inhibit ROS generation and cell proliferation through FOXO3a (77). This finding highlights the importance of the balance between PINK1 and ROS in normal cells and cancer cells (77). Additionally, luteolin, a common dietary flavonoid, can induce a lethal ER stress pathway and mitochondrial dysfunction by increasing the intracellular ROS levels (78). Dihydroartemisinin (DHA) induces cell apoptosis through mitochondrial membrane depolarization, cytochrome'C and caspase 9 (79). Furthermore, the cytotoxicity of DHA can increase the expression levels of GRP78, GADD153, eIF2α and caspase12 (79).

Hypoxia and ER stress

One of the main obstacles for tumor progression is that cancer cells grow in a low-oxygen environment (80). Hypoxia can stimulate the adaptive response to promote cell proliferation and survival, as well as angiogenesis (81). However, some researchers have illustrated that hypoxia can also cause cell apoptosis or necrosis (81,82). Consequently, under hypoxic conditions, understanding the decision-making process that regulates cell death, adaptation and resistance for treatment is crucial. Hypoxia has been demonstrated to induce ER stress (83). Additionally, hypoxia can induce the cell surface exposure of calreticulin, a hallmark of immunogenic cell death (84,85).

Studies have demonstrated that after knock out of the gene encoding endothelin-1, the expression levels of endothelin receptor type B, endothelin 1, endothelin converting enzyme 1 and endothelin receptor type A are upregulated, which will increase the sensitivity of GBM cells to hypoxia-induced ER stress (86,87). Furthermore, under hypoxic conditions, the expression levels of snail family transcriptional repressor 2 and mesoderm specific transcript are upregulated, which will amplify the inhibitory effect of IRE1 on various genes (88). Hypoxia also leads to upregulation of IGFBP6, IGFBP7, IGFBP10/CYR61, WISP1 and WISP2, and downregulation of IGFBP9/NOV at the mRNA level (89). IRE1 markedly downregulates IGFBP7, IGFBP10/CYR61, WISP1 and WISP2, and upregulates IGFBP9/NOV, which shows that ER stress is an essential part of malignant GBM cell proliferation (90).

Low glucose and ER stress

The glucose metabolism allows energy to be oxidized by its carbon bonds and then used in the form of ATP. The final product of glucose can be lactate or carbon dioxide (91). In the 1920s, Otto Warburg and his colleagues found that tumors were absorbing a lot of glucose compared with surrounding tissues (92). Furthermore, in the presence of oxygen, glucose can also be fermented to produce lactic acid through aerobic glycolysis (92). Subsequently, in 1929, the British biochemist Herbert Crabtree confirmed Warburg's findings and further revealed that the respiratory intensity of tumors was variable, and numerous tumors exhibited this phenomenon (93). Additionally, Racker developed his theory of the origin of the Warburg effect in terms of intracellular pH imbalance and ATPase activity defects (94). Research on genetics and pharmacology demonstrated that the Warburg effect is necessary for cancer cell proliferation (95). Studies have demonstrated that the direct and indirect result of cancer-causing mutations is the reprogramming of cancer cell metabolism (96,97). Obtaining the necessary nutrients from the nutrient-deficient environment is a common feature of tumor metabolism (98). Cancer cells can use these nutrients to maintain viability and build new biomass (98). Therefore, tumors are considered to be a metabolic disease (99).

Low glucose means that cancer cells collect less energy. GRP78 can be upregulated in a low-glycemic state, and enables GBM cells to survive by inducing autophagy (100). Additionally, in the low-glycemic stage, p-PERK and cleavage of ATF6 are upregulated, which indicates that low glucose can lead to ER stress, activation of caspase, cell dysfunction, cell arrest and cell death (101-103). Metformin is the first-line drug for type 2 diabetes, and it can induce cancer cell death via the ASK1/phorbol-12-myristate-13-acetate-induced protein 1 and ROS/ASK1/JNK signaling pathways (104). The aforementioned can also indicate that the ketogenic diet can partially inhibit cancer cell proliferation.

pH and ER stress

In tumor tissues, due to increased glycolytic activity, slow blood circulation and insufficient blood supply of cancer cells, the pH of certain tumors (astrocytoma and squamous cell carcinoma) is <6.0, while the normal physiological pH is 7.3 (105). The low outflow of potentially toxic metabolic waste and a low influx of metabolites further enhance acidic conditions where cancer cells are located (106). The metastasis of solid tumors to acidic sites can promote tumor growth, invasion, angiogenesis, immunosuppression and chemotherapy resistance (107). However, there are conflicting reports on the effects of the acidic environment on cancer cell physiology. For example, low pH can promote tumor development; however, certain studies have demonstrated that low pH can also induce cancer cell apoptosis (108,109). For example, acidic stress may lead to upregulation of Src activity, VEGF and MMPs, which is conducive to cancer cell survival and metastasis (110,111). Acid-mediated apoptosis is considered to be the result of caspase activation (109). The acidic environment can regulate GBM cell proliferation and radiosensitivity (112). Additionally, acidity induces the expression of eIF2α, IRE1α, ATF6, GADD153 and caspase12 (113,114). This indicates that, in the brain tissue, the acidic environment may kill cancer cells via the ER stress pathway.

Calcium activator or inhibitor and ER stress

Certain drugs that affect ER calcium balance, such as the thapsigargin (115), calcium ionophore A23187 (116) and calcium ion chelator EGTA (117), can lead to ER stress. Calcitonin C (Cal-C) is a photoactivated inhibitor (118). Cal-C binds to protein kinase C and other protein regulatory domains with diacylglycerol/phorbol ester binding sites (118). It may damage calcium imbalance in GBM cells, and then result in ER stress (119). 2,9-diazaspiro[5.5]undecane depletes intracellular Ca2+ stores (120). Cyclophilin/calcineurin inhibitor Cyclosporine A has certain apoptotic characteristics, causes numerous cytoplasmic vacuoles, and increases immunizing ER stress and autophagy markers, such as PERK, IRE1α, GRP78, GADD153 and LC3-II (118). Salinomycin is a polyether ionophore antibiotic and induces cell apoptosis through ER stress and autophagy (121). Nonsteroidal anti-inflammatory drugs can release additional Ca2+ to induce ER stress, thereby preventing cell transformation and slowing proliferation (122). Monensin can destroy calcium homeostasis and overcome TNF-related apoptosis-inducing ligand (TRAIL) resistance in GBM cells via ER stress, and thus it is currently considered an anticancer drug (123). Additionally, other polyether antibiotics, such as narasin, salinomycin, lasalocid A and nigericin, can also overcome TRAIL resistance in GBM cells via ER stress (123).

Lipid stress and ER stress

ER stress can give rise to changes in lipid metabolism; however, certain evidence suggests that dysfunctional lipid metabolism may activate UPR, regardless of whether there is a misfolded protein in the ER lumen (124). Subsequently, after UPR, genes involved in lipid metabolism will be upregulated (125). The brain is generally considered to be one of the most fat-rich organs, and the lipid content itself can affect various clinically relevant behavioral indicators (126). These bioactive lipids include steroids, diacyl glycerol, sphingolipids, phosphatidylinositol phosphate, phosphatidylcholine and polyunsaturated fatty acids (126). Furthermore, saturated fatty acids have the effect of promoting ER stress, while unsaturated fatty acids counteract this effect (126). Saturated fatty acids, such as palmitic acid and stearic acid, are known inducers of ER stress in various cell types, such as liver and breast cancer cells, and can regulate cell survival and apoptosis signals (126). SCD1, a downstream gene of sterol regulatory element-binding protein (SREBP), mediates lipid desaturation, which has also been found to be a critical determinant of cancer cell survival (127). SREBPs have important roles in regulating lipid metabolism and mediate lipid synthesis in GBM cells (127). Loss of SREBP and lipid synthesis can block GBM cell proliferation in xenograft models (128). In addition, SREBP ablation is also accompanied by the activation of IRE1α and PERK (129). These findings indicate that proliferating cells need to establish a balance between their proliferation rate and unsaturated lipid supply to prevent ER stress. Normal cells can regulate their proliferation rate in response to nutrient availability and retain a pool of unsaturated lipid, which allows cells to maintain homeostasis and avoid ER stress (129). However, with the rapid proliferation of cancer cells, if the exogenous unsaturated lipids are limited, the cancer cells will experience ER stress, eventually leading to cell death (129).

5. Potential targeted therapy for GBM via ER stress pathway

Due to the rapid proliferation of cancer cells, cancer cells need to synthesize a large amount of protein to support their own needs, resulting in misfolded proteins occurring in cancer cells. According to the ClinicalTrials.gov database (https://www.clinicaltrials.gov/), some studies have been carried out on the effect of ER stress on tumor treatment. Among them, a project investigating TN-TC11G (9-tetrahydrocannabinol + CBD) in combination with temozolomide (TMZ) and radiotherapy in patients with newly-diagnosed GBM is recruiting patients.

TMZ is the first-line drug for clinical glioma chemotherapy. It is an oral alkylated chemotherapy drug and effectively crosses the blood-brain barrier. However, over time, some GBM can gradually resist TMZ-induced damage. This resistance may be associated with the DNA repair pathway (O6-methylguanine DNA methyltransferase, DNA mismatch repair, base excision repair system), EGFR, MDM2 proto-oncogene, p53 mutation and PTEN (130). Therefore, researchers pay increasing attention to natural compounds, small molecules, viruses, bacteria, and calcium activators or inhibitors, and conduct basic research, aiming to one day treat glioma in clinical settings.

Natural compounds

GRP78 predominantly resides in the ER lumen within normal cells, and most of the research on GRP78 has focused on cytosolic or total GRP78 (131,132). However, in tumor microenvironments where GRP78 expression is upregulated, GRP78 also localizes to the surface of GBM cell membranes (133). GRP78 may influence not only GBM cells, but also the surrounding microenvironmental vasculature (133). Therefore, it has been gradually revealed that some compounds, such as epigallocatechin 3-gallate (EGCG), honokiol, celecoxib and bortemozib, can inhibit the growth of glioma by inhibiting GRP78 (133). IRE1 is the main mediator of the UPR (134). When cancer cells trigger ER stress in an unfavorable environment, the IRE1 signal can be an adaptive mechanism (135). However, the Food and Drug Administration-approved compounds methotrexate, cefoperazone, folinic acid and fludarabine phosphate, as inhibitors of IRE1, hinder the adaptation mechanism (135). In addition, flavokawain B, a natural kava chalcone, exhibits potent anti-tumor activity in various cancer types, such as lung cancer cells (136) and gastric cancer cells (134). It induces protective autophagy by targeting the ATF4-GADD153-AKT-mTOR signaling pathway (137). Piperlongumine preferentially kills high-grade glioma (HGG) cells but has little effect on normal brain cells (138). It induces ROS generation and disrupts protein folding in the ER by increasing the oxidative deactivation of peroxide reduction 4 to activate the ER stress pathway in HGG cells (138). Therefore, piperlongumine can be regarded as an effective drug for the treatment of GBM.

