Molecular mechanisms of drug resistance of glioblastoma part 1: ABC family proteins and inhibitors

Cover Page


Cite item

Full Text

Open Access Open Access
Restricted Access Access granted
Restricted Access Subscription or Fee Access

Abstract

The most common high-grade brain tumor in the adult population is glioblastoma. The life expectancy of patients with this tumor does not exceed 12-15 months, while relapses are observed in 100% of cases. One of the main reasons for the low efficiency of glioblastoma therapy is its multidrug resistance. In the development of the latter, transporter proteins of the ABC family play a key role. In this part, the emphasis is on the search for new molecular targets among growth factors, their receptors, signal transduction kinases, microRNAs, transcription factors, protooncogenes, and tumor suppressor genes involved in the regulation of proteins and genes of the ABC family and associated with the development of multidrug resistance in glioblastoma cells. The review also discusses the mechanisms of the cytotoxic action of inhibitors: ABC family proteins, tyrosine kinase receptors, non-receptor tyrosine kinases, vascular endothelial growth factor, kinases of signaling cascades, transcription factors, histone deacetylases, methyltransferases, replication and synthesis of DNA, microtubules and proteasome used in glioblastoma therapy or undergoing clinical trials.

Full Text

List of abbreviations

GBM, glioblastoma; BBB, blood–brain barrier; MDR, multidrug resistance; TMZ, temozolomide; ABCB1, ATP-binding cassette protein-1 subfamily B; ABCC1, ATP-binding cassette protein-1 subfamily C; BCRP, breast cancer resistance protein; EGFR, epidermal growth factor receptor; HDAC, histone deacetylases; MAPK, mitogen-activated protein kinase; MELK, maternal embryonic leucine zipper kinase; NF-kB, nuclear factor kappa B; PDGF, platelet-derived growth factor; PDGFR, platelet-derived growth factor receptor; P-gp, P-glycoprotein; VEGF, vascular endothelial growth factor; VEGFR, vascular endothelial growth factor receptor.

 

Glioblastoma (GBM) is the most common type of high-grade brain tumor in the adult population. It accounts for 77%–81% of all malignant primary brain tumors, and its incidence reaches 10 people per 100,000 [1]. The tumor can occur at any age but is especially common at 45–75 years [1]. The five-year overall survival rate of patients with GBM at the age of 20–44 years is 13%, and at the age of 55–64 years, it is only 1% [2]. The standard treatment protocol for GBM includes surgical resection of the tumor, followed by radiotherapy and chemotherapy [3]. With this treatment regimen, the life expectancy of patients is only 14.6 months, with 100% relapse [4].

Because of the deregulation of several physiological and pathological molecular, genetic, subcellular, and cellular mechanisms, the GBM cells become radio- and chemoresistance, which makes the therapy less effective. The key participants in the previously mentioned mechanisms are growth factors, cytokines and their receptors, proteins of signal transduction cascades, transcription factors, miRNAs, oncogenes, and tumor suppressor genes. All the external influences on the tumor cells are integrated using signaling cascades, transcription factors, miRNAs, and genes. Therefore, a detailed study on the molecular mechanisms in GBM cells will identify new key targets and nodes that can be used as markers of effective targeted therapy.

This review highlights the molecular mechanisms of GBM multidrug resistance (MDR) in pathophysiological cellular processes, with a focus on the transporter proteins of the ATP-binding cassette (ABC) family and the use of inhibitors in the treatment of GBM.

Drug resistance and the ABC family transporter proteins

The overexpression of ABC transporters plays a key role in the development of MDR in tumor cells, including GBM. In humans, 49 genes encoding ABC proteins were identified, which are divided into seven subfamilies: ABC1 (ABCA), MDR/TAP (ABCB), MRP (ABCC), ALD (ABCD), OABP (ABCE), GCN20 (ABCF), and white (ABCG) [5].

The ABCA subfamily includes 13 proteins, of which ABCA1, ABCG1, and ABCG4 are involved in excreting cholesterol and phospholipids from the cells [6].

Simultaneously, genes and proteins belonging to the subfamilies ABCB, ABCC, and ABCG are most often overexpressed in GBM. The structure of these proteins is well characterized [7], but the exact mechanisms of their regulation and protein and gene targets are underinvestigated.

ABCB1

Of the eleven ABCB proteins, protein-1 (ABCB1; P-glycoprotein; P-gp; Mdr1; CD243) is the most studied. This transmembrane protein, consisting of two ATP-binding and two transmembrane domains, is expressed on the apical membrane of the capillary endothelial cells forming the blood–brain barrier (BBB) and in the U87 human glioma stem cells [8]. The ABCB1 protein is involved in excreting the anticancer drugs (etoposide, doxorubicin, vinblastine, gefitinib, sunitinib, tacrolimus, and temozolomide), organic cations, carbohydrates, oligosaccharides, lipids, steroids, bilirubin, amino acids, peptides, antibiotics, xenobiotics, dexamethasone, and cardiac glycosides (digoxin) from cells [9, 10]. P-gp expression was found on the nuclear membrane of the cells [11]. However, the physiological significance of the previously mentioned phenomenon has not yet been established. The variety of ABCB1 substrates predetermines the variety of functions such as regulation of the bioavailability and distribution of chemotherapy drugs and restriction of their transfer through the BBB to the brain and the protection of tumor stem cells from toxins. Overexpression of ABCB1 in the intestinal enterocytes slows down the penetration of chemotherapy drugs into the bloodstream, which prevents them from reaching therapeutic concentrations in patients with GBM [12].

ABCB1 gene expression and activity are regulated by transcription factors, signal transduction kinases, miRNAs, and growth factors. For example, many transcription factors bind to the ABCB1 gene promoter, namely, protein p53, tyrosine domain containing protein-1 (YB-1), nuclear factor-kB (NF-kB), protein-containing cAMP binding element (CREB), transcription factor-1 containing a lysine domain and MADS domains (AP-1), and an enhancer of the subunit of the zeste 2 polycomb repressive complex 2 (EZH2) [12–14].

Simultaneously, in GBM, the previously mentioned transcription factors are activated by signaling cascades with the participation of phosphatidylinositol-3-kinase/protein kinase-B, the mechanistic target of rapamycin kinase (Pi3k/Akt/mTOR), Wnt5a-Frizzled, the receptor/kinase-3β-glycogen synthase (Wnt5a/Frizzled/Gsk-3β), Ras/Raf/mitogen-activated protein kinase (MAPK), and c-Jun/c-Jun N terminal kinase (JNK) [15–17]. Activation of the Pi3k/Akt/NF-kB cascade increases the expression of O-6-methylguanine-DNA-methyltransferase (MGMT) and; hence, the resistance of GBM cells to the main chemotherapy drug, temozolomide (TMZ) [18]. Therefore, upon activation of this pathway, the expression of the ABCB1 gene will increase the resistance of GBM to TMZ. More detailed studies have shown that the MAPK/Erk1/2 and p38MAPK cascades stimulate P-gp, and c-Jun/JNK inhibits ABCB1 expression (Figure) [19, 20]. In turn, p38MAPK is activated by CD133 membrane glycoprotein, which colocalizes on GBM membranes with the epidermal growth factor receptor (EGFR) [21]. The latter triggers the signal transducer and transcription activator 3 (STAT3) cascade, which also enhances tumor progression [22]. It is suggested that ABCB1 expression may be inhibited by the transcription factor O3A containing a forkhead domain (FOXO3a) since it is activated by PTEN- (deleted phosphatase and tensin homolog on chromosome 10) mediated inhibition of the Pi3k/Akt cascade [23].

 

Figure. Intracellular mechanisms of multidrug resistance of glioblastoma involving genes ABCB1 and ABCG2. See text for explanations / Рисунок. Внутриклеточные механизмы множественной лекарственной устойчивости глиобластомы с участием генов ABCB1 и ABCG2. Объяснения см. в тексте

 

MicroRNAs are also involved in the regulation of ABCB1 gene transcription [24, 25]. For example, has-miR-4261 inhibits P-gp expression through MGMT suppression, which increases cell sensitivity to TMZ [24]. MiR-200c also suppresses P-gp expression through the JNK2/c-Jun signaling pathway [26]. MiR-130a possibly activates ABCB1 through the Pi3k/Akt/PTEN/mTOR and Wnt/β-catenin signaling cascades (Figure) [25].

The level of long noncoding RNA SNHG15 correlates with high levels of β-catenin, EGFR, transcription factor-2 containing the SRY domain (SOX-2), and cell division kinase-6 (CDK6) in TMZ-resistant GBM cells. SNHG15 enhances tumor progression by inhibiting the miR-627-5p suppressor, leading to the activation of CDK6 and SOX-2 [27].

ABCB1 expression is regulated at the posttranscriptional level by degradation and intracellular redistribution of P-gp. For example, serine-threonine kinase PIM-1 prevents P-gp ubiquitination and its degradation by proteasome proteins [28]. In another case, the small GTPase of the RAS family RAB5 suppresses P-gp endocytosis and increases the number of its molecules on the cell membrane, while RAB4 activates endocytosis and reduces the number of molecules on the membrane [29]. In GBM, the ABCB1 promoter is methylated [30].

The activation of suppressor genes for amiloride- sensitive cation channels 3 and 4 (ACCN3 and ACCN4) suppresses EGFR expression and, through it ABCB1 activity [31]. In GBM U251 cells, the P-gp expression is suppressed by inhibiting Bcl-2 when exposed to bone morphogenetic protein-4 (BMP4) [32].