There are still numerous compounds that can induce cell apoptosis via the ER stress pathway in GBM cells. For example, Isochaihulactone, a natural compound extracted from the Chinese traditional herb Nan-Chai-Hu, can disrupt ER homeostasis in GBM cells (139). The novel resorcinol derivatives [2,4-bis (4-fluorophenylacetyl) resorcinol (BFP)] can increase some characteristic ER stress markers, such as GRP78, IRE1, eIF2α and GADD153, in human GBM cell lines (U251 and U87), and a mouse GBM cell line (C6 cells) (140). In addition, treatment with BFP can increase ROS generation and downstream caspase activation, such as caspase12, caspase9 and caspase7 (140). Cannabinoids can inhibit the epithelial-mesenchymal transition of several tumors in rats and mice, and enhance tumor immune surveillance (141). Therefore, cannabinoids may be considered as potential anticancer drugs. In 2003, cannabinoids were used to explore the anticancer mechanism (142). The results demonstrated that the main mediator of cannabinoid is the stress-regulating protein p8 (also designated as a candidate for metastasis 1) (142). Further research revealed that p8 has an apoptotic effect by upregulating ER stress-related genes ATF4 and GADD153 (142). Additionally, Shikonin (one of the main active ingredients of Chinese herbal medicine Lithospermum erythro-rhizon) (143), fatsioside A (a novel baccharane-type triterpene glycoside) (144), garlic compounds (diallyl sulfide and diallyl disulfide) (145), apigenin, (-)-epigallocatechin, and genistein (146,147), desipramine (a tricyclic antidepressant) (148), curcumin (149,150), xanthatin (a natural sesquiterpene lactone purified from Xanthium strumarium L.) (151), honokiol (a cell-wall component of M. grandiflora) (152), sinomenine hydrochloride (the main biologically active alkaloid isolated from Leymus chinensis) (152), radicol (a novel trinorguaiane type sesquiterpene) (153), phenyl isothiocyanate (a member of the isothiocyanate family) (154,155), redox organoruthenium compound 11 (RDC11; one of the most active compounds among the novel ruthenium-derived compounds) (156) and obtusaquinone [OBT; a natural compound from the heartwood of Dalbergia retusa (cocobolo)] (157) have also been reported to induce ER stress (Table I). Among them, after injecting GBM cells into the mouse cranial cavity, RDC11 and OBT can reduce tumor progression, and improve the survival rate of the mouse (156,157). Therefore, whether other natural compounds can pass through the blood-brain barrier at the individual level requires in-depth research. In addition, there are also numerous reports on other tumors, which suggest that natural compounds can induce cell apoptosis through ER stress. For example, aspirin can induce multiple myeloma (MM) cell apoptosis by inhibiting Blimp1, activating the ATF4/CHOP apoptotic pathway (158). Valosin containing protein (p97/VCP) is an ER-associated protein, and novel p97/VCP inhibitor induces ER stress and apoptosis in both bortezomib-sensitive and -resistant MM cells (159). Furthermore, there are numerous other compounds that can also affect cancer cell proliferation through ER stress in other tumors, such as 18βH (a semisynthetic derivative of -glycyrrhetinic acid) in breast cancer cells (MCF-7 and MDA-MBA-231) (160), resveratrol (a natural polyphenol compound) (161) and 2-pyrazine-PPD (a novel dammarane derivative) (162) in gastric cancer, sothiocyanates (natural compounds abundant in cruciferous vegetables) in non-small cell lung cancer cells (163), and honokiol (a hydroxylated biphenyl natural product) in prostate cancer, melanoma, lung cancer, leukemia and colorectal cancer (164). Therefore, whether these natural compounds can induce ER stress and be used for the treatment of GBM in vivo still requires basic verification and clinical trials.

Table I

Potential compounds for the treatment of glioblastoma.

Table I

Potential compounds for the treatment of glioblastoma.

Author, yearCompoundsImpact on UPR membersEfficacy for glioblastoma(Refs.)
Luet al, 2012BFPGRP78, GRP94, IRE1, eIF-2α, GADD153ROS generation; pro-apoptosis(140)
Guzmanet al, 2003; Carracedoet al, 2006CannabinoidsGRP78, ATF4, GADD153Pro-apoptosis; decreases the mitochondrial membrane potential(141,142)
Panet al, 2015Fatsioside APERK, eIF-2α, GADD153Induces apoptosis and death(144)
Daset al, 2007BFPCalpainROS generation; pro-apoptosis; increases in intracellular free [Ca2+]; release of cytochrome C(145)
Daset al, 2010ApigeninPARPReduces cell viability; pro-apoptosis(146)
Daset al, 2010; Djeriret al, 2018EGCG; GenisteinGRP78ROS generation; increase in intracellular free [Ca2+]; induces apoptosis; release of cytochrome C(146,147)
Maet al, 2011DMIGADD153, GADD34Pro-apoptosis(148)
Garrido-Armaset al, 2018; Sansaloneet al, 2019CurcuminIRE1, ATF6Induces apoptosis; decreases mitochondrial membrane potential(149,150)
Maet al, 2019XanthatinGRP78, PERK, IRE1, eIF-2α, GADD153, ATF4, ATF6, XBP1sPro-apoptosis; inhibits cell proliferation(151)
Martinet al, 2013HonokiolGRP78, ATF4Pro-apoptosis(152)
Liet al, 2017RADGRP78Induces apoptosis; blocks autophagy(153)
Chouet al, 2015; Chouet al, 2017PEITCGRP78, GADD153, XBP-1, IRE1α, calpain I, calpain IIInduces cell arrest and apoptosis; ROS generation; increases intracellular free [Ca2+](154,155)
Kimet al, 2014PiperlongumineeIF-2α, GADD153Increases ROS levels(138)
Menget al, 2009RDC11GRP78, XBP1, GADD153Induces DNA damage; inhibits cell proliferation(156)
Badret al, 2013ObtusaquinoneC-junInduces DNA damage; pro-apoptosis; ROS generation(157)
Choet al, 2019NEO214GRP78, GADD153Induces apoptosis; decreases glioma progression(175)
Choet al, 2017NEO212GRP78, GADD153Induces cell death; blocks autophagy(177)
Marin-Ramoset al, 2019NEO100Calpain-1RhoA activation; induces apoptosis; reduces GSC invasion(178)
McCubreyet al, 2006 2-amino-N-acetamideGRP78Induces apoptosis(176)
Linet al, 2019AirateroneIRE1αROS generation; inhibits cell proliferation(66)
Chenet al, 2019Compound-7gGRP78, eIF2α, Ire1α, GADD153Suppresses cell proliferation and viability; induces apoptosis(179)
Koncarevicet al, 2009TPCGADD45Induces cell arrest(180)
Liet al, 2017EMAP IIGRP78, GADD153Induces apoptosis; decreases GBM-induced angiogenesis(181)
Eomet al, 2010BerberineGRP78, PERK, eIF2α, GADD153ROS generation; pro-apoptosis; increases intracellular free [Ca2+](182)
Wanget al, 2017NIM811ATF4, eIF2αBlocks autophagy; promotes cell death(183)
Suzukiet al, 2013CelecoxibGRP78, GADD153Induces cell autophagy and cell arrest; delays cell proliferation; pro-apoptosis(184)
Yeet al, 2019SRT2183GRP78, PERK, IRE1α, eIF2α, GADD153Inhibits cell proliferation; induces cell arrest; pro-apoptosis(185)
Jiaet al, 201017α-AEDGRP78, eIF2α, XBP1, ATF6, GADD153Induces autophagy and apoptosis(186)
Liuet al, 2013
Shenet al, 2014
Bownet al, 2000
MinocyclineGRP78, eIF2α, GADD153Induces autophagy and apoptosis(187)
Parket al, 2019UAGRP78, GRP94, PERK, ATF6, eIF2α, GADD153, IRE1Increases intracellular free [Ca2+]; induces autophagy andapoptosis; loss of the mitochondrial membrane potential(188-190)
Kavithaet al, 2015AsAGRP78, calpain, IRE1α, calnexinInduces apoptosis; increases intracellular free [Ca2+](12)

B, Viruses, bacteria, calcium activators or inhibitors
Kimet al, 2017OP-AGADD153Induces apoptosis(191)
Qaisiyaet al, 2017UCBGADD153Induces inflammation and apoptosis(192)
Mahoneyet al, 2011RhabdovirusIRE1αPromotes cell death(193)
Abrahamet al, 2013CHIKVXBP1, eIF2αDNA fragmentation; PARP cleavage; loss of mitochondrial membrane potential(194)
Kusaczuket al, 2018SiNPsGRP78, GRP94Impaired mitochondria function; proinflammatory response(195)
Rubioloet al, 2014YessotoxinPERK, eIF2α, XBP1Induces autophagy, apoptosis and cell arrest(196)
Kimet al, 2011AmiodaroneGADD153Increases intracellular free [Ca2+](197)

[i] BFP, 2,4-bis (4-fluorophenylacetyl) resorcinol; DAS, diallyl sulfide; DADS, diallyl disulfide; EGCG, epigallocatechin 3-gallate; DMI, desipramine; SH, sinomenine hydrochloride; RAD, radicol; PEITC, phenyl isothiocyanate; RDC11, ruthenium-derived compounds 11; OSU-03012, 2-amino-N-acetamide; NEO214, rolipram-perillyl alcohol conjugate; NEO212, perillyl alcohol conjugate; NEO100, enriched perillyl alcohol manufactured under cGMP conditions; TPC, platinum thiopyridine(II) complex; EMAP II, endothelial monocyte activating polypeptide II; NIM811, N-methyl-4-isoleucine-cyclosporine; SRT2183, (R)-N-(2-(3-((3-hydroxypyrrolidin-1-yl)Methyl)iMidazo[2,1-b]thiazol-6-yl)phenyl)-2-naphthaMide; 17α-AED, neuro-steroid, 5-androstene 3β,17α diol; UA, ursolic acid; AsA, asiatic acid; OP-A, ovitriol A; UCB, unconjugated bilirubin; CHIKV, Chikungunya virus; SiNPs, silica nanoparticles; ROS, reactive oxygen species; GSC, glioblastoma stem cell; GBM, glioblastoma.

Small molecule compounds

GBM stem cells (GSCs) can be considered key drivers of tumor growth, aggressiveness and therapy resistance in GBM (8). PERK is a well-characterized switch between survival and death during persistent ER stress and mediates cell death through induction of GADD153 (165). A previous study has demonstrated that ER stress aggravation targets GSCs, and PERK directly regulates SOX2 downregulation at the protein level to induce GSC differentiation, independent from eIF2α/ATF4 signaling (165). Furthermore, ionizing radiation potentiates ER stress, which reduces proliferation in a PERK-dependent manner (166). Adding PERK inhibitor, ER stress inducer (2DG) and GADD34 phosphatase inhibitor (Sal003) to irradiated GBM cells can reduce cell viability (166). In addition, other highly selective PERK inhibitors may provide a ground-breaking, anticancer treatment strategy in a PERK-dependent manner. 7-Methyl-5-(1-2,3-dihydro-1H-indo l-5-yl)-7H-pyrrolo[2,3-d]pyrimidin-4-amine (GSK2606414) is an oral, effective and selective PERK inhibitor, which inhibited the growth of a human tumor xenograft in mice by inducing ER stress (167). Therefore, GSK2606414 treatment may be considered as an effective drug therapy for GBM. Small-molecule inhibitor 42215 of PERK can markedly induce apoptosis after treatment of cancer cells (168). ATF4 is the master regulator of the cellular stress response and the core regulator of the PERK-eIF2α signaling pathway (169). Kurarinone (Extract of S. flavescens Roots) activates ATF4 to induce cancer cell apoptosis (169). A mixture of sixteen previously selected small molecules ('active mixture'; AM16: l-arginine, l-tyrosine, l-histidine, l-tryptophan, l-methionine, l-phenylalanine, adenine, l-(-)-malic acid, 2-deoxy-d-ribose, orotic acid, d-(+)-mannose, hippuric acid, pyridoxine, d-biotin, (-)-riboflavin and l-ascorbic acid) can upregulate ATF4 and GADD153 to induce cell apoptosis (170). Furthermore, salicylaldehyde analogs (MK0186893) and umbelliferones (4 µ8c) can be used as IRE1 RNase inhibitors, and have shown promise as a potential therapeutic strategy to counteract disease pathogenesis associated with overactive IRE1 signaling (171,172). Palmitoylation inhibitors (2-bromopalmitate, cerulenin and tunicamycin) induce cell death by promoting the accumulation of XBP1 (173). In addition, >10,000 non-toxic compounds that may activate IRE1-dependent XBP1 splicing through a mechanism independent of binding the IRE1 kinase active site have been identified by a high-throughput screening approach in July 2020 (174). This undoubtedly provides more drugs for the treatment of tumors via ER stress and requires in-depth exploration and screening.

In addition to specific small molecule modulators of UPR pathways and their potential use in developing anticancer therapies, numerous small molecule compounds have been demonstrated to induce GBM cell apoptosis (Table I). For example, NEO214 (rolipram-perillyl alcohol conjugate) is produced by the covalent linkage of carbamate linkage and cyclopropanol (175). It can cross the blood-brain barrier, and induce cell death via ER stress and the death receptor 5/TRAIL/TNF superfamily member 10 signaling pathway (175). Therefore, NEO214 may be considered as a potential clinical antitumor drug. Asiatic acid (AsA) is a natural small molecule, and it is widely used to cure various neurological disorders. AsA also induces ER stress by increasing GRP78, IRE1α and calpain, to damage a cellular organization in human GBM cells (LN18, U87MG and U118MG) (176). Notably, AsA can cross the blood-brain barrier (12). Therefore, AsA may be considered as a potential clinical antitumor drug. In addition, 2-amino-N-acetamide (176), NEO212 (a combination of TMZ and perillyl alcohol) (177), NEO100 (a high-purity and high-quality polychlorohydrin) (178), Compound-7g (179), platinum thiopyridine (II) complex (180), endothelial monocyte activating polypeptide II (181), berberine (an isoquinoline quaternary alkaloid isolated from a variety of medicinal plants) (182), N-methyl-4-isoleucine-cyclosporine (a small molecule cyclophilin binding inhibitor) (183), celecoxib (a non-steroidal anti-inflammatory drug) (184), the silent mating type information regulation 2 homolog activator (R)-N-(2-(3-((3-Hydroxypyrrolidin-1-yl)methyl)imidazo[2,1-b] thiazol-6-yl)phenyl)-2-naphthamide (185), neuro-steroid, 5-androstene 3β,17α diol (186), minocycline (187) and ursolic acid (188-190) can also induce GBM cell apoptosis via the ER stress pathway. Whether these small molecules can be used in clinical trials needs to be examined in animal models to ensure that they can cross the blood-brain barrier and reduce damage to normal cells.

Viruses, bacteria, and calcium activators or inhibitors

In addition to compounds, some viruses, bacteria, and calcium activators or inhibitors have been reported to induce GBM cell apoptosis via ER stress (Table I). For instance, ovitriol A (a fungal sesterterpene from Bipolaris oryzae) can induce paralysis-like cell death in human glioma cell lines (T98G, U251MG, U343, U373MG and A172), accompanied by the expansion of the ER (191). Unconjugated bilirubin may cause bilirubin neurotoxicity (192). It can also cause cell death t by increasing GADD153 in U87MG cells (192). Rhabdovirus is an important regulator of rhabdovirus-mediated cytotoxicity and mediates the ER stress response pathway to induce cell apoptosis (193). Chikungunya virus (CHIKV), an old-world alphavirus, can induce DNA fragmentation, loss of mitochondrial membrane potential, poly(ADP-ribose) polymerase (PARP) cleavage, nuclear enrichment and visible cytopathic effects in a dose- and time-dependent manner (194).

The focus of tumor research has always been the optimization of treatment strategies for malignant tumors. Silica nanoparticles (SiNPs) are such a strategy, which is rapidly developing into a promising tool for cancer diagnosis, imaging and treatment (195). SiNPs lead to impaired mitochondrial function, ROS generation and cell death by elevating levels of ER stress genes, including GRP94, GRP78, GADD153 and cyclooxygenase-2 (COX2) (195). Yessotoxin (YTX) is a polycyclic ether compound produced by dinoflagellate and accumulates in filter-fed shellfish (196). YTX can upregulate p-PERK, p-eIF2α, s-XBP1 and GADD153 in human glioma cell lines (SF539, SF295 and SNB75) (196). Additionally, YTX induces cell cycle arrest and increases cholesterol and polar lipid content in glioma cells (196). In addition, amiodarone is a widely used antiarrhythmic drug (197). Amiodarone can inhibit a variety of ion channels, including Na+/Ca2+ exchangers, L-type Ca2+ channels and Na+ channels, and increases the intracellular Ca(2+) level and GADD153 expression (197). Although viruses, bacteria, and calcium activators or inhibitors can induce cancer cell apoptosis through ER stress, their safety still needs to be considered.

Combined application

Various studies on GRP78 have also demonstrated that GRP78 has an important role in recurrent GBM and tumor progression after initial treatment (198,199). Of particular importance is TMZ, the standard-of-care chemotherapeutic treatment for GBM. TMZ has been demonstrated to result in activation of the UPR in GBM cells, inducing increased levels of UPR markers, GRP78 and GADD153 (100). Therefore, certain drugs can be used in combination with TMZ to enhance the sensitivity of GBM to TMZ by increasing ER stress (Table II). For example, bufothionine is extracted from the skins and parotid venom glands of the toad Bufo bufo gargarizans Cantor (200). Bufothionine can synergize with TMZ to exert an anti-growth effect by triggering ER stress (200). PI3K/mTOR signaling is ubiquitous in GBM (201). XL765 (Voxtalisib/SAR245409), an effective dual inhibitor of PI3Ks and mTOR, inhibits the proliferation of GBM cells by inducing ER stress-dependent apoptosis (201). The combination of XL765 and TMZ can achieve improved therapeutic effects in A172, U87MG and T98G cells (201). Fluoxetine (FLT), as a drug widely used in cancer-related depression, has strong anticancer effects in different types of cancer cells, such as human ovarian granulosa tumor COV434 cells, and SKBR3 and MCF-7 breast cancer cells (202). The combination of FLT and TMZ can also induce the activation of ATF6, PERK, eIF2α, ATF4 and GADD153 (202). The upregulation of prolyl 4-hydroxylase-β polypeptide (P4HB) expression is associated with the increase of the IC50 of TMZ, and is relatively upregulated in resistant GBM cells (203). Targeting P4HB blocks its protective function and makes GBM cells sensitive to TMZ (203). Chloroquine (CQ), a quinoline-based antimalarial drug, can kill the plasmodium falciparum parasite in the red blood cell stage by blocking the acidic food vacuole heme to detoxify (198). Under physiological pH conditions, CQ has unique chemical properties and is a weak base that easily crosses the lipid bilayer of cells (198). The combination of TMZ and CQ can trigger cell death by enhancing the formation of LC3B-II, the accumulation of polyubiquitinated proteins, GADD153 and the cleavage of PARP (198,204,205). N,N-[(8-hydroxyquinoline)methyl]-substituted benzylamine (JLK1486) is a novel type of ER stress inducer (206). The combined use of TMZ and JLK1486 can cause long-term ER stress in human GBM cell lines (U87MG, A172 and T98G) by increasing the levels of GRP78, ATF4 and GADD153 (206). In addition, celecoxib is a selective inhibitor of COX2. Increasing reports have described that this drug has powerful anti-proliferation and pro-apoptotic effects without the obvious involvement of COX2 (207,208). Celecoxib causes ER stress by leaking calcium from the ER into the cytoplasm (207). The combination of bortezomib and celecoxib can increase the expression levels of ER stress markers GRP78 and GADD153, and cause the activation of c-jun (207). In addition, whether other compounds enhance the sensitivity of GBM to TMZ by enhancing ER stress needs to be verified.

Table II

Potential treatments for glioblastoma in combination with TMZ.

Table II

Potential treatments for glioblastoma in combination with TMZ.

Author, yearCombined drugsImpact on UPR membersEfficacy for glioblastoma(Refs.)
Sunet al, 2019BufothionineGADD153, PERK, eIF2α, ATF6Pro-apoptosis; inhibiting cell proliferation(200)
Zhaoet al, 2019XL765PERK, eIF2α, GADD153Pro-apoptosis; inhibiting cell viability(201)
Maet al, 2016FluoxetinePERK, eIF2α, ATF4, ATF6Pro-apoptosis(202)
Sunet al, 2013P4HB inhibitionATF4, GADD34Pro-apoptosis(203)
Goldenet al, 2014; Goldenet al, 2015; Shteingauzet al, 2018ChloroquineGRP78, GADD153Block autophagy(198,204,205)
Weatherbeeet al, 2016JLK1486GRP78, ATF4, GADD153Decreasing cell proliferation; inducing cell death and the formation of DNA DSBs(206)
Kardoshet al, 2013DMCGRP78, GADD153Pro-apoptosis(207)

[i] XL765, voxtalisib; P4HB, prolyl 4-hydroxylase-β polypeptide; JLK1486, N,N-[(8-hydroxyquinoline)methyl]-substituted benzylamine; DMC, 2,5-dimethyl-celecoxib; DNA DSBs, DNA double stranded breaks.

6. Conclusions and perspectives

The resistance of GBM to TMZ treatment is the bottleneck of clinical treatment of this disease. Numerous cellular processes, including inflammation, autophagy and apoptosis, are regulated by the ER stress pathway. Furthermore, ER stress is a key regulator of TMZ sensitivity and is more likely to function in a cell-specific manner. Under low-dose and short-term TMZ treatment, ER stress may have cytoprotective effects. However, persistent ER stress can induce cell apoptosis. Therefore, numerous external factors and internal changes can induce ER stress in GBM cells. Although internal factors cannot be directly influenced, external factors can be used to delay GBM cell proliferation. For example, in a daily diet, patients with glioma can eat foods with less sugar and saturated fatty acids to inhibit cancer cell proliferation.

In addition, this review also summarizes some natural compounds and small molecule compounds, which are expected to treat GBM via ER stress. These compounds are also expected to be combined with TMZ. Additionally, there are numerous reports on other tumors that certain specific small-molecule regulators and natural compounds can specifically induce ER stress (165,209,210). Whether they can cross the blood-brain barrier and induce GBM cell apoptosis still requires verification. Additionally, the safety of the drug should also be worthy of consideration. Overall, although limited research has explored the pro-apoptosis function of ER stress for GBM cells, it has demonstrated that induced ER stress appears to be a potential treatment for GBM in the future.

Availability of data and materials

The datasets used and/or analyzed during the current study are available from the corresponding author on reasonable request.

Authors' contributions

PS and ZZ wrote this manuscript. JX prepared the table and figure. LZ revised grammar and polished vocabulary after the first review. HC drafted the manuscript. Data authentication is not applicable. All authors read and approved the final manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Acknowledgments

The authors would like to thank Dr Zhen Dong (State Key Laboratory of Silkworm Genome Biology, Southwest University, Chongqing, China) for revising the manuscript and providing kind suggestions.

Funding

This work was supported by the National Science Foundation of China (grant nos. 81872071 and 81902664), the Natural Science Foundation of Chongqing of China (grant no. cstc2019jcyj-zdxmX0033), the Graduate Research and Innovation Project of Chongqing of China (grant no. 2019CYB19117), the Fundamental Research Funds for the Central Universities (grant no. XYDS201912) and the State Key Laboratory of Silkworm Genome Biology (grant no. SKLSGB-ORP202002).

References

1 

Zhao Y, He J, Li Y, Lv S and Cui H: NUSAP1 potentiates chemoresistance in glioblastoma through its SAP domain to stabilize ATR. Signal Transduct Target Ther. 5:442020. View Article : Google Scholar : PubMed/NCBI

2 

Weller M, Wick W, Aldape K, Brada M, Berger M, Pfister SM, Nishikawa R, Rosenthal M, Wen PY, Stupp R and Reifenberger G: Glioma. Nat Rev Dis Primers. 1:150172015. View Article : Google Scholar : PubMed/NCBI

3 

Reimunde P, Pensado-López A, Carreira Crende M, Lombao Iglesias V, Sánchez L, Torrecilla-Parra M, Ramírez CM, Anfray C and Torres Andón F: Cellular and molecular mechanisms underlying glioblastoma and zebrafish models for the discovery of new treatments. Cancers (Basel). 13:10872021. View Article : Google Scholar

4 

Hoffman LM, Veldhuijzen van Zanten SEM, Colditz N, Baugh J, Chaney B, Hoffmann M, Lane A, Fuller C, Miles L, Hawkins C, et al: Clinical, radiologic, pathologic, and molecular characteristics of long-term survivors of diffuse intrinsic pontine glioma (DIPG): A collaborative report from the International and european society for pediatric oncology DIPG registries. J Clin Oncol. 36:1963–1972. 2018. View Article : Google Scholar : PubMed/NCBI

5 

Fareh M, Almairac F, Turchi L, Burel-Vandenbos F, Paquis P, Fontaine D, Lacas-Gervais S, Junier MP, Chneiweiss H and Virolle T: Cell-based therapy using miR-302-367 expressing cells represses glioblastoma growth. Cell Death Dis. 8:e27132017. View Article : Google Scholar : PubMed/NCBI

6 

Brandes AA, Tosoni A, Franceschi E, Reni M, Gatta G and Vecht C: Glioblastoma in adults. Crit Rev Oncol Hematol. 67:139–152. 2008. View Article : Google Scholar : PubMed/NCBI

7 

Tanaka S, Louis DN, Curry WT, Batchelor TT and Dietrich J: Diagnostic and therapeutic avenues for glioblastoma: No longer a dead end? Nat Rev Clin Oncol. 10:14–26. 2013. View Article : Google Scholar

8 

Yang L, Shi P, Zhao G, Xu J, Peng W, Zhang J, Zhang G, Wang X, Dong Z, Chen F and Cui H: Targeting cancer stem cell pathways for cancer therapy. Signal Transduct Target Ther. 5:82020. View Article : Google Scholar : PubMed/NCBI

9 

Aoyama-Ishiwatari S and Hirabayashi Y: Endoplasmic reticulum-mitochondria contact sites-emerging intracellular signaling hubs. Front Cell Dev Biol. 9:6538282021. View Article : Google Scholar : PubMed/NCBI

10 

Gonzalez-Gronow M, Gopal U, Austin RC and Pizzo SV: Glucose-regulated protein (GRP78) is an important cell surface receptor for viral invasion, cancers, and neurological disorders. IUBMB Life. 73:843–854. 2021. View Article : Google Scholar : PubMed/NCBI