ABCB1 expression in GBM cells is activated in cyclic hypoxia due to exposure to hypoxia-inducible factor-1α (HIF-1α), which reduces the sensitivity of tumor cells to doxorubicin [33]. In turn, HIF-1α stimulates the synthesis of carboxylic anhydrase-9 (CA9) in GBM cells under hypoxic conditions, which further reduces the pH of the tumor microenvironment, enhancing its resistance to chemotherapy [34]. Therefore, CA9 expression is expected to be positively correlated with the activity of the P-gp protein. HIF-1α also activates the expression of erythroid-like nuclear transcription factor-2 (NRF2), which enhances the resistance of GBM to chemo- and radiotherapy [35]. In this regard, NRF2 expression can correlate with the activity of ABCB1 and P-gp. On the contrary, in aerobic glycolysis, CDC-like kinase-1 (CLK1) enhances glucose utilization and suppression of lactate formation in GL261 glioma cells. When CLK1 is activated through the AMP-activated protein kinase (AMPK)/mTOR signaling cascade, HIF-1α expression is inhibited [36]. CLK1 can be considered a new target of action aimed at ABCB1 and P-gp.

Currently, suppressing ABCB1 expression and P-gp activity is the mechanism of action of many drugs, such as amiodarone, azithromycin, captopril, clarithromycin, cyclosporine, piperine, quercetin, quinidine, quinine, reserpine, ritonavir, tariquidar, and verapamil.

ABCC

Expression of ABCC proteins (MRP) including the subfamilies MRP1 (ABCC1; multidrug resistance protein 1), MRP3 (ABCC3), MRP4 (ABCC4), and MRP5 (ABCC5) is also observed in GBM cells [37]. ABCC1 is expressed in the tumor stem cells and is weakly expressed in adherent GBM cells [38]. ABCC1 removes chemotherapy drugs (vincristine, etoposide, doxorubicin, methotrexate, cisplatin, and mitoxantrone), C4 leukotriene, conjugates of estrogen, glucuronides, sulfate conjugates of steroid hormones, heavy metals, organic amines, and lipids from the tumor cells [39]. Proteins MRP3 and MRP4 are involved in excreting glucocorticoids and prostaglandins E1 and E2, respectively, and MRP5 is involved in excreting chemotherapy drugs (thiopurine, 6-mercaptopurine, and thioguanine) and their conjugates with glutathione and glucosyl- and sulfatidylsteroids [40]. The proteins ABCC4 and ABCC5 regulate intracellular signaling to the nucleus through cyclic adenosine monophosphate (cAMP). Additionally, ABCC5 promotes the degradation of phosphodiesterases and the elimination of cyclic nucleotides [41]. Proteins ABCC8 and ABCC9, being sulfonylurea receptors, form ATP-binding subunits of the potassium channel and inhibit the activity of GBM cells [42]. All these proteins can be considered new targets for targeted anticancer drugs.

Many mechanisms are used to regulate ABCC genes. For example, the expression of ABCC2 and ABCC4 is suppressed by the action of secreted frizzled-like protein-4 (sFRP4) and tacrolimus and the expression of ABCC1 in stem and GBM cells [43]. Activation of ABCC1 and ABCC3 is stimulated by MGMT through the insulin-like growth factor 1 (IGF1R) and Pi3k/Akt/MYC cascade and the transcription factor EZH2, respectively; therefore, MRP1 and MRP3 proteins may be involved in the development of GBM resistance to TMZ [44].

The large vaulted ABCC subfamily protein (MVP/LRP) is activated through the EGFR and SHH/GLI signaling cascade but is inhibited by the maternal embryonic leucine zipper domain containing kinase (MELK) and PTEN [45]. In turn, EGFR activation in GBM enhances the expression of the c-MET receptor for hepatocyte growth factor (HGF) [46]. c-MET and HGF can activate the expression of ABCC and ABCB1. Additionally, the translation of c-MET and PTEN is regulated by Musashi RNA binding proteins 1 and 2 (MSI1; MSI2) [47]. For this reason, MSI1 and MSI2 may be involved in the regulation of ABCC genes. Another study examined the effect of the hedgehog (HH) pathway on the sensitivity of glioma cells to vincristine, depending on the MRP1 gene expression. The inhibition of the HH cascade through suppression of the MRP1 gene enhances the chemotherapy drug cytotoxicity [48]. This indicates the involvement of the HH cascade in the regulation of MDR genes in gliomas. An interesting study investigated the expression of ABCA1, MRP4, and MRP5 in GBM stem cells during differentiation and suggested that differentiation enhances MDR in GBM cells. This hypothesis was confirmed by detecting overexpression of ABC transporters in differentiated GBM cells compared with stem cells [49, 50].

ABCE

Of the ABCE subfamily, ABCE1, an inhibitor of ribonuclease L, may be involved in the development of MDR. This enzyme binds to 5′-phosphorylated 2′,5′-linked oligoadenylates and inhibits the 2-5L A/RNA signaling pathway. The ABCE1 protein, together with eukaryotic translation initiation factors (eIF2, eIF5, and eIF3), purifies the 40S subunits of ribosomes, participating in their biosynthesis and transport from the nucleus [51]. Thus, ABCE1 promotes protein biosynthesis and the development of MDR.

ABCG2

GBM tumors and stem cells often overexpress the ABCG2 gene and breast cancer resistance protein (BCRP), which is involved in the development of tumor resistance to chemotherapy drugs [52]. BCRP is found on the nuclear membrane of LN229 GBM cells, the significance of which is still unclear [14]. It is suggested that ABCG2 overexpression is the result of gene rearrangement or amplification, increasing the resistance of GBM to mitoxantrone, topotecan, irinotecan, epirubicin, camptothecin, daunorubicin, doxorubicin, and anthracyclines [53, 54]. ABCG2 expression is inhibited by sFRP4 and LRIG1 through their suppression of EGFR expression and activation of the eukaryotic translation initiation factor 2 alpha kinase-3 (PERK)/activation transcription factor-4 (ATF4) cascade [31, 43]. Expression of the ABCG2 gene is also inhibited by miR-145 and activated by the transcription factors NRF2 and EZH2 (Figure) [55, 56].

Coexpression of P-gp and BCRP is observed in GBM cells and BBB epithelial cells due to their joint functioning [57]. A correlation has been established between the expression of BCRP proteins and tyrosine kinase receptor-1 with immunoglobulin-like and EGF-like domains (Tie), confirming the association of BCRP with angiogenesis [58]. The suppression of BCRP activity underlies the mechanism of action of many drugs (vinblastine, vincristine, temozolomide, topotecan, irinotecan, mitoxantrone, camptothecins, anthracyclines, elacridar, and celecoxib) [31, 59, 60].

Drug resistance inhibitors in glioblastoma

Many chemotherapy drugs or targeted therapies are used (or studied in clinical trials) to treat GBM. Most of them are inhibitors. According to their mechanism of action, they can be divided into the following: inhibitors of ABC transporters, heat shock proteins, tyrosine kinase receptors, signaling cascade kinases, enzymes, microtubules, proteasomes, transcription factors, and DNA synthesis (Table).

 

Table. Clinical trials of targeted drugs for glioblastoma therapy / Таблица. Клинические испытания таргетных препаратов для терапии глиобластомы

Drug

Target

Trial phase

Reference

AEE788

VEGFR, EGFR

I

61

Aflibercept

VEGF-A, VEGF-B, PLGF

I

62

Bevacizumab

VEGF

BEV + IR, III

63, 64

Vandetanib

VEGFR2, EGFR

I

65

Vatalanib (PTK787)

VEGFR1-3, PDGFRβ, c-kit TKI

I

66

Gefitinib

EGFR

II

67

Golvatinib (E7050)

MET/HGF

I

68

Depatuxizumab (ABT-414, ABT-806)

EGFR

II

69

Cabozantinib (XL-184)

VEGFR2, c-MET

ТМЗ + radiotherapy

II

70

Lapatinib

EGFR, HER-2

II

71

Lenvatinib (E7080)

VEGFR2, VEGFR3, FGFR1, c-kit, PDGFRβ

In vivo

72

Nimotuzumab

nimotuzumab EGFR antibodies in combination with TMZ

III

73

Olaratumab (IMC-3G3)

PDGFRα

II

74

Onartuzumab (MetMAb)

MET/HGF

II

75, 76

Pazopanib (GW786034)

VEGFR1-3, PDGFRα, PDGFRβ, c-Kit TKI

II

77

Panitumumab (ABX–EGF)

EGFR

II

78

Pertuzumab

HER2

FDA approved

79

Ramucirumab (IMC-1121B)

VEGFR2

II

80

Rilotumumab (AMG 102)

MET/HGF

II

81

Rindopepimuth (CDX-110)

EGFRvIII

III

82

Sorafenib

VEGFR2, Raf1, PDGFR, c-Kit, Flt3

I

83

Sunitinib

VEGR2, PDGFRα, PDGFRβ,

c-Kit, Flt3

II

84

Tacrolimus (FK506)

FK506-binding protein 12 (FKBP12)

In vitro

85

Temsirolimus

mTOR

I/II

86

Tivantinib (ARQ197)

MET

In vitro, U251, T98MG, I

87

Tivozanib

VEGR3

II

88

Trastuzumab

HER2

In vivo

89

Ficlatuzumab (AV-299)

MET/HGF

I

90

Cediranib

VEGFR1-3, PDGFRβ, c-kit

II

91

Cetuximab (C225)

EGFR

I

92

Cilengitide

Integrins αvβ3 αvβ5

II/III

93

Everolimus

mTOR

I/II

94, 95

Enzastaurin (LY317615)

PKCβ, Pi3k/Akt/mTOR

III

96

Erlotinib (OSI-774)

EGFR

II

97

Zetakin

IL13Rα2

I

98

125I-MAb

EGFR

II

99

INCB28060 (INC280, capmatinib)

MET/HGF

Ib/II

100

mAb 806 (ABT-806)

ΔEGFR

I

101

MK0752

γ-Secretase

I

102

RO5323441

PLGF

I

103

Tf-CRM107

Transferrin

I

104

scFvM58-sTRAIL

MRP3, TRAIL-R1, TRAIL-R2

MRP3

105

XL765 (SAR245409, voxtalisib)

PI3K/mTOR

In vitro, in vivo

106

Note: GM-CSF, granulocyte-macrophage colony-stimulating factor; HGF, hepatocyte growth factor; FGFR1, fibroblast growth factor receptor; mTOR, mechanistic target of rapamycin kinase; PKCβ, protein kinase Cβ; PLGF, placental growth factor; TMZ, temozolomide.