11 

Yadav RK, Chae SW, Kim HR and Chae HJ: Endoplasmic reticulum stress and cancer. J Cancer Prev. 19:75–88. 2014. View Article : Google Scholar : PubMed/NCBI

12 

Kavitha CV, Jain AK, Agarwal C, Pierce A, Keating A, Huber KM, Serkova NJ, Wempe MF, Agarwal R and Deep G: Asiatic acid induces endoplasmic reticulum stress and apoptotic death in glioblastoma multiforme cells both in vitro and in vivo. Mol Carcinog. 54:1417–1429. 2015. View Article : Google Scholar :

13 

Zhang D, Wang F, Pang Y, Ke XX, Zhu S, Zhao E, Zhang K, Chen L and Cui H: Down-regulation of CHERP inhibits neuroblastoma cell proliferation and induces apoptosis through ER stress induction. Oncotarget. 8:80956–80970. 2017. View Article : Google Scholar : PubMed/NCBI

14 

McGrath EP, Centonze FG, Chevet E, Avril T and Lafont E: Death sentence: The tale of a fallen endoplasmic reticulum. Biochim Biophys Acta Mol Cell Res. 1868:1190012021. View Article : Google Scholar : PubMed/NCBI

15 

Wei J and Fang D: Endoplasmic reticulum stress signaling and the pathogenesis of hepatocarcinoma. Int J Mol Sci. 22:17992021. View Article : Google Scholar : PubMed/NCBI

16 

Bernales S, Papa FR and Walter P: Intracellular signaling by the unfolded protein response. Annu Rev Cell Dev Biology. 22:487–508. 2006. View Article : Google Scholar

17 

Wang M and Kaufman RJ: The impact of the endoplasmic reticulum protein-folding environment on cancer development. Nat Rev Cancer. 14:581–597. 2014. View Article : Google Scholar : PubMed/NCBI

18 

Stöhr D, Jeltsch A and Rehm M: TRAIL receptor signaling: From the basics of canonical signal transduction toward its entanglement with ER stress and the unfolded protein response. Int Rev Cell Mol Biol. 351:57–99. 2020. View Article : Google Scholar : PubMed/NCBI

19 

Kim C and Kim B: Anti-cancer natural products and their bioactive compounds inducing ER stress-mediated apoptosis: A review. Nutrients. 10:10212018. View Article : Google Scholar :

20 

Wang XZ, Harding HP, Zhang Y, Jolicoeur EM, Kuroda M and Ron D: Cloning of mammalian Ire1 reveals diversity in the ER stress responses. EMBO J. 17:5708–5717. 1998. View Article : Google Scholar : PubMed/NCBI

21 

Walter P and Ron D: The unfolded protein response: From stress pathway to homeostatic regulation. Science. 334:1081–1086. 2011. View Article : Google Scholar : PubMed/NCBI

22 

Haze K, Yoshida H, Yanagi H, Yura T and Mori K: Mammalian transcription factor ATF6 is synthesized as a transmembrane protein and activated by proteolysis in response to endoplasmic reticulum stress. Mol Biol Cell. 10:3787–3799. 1999. View Article : Google Scholar : PubMed/NCBI

23 

Elmore JM, Griffin BD and Walley JW: Advances in functional proteomics to study plant-pathogen interactions. Curr Opin Plant Biol. 63:1020612021. View Article : Google Scholar : PubMed/NCBI

24 

Liu Y, Chen DQ, Han JX, Zhao TT and Li SJ: A review of traditional Chinese medicine in treating renal interstitial fibrosis via endoplasmic reticulum stress-mediated apoptosis. Biomed Res Int. 2021:66677912021.PubMed/NCBI

25 

Fusée LTS, Marín M, Fåhraeus R and López I: Alternative mechanisms of p53 action during the unfolded protein response. Cancers(Basel). 12:4012020.

26 

Storchova R, Burdova K, Palek M, Medema RH and Macurek L: A novel assay for screening WIP1 phosphatase substrates in nuclear extracts. FEBS J. May 13–2021.Epub ahead of prin. View Article : Google Scholar : PubMed/NCBI

27 

Vodicka P, Andera L, Opattova A and Vodickova L: The interactions of DNA repair, telomere homeostasis, and p53 mutational status in solid cancers: Risk, prognosis, and prediction. Cancers (Basel). 13:4792021. View Article : Google Scholar

28 

Yukimoto A, Watanabe T, Sunago K, Nakamura Y, Tanaka T, Koizumi Y, Yoshida O, Tokumoto Y, Hirooka M, Abe M and Hiasa Y: The long noncoding RNA of RMRP is downregulated by PERK, which induces apoptosis in hepatocellular carcinoma cells. Sci Rep. 11:79262021. View Article : Google Scholar : PubMed/NCBI

29 

Huang T, Xu T, Wang Y, Zhou Y, Yu D, Wang Z, He L, Chen Z, Zhang Y, Davidson D, et al: Cannabidiol inhibits human glioma by induction of lethal mitophagy through activating TRPV4. Autophagy. Feb 25–2021.Epub ahead of prin. View Article : Google Scholar

30 

Kuran D, Flis S, Antoszczak M, Piskorek M and Huczyński A: Ester derivatives of salinomycin efficiently eliminate breast cancer cells via ER-stress-induced apoptosis. Eur J Pharmacol. 893:1738242021. View Article : Google Scholar

31 

Bressler KR, Ross JA, Ilnytskyy S, Vanden Dungen K, Taylor K, Patel K, Zovoilis A, Kovalchuk I and Thakor N: Depletion of eukaryotic initiation factor 5B (eIF5B) reprograms the cellular transcriptome and leads to activation of endoplasmic reticulum (ER) stress and c-Jun N-terminal kinase (JNK). Cell Stress Chaperones. 26:253–264. 2021. View Article : Google Scholar

32 

González-Quiroz M, Blondel A, Sagredo A, Hetz C, Chevet E and Pedeux R: When endoplasmic reticulum proteostasis meets the DNA damage response. Trends Cell Biol. 30:881–891. 2020. View Article : Google Scholar : PubMed/NCBI

33 

Ventura JJ, Hübner A, Zhang C, Flavell RA, Shokat KM and Davis RJ: Chemical genetic analysis of the time course of signal transduction by JNK. Mol Cell. 21:701–710. 2006. View Article : Google Scholar : PubMed/NCBI

34 

Tabas I and Ron D: Integrating the mechanisms of apoptosis induced by endoplasmic reticulum stress. Nat Cell Biol. 13:184–190. 2011. View Article : Google Scholar : PubMed/NCBI

35 

Farshbaf M, Khosroushahi AY, Mojarad-Jabali S, Zarebkohan A, Valizadeh H and Walker PR: Cell surface GRP78: An emerging imaging marker and therapeutic target for cancer. J Control Release. 328:932–941. 2020. View Article : Google Scholar : PubMed/NCBI

36 

Otero JH, Lizak B and Hendershot LM: Life and death of a BiP substrate. Semin Cell Dev Biol. 21:472–478. 2010. View Article : Google Scholar :

37 

Bertolotti A, Zhang Y, Hendershot LM, Harding HP and Ron D: Dynamic interaction of BiP and ER stress transducers in the unfolded-protein response. Nat Cell Biol. 2:326–332. 2000. View Article : Google Scholar : PubMed/NCBI

38 

Shen J, Snapp EL, Lippincott-Schwartz J and Prywes R: Stable binding of ATF6 to BiP in the endoplasmic reticulum stress response. Mol Cell Biol. 25:921–932. 2005. View Article : Google Scholar : PubMed/NCBI

39 

Pincus D, Chevalier MW, Aragon T, van Anken E, Vidal SE, El-Samad H and Walter P: BiP binding to the ER-stress sensor Ire1 tunes the homeostatic behavior of the unfolded protein response. PLoS Biol. 8:e10004152010. View Article : Google Scholar : PubMed/NCBI

40 

Harding HP, Zhang Y, Bertolotti A, Zeng H and Ron D: Perk is essential for translational regulation and cell survival during the unfolded protein response. Mol Cell. 5:897–904. 2000. View Article : Google Scholar : PubMed/NCBI

41 

Harding HP, Zhang Y and Ron D: Protein translation and folding are coupled by an endoplasmic-reticulum-resident kinase. Nature. 397:271–274. 1999. View Article : Google Scholar : PubMed/NCBI

42 

Kadowaki H and Nishitoh H: Signaling pathways from the endoplasmic reticulum and their roles in disease. Genes. 4:306–333. 2013. View Article : Google Scholar : PubMed/NCBI

43 

Hetz C and Mollereau B: Disturbance of endoplasmic reticulum proteostasis in neurodegenerative diseases. Nat Rev Neurosci. 15:233–249. 2014. View Article : Google Scholar : PubMed/NCBI

44 

Dai C, Li J, Tang S, Li J and Xiao X: Colistin-induced nephrotoxicity in mice involves the mitochondrial, death receptor, and endoplasmic reticulum pathways. Antimicrob Agents Chemother. 58:4075–4085. 2014. View Article : Google Scholar : PubMed/NCBI

45 

McCullough KD, Martindale JL, Klotz LO, Aw TY and Holbrook NJ: Gadd153 sensitizes cells to endoplasmic reticulum stress by down-regulating Bcl2 and perturbing the cellular redox state. Mol Cell Biol. 21:1249–1259. 2001. View Article : Google Scholar : PubMed/NCBI

46 

Kroemer G and Reed JC: Mitochondrial control of cell death. Nat Med. 6:513–519. 2000. View Article : Google Scholar : PubMed/NCBI

47 

Hengartner MO: The biochemistry of apoptosis. Nature. 407:770–776. 2000. View Article : Google Scholar : PubMed/NCBI

48 

Lien JC, Huang CC, Lu TJ, Tseng CH, Sung PJ, Lee HZ, Bao BY, Kuo YH and Lu TL: Naphthoquinone derivative PPE8 induces endoplasmic reticulum stress in p53 null H1299 cells. Oxid Med Cell Longev. 2015:4536792015. View Article : Google Scholar : PubMed/NCBI

49 

Kimata Y, Kimata YI, Shimizu Y, Abe H, Farcasanu IC, Takeuchi M, Rose MD and Kohno K: Genetic evidence for a role of BiP/Kar2 that regulates Ire1 in response to accumulation of unfolded proteins. Mol Biol Cell. 14:2559–2569. 2003. View Article : Google Scholar : PubMed/NCBI

50 

Jabouille A, Delugin M, Pineau R, Dubrac A, Soulet F, Lhomond S, Pallares-Lupon N, Prats H, Bikfalvi A, Chevet E, et al: Glioblastoma invasion and cooption depend on IRE1alpha eSndoribonuclease activity. Oncotarget. 6:24922–24934. 2015. View Article : Google Scholar : PubMed/NCBI

51 

Korennykh AV, Egea PF, Korostelev AA, Finer-Moore J, Zhang C, Shokat KM, Stroud RM and Walter P: The unfolded protein response signals through high-order assembly of Ire1. Nature. 457:687–693. 2009. View Article : Google Scholar

52 

Yoshida H, Matsui T, Yamamoto A, Okada T and Mori K: XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell. 107:881–891. 2001. View Article : Google Scholar

53 

Lim R, Barker G and Lappas M: TRADD, TRAF2, RIP1 and TAK1 are required for TNF-alpha-induced pro-labour mediators in human primary myometrial cells. Am J Reprod Immunol. Mar 24–2017.Epub ahead of print. View Article : Google Scholar

54 

Bluher M, Bashan N, Shai I, Harman-Boehm I, Tarnovscki T, Avinaoch E, Stumvoll M, Dietrich A, Klöting N and Rudich A: Activated Ask1-MKK4-p38MAPK/JNK stress signaling pathway in human omental fat tissue may link macrophage infiltration to whole-body Insulin sensitivity. J Clin Endocrinol Metab. 94:2507–2515. 2009. View Article : Google Scholar : PubMed/NCBI

55 

Syc-Mazurek SB, Rausch RL, Fernandes KA, Wilson MP and Libby RT: MKK4 and MKK7 are important for retinal development and axonal injury-induced retinal ganglion cell death. Cell Death Dis. 9:10952018. View Article : Google Scholar : PubMed/NCBI

56 

Sujitha S, Dinesh P and Rasool M: Berberine modulates ASK1 signaling mediated through TLR4/TRAF2 via upregulation of miR-23a. Toxicol Appl Pharmacol. 359:34–46. 2018. View Article : Google Scholar : PubMed/NCBI

57 

Matsui Y, Kuwabara T, Eguchi T, Nakajima K and Kondo M: Acetylation regulates the MKK4-JNK pathway in T cell receptor signaling. Immunol Lett. 194:21–28. 2018. View Article : Google Scholar

58 

Cao S, Tang J, Huang Y, Li G, Li Z, Cai W, Yuan Y, Liu J, Huang X and Zhang H: The road of solid tumor survival: From drug-induced endoplasmic reticulum stress to drug resistance. Front Mol Biosci. 8:6205142021. View Article : Google Scholar : PubMed/NCBI

59 

Bagchi AK, Malik A, Akolkar G, Zimmer A, Belló-Klein A, De Angelis K, Jassal DS, Fini MA, Stenmark KR and Singal PK: Study of ER stress and apoptotic proteins in the heart and tumor exposed to doxorubicin. Biochim Biophys Acta Mol Cell Res. 119039:20211868.

60 

Lynch JM, Maillet M, Vanhoutte D, Schloemer A, Sargent MA, Blair NS, Lynch KA, Okada T, Aronow BJ, Osinska H, et al: A thrombospondin-dependent pathway for a protective ER stress response. Cell. 149:1257–1268. 2012. View Article : Google Scholar : PubMed/NCBI

61 

Adachi Y, Yamamoto K, Okada T, Yoshida H, Harada A and Mori K: ATF6 is a transcription factor specializing in the regulation of quality control proteins in the endoplasmic reticulum. Cell Struct Funct. 33:75–89. 2008. View Article : Google Scholar : PubMed/NCBI

62 

Bommiasamy H, Back SH, Fagone P, Lee K, Meshinchi S, Vink E, Sriburi R, Frank M, Jackowski S, Kaufman RJ and Brewer JW: ATF6alpha induces XBP1-independent expansion of the endoplasmic reticulum. J Cell Sci. 122:1626–1636. 2009. View Article : Google Scholar : PubMed/NCBI

63 

Pinkham K, Park DJ, Hashemiaghdam A, Kirov AB, Adam I, Rosiak K, da Hora CC, Teng J, Cheah PS, Carvalho L, et al: Stearoyl CoA desaturase is essential for regulation of endoplasmic reticulum homeostasis and tumor growth in glioblastoma cancer stem cells. Stem Cell Reports. 12:712–727. 2019. View Article : Google Scholar : PubMed/NCBI

64 

Shakhparonov MI, Antipova NV, Shender VO, Shnaider PV, Arapidi GP, Pestov NB and Pavlyukov MS: Expression and intracellular localization of paraoxonase 2 in different types of malignancies. Acta Naturae. 10:92–99. 2018. View Article : Google Scholar : PubMed/NCBI

65 

Le Mercier M, Lefranc F, Mijatovic T, Debeir O, Haibe-Kains B, Bontempi G, Decaestecker C, Kiss R and Mathieu V: Evidence of galectin-1 involvement in glioma chemoresistance. Toxicol Appl Pharmacol. 229:172–183. 2008. View Article : Google Scholar : PubMed/NCBI

66 

Lin HY, Ko CY, Kao TJ, Yang WB, Tsai YT, Chuang JY, Hu SL, Yang PY, Lo WL and Hsu TI: CYP17A1 Maintains the survival of glioblastomas by regulating SAR1-mediated endoplasmic reticulum health and redox homeostasis. Cancers (Basel). 11:13782019. View Article : Google Scholar

67 

Jakubowicz-Gil J, Langner E, Badziul D, Wertel I and Rzeski W: Silencing of Hsp27 and Hsp72 in glioma cells as a tool for programmed cell death induction upon temozolomide and quercetin treatment. Toxicol Appl Pharmacol. 273:580–589. 2013. View Article : Google Scholar : PubMed/NCBI

68 

Horibe T, Torisawa A, Kohno M and Kawakami K: Molecular mechanism of cytotoxicity induced by Hsp90-targeted Antp-TPR hybrid peptide in glioblastoma cells. Mol Cancer. 11:592012. View Article : Google Scholar : PubMed/NCBI

69 

Choi JW, Schroeder MA, Sarkaria JN and Bram RJ: Cyclophilin B supports Myc and mutant p53-dependent survival of glioblastoma multiforme cells. Cancer Res. 74:484–496. 2014. View Article : Google Scholar

70 

Hu Y, Chu L, Liu J, Yu L, Song SB, Yang H and Han F: Knockdown of CREB3 activates endoplasmic reticulum stress and induces apoptosis in glioblastoma. Aging (Albany NY). 11:8156–8168. 2019. View Article : Google Scholar

71 

Chen Y, Tsai YH and Tseng SH: HDAC inhibitors and RECK modulate endoplasmic reticulum stress in tumor cells. Int J Mol Sci. 18:2582017. View Article : Google Scholar :

72 

Nguyen TTT, Ishida CT, Shang E, Shu C, Bianchetti E, Karpel-Massler G and Siegelin MD: Activation of LXR receptors and inhibition of TRAP1 causes synthetic lethality in solid tumors. Cancers (Basel). 11:7882019. View Article : Google Scholar

73 

Stock K, Kumar J, Synowitz M, Petrosino S, Imperatore R, Smith ES, Wend P, Purfürst B, Nuber UA, Gurok U, et al: Neural precursor cells induce cell death of high-grade astrocytomas through stimulation of TRPV1. Nat Med. 18:1232–1238. 2012. View Article : Google Scholar : PubMed/NCBI

74 

Zeeshan HM, Lee GH, Kim HR and Chae HJ: Endoplasmic reticulum stress and associated ROS. Int J Mol Sci. 17:3272016. View Article : Google Scholar : PubMed/NCBI

75 

Kyani A, Tamura S, Yang S, Shergalis A, Samanta S, Kuang Y, Ljungman M and Neamati N: Discovery and mechanistic elucidation of a class of protein disulfide isomerase inhibitors for the treatment of glioblastoma. ChemMedChem. 13:164–177. 2018. View Article : Google Scholar :

76 

Chen WL, Wang CC, Lin YJ, Wu CP and Hsieh CH: Cycling hypoxia induces chemoresistance through the activation of reactive oxygen species-mediated B-cell lymphoma extra-long pathway in glioblastoma multiforme. J Transl Med. 13:3892015. View Article : Google Scholar : PubMed/NCBI

77 

Agnihotri S, Golbourn B, Huang X, Remke M, Younger S, Cairns RA, Chalil A, Smith CA, Krumholtz SL, Mackenzie D, et al: PINK1 is a negative regulator of growth and the warburg effect in glioblastoma. Cancer Res. 76:4708–4719. 2016. View Article : Google Scholar : PubMed/NCBI

78 

Wang Q, Wang H, Jia Y, Pan H and Ding H: Luteolin induces apoptosis by ROS/ER stress and mitochondrial dysfunction in gliomablastoma. Cancer Chemother Pharmacol. 79:1031–1041. 2017. View Article : Google Scholar : PubMed/NCBI

79 

Qu C, Ma J, Liu X, Xue Y, Zheng J, Liu L, Liu J, Li Z, Zhang L and Liu Y: Dihydroartemisinin exerts anti-tumor activity by inducing mitochondrion and endoplasmic reticulum apoptosis and autophagic cell death in human glioblastoma cells. Front Cell Neurosci. 11:3102017. View Article : Google Scholar : PubMed/NCBI

80 

Schito L and Semenza GL: Hypoxia-inducible factors: Master regulators of cancer progression. Trends Cancer. 2:758–770. 2016. View Article : Google Scholar

81 

Zhou J, Schmid T, Schnitzer S and Brune B: Tumor hypoxia and cancer progression. Cancer Lett. 237:10–21. 2006. View Article : Google Scholar

82 

Wang GL, Jiang BH, Rue EA and Semenza GL: Hypoxia-inducible factor 1 is a basic-helix-loop-helix-PAS heterodimer regulated by cellular O2 tension. Proc Natl Acad Sci USA. 92:5510–5514. 1995. View Article : Google Scholar : PubMed/NCBI

83 

Bischoff FC, Werner A, John D, Boeckel JN, Melissari MT, Grote P, Glaser SF, Demolli S, Uchida S, Michalik KM, et al: Identification and functional characterization of hypoxia-induced endoplasmic reticulum stress regulating lncRNA (HypERlnc) in pericytes. Circ Res. 121:368–375. 2017. View Article : Google Scholar : PubMed/NCBI

84 

Han YK, Park GY, Bae MJ, Kim JS, Jo WS and Lee CG: Hypoxia induces immunogenic cell death of cancer cells by enhancing the exposure of cell surface calreticulin in an endoplasmic reticulum stress-dependent manner. Oncol Lett. 18:6269–6274. 2019.PubMed/NCBI

85 

Minchenko DO, Riabovol OO, Ratushna OO and Minchenko OH: Hypoxic regulation of the expression of genes encoded estrogen related proteins in U87 glioma cells: Effect of IRE1 inhibition. Endocr Regul. 51:8–19. 2017. View Article : Google Scholar : PubMed/NCBI

86 

Minchenko DO, Tsymbal DO, Riabovol OO, Viletska YM, Lahanovska YO, Sliusar MY, Bezrodnyi BH and Minchenko OH: Hypoxic regulation of EDN1, EDNRA, EDNRB, and ECE1 gene expressions in ERN1 knockdown U87 glioma cells. Endocr Regul. 53:250–262. 2019. View Article : Google Scholar : PubMed/NCBI

87 

Minchenko OH, Tsymbal DO, Minchenko DO, Kovalevska OV, Karbovskyi LL and Bikfalvi A: Inhibition of ERN1 signaling enzyme affects hypoxic regulation of the expression of E2F8, EPAS1, HOXC6, ATF3, TBX3 and FOXF1 genes in U87 glioma cells. Ukr Biochem J. 87:76–87. 2015. View Article : Google Scholar : PubMed/NCBI

88 

Minchenko OH, Tsymbal DO, Minchenko DO and Kubaychuk OO: Hypoxic regulation of MYBL1, MEST, TCF3, TCF8, GTF2B, GTF2F2 and SNAI2 genes expression in U87 glioma cells upon IRE1 inhibition. Ukr Biochem J. 88:52–62. 2016. View Article : Google Scholar

89 

Minchenko DO, Kharkova AP, Tsymbal DO, Karbovskyi LL and Minchenko OH: IRE1 inhibition affects the expression of insulin-like growth factor binding protein genes and modifies its sensitivity to glucose deprivation in U87 glioma cells. Endocr Regul. 49:185–197. 2015. View Article : Google Scholar : PubMed/NCBI

90 

Minchenko OH, Kharkova AP, Minchenko DO and Karbovskyi LL: Effect of hypoxia on the expression of genes that encode some IGFBP and ccn proteins in U87 glioma cells depends on IRE1 signaling. Ukr Biochem J. 87:52–63. 2015. View Article : Google Scholar

91 

Minami N, Tanaka K, Sasayama T, Kohmura E, Saya H and Sampetrean O: Lactate reprograms energy and lipid metabolism in glucose-deprived oxidative glioma stem cells. Metabolites. 11:3252021. View Article : Google Scholar : PubMed/NCBI

92 

Liberti MV and Locasale JW: The warburg effect: How does it benefit cancer cells? Trends Biochem Sci. 41:211–218. 2016. View Article : Google Scholar : PubMed/NCBI

93 

Crabtree HG: Observations on the carbohydrate metabolism of tumours. Biochem J. 23:536–545. 1929. View Article : Google Scholar : PubMed/NCBI

94 

Racker E: Bioenergetics and the problem of tumor growth. Am Sci. 60:56–63. 1972.PubMed/NCBI

95 

Shim H, Chun YS, Lewis BC and Dang CV: A unique glucose-dependent apoptotic pathway induced by c-Myc. Proc Natl Acad Sci USA. 95:1511–1516. 1998. View Article : Google Scholar : PubMed/NCBI

96 

Gu M, Gao Y and Chang P: KRAS mutation dictates the cancer immune environment in pancreatic ductal adenocarcinoma and other adenocarcinomas. Cancers (Basel). 13:24292021. View Article : Google Scholar

97 

Khrabrova DA, Yakubovskaya MG and Gromova ES: AML-associated mutations in DNA methyltransferase DNMT3A. Biochemistry (Mosc). 86:307–318. 2021. View Article : Google Scholar

98 

Pavlova NN and Thompson CB: The emerging hallmarks of cancer metabolism. Cell Metab. 23:27–47. 2016. View Article : Google Scholar : PubMed/NCBI

99 

Mechelli R, Romano S, Romano C, Morena E, Buscarinu MC, Bigi R, Bellucci G, Reniè R, Pellicciari G, Salvetti M and Ristori G: MAIT cells and microbiota in multiple sclerosis and other autoimmune diseases. Microorganisms. 9:11322021. View Article : Google Scholar : PubMed/NCBI

100 

Pyrko P, Schonthal AH, Hofman FM, Chen TC and Lee AS: The unfolded protein response regulator GRP78/BiP as a novel target for increasing chemosensitivity in malignant gliomas. Cancer Res. 67:9809–9816. 2007. View Article : Google Scholar : PubMed/NCBI

101 

Soejima E, Ohki T, Kurita Y, Yuan X, Tanaka K, Kakino S, Hara K, Nakayama H, Tajiri Y and Yamada K: Protective effect of 3-hydroxybutyrate against endoplasmic reticulum stress-associated vascular endothelial cell damage induced by low glucose exposure. PLoS One. 13:e01911472018. View Article : Google Scholar : PubMed/NCBI

102 

Zhang Y, Ishida CT, Ishida W, Lo SL, Zhao J, Shu C, Bianchetti E, Kleiner G, Sanchez-Quintero MJ, Quinzii CM, et al: Combined HDAC and bromodomain protein inhibition reprograms tumor cell metabolism and elicits synthetic lethality in glioblastoma. Clin Cancer Res. 24:3941–3954. 2018. View Article : Google Scholar : PubMed/NCBI

103 

Minchenko DO, Hubenya OV, Terletsky BM, Moenner M and Minchenko OH: Effect of glutamine or glucose deprivation on the expression of cyclin and cyclin-dependent kinase genes in glioma cell line U87 and its subline with suppressed activity of signaling enzyme of endoplasmic reticulum-nuclei-1. Ukr Biokhim Zh (1999). 83:18–29. 2011.