 

ABC transporter inhibitors

In clinical practice, drugs suppressing the expression of P-gp and BCRP are used or studied in a phase of clinical trials. These include sunitinib malate (SU11248), an inhibitor of platelet-derived growth factor tyrosine kinase receptor (PDGFR), and vascular endothelial growth factor receptor (VEGFR), which inhibits the activity of P-gp and BCRP proteins through ATP hydrolysis [107]. Imatinib inhibits these proteins and PDGFR by suppressing the activity of cytochrome P4503A (CYP3A) [108]. However, imatinib enters nerve cells through the P-gp and BCRP transporters; therefore, the penetration of the drug into cells overexpressing these proteins is significantly restricted. With an increase in the concentration (0.5–50.0 μM) of imatinib, its accumulation in C6 glioma cells proportionally increases [109]. Other inhibitors of P-gp and BCRP include elacridar and pantoprazole, which increase the permeability of GBM cells to imatinib in mice by 1.8-fold and 4.2-fold, respectively [110].

Moreover, statins can be used as inhibitors of ABC transporters, since they stimulate the synthesis of nitric oxide and; therefore, participate in the tyrosine nitration of ABC transporters, reducing their activity [111].

Other P-gp inhibitors include verapamil and cyclosporin A, which inhibit calcium channels [112]. The following drugs are at the clinical trial phase: phosphodiesterase-5 inhibitors fumitremrergin, indolyl diketopiperazine, and Ko143 [(3S,6S,12aS)-1,2,3,4,6,7,12,12a-octahydro-9-methoxy-6- (2 methylpropyl)-1,4-dioxopyrazino[1′,2′:1,6]pyrido[3, 4-b]indole-3-propionic acid 1,1-dimethylethyl ether], which inhibit ABCG2 [113].

A study by Spanish scientists showed that melatonin, a hormone of the pineal gland, stimulates promoter methylation by suppressing the expression of the ABCG2 gene and BCRP protein, thereby synergistically enhancing the effect of TMZ on A172 human GBM cells [113]. Additionally, proteins containing methyl CpG binding domains (MBD2 and MeCP2) can methylate the ABCG2 gene [114]. The activity of the ABCG2 gene also decreases upon histone acetylation [115].

Tyrosine kinase receptor inhibitors

In GBM, EGFR and EGFRvIII are overexpressed, and therefore, their inhibitors (monoclonal antibodies and small molecules) are used for treatment. The monoclonal antibodies group includes cetuximab and panitumumab, while the small molecules group includes gefitinib, erlotinib, and lapatinib (Table) [66, 67]. Erlotinib inhibits the proliferation of stem cells and GBM cells, where EGFR gene amplification or EGFR overexpression is noted [97]. In patients with GBM, who experienced EGFR amplification but not EGFRvIII expression, the use of cetuximab increased progression-free survival and overall survival rates to 3.03 and 5.57 months, respectively, compared with 1.63 and 3.97 months in the group that was not using the drug [116].

Nonreceptor tyrosine kinase inhibitors

In GBM, overexpression of nonreceptor tyrosine kinase proto-oncogenes 1 and 2 (c-Abl, Arg) is often noted, which enhances tumor progression. Imatinib inhibits the expression of ABL1 and ABL2 kinases through the STAT3/HSP27/AKT/NF-kB signaling cascade and the expression of NF-kB target genes, induction of apoptosis, and cessation of cells in the G2/M phase [117].

Vascular endothelial growth factor inhibitors

The monoclonal antibody bevacizumab is a VEGF inhibitor widely used in clinical practice. Its efficiency in treating GBM was confirmed in a sample of 637 patients who had an increase in progression-free survival rates (10.7 months) compared with the placebo group (7.3 months) [118]. A phase II clinical trial was conducted in 22 centers in Germany on 182 patients with GBM. The patients were distributed randomly into two groups and received bevacizumab and irinotecan with radiotherapy or daily TMZ. The trial showed an increase in the six-month progression-free survival rate (p < 0.001) to 79.3% in the group receiving bevacizumab and irinotecan with radiotherapy compared with the group receiving TMZ (42.6%). In absolute terms, the progression-free survival rate was increased from 5.99 to 9.7 months (p < 0.001). However, the overall survival rate did not change and amounted to 16.6 months in the group receiving bevacizumab and irinotecan with radiotherapy and 17.5 months in the group receiving TMZ [63].

Kinase signaling cascade inhibitors

Pi3k/mTOR inhibitors XL765 (SAR245409, voxtalizib), temsirolimus, and tacrolimus (target FKBP12; Table) are currently under clinical trials [85, 86]. Voxtalizib reduces the lactate/pyruvate ratio in U87MG GBM cells, resulting in the inhibition of glycolysis, acidosis, and hypoxia [119]. Tacrolimus forms a complex with the FKBP12 protein, which suppresses the formation of calcineurin and the expression of ABCC1 in T98G cells, increasing their sensitivity to vincristine, etoposide, and taxol [120].

Transcription factor inhibitors

American researchers from Duke University revealed a synergistic cytotoxic effect of JSI-124 inhibitors (STAT3 target) and gefitinib on GBM cells. They increase the sensitivity of glioma cells to TMZ, 1,3-bis(2-chloroethyl) nitrosourea, and cisplatin [121]. Another STAT3 inhibitor, STX-0119, suppresses the expression of mTOR, S6, and protein-1 binding translation initiation factor 4E (4E-BP1) through the regulation of the expression of CHI3L1 chitinase-3-like protein-1 (YKL-40) in U87 GBM cells [122].

Inhibitors BAY117082, parthenolide, and MG132 (targeting NF-kB) have a fundamentally different mechanism of action, which is arresting the U138MG, U87, and U373 GBM cell cycle in the G2/M phase, depolarizing the mitochondrial membranes, releasing cytochrome c, and inhibiting the BCL-xL activity [123]. Tetra-O-methyl-nordihydroguaiaretic acid, a repressor of Survivin and CDK1, which are the target genes of transcription factor SP1, also induces cycle arrest in the G2/M phase and apoptosis of GBM cells, suppressing their proliferation [124].

Histone deacetylase inhibitors

HDAC histone deacetylases are involved in the deacetylation of histones H3, H4, H2A, and H2B of ε-N-acetyl-lysine, thereby changing chromatin conformation and suppressing the gene expression. In contrast, HDAC inhibitors (HDACi), such as sodium valproate, enhance gene transcription [125]. Suberoylanilide hydroxamic acid (SAHA) prevents the acetylation and ubiquitination of nucleolin (NCL) (involved in ribosome biogenesis and RNA maturation) through suppression of the JNK/STAT3 signaling pathway, which inhibits the expression of SOX2, OCT4, BMI1, and CD133 genes in GBM stem cells, inhibiting cell proliferation [126, 127]. At high concentrations (more than 5 μM), SAHA induces the activity of caspases 8 and 9 and p53 protein, which trigger apoptosis in GBM stem cells [127]. Another inhibitor, romidepsin (FK228) stimulates the activity of caspase 3, Bax, and Parp proteins and inhibits the Pi3k/Akt/mTOR cascade and BCL2 protein expression, thereby inducing the programmed death of GBM cells. These events enhance the cytotoxic effect of TMZ on GBM cells [128].

RGFP109 inhibitor suppresses the formation of the NF-kB/p65 complex with the coactivators p300- and p30/CBP-associated factor PCAF, which enhances the expression of the growth inhibitor suppressor gene 4 (ING4) and suppresses the expression of NF-kB target genes that stimulate progression of GBM [129].

Histone lysine demethylase inhibitors

Inhibitors JIB04 and CPI-455 of histone lysine demethylase (KDM) dephosphorylate AKT, triggering the arrest of the G2 phase of the cell cycle, autophagy, and apoptosis of GBM cells [130, 131].

Methyltransferase inhibitors

TMZ is a widely used MGMT inhibitor in GBM chemotherapy. The susceptibility of glioma cells to TMZ increases the miR-198 level, which suppresses MGMT expression [132]. miR-198 methylates the WNT3 promoter, inhibits the Wnt3/Gsk-3β/β-catenin cascade, and binds β-catenin to the ABCB1 promoter, which suppresses its expression. These events enhance the sensitivity of GBM cells to doxorubicin, vinblastine, and topotecan (P-gp substrates) [133]. NCL protein expression is associated with increased sensitivity of gliomas to TMZ [126]; this chemotherapeutic drug induces autophagy in MOGGCCM glioma cells through activation of the expression of Beclin protein (ATG6) and light chain 3α of microtubule-binding protein-1 (LC3II) and decreasing the level of p62 protein [134]. Meanwhile, overexpression of YB-1, AKT3, MELK, EZH2, and MVP/LRP proteins enhances the resistance of GBM cells to TMZ [135]. AKT hyperactivation induces the expression of SPARC/osteonectin proteoglycan-1 containing cwcv- and kazal-like domains (SPOCK1), which enhances GBM invasion and its resistance to TMZ [136]. This is also facilitated by the action of connective tissue growth factor, which activates the TGF-β1/ERK1/2/Smad cascade and overexpression of the SOX2, SOX9, and HOXA9 genes and the mTOR protein [137]. It is suggested that the transcription factor ZEB1 helps GBM cells resist TMZ, as evidenced by data on the regulation of its expression by miR-200 and ROBO1, c-MYB, and MGMT proteins [138].

Recently, histone methyltransferase G9a inhibitors, such as BIX01294, have been used to treat GBM. This compound suppresses the expression of autophagy proteins LC3B, LC3II, protein 1 containing phosphoinositide-interacting WD repeat domain (WIPI1), and differentiation markers, glial fibrillary acidic protein, tubulin III (TUBB3) in GBM cells, which prevent tumor malignancy [139].

Topoisomerase inhibitors

One of the topoisomerase II inhibitors used in the chemotherapy of brain tumors is etoposide which suppresses livin-α mRNA expression in U251 GBM cells and activates apoptosis [140]. Mitoxantrone, which inhibits the activity of BCRP in LN229 GBM cells, also reduces the expression of topoisomerase II. In turn, the BCRP inhibitor fumitremrergin C can enhance the antitumor effect of mitoxantrone [11]. The inhibitory action of topotecan and irinotecan targets topoisomerase I. Irinotecan translocates the transcription factor YB-1 into the nucleus and enhances the effect of trichostatin A, an inhibitor of histone deacetylase [141].

DNA replication and synthesis inhibitors

Chemotherapy protocols for brain tumors may include the platinum chemotherapy drugs carboplatin and cisplatin. The platinum atom of these chemotherapy drugs forms coordination bonds with guanine bases and alkylating adducts of DNA, which prevent its replication and synthesis, initiating the apoptosis of tumor cells. An increase in the sensitivity of, for example, U373 glioma cells to carboplatin can be noted when the expression of the tyrosinase-related protein (TRP2), BCRP, MGMT, P-gp, MRP1, and MRP3 is inhibited [142]. The use of the JAK2/STAT3 inhibitor JSI-124 enhances the efficiency of the cytotoxic effect of cisplatin on GBM cells [121].

Besides the platinum compounds, bis-chloroethyl nitrosourea (BCNU) and hydroxycarbamide have an alkylating effect on DNA. The expression of the proapoptotic protein lipocalin-2 (LCN2) and Clk1 and the dephosphorylation of Akt increase the sensitivity of glioma cells to BCNU [36, 143]. On the contrary, the overexpression of protein mRNA of the guanine nucleotide exchange factor Rex-1 enhances the proliferation and resistance of GBM cells to BCNU by activating the p38MAPK/JNK and Pi3k/Akt/Gsk-3β signaling cascades, ABCG2 expression, and inhibition of apoptosis [144]. Doxorubicin, which induces the formation of free radicals in the membranes of tumor cells, suppresses DNA synthesis. However, during cyclic hypoxia, overexpression of the ABCB1 and ABCG2 genes and hyperactivation of P-gp, BCRP, and HIF-1α enhance the resistance of GBM cells to the chemotherapy drug [145, 146].

Microtubule and proteasome inhibitors

Microtubule and proteasome inhibitors include vincristine and vinblastine which bind to cleavage spindle microtubule tubulin, causing it to rupture and stop tumor cell mitosis. Vincristine at the doses of 10 and 60 nM for one to three days reduces the expression of ABCB1 mRNA in SF188 cells of GBM [147].

In turn, DSF-Cu inhibits the function of proteasomes and DNA repair, thereby potentiating the effect of DNA alkylating chemotherapy drugs and TMZ on brain tumor stem cells [148].

Conclusion

The overexpression of ABC family transporter proteins causes the GBM cell to resist the targeted and chemotherapy drugs. Targeted drugs used to treat GBM or those under clinical trials are inhibitors of ABC transporters (sunitinib malate, imatinib, elacridar, pantoprazole, statins, verapamil, cyclosporine A, ONT-093, XR9576, and flavonoids), phosphodiesterase-5 (fumitremrergin, indolyl diketopiperazine, and OSU-03012), tyrosine kinase EGFR receptors (cetuximab, panitumumab, gefitinib, erlotinib, and lapatinib), vascular endothelial growth factor (bevacizumab and everolimus), signaling pathway proteins (γ-secretase, voxtalisib, and temsirolimus), transcription factors NF-kB (BAY117082, parthenolide, and MG132) and STAT3 (JSI-124 and STX-0119), histone deacetylases (sodium valproate, SAHA, trichostatin A, romidepsin, and RGFP109), methyltransferase (MGMT) (TMZ), O-6-benzylguanine and histone methyltransferase G9a (BIX01294), histone lysine demethylase (JIB04 and CPI-455), topoisomerases I (topotecan and irinotecan) and II (etoposide), DNA replication and synthesis (carboplatin and cisplatin, BCNU, hydroxycarbamide, and doxorubicin), microtubules (vincristine and vinblastine), and proteasome (DSF-Cu). To develop and create new highly selective targeted drugs, it is necessary to search for and identify new selective molecular targets in GBM cells. The following can serve as new targets for the action of targeted drugs on ABC transporters: proteins MSI1 and MSI2, p53, TRAIL, MGMT, GRP78, RAB4, RAB5, p19; receptors frizzled, RET, and EGFR; signal transduction and cell cycle kinases JAK1/2, AMPK, WNT5а, MELK, CLK1, GSK-3β, and CDK6; miRNA miR-130a, miR-145, miR-27a, miR-328, miR-200c; transcription factors NRF2, FOXO3a, YB-1, NF-kB, ATF4, CREB, AP-1, ZEB1, and ZEB2; the translation factors eIF2, eIF5, eIF3, and PERK; genes and oncogenes c-MET, HGF, PTEN, ACCN3, and ACCN4.

×

About the authors

Alexander N. Chernov

Institute of Experimental Medicine

Author for correspondence.
Email: al.chernov@mail.ru
ORCID iD: 0000-0003-2464-7370
Scopus Author ID: 26649406700

Research Associate, Department of General Pathology and Pathological Physiology

Russian Federation, 12, Academician Pavlov Str., Saint Petersburg, 197376

Olga V. Shamova

Institute of Experimental Medicine; Saint Petersburg State University

Email: oshamova@yandex.ru
ORCID iD: 0000-0002-5168-2801
Scopus Author ID: 6603643804
ResearcherId: F-6743-2013

Dr. Sci. (Biol.), Associate Professor, Corresponding Member of the Russian Academy of Sciences, Head of the Department of General Pathology and Pathological Physiology

Russian Federation, 12, Academician Pavlov Str., Saint Petersburg, 197376; Saint Petersburg