104 

Ma L, Wei J, Wan J, Wang W, Wang L, Yuan Y, Yang Z, Liu X and Ming L: Low glucose and metformin-induced apoptosis of human ovarian cancer cells is connected to ASK1 via mitochondrial and endoplasmic reticulum stress-associated pathways. J Exp Clin Cancer Res. 38:772019. View Article : Google Scholar : PubMed/NCBI

105 

Schornack PA and Gillies RJ: Contributions of cell metabolism and H+ diffusion to the acidic pH of tumors. Neoplasia. 5:135–145. 2003. View Article : Google Scholar : PubMed/NCBI

106 

Tannock IF and Rotin D: Acid pH in tumors and its potential for therapeutic exploitation. Cancer Res. 49:4373–4384. 1989.PubMed/NCBI

107 

Rofstad EK, Mathiesen B, Kindem K and Galappathi K: Acidic extracellular pH promotes experimental metastasis of human melanoma cells in athymic nude mice. Cancer Res. 66:6699–6707. 2006. View Article : Google Scholar : PubMed/NCBI

108 

Hjelmeland AB, Wu Q, Heddleston JM, Choudhary GS, MacSwords J, Lathia JD, McLendon R, Lindner D, Sloan A and Rich JN: Acidic stress promotes a glioma stem cell phenotype. Cell Death Differ. 18:829–840. 2011. View Article : Google Scholar :

109 

Park HJ, Lyons JC, Ohtsubo T and Song CW: Acidic environment causes apoptosis by increasing caspase activity. Br J Cancer. 80:1892–1897. 1999. View Article : Google Scholar : PubMed/NCBI

110 

Kato Y, Lambert CA, Colige AC, Mineur P, Noël A, Frankenne F, Foidart JM, Baba M, Hata R, Miyazaki K and Tsukuda M: Acidic extracellular pH induces matrix metalloproteinase-9 expression in mouse metastatic melanoma cells through the phospholipase D-mitogen-activated protein kinase signaling. J Biol Chem. 280:10938–10944. 2005. View Article : Google Scholar : PubMed/NCBI

111 

Xu L, Fukumura D and Jain RK: Acidic extracellular pH induces vascular endothelial growth factor (VEGF) in human glioblastoma cells via ERK1/2 MAPK signaling pathway: Mechanism of low pH-induced VEGF. J Biol Chemistry. 277:11368–11374. 2002. View Article : Google Scholar

112 

Reichert M, Steinbach JP, Supra P and Weller M: Modulation of growth and radiochemosensitivity of human malignant glioma cells by acidosis. Cancer. 95:1113–1119. 2002. View Article : Google Scholar : PubMed/NCBI

113 

Xie ZY, Chen L, Wang F, Liu L, Zhang C, Wang K, Cai F, Sinkemanni A, Hong X and Wu XT: Endoplasmic reticulum stress is involved in nucleus pulposus degeneration and attenuates low pH-Induced apoptosis of rat nucleus pulposus cells. DNA Cell Biol. 36:627–637. 2017. View Article : Google Scholar : PubMed/NCBI

114 

Dong L, Krewson EA and Yang LV: Acidosis Activates endoplasmic reticulum stress pathways through GPR4 in human vascular endothelial cells. Int J Mol Sci. 18:2782017. View Article : Google Scholar :

115 

Christensen SB, Andersen A, Kromann H, Treiman M, Tombal B, Denmeade S and Isaacs JT: Thapsigargin analogues for targeting programmed death of androgen-independent prostate cancer cells. Bioorg Med Chem. 7:1273–1280. 1999. View Article : Google Scholar : PubMed/NCBI

116 

Jansen S, Arning J and Beyersmann D: Effects of the Ca ionophore a23187 on zinc-induced apoptosis in C6 glioma cells. Biol Trace Elem Res. 96:133–142. 2003. View Article : Google Scholar

117 

Grant SK, Bansal A, Mitra A, Feighner SD, Dai G, Kaczorowski GJ and Middleton RE: Delay of intracellular calcium transients using a calcium chelator: Application to high-throughput screening of the capsaicin receptor ion channel and G-protein-coupled receptors. Anal Biochem. 294:27–35. 2001. View Article : Google Scholar : PubMed/NCBI

118 

Ciechomska IA, Gabrusiewicz K, Szczepankiewicz AA and Kaminska B: Endoplasmic reticulum stress triggers autophagy in malignant glioma cells undergoing cyclosporine a-induced cell death. Oncogene. 32:1518–1529. 2013. View Article : Google Scholar

119 

Kaul A and Maltese WA: Killing of cancer cells by the photoactivatable protein kinase C inhibitor, calphostin C, involves induction of endoplasmic reticulum stress. Neoplasia. 11:823–834. 2009. View Article : Google Scholar : PubMed/NCBI

120 

Martinez NJ, Rai G, Yasgar A, Lea WA, Sun H, Wang Y, Luci DK, Yang SM, Nishihara K, Takeda S, et al: A high-throughput screen identifies 2,9-diazaspiro[5.5]Undecanes as inducers of the endoplasmic reticulum stress response with cytotoxic activity in 3D glioma cell models. PLoS One. 11:e01614862016. View Article : Google Scholar

121 

Yu SN, Kim SH, Kim KY, Ji JH, Seo YK, Yu HS and Ahn SC: Salinomycin induces endoplasmic reticulum stressmediated autophagy and apoptosis through generation of reactive oxygen species in human glioma U87MG cells. Oncol Rep. 37:3321–3328. 2017. View Article : Google Scholar : PubMed/NCBI

122 

White MC, Johnson GG, Zhang W, Hobrath JV, Piazza GA and Grimaldi M: Sulindac sulfide inhibits sarcoendoplasmic reticulum Ca2+ ATPase, induces endoplasmic reticulum stress response, and exerts toxicity in glioma cells: Relevant similarities to and important differences from celecoxib. J Neurosci Res. 91:393–406. 2013. View Article : Google Scholar : PubMed/NCBI

123 

Yoon MJ, Kang YJ, Kim IY, Kim EH, Lee JA, Lim JH, Kwon TK and Choi KS: Monensin, a polyether ionophore antibiotic, overcomes TRAIL resistance in glioma cells via endoplasmic reticulum stress, DR5 upregulation and c-FLIP downregulation. Carcinogenesis. 34:1918–1928. 2013. View Article : Google Scholar : PubMed/NCBI

124 

Han J and Kaufman RJ: The role of ER stress in lipid metabolism and lipotoxicity. J Lipid Res. 57:1329–1338. 2016. View Article : Google Scholar : PubMed/NCBI

125 

Thibault G, Shui G, Kim W, McAlister GC, Ismail N, Gygi SP, Wenk MR and Ng DT: The membrane stress response buffers lethal effects of lipid disequilibrium by reprogramming the protein homeostasis network. Mol Cell. 48:16–27. 2012. View Article : Google Scholar : PubMed/NCBI

126 

Guo W, Wong S, Xie W, Lei T and Luo Z: Palmitate modulates intracellular signaling, induces endoplasmic reticulum stress, and causes apoptosis in mouse 3T3-L1 and rat primary preadipocytes. Am J Physiol Endocrinol Metab. 293:E576–E586. 2007. View Article : Google Scholar : PubMed/NCBI

127 

Griffiths B, Lewis CA, Bensaad K, Ros S, Zhang Q, Ferber EC, Konisti S, Peck B, Miess H, East P, et al: Sterol regulatory element binding protein-dependent regulation of lipid synthesis supports cell survival and tumor growth. Cancer Metab. 1:32013. View Article : Google Scholar : PubMed/NCBI

128 

Williams KJ, Argus JP, Zhu Y, Wilks MQ, Marbois BN, York AG, Kidani Y, Pourzia AL, Akhavan D, Lisiero DN, et al: An essential requirement for the SCAP/SREBP signaling axis to protect cancer cells from lipotoxicity. Cancer Res. 73:2850–2862. 2013. View Article : Google Scholar : PubMed/NCBI

129 

Ackerman D and Simon MC: Hypoxia, lipids, and cancer: Surviving the harsh tumor microenvironment. Trends Cell Biol. 24:472–478. 2014. View Article : Google Scholar : PubMed/NCBI

130 

He Y, Su J, Lan B, Gao Y and Zhao J: Targeting off-target effects: Endoplasmic reticulum stress and autophagy as effective strategies to enhance temozolomide treatment. Onco Targets Ther. 12:1857–1865. 2019. View Article : Google Scholar : PubMed/NCBI

131 

Zhang Y, Tseng CC, Tsai YL, Fu X, Schiff R and Lee AS: Cancer cells resistant to therapy promote cell surface relocalization of GRP78 which complexes with PI3K and enhances PI(3,4,5)P3 production. PLoS One. 8:e800712013. View Article : Google Scholar : PubMed/NCBI

132 

Misra UK, Deedwania R and Pizzo SV: Binding of activated alpha2-macroglobulin to its cell surface receptor GRP78 in 1-LN prostate cancer cells regulates PAK-2-dependent activation of LIMK. J Biol Chem. 280:26278–26286. 2005. View Article : Google Scholar : PubMed/NCBI

133 

Liu K, Tsung K and Attenello FJ: Characterizing cell stress and GRP78 in glioma to enhance tumor treatment. Front Oncol. 10:6089112020. View Article : Google Scholar : PubMed/NCBI

134 

Hseu YC, Lin RW, Shen YC, Lin KY, Liao JW, Thiyagarajan V and Yang HL: Flavokawain B and doxorubicin work synergistically to impede the propagation of gastric cancer cells via ROS-mediated apoptosis and autophagy Pathways. Cancers (Basel). 12:4752020. View Article : Google Scholar

135 

Doultsinos D, Carlesso A, Chintha C, Paton JC, Paton AW, Samali A, Chevet E and Eriksson LA: Peptidomimetic-based identification of FDA-approved compounds inhibiting IRE1 activity. FEBS J. 288:945–960. 2021. View Article : Google Scholar

136 

Hua R, Pei Y, Gu H, Sun Y and He Y: Antitumor effects of flavokawain-B flavonoid in gemcitabine-resistant lung cancer cells are mediated via mitochondrial-mediated apoptosis, ROS production, cell migration and cell invasion inhibition and blocking of PI3K/AKT Signaling pathway. J BUON. 25:262–267. 2020.PubMed/NCBI

137 

Wang J, Qi Q, Zhou W, Feng Z, Huang B, Chen A, Zhang D, Li W, Zhang Q, Jiang Z, et al: Inhibition of glioma growth by flavokawain B is mediated through endoplasmic reticulum stress induced autophagy. Autophagy. 14:2007–2022. 2018. View Article : Google Scholar : PubMed/NCBI

138 

Kim TH, Song J, Kim SH, Parikh AK, Mo X, Palanichamy K, Kaur B, Yu J, Yoon SO, Nakano I and Kwon CH: Piperlongumine treatment inactivates peroxiredoxin 4, exacerbates endoplasmic reticulum stress, and preferentially kills high-grade glioma cells. Neuro Oncol. 16:1354–1364. 2014. View Article : Google Scholar : PubMed/NCBI