References

  1. Hanif F, Muzaffar K, Perveen K, et al. Glioblastoma multiforme: A review of its epidemiology and pathogenesis through clinical presentation and treatment. Asian Pac J Cancer Prev. 2017;18(1):3–9. doi: 10.22034/APJCP.2017.18.1.3
  2. Merabishvili VM. Oncological statistics (traditional methods, new information technologies): Guidelines for physicians. Part I. Saint Petersburg: Costa; 2015. (In Russ.)
  3. Stupp R, Mason WP, van den Bent MJ, et al. Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma. N Engl J Med. 2005;352(10):987–96. doi: 10.1056/NEJMoa043330
  4. Johnson DR, O’Neill BP. Glioblastoma survival in the United States before and during the temozolomide era. J Neurooncol. 2012;107(2):359–364. doi: 10.1007/s11060-011-0749-4
  5. Dréan A, Rosenberg S, Lejeune FX, et al. ATP binding cassette (ABC) transporters: expression and clinical value in glioblastoma. J Neurooncol. 2018;138(3):479–486. doi: 10.1007/s11060-018-2819-3
  6. Demina EP, Miroshnikova VV, Schwarzman AL. Role of the ABC transporters A1 and G1, key reverse cholesterol transport proteins, in atherosclerosis. Mol Biol (Mosk). 2016;50(2):223–230. doi: 10.7868/S002689841602004X
  7. Zolnerciks JK, Andress EJ, Nicolaou M, Linton KJ. Structure of ABC transporters. Essays Biochem. 2011;50(1):43–61. doi: 10.1042/bse0500043
  8. Gomez-Zepeda D, Taghi M, Scherrmann JM. ABC transporters at the blood-brain interfaces, their study models, and drug delivery implications in gliomas. Pharmaceutics. 2019;12(1):20. doi: 10.3390/pharmaceutics12010020
  9. Liu X. ABC Family Transporters. Adv Exp Med Biol. 2019;1141:13–100. doi: 10.1007/978-981-13-7647-4_2
  10. Cascorbi I, Haenisch S. Pharmacogenetics of ATP-binding cassette transporters and clinical implications. Methods Mol Biol. 2010;596:95–121. doi: 10.1007/978-1-60761-416-6_6
  11. Bhatia P, Bernier M, Sanghvi M, et al. Breast cancer resistance protein (BCRP/ABCG2) localises to the nucleus in glioblastoma multiforme cells. Xenobiotica. 2012;42(8):748–755. doi: 10.3109/00498254.2012.662726
  12. Dean M, Rzhetsky A, Allikmets R. The human ATP-binding cassette (ABC) transporter superfamily. Genome Res. 2001;11(7):1156–1166. doi: 10.1101/gr.184901
  13. Hientz K, Mohr A, Bhakta-Guha D, Efferth T. The role of p53 in cancer drug resistance and targeted chemotherapy. Oncotarget. 2017;8(5):8921–8946. doi: 10.18632/oncotarget.13475
  14. Zhang P, de Gooijer MC, Buil LCM, et al. ABCB1 and ABCG2 restrict the brain penetration of a panel of novel EZH2-Inhibitors. Int J Cancer. 2015;137(8):2007–2018. doi: 10.1002/ijc.29566
  15. Colardo M, Segatto M, Di Bartolomeo S. Targeting RTK-PI3K-mTOR axis in gliomas: an update. Int J Mol Sci. 2021;22(9):4899. doi: 10.3390/ijms22094899
  16. Latour M, Her N-G, Kesari S, Nurmemmedov E. WNT Signaling as a therapeutic target for glioblastoma. Int J Mol Sci. 2021;22(16):8428. doi: 10.3390/ijms22168428
  17. Healy FM, Prior IA, MacEwan DJ. The importance of Ras in drug resistance in cancer. Br J Pharmacol. 2021. doi: 10.1111/bph.15420
  18. Avci NG, Ebrahimzadeh-Pustchi S, Akay YM, et al. NF-κB inhibitor with temozolomide results in significant apoptosis in glioblastoma via the NF-κB(p65) and actin cytoskeleton regulatory pathways. Sci Rep. 2020;10(1):13352. doi: 10.1038/s41598-020-70392-5
  19. Xu P, Zhang G, Hou S, Sha LG. MAPK8 mediates resistance to temozolomide and apoptosis of glioblastoma cells through MAPK signaling pathway. Biomed Pharmacother. 2018;106:1419–1427. doi: 10.1016/j.biopha.2018.06.084
  20. Chen X, Hao A, Li X, et al. Activation of JNK and p38 MAPK mediated by ZDHHC17 drives glioblastoma multiforme development and malignant progression. Theranostics. 2020;10(3):998–1015. doi: 10.7150/thno.40076
  21. Lin SP, Lee YT, Wang JY, et al. Survival of cancer stem cells under hypoxia and serum depletion via decrease in PP2A activity and activation of p38-MAPKAPK2-Hsp27. PLoS One. 2012;7(11):e49605. doi: 10.1371/journal.pone.0049605
  22. Ouédraogo ZG, Biau J, Kemeny J-L, et al. Role of STAT3 in genesis and progression of human malignant gliomas. Mol Neurobiol. 2017;54(8):5780–5797. doi: 10.1007/s12035-016-0103-0
  23. Aroui S, Dardevet L, Najlaoui F, et al. PTEN-regulated AKT/FoxO3a/Bim signaling contributes to human cell glioblastoma apoptosis by platinum-maurocalcin conjugate. Int J Biochem Cell Biol. 2016;77(Pt A):15–22. doi: 10.1016/j.biocel.2016.05.013
  24. Medarova Z, Pantazopoulos P, Yoo B. Screening of potential miRNA therapeutics for the prevention of multi-drug resistance in cancer cells. Sci Rep. 2020;10(1):1970. doi: 10.1038/s41598-020-58919-2
  25. Zhang HD, Jiang LH, Sun DW, et al. The role of miR-130a in cancer. Breast Cancer. 2017;24(4):521–527. doi: 10.1007/s12282-017-0776-x
  26. Sui H, Cai GX, Pan SF, et al. miR200c attenuates P-gp-mediated MDR and metastasis by targeting JNK2/c-Jun signaling pathway in colorectal cancer. Mol Cancer Ther. 2014;13(12):3137–3151. doi: 10.1158/1535-7163.MCT-14-0167
  27. Li Z, Zhang J, Zheng H, et al. Modulating lncRNA SNHG15/CDK6/miR-627 circuit by palbociclib, overcomes temozolomide resistance and reduces M2-polarization of glioma associated microglia in glioblastoma multiforme. J Exp Clin Cancer Res. 2019;38(1):380. doi: 10.1186/s13046-019-1371-0
  28. Tursynbay Y, Zhang J, Li Z, et al. Pim-1 kinase as cancer drug target: An update. Biomed Rep. 2016;4(2):140–146. doi: 10.3892/br.2015.561
  29. Katayama K, Noguchi K, Sugimoto Y. Regulations of P-Glycoprotein/ABCB1/MDR1 in human cancer cells. N J Sci. 2014(2):1–10. doi: 10.1155/2014/476974
  30. Oberstadt MC, Bien-Möller S, Weitmann K, et al. Epigenetic modulation of the drug resistance genes MGMT, ABCB1 and ABCG2 in glioblastoma multiforme. BMC Cancer. 2013;13:617. doi: 10.1186/1471-2407-13-617
  31. Liu B, Guo Z, Dong H, et al. LRIG1, human EGFR inhibitor, reverses multidrug resistance through modulation of ABCB1 and ABCG2. Brain Res. 2015;1611:93–100. doi: 10.1016/j.brainres.2015.03.023
  32. Xi G, Best B, Mania-Farnell B, et al. therapeutic potential for bone morphogenetic protein 4 in human malignant glioma. Neoplasia. 2017;19(4):261–270. doi: 10.1016/j.neo.2017.01.006
  33. Zhang X, Ding K, Wang J, et al. Chemoresistance caused by the microenvironment of glioblastoma and the corresponding solutions. Biomed Pharmacother. 2019;109:39–46. doi: 10.1016/j.biopha.2018.10.063
  34. Said HM, Hagemann C, Carta F, et al. Hypoxia induced CA9 inhibitory targeting by two different sulfonamide derivatives including acetazolamide in human glioblastoma. Bioorg Med Chem. 2013;21(13):3949–3957. doi: 10.1016/j.bmc.2013.03.068
  35. Pölönen P, Jawahar Deen A, Leinonen HM, et al. Nrf2 and SQSTM1/p62 jointly contribute to mesenchymal transition and invasion in glioblastoma. Oncogene. 2019;38(50):7473–7490. doi: 10.1038/s41388-019-0956-6
  36. Zhang L, Yang H, Zhang W, et al. Clk1-regulated aerobic glycolysis is involved in gliomas chemoresistance. J Neurochem. 2017;142(4):574–588. doi: 10.1111/jnc.14096
  37. Tivnan A, Zakaria Z, O’Leary C, et al. Inhibition of multidrug resistance protein 1 (MRP1) improves chemotherapy drug response in primary and recurrent glioblastoma multiforme. Front Neurosci. 2015;9:218. doi: 10.3389/fnins.2015.00218
  38. Begicevic R-R, Falasca M. ABC transporters in cancer stem cells: beyond chemoresistance. Int J Mol Sci. 2017;18(11):2362. doi: 10.3390/ijms18112362
  39. Johnson ZL, Chen J. Structural basis of substrate recognition by the multidrug resistance protein MRP1. Cell. 2017;168(6):1075–1085.e9. doi: 10.1016/j.cell.2017.01.041
  40. Zhang YK, Wang YJ, Gupta P, Chen ZS. Multidrug resistance proteins (MRPs) and cancer therapy. AAPS J. 2015;17(4):802–812. doi: 10.1208/s12248-015-9757-1
  41. Pattabiraman PP, Pecen PE, Rao PV. MRP4-mediated regulation of intracellular cAMP and cGMP levels in trabecular meshwork cells and homeostasis of intraocular pressure. Invest Ophthalmol Vis Sci. 2013;54(3):1636–1649. doi: 10.1167/iovs.12-11107
  42. Mao X, He Z, Zhou F, et al. Prognostic significance and molecular mechanisms of adenosine triphosphate-binding cassette subfamily C members in gastric cancer. Medicine (Baltimore). 2019;98(50):e18347. doi: 10.1097/MD.0000000000018347
  43. Bhuvanalakshmi G, Arfuso F, Millward M, et al. Secreted frizzled-related protein 4 inhibits glioma stem-like cells by reversing epithelial to mesenchymal transition, inducing apoptosis and decreasing cancer stem cell properties. PLoS One. 2015;10(6):e0127517. doi: 10.1371/journal.pone.0127517
  44. Kosalai ST, Abdelrazak Morsy MH, Papakonstantinou N, et al. EZH2 upregulates the PI3K/AKT pathway through IGF1R and MYC in clinically aggressive chronic lymphocytic leukaemia. Epigenetics. 2019;14(11):1125–1140. doi: 10.1080/15592294.2019.1633867
  45. Navarro L, Gil-Benso R, Megías J, et al. Alteration of major vault protein in human glioblastoma and its relation with EGFR and PTEN status. Neuroscience. 2015;297:243–251. doi: 10.1016/j.neuroscience.2015.04.005
  46. Guo G, Narayan RN, Horton L, et al. The Role of EGFR-Met interactions in the pathogenesis of glioblastoma and resistance to treatment. Curr Cancer Drug Targets. 2017;17(3):297–302. doi: 10.2174/1568009616666161215162515
  47. Kudinov AE, Karanicolas J, Golemis EA, Boumber Y. Musashi RNA-Binding proteins as cancer drivers and novel therapeutic targets. Clin Cancer Res. 2017;23(9):2143–2153. doi: 10.1158/1078-0432.CCR-16-2728
  48. Shahi MH, Farheen S, Mariyath MP, Castresana JS. Potential role of Shh-Gli1-BMI1 signaling pathway nexus in glioma chemoresistance. Tumour Biol. 2016;37(11):15107–15114. doi: 10.1007/s13277-016-5365-7
  49. Rama AR, Alvarez PJ, Madeddu R, Aranega A. ABC transporters as differentiation markers in glioblastoma cells. Mol Biol Rep. 2014;41(8):4847–4851. doi: 10.1007/s11033-014-3423-z
  50. Uribe D, Torres Á, Rocha JD, et al. Multidrug resistance in glioblastoma stem-like cells: Role of the hypoxic microenvironment and adenosine signaling. Mol Aspects Med. 2017;55:140–151. doi: 10.1016/j.mam.2017.01.009
  51. Navarro-Quiles C, Mateo-Bonmatí E, Micol JL. ABCE proteins: from molecules to development. Front Plant Sci. 2018;9:1125. doi: 10.3389/fpls.2018.01125
  52. Chen L, Shi L, Wang W, Zhou Y. ABCG2 downregulation in glioma stem cells enhances the therapeutic efficacy of demethoxycurcumin. Oncotarget. 2017;8(26):43237–43247. doi: 10.18632/oncotarget.18018
  53. Nakanishi T, Ross D. Breast cancer resistance protein (BCRP/ABCG2): its role in multidrug resistance and regulation of its gene expression. Chin J Cancer. 2012;31(2):73–99. doi: 10.5732/cjc.011.10320
  54. Goncalves J, Bicker J, Alves G, et al. Relevance of breast cancer resistance protein to brain distribution and central acting drugs: A pharmacokinetic perspective. Curr Drug Metab. 2018;19(12):1021–1041. doi: 10.2174/1389200219666180629121033
  55. Shi L, Wang Z, Sun G, et al. miR-145 inhibits migration and invasion of glioma stem cells by targeting ABCG2. Neuromolecular Med. 2014;16(2):517–528. doi: 10.1007/s12017-014-8305-y
  56. Tian S, Yong M, Zhu J, et al. Enhancement of the effect of Methyl Pyropheophorbide-a-Mediated photodynamic therapy was achieved by increasing ROS through inhibition of Nrf2-HO-1 or Nrf2-ABCG2 signaling. Anticancer Agents Med Chem. 2017;17(13):1824–1836. doi: 10.2174/1871520617666170327145857
  57. Agarwal S, Hartz AMS, Elmquist WF, Bauer B. Breast cancer resistance protein and P-glycoprotein in brain cancer: Two gatekeepers team up. Curr Pharm Des. 2011;17(26):2793–2802. doi: 10.2174/138161211797440186
  58. Martin V, Xu J, Pabbisetty SK, et al. Tie2-mediated multidrug resistance in malignant gliomas is associated with upregulation of ABC transporters. Oncogene. 2009;28(24):2358–2363. doi: 10.1038/onc.2009.103
  59. Jin Y, Bin ZQ, Qiang H, et al. ABCG2 is related with the grade of glioma and resistance to mitoxantone, a chemotherapeutic drug for glioma. J Cancer Res Clin Oncol. 2009;135(10):1369–1376. doi: 10.1007/s00432-009-0578-4
  60. Wijaya J, Fukuda Y, Schuetz JD. Obstacles to brain tumor therapy: Key ABC transporters. Int J Mol Sci. 2017;18(12):2544. doi: 10.3390/ijms18122544
  61. Reardon DA, Conrad CA, Cloughesy T, et al. Phase I study of AEE788, a novel multitarget inhibitor of ErbB- and VEGF-receptor-family tyrosine kinases, in recurrent glioblastoma patients. Cancer Chemother Pharmacol. 2012;69(6):1507–1518. doi: 10.1007/s00280-012-1854-6
  62. Nayak L, de Groot J, Wefel JS, et al. Phase I trial of aflibercept (VEGF trap) with radiation therapy and concomitant and adjuvant temozolomide in patients with high-grade gliomas. J Neurooncol. 2017;132(1):181–188. doi: 10.1007/s11060-016-2357-9
  63. Herrlinger U, Schäfer N, Steinbach JP, et al. bevacizumab plus irinotecan versus temozolomide in newly diagnosed O6-Methylguanine-DNA methyltransferase nonmethylated glioblastoma: The Randomized GLARIUS Trial. J Clin Oncol. 2016;34(14):1611–1619. doi: 10.1200/JCO.2015.63.4691
  64. Lu-Emerson C, Duda DG, Emblem KE, et al. Lessons from anti-vascular endothelial growth factor and anti-vascular endothelial growth factor receptor trials in patients with glioblastoma. J Clin Oncol. 2015;33(10):1197–1213. doi: 10.1200/JCO.2014.55.9575
  65. Chheda MG, Wen PY, Hochberg FH, et al. Vandetanib plus sirolimus in adults with recurrent glioblastoma: results of a phase I and dose expansion cohort study. J Neurooncol. 2015;121(3):627–634. doi: 10.1007/s11060-014-1680-2
  66. Pearson J, Regad T. Targeting cellular pathways in glioblastoma multiforme. Sig Transduct Target Ther. 2017;2:17040. doi: 10.1038/sigtrans.2017.40
  67. Arif SH, Pandith AA, Tabasum R, et al. Significant effect of anti-tyrosine kinase inhibitor (gefitinib) on overall survival of the glioblastoma multiforme patients in the backdrop of mutational status of epidermal growth factor receptor and PTEN genes. Asian J Neurosurg. 2018;13(1):46–52. doi: 10.4103/ajns.AJNS_95_17
  68. Molife LR, Dean EJ, Blanco-Codesido M, et al. A phase I, dose-escalation study of the multitargeted receptor tyrosine kinase inhibitor, golvatinib, in patients with advanced solid tumors. Clin Cancer Res. 2014;20(24):6284–6294. doi: 10.1158/1078-0432.CCR-14-0409
  69. Padovan M, Eoli M, Pellerino A, et al. Depatuxizumab mafodotin (Depatux-M) plus temozolomide in recurrent glioblastoma patients: Real-world experience from a multicenter study of Italian Association of Neuro-Oncology (AINO). Cancers (Basel). 2021;13(11):2773. doi: 10.3390/cancers13112773
  70. Wen PY, Drappatz J, de Groot J, et al. Phase II study of cabozantinib in patients with progressive glioblastoma: subset analysis of patients naive to antiangiogenic therapy. Neuro Oncol. 2018;20(2):249–258. doi: 10.1093/neuonc/nox154
  71. Yu A, Faiq N, Green S, et al. Report of safety of pulse dosing of lapatinib with temozolomide and radiation therapy for newly-diagnosed glioblastoma in a pilot phase II study. J Neurooncol. 2017;134(2):357–362. doi: 10.1007/s11060-017-2533-6
  72. Li J, Zou C-L, Zhang Z-M, et al. A multi-targeted tyrosine kinase inhibitor lenvatinib for the treatment of mice with advanced glioblastoma. Mol Med Rep. 2017;16(5):7105–7111. doi: 10.3892/mmr.2017.7456
  73. Westphal M, Heese O, Steinbach JP, et al. A randomised, open label phase III trial with nimotuzumab, an anti-epidermal growth factor receptor monoclonal antibody in the treatment of newly diagnosed adult glioblastoma. Eur J Cancer. 2015;51(4):522–532. doi: 10.1016/j.ejca.2014.12.019
  74. Olaratumab Completed Phase 2 Trials for Glioblastoma Multiforme, Adult Treatment. NCT00895180. https://go.drugbank.com/drugs/DB06043/clinical_trials?conditions=DBCOND0088047&phase=2&purpose=treatment&status=completed&__cf_chl_jschl_tk__=pmd_cEbWshGFMGoqes6n_B7D0tXCsfTuqIh_en6wF52PFuw-1629969594-0-gqNtZGzNApCjcnBszQiR
  75. Morley R, Cardenas A, Hawkins P, et al. Safety of onartuzumab in patients with solid tumors: Experience to date from the Onartuzumab Clinical Trial Program. PLoS One. 2015;10(10):e0139679. doi: 10.1371/journal.pone.0139679
  76. Cloughesy T, Finocchiaro G, Belda-Iniesta C, et al. Randomized, double-blind, placebo-controlled, multicenter phase II study of onartuzumab plus bevacizumab versus placebo plus bevacizumab in patients with recurrent glioblastoma: efficacy, safety, and hepatocyte growth factor and O6-Methylguanine-DNA methyltransferase biomarker analyses. J Clin Oncol. 2017;35(3):343–351. doi: 10.1200/JCO.2015.64.7685
  77. Pazopanib Completed Phase 2 Trials for Glioblastoma Multiforme (GBM) / Central Nervous System Neoplasms / Neoplasms, Brain / Gliosarcoma Treatment. NCT01931098 [Internet]. Available from: https://go.drugbank.com/drugs/DB06589/clinical_trials?conditions=DBCOND0032525%2CDBCOND0046976%2CDBCOND0002894%2CDBCOND0054211&phase=2&purpose=treatment&status=completed. Accessed: Dec 18, 2021.
  78. Panitumumab and Irinotecan for Malignant Gliomas [Internet]. Available from: https://clinicaltrials.gov/ct2/show/NCT01017653. Accessed: Dec 18, 2021.