139 

Tsai SF, Tao M, Ho LI, Chiou TW, Lin SZ, Su HL and Harn HJ: Isochaihulactone-induced DDIT3 causes ER stress-PERK independent apoptosis in glioblastoma multiforme cells. Oncotarget. 8:4051–4061. 2017. View Article : Google Scholar :

140 

Lu DY, Chang CS, Yeh WL, Tang CH, Cheung CW, Leung YM, Liu JF and Wong KL: The novel phloroglucinol derivative BFP induces apoptosis of glioma cancer through reactive oxygen species and endoplasmic reticulum stress pathways. Phytomedicine. 19:1093–1100. 2012. View Article : Google Scholar : PubMed/NCBI

141 

Guzman M: Cannabinoids: Potential anticancer agents. Nat Rev Cancer. 3:745–755. 2003. View Article : Google Scholar : PubMed/NCBI

142 

Carracedo A, Lorente M, Egia A, Blázquez C, García S, Giroux V, Malicet C, Villuendas R, Gironella M, González-Feria L, et al: The stress-regulated protein p8 mediates cannabinoid-induced apoptosis of tumor cells. Cancer Cell. 9:301–312. 2006. View Article : Google Scholar : PubMed/NCBI

143 

Ma X, Yu M, Hao C and Yang W: Shikonin induces tumor apoptosis in glioma cells via endoplasmic reticulum stress, and Bax/Bak mediated mitochondrial outer membrane permeability. J Ethnopharmacol. 263:1130592020. View Article : Google Scholar : PubMed/NCBI

144 

Pan JM, Zhou L, Wang GB, Xia GW, Xue K, Cui XG, Shi HZ, Liu JH and Hu J: Fatsioside A inhibits the growth of glioma cells via the induction of endoplasmic reticulum stress-mediated apoptosis. Mol Med Rep. 11:3493–3498. 2015. View Article : Google Scholar : PubMed/NCBI

145 

Das A, Banik NL and Ray SK: Garlic compounds generate reactive oxygen species leading to activation of stress kinases and cysteine proteases for apoptosis in human glioblastoma T98G and U87MG cells. Cancer. 110:1083–1095. 2007. View Article : Google Scholar : PubMed/NCBI

146 

Das A, Banik NL and Ray SK: Flavonoids activated caspases for apoptosis in human glioblastoma T98G and U87MG cells but not in human normal astrocytes. Cancer. 116:164–176. 2010.

147 

Djerir D, Iddir M, Bourgault S, Lamy S and Annabi B: Biophysical evidence for differential gallated green tea catechins binding to membrane type-1 matrix metalloproteinase and its interactors. Biophys Chem. 234:34–41. 2018. View Article : Google Scholar : PubMed/NCBI

148 

Ma J, Qiu Y, Yang L, Peng L, Xia Z, Hou LN, Fang C, Qi H and Chen HZ: Desipramine induces apoptosis in rat glioma cells via endoplasmic reticulum stress-dependent CHOP pathway. J Neurooncol. 101:41–48. 2011. View Article : Google Scholar

149 

Garrido-Armas M, Corona JC, Escobar ML, Torres L, Ordóñez-Romero F, Hernández-Hernández A and Arenas-Huertero F: Paraptosis in human glioblastoma cell line induced by curcumin. Toxicol In Vitro. 51:63–73. 2018. View Article : Google Scholar : PubMed/NCBI

150 

Sansalone L, Veliz EA, Myrthil NG, Stathias V, Walters W, Torrens II, Schürer SC, Vanni S, Leblanc RM and Graham RM: Novel Curcumin Inspired Bis-chalcone promotes endoplasmic reticulum stress and glioblastoma neurosphere cell death. Cancers (Basel). 11:3572019. View Article : Google Scholar

151 

Ma YY, Di ZM, Cao Q, Xu WS, Bi SX, Yu JS, Shen YJ, Yu YQ, Shen YX and Feng LJ: Xanthatin induces glioma cell apoptosis and inhibits tumor growth via activating endoplasmic reticulum stress-dependent CHOP pathway. Acta Pharmacol Sin. 41:404–414. 2020. View Article : Google Scholar :

152 

Martin S, Lamb HK, Brady C, Lefkove B, Bonner MY, Thompson P, Lovat PE, Arbiser JL, Hawkins AR and Redfern CP: Inducing apoptosis of cancer cells using small-molecule plant compounds that bind to GRP78. Br J Cancer. 109:433–443. 2013. View Article : Google Scholar : PubMed/NCBI

153 

Li ZY, Zhang C, Chen L, Chen BD, Li QZ, Zhang XJ and Li WP: Radicol, a novel trinorguaiane-type sesquiterpene, induces temozolomide-resistant glioma cell apoptosis via ER stress and Akt/mTOR pathway blockade. Phytother Res. 31:729–739. 2017. View Article : Google Scholar : PubMed/NCBI

154 

Chou YC, Chang MY, Wang MJ, Harnod T, Hung CH, Lee HT, Shen CC and Chung JG: PEITC induces apoptosis of human brain glioblastoma GBM8401 cells through the extrinsic- and intrinsic-signaling pathways. Neurochem Int. 81:32–40. 2015. View Article : Google Scholar : PubMed/NCBI

155 

Chou YC, Chang MY, Wang MJ, Liu HC, Chang SJ, Harnod T, Hung CH, Lee HT, Shen CC and Chung JG: Phenethyl isothiocyanate alters the gene expression and the levels of protein associated with cell cycle regulation in human glioblastoma GBM 8401 cells. Environ Toxicol. 32:176–187. 2017. View Article : Google Scholar

156 

Meng X, Leyva ML, Jenny M, Gross I, Benosman S, Fricker B, Harlepp S, Hébraud P, Boos A, Wlosik P, et al: A ruthenium-containing organometallic compound reduces tumor growth through induction of the endoplasmic reticulum stress gene CHOP. Cancer Res. 69:5458–5466. 2009. View Article : Google Scholar : PubMed/NCBI

157 

Badr CE, Van Hoppe S, Dumbuya H, Tjon-Kon-Fat LA and Tannous BA: Targeting cancer cells with the natural compound obtusaquinone. J Natl Cancer Inst. 105:643–653. 2013. View Article : Google Scholar : PubMed/NCBI

158 

Liu H, Xiong C, Liu J, Sun T, Ren Z, Li Y, Geng J and Li X: Aspirin exerts anti-tumor effect through inhibiting Blimp1 and activating ATF4/CHOP pathway in multiple myeloma. Biomed Pharmacother. 125:1100052020. View Article : Google Scholar : PubMed/NCBI

159 

Nishimura N, Radwan MO, Amano M, Endo S, Fujii E, Hayashi H, Ueno S, Ueno N, Tatetsu H, Hata H, et al: Novel p97/VCP inhibitor induces endoplasmic reticulum stress and apoptosis in both bortezomib-sensitive and -resistant multiple myeloma cells. Cancer Sci. 110:3275–3287. 2019. View Article : Google Scholar : PubMed/NCBI

160 

Alper P, Salomatina OV, Salakhutdinov NF, Ulukaya E and Ari F: Soloxolone methyl, as a 18βH-glycyrrhetinic acid derivate, may result in endoplasmic reticulum stress to induce apoptosis in breast cancer cells. Bioorg Med Chem. 30:1159632021. View Article : Google Scholar

161 

Ren M, Zhou X, Gu M, Jiao W, Yu M, Wang Y, Liu S, Yang J and Ji F: Resveratrol synergizes with cisplatin in antineoplastic effects against AGS gastric cancer cells by inducing endoplasmic reticulum stress-mediated apoptosis and G2/M phase arrest. Oncol Rep. 44:1605–1615. 2020.PubMed/NCBI

162 

De Wang X, Li T, Li Y, Yuan WH and Zhao YQ: 2-Pyrazine-PPD, a novel dammarane derivative, showed anticancer activity by reactive oxygen species-mediate apoptosis and endoplasmic reticulum stress in gastric cancer cells. Eur J Pharmacol. 881:1732112020. View Article : Google Scholar : PubMed/NCBI

163 

Zhang Q, Chen M, Cao L, Ren Y, Guo X, Wu X and Xu K: Phenethyl isothiocyanate synergistically induces apoptosis with Gefitinib in non-small cell lung cancer cells via endoplasmic reticulum stress-mediated degradation of Mcl-1. Mol Carcinog. 59:590–603. 2020. View Article : Google Scholar : PubMed/NCBI

164 

Zhu J, Xu S, Gao W, Feng J and Zhao G: Honokiol induces endoplasmic reticulum stress-mediated apoptosis in human lung cancer cells. Life Sci. 221:204–211. 2019. View Article : Google Scholar : PubMed/NCBI

165 

Peñaranda-Fajardo NM, Meijer C, Liang Y, Dijkstra BM, Aguirre-Gamboa R, den Dunnen WFA and Kruyt FAE: ER stress and UPR activation in glioblastoma: Identification of a noncanonical PERK mechanism regulating GBM stem cells through SOX2 modulation. Cell Death Dis. 10:6902019. View Article : Google Scholar : PubMed/NCBI

166 

Dadey DYA, Kapoor V, Khudanyan A, Thotala D and Hallahan DE: PERK regulates glioblastoma sensitivity to ER stress although promoting radiation resistance. Mol Cancer Res. 16:1447–1453. 2018. View Article : Google Scholar : PubMed/NCBI

167 

Axten JM, Medina JR, Feng Y, Shu A, Romeril SP, Grant SW, Li WH, Heerding DA, Minthorn E, Mencken T, et al: Discovery of 7-methyl-5-(1-{[3-(trifluoromethyl)phenyl]acetyl}-2,3-dihydro-1H-indol-5-yl)-7H-pyrrolo[2,3-d]pyrimidin-4-amine (GSK2606414), a potent and selective first-in-class inhibitor of protein kinase R (PKR)-like endoplasmic reticulum kinase (PERK). J Med Chem. 55:7193–7207. 2012. View Article : Google Scholar : PubMed/NCBI

168 

Rozpędek W, Pytel D, Wawrzynkiewicz A, Siwecka N, Dziki A, Dziki Ł, Diehl JA and Majsterek I: Use of Small-molecule inhibitory compound of PERK-dependent signaling pathway as a promising target-based therapy for colorectal cancer. Curr Cancer Drug Targets. 20:223–238. 2020. View Article : Google Scholar

169 

Nishikawa S, Itoh Y, Tokugawa M, Inoue Y, Nakashima KI, Hori Y, Miyajima C, Yoshida K, Morishita D, Ohoka N, et al: Kurarinone from Sophora Flavescens roots triggers ATF4 activation and cytostatic effects through PERK phosphorylation. Molecules. 24:31102019. View Article : Google Scholar :

170 

Scheffer D, Kulcsár G, Nagyéri G, Kiss-Merki M, Rékási Z, Maloy M and Czömpöly T: Active mixture of serum-circulating small molecules selectively inhibits proliferation and triggers apoptosis in cancer cells via induction of ER stress. Cell Signal. 65:1094262020. View Article : Google Scholar

171 

Guirao-Abad JP, Weichert M, Albee A, Deck K and Askew DS: A Human IRE1 inhibitor blocks the unfolded protein response in the pathogenic fungus aspergillus fumigatus and suggests noncanonical functions within the pathway. mSphere. 5:e00879–20. 2020. View Article : Google Scholar : PubMed/NCBI

172 

Cross BC, Bond PJ, Sadowski PG, Jha BK, Zak J, Goodman JM, Silverman RH, Neubert TA, Baxendale IR, Ron D and Harding HP: The molecular basis for selective inhibition of unconventional mRNA splicing by an IRE1-binding small molecule. Proc Natl Acad Sci USA. 109:E869–E878. 2012. View Article : Google Scholar : PubMed/NCBI

173 

Chen X, Li H, Fan X, Zhao C, Ye K, Zhao Z, Hu L, Ma H, Wang H and Fang Z: Protein palmitoylation regulates cell survival by modulating XBP1 activity in glioblastoma multiforme. Mol Ther Oncolytics. 17:518–530. 2020. View Article : Google Scholar : PubMed/NCBI

174 

Grandjean JMD, Madhavan A, Cech L, Seguinot BO, Paxman RJ, Smith E, Scampavia L, Powers ET, Cooley CB, Plate L, et al: Pharmacologic IRE1/XBP1s activation confers targeted ER proteostasis reprogramming. Nat Chem Biol. 16:1052–1061. 2020. View Article : Google Scholar : PubMed/NCBI