  79. Dean L, Kane M, Pratt VM, et al. Pertuzumab Therapy and ERBB2 Genotype. In: Medical Genetics Summaries [Internet]. Bethesda (MD): National Center for Biotechnology Information (US); 2012–2015.
  80. Adult Glioblastoma Multiforme Completed Phase 2 Trials for Ramucirumab (DB05578). NCT00895180 [Internet]. Available from: https://go.drugbank.com/indications/DBCOND0088047/clinical_trials/DB05578?phase=2&status=completed. Accessed: Dec 18, 2021.
  81. Affronti ML, Jackman JG, McSherry F, et al. Phase II study to evaluate the efficacy and safety of rilotumumab and bevacizumab in subjects with recurrent malignant glioma. Oncologist. 2018;23(8):889–e98. doi: 10.1634/theoncologist.2018-0149
  82. Weller M, Butowski N, Tran DD, et al. Rindopepimut with temozolomide for patients with newly diagnosed, EGFRvIII-expressing glioblastoma (ACT IV): a randomised, double-blind, international phase 3 trial. Lancet Oncol. 2017;18(10):1373–1385. doi: 10.1016/S1470-2045(17)30517-X
  83. Nghiemphu PL, Ebiana VA, Wen P, et al. Phase I study of sorafenib and tipifarnib for recurrent glioblastoma: NABTC 05-02. J Neurooncol. 2018;136(1):79–86. doi: 10.1007/s11060-017-2624-4
  84. Grisanti S, Ferrari VD, Buglione M, et al. Second line treatment of recurrent glioblastoma with sunitinib: results of a phase II study and systematic review of literature. J Neurosurg Sci. 2019;63(4):458–467. doi: 10.23736/S0390-5616.16.03874-1
  85. Torres Á, Arriagada V, Erices JI, et al. FK506 Attenuates the MRP1-mediated chemoresistant phenotype in glioblastoma stem-like cells. Int J Mol Sci. 2018;19(9):2697. doi: 10.3390/ijms19092697
  86. Schiff D, Jaeckle KA, Anderson SK, et al. Phase I/II trial of temsirolimus and sorafenib in treatment of patients with recurrent glioblastoma: North Central Cancer Treatment Group Study/Alliance N0572. Cancer. 2018;124(7):1455–1463. doi: 10.1002/cncr.31219
  87. Wu Y, Li Z, Zhang L, Liu G. Tivantinib hampers the proliferation of glioblastoma cells via PI3K/Akt/Mammalian target of rapamycin (mTOR) signaling. Med Sci Monit. 2019;25:7383–7390. doi: 10.12659/MSM.919319
  88. Kalpathy-Cramer J, Chandra V, Da X, et al. Phase II study of tivozanib, an oral VEGFR inhibitor, in patients with recurrent glioblastoma. J Neurooncol. 2017;131(3):603–610. doi: 10.1007/s11060-016-2332-5
  89. Askoxylakis V, Ferraro GB, Kodack DP, et al. Preclinical efficacy of Ado-trastuzumab Emtansine in the brain microenvironment. J Natl Cancer Inst. 2015;108(2):djv313. doi: 10.1093/jnci/djv313
  90. Bauman JE, Ohr J, Gooding WE, et al. Phase I study of ficlatuzumab and cetuximab in cetuximab-resistant, recurrent/metastatic head and neck cancer. Cancers (Basel). 2020;12(6):1537. doi: 10.3390/cancers12061537
  91. Brown N, McBain C, Nash S, et al. Multi-center randomized phase II study comparing cediranib plus gefitinib with cediranib plus placebo in subjects with recurrent/progressive glioblastoma. PLoS One. 2016;11(5):e0156369. doi: 10.1371/journal.pone.0156369
  92. Super-Selective Intraarterial Cerebral Infusion of Cetuximab (Erbitux) for Treatment of Relapsed/Refractory GBM and AA. NCT01238237 [Internet]. Available from: https://clinicaltrials.gov/ct2/show/NCT01238237. Accessed: Dec 18, 2021.
  93. Stupp R, Hegi ME, Gorlia T, et al. Cilengitide combined with standard treatment for patients with newly diagnosed glioblastoma with methylated MGMT promoter (CENTRIC EORTC 26071-22072 study): a multicentre, randomised, open-label, phase 3 trial. Lancet Oncol. 2014;15(10):1100–1108. doi: 10.1016/S1470-2045(14)70379-1
  94. Miklja Z, Yadav VN, Cartaxo RT, et al. Everolimus improves the efficacy of dasatinib in PDGFRα-driven glioma. J Clin Invest. 2020;130(10):5313–5325. doi: 10.1172/JCI133310
  95. Chinnaiyan P, Won M, Wen PY, et al. A randomized phase II study of everolimus in combination with chemoradiation in newly diagnosed glioblastoma: results of NRG Oncology RTOG 0913. Neuro Oncol. 2018;20(5):666–673. doi: 10.1093/neuonc/nox209
  96. Tucker N. Enzastaurin Dosed in first phase 3 Study of newly diagnosed glioblastoma multiforme [Internet]. Available from: https://www.targetedonc.com/view/enzastaurin-dosed-in-first-phase-3-study-of-newly-diagnosed-glioblastoma-multiforme. Accessed: Dec 18, 2021.
  97. Erlotinib in treating patients with recurrent or progressive glioblastoma multiforme. NCT00054496 [Internet]. Available from: https://clinicaltrials.gov/ct2/show/NCT00054496. Accessed: Dec 18, 2021.
  98. Brown CE, Badie B, Barish ME, et al. Bioactivity and safety of IL13Rα2-Redirected Chimeric Antigen Receptor CD8+ T Cells in patients with recurrent glioblastoma. Clin Cancer Res. 2015;21(18):4062–4072. doi: 10.1158/1078-0432.CCR-15-0428
  99. Li L, Quang TS, Gracely EJ, et al. A Phase II study of anti-epidermal growth factor receptor radioimmunotherapy in the treatment of glioblastoma multiforme. J Neurosurg. 2010;113(2):192–198. doi: 10.3171/2010.2.JNS091211
  100. Van den Bent M, Azaro A, De Vos F, et al. A Phase Ib/II, open-label, multicenter study of INC280 (capmatinib) alone and in combination with buparlisib (BKM120) in adult patients with recurrent glioblastoma. J Neurooncol. 2020;146(1):79–89. doi: 10.1007/s11060-019-03337-2
  101. Cleary JM, Reardon DA, Azad N, et al. A phase 1 study of ABT-806 in subjects with advanced solid tumors. Invest New Drugs. 2015;33(3):671–678. doi: 10.1007/s10637-015-0234-6
  102. Hoffman LM, Fouladi M, Olson J, et al. Phase I trial of weekly MK-0752 in children with refractory central nervous system malignancies: A Pediatric Brain Tumor Consortium Study. Childs Nerv Syst. 2015;31(8):1283–1289. doi: 10.1007/s00381-015-2725-3
  103. Lassen U, Chinot OL, McBain C, et al. Phase 1 dose-escalation study of the antiplacental growth factor monoclonal antibody RO5323441 combined with bevacizumab in patients with recurrent glioblastoma. Neuro Oncol. 2015;17(7):1007–1015. doi: 10.1093/neuonc/nov019
  104. Tortorella S, Karagiannis TC. Transferrin receptor-mediated endocytosis: a useful target for cancer therapy. J Membr Biol. 2014;247(4):291–307. doi: 10.1007/s00232-014-9637-0
  105. Yaylim I, Azam S, Farooqi AA, et al. Critical molecular and genetic markers in primary brain tumors with their clinical importance. In: Neurooncology – Newer Developments. Chapter 6. IntechOpen; 2016. doi: 10.5772/63550
  106. Zhao H, Chen G, Liang H, Dual PI3K/mTOR Inhibitor, XL765, suppresses glioblastoma growth by inducing ER stress-dependent apoptosis. Onco Targets Ther. 2019;12:5415–5424. doi: 10.2147/OTT.S210128
  107. Shukla S, Robey RW, Bates SE, Ambudkar SV. Sunitinib (Sutent, SU11248), a small-molecule receptor tyrosine kinase inhibitor, blocks function of the ATP-binding cassette (ABC) transporters P-glycoprotein (ABCB1) and ABCG2. Drug Metab Dispos. 2009;37(2):359–365. doi: 10.1124/dmd.108.024612
  108. Englund G, Lundquist P, Skogastierna C, et al. Cytochrome p450 inhibitory properties of common efflux transporter inhibitors. Drug Metab Dispos. 2014;42(3):441–447. doi: 10.1124/dmd.113.054932
  109. Declèves X, Bihorel S, Debray M, et al. ABC transporters and the accumulation of imatinib and its active metabolite CGP74588 in rat C6 glioma cells. Pharmacol Res. 2008;57(3):214–222. doi: 10.1016/j.phrs.2008.01.006
  110. Eadie LN, Hughes TP, White DL. Interaction of the efflux transporters ABCB1 and ABCG2 with imatinib, nilotinib, and dasatinib. Clin Pharmacol Ther. 2014;95(3):294–306. doi: 10.1038/clpt.2013.208
  111. Pun NT, Jeong C-H. Statin as a potential chemotherapeutic agent: current updates as a monotherapy, combination therapy, and treatment for anti-cancer drug resistance. Pharmaceuticals (Basel). 2021;14(5):470. doi: 10.3390/ph14050470
  112. Nguyen TT, Duong VA, Maeng HJ. Pharmaceutical formulations with P-Glycoprotein inhibitory effect as promising approaches for enhancing oral drug absorption and bioavailability. Pharmaceutics. 2021;13(7):1103. doi: 10.3390/pharmaceutics13071103
  113. Toyoda Y, Takada T, Suzuki H. Inhibitors of human ABCG2: from technical background to recent updates with clinical implications. Front Pharmacol. 2019;10:208. doi: 10.3389/fphar.2019.00208
  114. Martín V, Sanchez-Sanchez AM, Herrera F, et al. Melatonin-induced methylation of the ABCG2/BCRP promoter as a novel mechanism to overcome multidrug resistance in brain tumour stem cells. Br J Cancer. 2013;108(10):2005–2012. doi: 10.1038/bjc.2013.188
  115. You D, Richardson JR, Aleksunes LM. Epigenetic regulation of multidrug resistance protein 1 and breast cancer resistance protein transporters by histone deacetylase inhibition. Drug Metab Dispos. 2020;48(6):459–480. doi: 10.1124/dmd.119.089953
  116. Lv S, Teugels E, Sadones J, et al. Correlation of EGFR, IDH1 and PTEN status with the outcome of patients with recurrent glioblastoma treated in a phase II clinical trial with the EGFR-blocking monoclonal antibody cetuximab. Int J Oncol. 2012;41(3):1029–1035. doi: 10.3892/ijo.2012.1539
  117. Lamballe F, Toscano S, Conti F, et al. Coordination of signalling networks and tumorigenic properties by ABL in glioblastoma cells. Oncotarget. 2016;7(46):74747–74767. doi: 10.18632/oncotarget.12546
  118. Gilbert MR, Dignam JJ, Armstrong TS, et al. A randomized trial of bevacizumab for newly diagnosed glioblastoma. N Engl J Med. 2014;370(8):699–708. doi: 10.1056/NEJMoa1308573
  119. Radoul M, Chaumeil MM, Eriksson P, et al. MR studies of glioblastoma models treated with dual PI3K/mTOR inhibitor and temozolomide: metabolic changes are associated with enhanced survival. Mol Cancer Ther. 2016;15(5):1113–1122. doi: 10.1158/1535-7163.MCT-15-0769
  120. Garrido W, Muñoz M, San Martín R, Quezada C. FK506 confers chemosensitivity to anticancer drugs in glioblastoma multiforme cells by decreasing the expression of the multiple resistance-associated protein-1. Biochem Biophys Res Commun. 2011;411(1):62–68. doi: 10.1016/j.bbrc.2011.06.087
  121. Lo H-W, Cao X, Zhu H, Ali-Osman F. Constitutively activated STAT3 frequently coexpresses with epidermal growth factor receptor in high-grade gliomas and targeting STAT3 sensitizes them to iressa and alkylators. Clin Cancer Res. 2008;14(19):6042–6054. doi: 10.1158/1078-0432.CCR-07-4923
  122. Miyata H, Ashizawa T, Iizuka A, et al. Combination of a STAT3 inhibitor and an mTOR inhibitor against a temozolomide-resistant glioblastoma cell line. Cancer Genomics Proteomics. 2017;14(1):83–91. doi: 10.21873/cgp.20021
  123. Zanotto-Filho A, Braganhol E, Schröder R, et al. NF-κB inhibitors induce cell death in glioblastomas. Biochem Pharmacol. 2011;81(3):412–424. doi: 10.1016/j.bcp.2010.10.014
  124. Castro-Gamero AM, Borges KS, Moreno DA, et al. Tetra-O-methyl nordihydroguaiaretic acid, an inhibitor of Sp1-mediated survivin transcription, induces apoptosis and acts synergistically with chemo-radiotherapy in glioblastoma cells. Invest New Drugs. 2013;31(4):858–870. doi: 10.1007/s10637-012-9917-4
  125. Chen R, Zhang M, Zhou Y, et al. The application of histone deacetylases inhibitors in glioblastoma. J Exp Clin Cancer Res. 2020;39(1):138. doi: 10.1186/s13046-020-01643-6
  126. Ko CY, Lin CH, Chuang JY, et al. MDM2 degrades deacetylated nucleolin through ubiquitination to promote glioma stem-like cell enrichment for chemotherapeutic resistance. Mol Neurobiol. 2018;55(4):3211–3223. doi: 10.1007/s12035-017-0569-4
  127. Hsu CC, Chang WC, Hsu TI, et al. Suberoylanilide hydroxamic acid represses glioma stem-like cells. J Biomed Sci. 2016;23(1):81. doi: 10.1186/s12929-016-0296-6
  128. Wu Y, Dong L, Bao S, et al. FK228 augmented temozolomide sensitivity in human glioma cells by blocking PI3K/AKT/mTOR signal pathways. Biomed Pharmacother. 2016;84:462–469. doi: 10.1016/j.biopha.2016.09.051
  129. Li ZY, Li QZ, Chen L, et al. Histone deacetylase inhibitor RGFP109 overcomes temozolomide resistance by blocking NF-κB-dependent transcription in glioblastoma cell lines. Neurochem Res. 2016;41(12):3192–3205. doi: 10.1007/s11064-016-2043-5
  130. Banelli B, Daga A, Forlani A, et al. Small molecules targeting histone demethylase genes (KDMs) inhibit growth of temozolomide-resistant glioblastoma cells. Oncotarget. 2017;8(21):34896–34910. doi: 10.18632/oncotarget.16820
  131. Romani M, Daga A, Forlani A, et al. Targeting of histone demethylases KDM5A and KDM6B inhibits the proliferation of temozolomide-resistant glioblastoma cells. Cancers (basel). 2019;11(6):878. doi: 10.3390/cancers11060878
  132. Nie E, Jin X, Wu W, et al. MiR-198 enhances temozolomide sensitivity in glioblastoma by targeting MGMT. J Neurooncol. 2017;133(1):59–68. doi: 10.1007/s11060-017-2425-9
  133. Riganti C, Salaroglio IC, Pinzòn-Daza ML, et al. Temozolomide down-regulates P-glycoprotein in human blood-brain barrier cells by disrupting Wnt3 signaling. Cell Mol Life Sci. 2014;71(3):499–516. doi: 10.1007/s00018-013-1397-y
  134. Jakubowicz-Gil J, Bądziul D, Langner E, et al. Temozolomide and sorafenib as programmed cell death inducers of human glioma cells. Pharmacol Rep. 2017;69(4):779–787. doi: 10.1016/j.pharep.2017.03.008
  135. Stavrovskaya AA, Shushanov SS, Rybalkina EY. Problems of glioblastoma multiforme drug resistance. Biochemistry (Mosc). 2016;81(2):91–100. doi: 10.1134/S0006297916020036
  136. Yu F, Li G, Gao J, et al. SPOCK1 is upregulated in recurrent glioblastoma and contributes to metastasis and temozolomide resistance. Cell Prolif. 2016;49(2):195–206. doi: 10.1111/cpr.12241
  137. Garros-Regulez L, Aldaz P, Arrizabalaga O, et al. mTOR inhibition decreases SOX2-SOX9 mediated glioma stem cell activity and temozolomide resistance. Expert Opin Ther Targets. 2016;20(4):393–405. doi: 10.1517/14728222.2016.1151002
  138. Siebzehnrubl FA, Silver DJ, Tugertimur B, et al. The ZEB1 pathway links glioblastoma initiation, invasion and chemoresistance. EMBO Mol Med. 2013;5(8):1196–1212. doi: 10.1002/emmm.201302827
  139. Ciechomska IA, Przanowski P, Jackl J, et al. BIX01294, an inhibitor of histone methyltransferase, induces autophagy-dependent differentiation of glioma stem-like cells. Sci Rep. 2016;6:38723. doi: 10.1038/srep38723
  140. Jin F, Zhao L, Guo Y-J, et al. Influence of Etoposide on anti-apoptotic and multidrug resistance-associated protein genes in CD133 positive U251 glioblastoma stem-like cells. Brain Res. 2010;1336:103–111. doi: 10.1016/j.brainres.2010.04.005
  141. Bieler A, Mantwill K, Dravits T, et al. Novel three-pronged strategy to enhance cancer cell killing in glioblastoma cell lines: histone deacetylase inhibitor, chemotherapy, and oncolytic adenovirus dl520. Hum Gene Ther. 2006;17(1):55–70. doi: 10.1089/hum.2006.17.55
  142. Liu G, Akasaki Y, Khong HT, et al. Cytotoxic T cell targeting of TRP-2 sensitizes human malignant glioma to chemotherapy. Oncogene. 2005;24(33):5226–5234. doi: 10.1038/sj.onc.1208519
  143. Zheng LT, Lee S, Yin GN, et al. Down-regulation of lipocalin 2 contributes to chemoresistance in glioblastoma cells. J Neurochem. 2009;111(5):1238–1251. doi: 10.1111/j.1471-4159.2009.06410.x
  144. Kim BS, Kang KS, Choi JI, et al. Knockdown of the potential cancer stem-like cell marker Rex-1 improves chemotherapeutic effects in gliomas. Hum Gene Ther. 2011;22(12):1551–1562. doi: 10.1089/hum.2011.096
  145. Chou CW, Wang CC, Wu CP, et al. Tumor cycling hypoxia induces chemoresistance in glioblastoma multiforme by upregulating the expression and function of ABCB1. Neuro Oncol. 2012;14(10):1227–1238. doi: 10.1093/neuonc/nos195
  146. Pinzón-Daza M, Garzón R, Couraud P, et al. The association of statins plus LDL receptor-targeted liposome-encapsulated doxorubicin increases in vitro drug delivery across blood-brain barrier cells. Br J Pharmacol. 2012;167(7):1431–1447. doi: 10.1111/j.1476-5381.2012.02103.x
  147. Valera ET, de Freitas Cortez MA, de Paula Queiroz RG, et al. Pediatric glioblastoma cell line shows different patterns of expression of transmembrane ABC transporters after in vitro exposure to vinblastine. Childs Nerv Syst. 2009;25(1):39–45. doi: 10.1007/s00381-008-0740-3
  148. Lun X, Wells JC, Grinshtein N, et al. Disulfiram when combined with copper enhances the therapeutic effects of temozolomide for the treatment of glioblastoma. Clin Cancer Res. 2016;22(15):3860–3875. doi: 10.1158/1078-0432.CCR-15-1798

Supplementary files

Supplementary Files
Action
1. JATS XML
2. Figure. Intracellular mechanisms of multidrug resistance of glioblastoma involving genes ABCB1 and ABCG2. See text for explanations

Download (376KB)

Copyright (c) 2021 Chernov A.N., Shamova O.V.

Creative Commons License
This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License.

СМИ зарегистрировано Федеральной службой по надзору в сфере связи, информационных технологий и массовых коммуникаций (Роскомнадзор).
Регистрационный номер и дата принятия решения о регистрации СМИ: серия ПИ № ФС 77 - 74760 от 29.12.2018 г.


This website uses cookies

You consent to our cookies if you continue to use our website.

About Cookies