175 

Cho HY, Thein TZ, Wang W, Swenson SD, Fayngor RA, Ou M, Marín-Ramos NI, Schönthal AH, Hofman FM and Chen TC: The Rolipram-Perillyl alcohol conjugate (NEO214) is a mediator of cell death through the death receptor pathway. Mol Cancer Ther. 18:517–530. 2019. View Article : Google Scholar : PubMed/NCBI

176 

McCubrey JA, Lahair MM and Franklin RA: OSU-03012 in the treatment of glioblastoma. Mol Pharmacol. 70:437–439. 2006. View Article : Google Scholar : PubMed/NCBI

177 

Cho HY, Wang W, Jhaveri N, Lee DJ, Sharma N, Dubeau L, Schönthal AH, Hofman FM and Chen TC: NEO212, temozolomide conjugated to perillyl alcohol, is a novel drug for effective treatment of a broad range of temozolomide-resistant gliomas. Mol Cancer Ther. 13:2004–2017. 2014. View Article : Google Scholar : PubMed/NCBI

178 

Marin-Ramos NI, Perez-Hernandez M, Tam A, Swenson SD, Cho HY, Thein TZ, Hofman FM and Chen TC: Inhibition of motility by NEO100 through the calpain-1/RhoA pathway. J Neurosurgery. Aug 16–2019.Epub ahead of print. View Article : Google Scholar

179 

Chen HY, He LJ, Li SQ, Zhang YJ, Huang JH, Qin HX, Wang JL, Li QY and Yang DL: A derivate of benzimidazole-isoquinolinone induces SKP2 transcriptional inhibition to exert anti-tumor activity in glioblastoma cells. Molecules. 24:27222019. View Article : Google Scholar :

180 

Koncarevic S, Urig S, Steiner K, Rahlfs S, Herold-Mende C, Sueltmann H and Becker K: Differential genomic and proteomic profiling of glioblastoma cells exposed to terpyridineplatinum(II) complexes. Free Radic Biol Med. 46:1096–1108. 2009. View Article : Google Scholar : PubMed/NCBI

181 

Li Z, Ma J, Liu L, Liu X, Wang P, Liu Y, Li Z, Zheng J, Chen J, Tao W and Xue Y: Endothelial-monocyte activating polypeptide II suppresses the in vitro glioblastoma-induced angiogenesis by inducing autophagy. Front Mol Neurosci. 10:2082017. View Article : Google Scholar :

182 

Eom KS, Kim HJ, So HS, Park R and Kim TY: Berberine-induced apoptosis in human glioblastoma T98G cells is mediated by endoplasmic reticulum stress accompanying reactive oxygen species and mitochondrial dysfunction. Biol Pharm Bull. 33:1644–1649. 2010. View Article : Google Scholar : PubMed/NCBI

183 

Wang L, Gundelach JH and Bram RJ: Cycloheximide promotes paraptosis induced by inhibition of cyclophilins in glioblastoma multiforme. Cell Death Dis. 8:e28072017. View Article : Google Scholar : PubMed/NCBI

184 

Suzuki K, Gerelchuluun A, Hong Z, Sun L, Zenkoh J, Moritake T and Tsuboi K: Celecoxib enhances radiosensitivity of hypoxic glioblastoma cells through endoplasmic reticulum stress. Neuro Oncol. 15:1186–1199. 2013. View Article : Google Scholar : PubMed/NCBI

185 

Ye T, Wei L, Shi J, Jiang K, Xu H, Hu L, Kong L, Zhang Y, Meng S and Piao H: Sirtuin1 activator SRT2183 suppresses glioma cell growth involving activation of endoplasmic reticulum stress pathway. BMC Cancer. 19:7062019. View Article : Google Scholar : PubMed/NCBI

186 

Jia W, Loria RM, Park MA, Yacoub A, Dent P and Graf MR: The neuro-steroid, 5-androstene 3β,17α diol; induces endoplasmic reticulum stress and autophagy through PERK/eIF2α signaling in malignant glioma cells and transformed fibroblasts. Int J Biochem Cell Biol. 42:2019–2029. 2010. View Article : Google Scholar : PubMed/NCBI

187 

Liu WT, Huang CY, Lu IC and Gean PW: Inhibition of glioma growth by minocycline is mediated through endoplasmic reticulum stress-induced apoptosis and autophagic cell death. Neuro Oncol. 15:1127–1141. 2013. View Article : Google Scholar : PubMed/NCBI

188 

Shen S, Zhang Y, Zhang R, Tu X and Gong X: Ursolic acid induces autophagy in U87MG cells via ROS-dependent endoplasmic reticulum stress. Chem Biol Interact. 218:28–41. 2014. View Article : Google Scholar : PubMed/NCBI

189 

Bown CD, Wang JF and Young LT: Increased expression of endoplasmic reticulum stress proteins following chronic valproate treatment of rat C6 glioma cells. Neuropharmacology. 39:2162–2169. 2000. View Article : Google Scholar : PubMed/NCBI

190 

Park E, Gim J, Kim DK, Kim CS and Chun HS: Protective effects of alpha-lipoic acid on glutamate-induced cytotoxicity in C6 glioma cells. Biol Pharm Bull. 42:94–102. 2019. View Article : Google Scholar : PubMed/NCBI

191 

Kim IY, Kwon M, Choi MK, Lee D, Lee DM, Seo MJ and Choi KS: Ophiobolin A kills human glioblastoma cells by inducing endoplasmic reticulum stress via disruption of thiol proteostasis. Oncotarget. 8:106740–106752. 2017. View Article : Google Scholar :

192 

Qaisiya M, Brischetto C, Jasprova J, Vitek L, Tiribelli C and Bellarosa C: Bilirubin-induced ER stress contributes to the inflammatory response and apoptosis in neuronal cells. Arch Toxicol. 91:1847–1858. 2017. View Article : Google Scholar

193 

Mahoney DJ, Lefebvre C, Allan K, Brun J, Sanaei CA, Baird S, Pearce N, Grönberg S, Wilson B, Prakesh M, et al: Virus-tumor interactome screen reveals ER stress response can reprogram resistant cancers for oncolytic virus-triggered caspase-2 cell death. Cancer Cell. 20:443–456. 2011. View Article : Google Scholar : PubMed/NCBI

194 

Abraham R, Mudaliar P, Padmanabhan A and Sreekumar E: Induction of cytopathogenicity in human glioblastoma cells by chikungunya virus. PLoS One. 8:e758542013. View Article : Google Scholar : PubMed/NCBI

195 

Kusaczuk M, Kretowski R, Naumowicz M, Stypulkowska A and Cechowska-Pasko M: Silica nanoparticle-induced oxidative stress and mitochondrial damage is followed by activation of intrinsic apoptosis pathway in glioblastoma cells. Int J Nanomedicine. 13:2279–2294. 2018. View Article : Google Scholar : PubMed/NCBI

196 

Rubiolo JA, Lopez-Alonso H, Martínez P, Millán A, Cagide E, Vieytes MR, Vega FV and Botana LM: Yessotoxin induces ER-stress followed by autophagic cell death in glioma cells mediated by mTOR and BNIP3. Cell Signal. 26:419–432. 2014. View Article : Google Scholar : PubMed/NCBI

197 

Kim IY, Kang YJ, Yoon MJ, Kim EH, Kim SU, Kwon TK, Kim IA and Choi KS: Amiodarone sensitizes human glioma cells but not astrocytes to TRAIL-induced apoptosis via CHOP-mediated DR5 upregulation. Neuro Oncol. 13:267–279. 2011. View Article : Google Scholar : PubMed/NCBI

198 

Golden EB, Cho HY, Jahanian A, Hofman FM, Louie SG, Schönthal AH and Chen TC: Chloroquine enhances temozolomide cytotoxicity in malignant gliomas by blocking autophagy. Neurosurg Focus. 37:E122014. View Article : Google Scholar : PubMed/NCBI

199 

Przystal JM, Waramit S, Pranjol MZI, Yan W, Chu G, Chongchai A, Samarth G, Olaciregui NG, Tabatabai G, Carcaboso AM, et al: Efficacy of systemic temozolomide-activated phage-targeted gene therapy in human glioblastoma. EMBO Mol Med. 11:e84922019. View Article : Google Scholar : PubMed/NCBI

200 

Sun Y and Zhang X: Bufothionine promotes apoptosis via triggering ER stress and synergizes with temozolomide in glioblastoma multiforme cells. Anat Rec (Hoboken). 302:1950–1957. 2019. View Article : Google Scholar

201 

Zhao H, Chen G and Liang H: Dual PI3K/mTOR inhibitor, XL765, suppresses glioblastoma growth by inducing ER stress-dependent apoptosis. Onco Targets Ther. 12:5415–5424. 2019. View Article : Google Scholar : PubMed/NCBI

202 

Ma J, Yang YR, Chen W, Chen MH, Wang H, Wang XD, Sun LL, Wang FZ and Wang DC: Fluoxetine synergizes with temozolomide to induce the CHOP-dependent endoplasmic reticulum stress-related apoptosis pathway in glioma cells. Oncol Rep. 36:676–684. 2016. View Article : Google Scholar : PubMed/NCBI

203 

Sun S, Lee D, Ho AS, Pu JK, Zhang XQ, Lee NP, Day PJ, Lui WM, Fung CF and Leung GK: Inhibition of prolyl 4-hydroxylase, beta polypeptide (P4HB) attenuates temozolomide resistance in malignant glioma via the endoplasmic reticulum stress response (ERSR) pathways. Neuro Pncol. 15:562–577. 2013.

204 

Golden EB, Cho HY, Hofman FM, Louie SG, Schonthal AH and Chen TC: Quinoline-based antimalarial drugs: A novel class of autophagy inhibitors. Neurosurg Focus. 38:E122015. View Article : Google Scholar : PubMed/NCBI

205 

Shteingauz A, Porat Y, Voloshin T, Schneiderman RS, Munster M, Zeevi E, Kaynan N, Gotlib K, Giladi M, Kirson ED, et al: AMPK-dependent autophagy upregulation serves as a survival mechanism in response to Tumor Treating Fields (TTFields). Cell Death Dis. 9:10742018. View Article : Google Scholar : PubMed/NCBI

206 

Weatherbee JL, Kraus JL and Ross AH: ER stress in temozolomide-treated glioblastomas interferes with DNA repair and induces apoptosis. Oncotarget. 7:43820–43834. 2016. View Article : Google Scholar : PubMed/NCBI

207 

Kardosh A, Golden EB, Pyrko P, Uddin J, Hofman FM, Chen TC, Louie SG, Petasis NA and Schönthal AH: Aggravated endoplasmic reticulum stress as a basis for enhanced glioblastoma cell killing by bortezomib in combination with celecoxib or its non-coxib analogue, 2,5-dimethyl-celecoxib. Cancer Res. 68:843–851. 2008. View Article : Google Scholar : PubMed/NCBI

208 

Wang D, Fu L, Sun H, Guo L and DuBois RN: Prostaglandin E2 promotes colorectal cancer stem cell expansion and metastasis in mice. Gastroenterology. 149:1884–1895.e4. 2015. View Article : Google Scholar : PubMed/NCBI

209 

Grandjean JM and Wiseman RL: Small molecule strategies to harness the unfolded protein response: Where do we go from here? J Biol Chemistry. 295:15692–15711. 2020. View Article : Google Scholar

210 

Bian T, Tagmount A, Vulpe C, Vijendra KC and Xing C: CXL146, a novel 4H-chromene derivative, targets GRP78 to selectively eliminate multidrug-resistant cancer cells. Mol Pharmacol. 97:402–408. 2020. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

August-2021
Volume 59 Issue 2

Print ISSN: 1019-6439
Online ISSN:1791-2423

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Shi P, Zhang Z, Xu J, Zhang L and Cui H: Endoplasmic reticulum stress‑induced cell death as a potential mechanism for targeted therapy in glioblastoma (Review). Int J Oncol 59: 60, 2021
APA
Shi, P., Zhang, Z., Xu, J., Zhang, L., & Cui, H. (2021). Endoplasmic reticulum stress‑induced cell death as a potential mechanism for targeted therapy in glioblastoma (Review). International Journal of Oncology, 59, 60. https://doi.org/10.3892/ijo.2021.5240
MLA
Shi, P., Zhang, Z., Xu, J., Zhang, L., Cui, H."Endoplasmic reticulum stress‑induced cell death as a potential mechanism for targeted therapy in glioblastoma (Review)". International Journal of Oncology 59.2 (2021): 60.
Chicago
Shi, P., Zhang, Z., Xu, J., Zhang, L., Cui, H."Endoplasmic reticulum stress‑induced cell death as a potential mechanism for targeted therapy in glioblastoma (Review)". International Journal of Oncology 59, no. 2 (2021): 60. https://doi.org/10.3892/ijo.2021.5240