Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Acta Pharm Sin B. 2017 Jan; 7(1): 38–51.
Published online 2016 Dec 3. doi: 10.1016/j.apsb.2016.09.002
PMCID: PMC5237711
PMID: 28119807

Regulation of multidrug resistance by microRNAs in anti-cancer therapy

Xin An,a,b Cesar Sarmiento,b Tao Tan,b, and Hua Zhub,

Abstract

Multidrug resistance (MDR) remains a major clinical obstacle to successful cancer treatment. Although diverse mechanisms of MDR have been well elucidated, such as dysregulation of drugs transporters, defects of apoptosis and autophagy machinery, alterations of drug metabolism and drug targets, disrupti on of redox homeostasis, the exact mechanisms of MDR in a specific cancer patient and the cross-talk among these different mechanisms and how they are regulated are poorly understood. MicroRNAs (miRNAs) are a new class of small noncoding RNAs that could control the global activity of the cell by post-transcriptionally regulating a large variety of target genes and proteins expression. Accumulating evidence shows that miRNAs play a key regulatory role in MDR through modulating various drug resistant mechanisms mentioned above, thereby holding much promise for developing novel and more effective individualized therapies for cancer treatment. This review summarizes the various MDR mechanisms and mainly focuses on the role of miRNAs in regulating MDR in cancer treatment.

KEY WORDS: Multidrug resistance, miRNA, Cancer, Therapy, Autophagy, Redox Homeostasis

Graphical abstract

In this review article, we summarized known mechanisms for multidrug resistance (MDR) in cancer cells and the role of microRNAs in regulating these mechanisms. We hope this review can shed light on our understanding of MDR and developing effective means for treatment of MDR in cancers.An external file that holds a picture, illustration, etc.
Object name is fx1.gif

1. Introduction

Although the use of chemotherapeutic agents has substantially improved the anti-tumor efficacy during the last decades, the development of multidrug resistance (MDR) remains the largest obstacle to the success of cancer chemotherapies. MDR is defined as the resistance of cancer cells to a diverse panel of structurally and functionally unassociated drugs1. This resistance can occur naturally (inherent resistance) or acquired during the course of chemotherapy or upon recurrence after successful chemotherapy2, 3.

The development of MDR is a complicated and multifactorial process. During the last few decades, diverse mechanisms have been implicated in the development of both intrinsic and acquired MDR. The main mechanisms include: (1) Overexpression of MDR transporters; (2) Defects in the apoptotic machinery; (3) Induction of autophagy; (4) Alteration of drug metabolism; (5) Alteration in drug targets and DNA repair; and (6) Disruption of redox homeostasis. However, the exact mechanisms of MDR in a specific cancer patient and the specific biomarker to define it, the cross-talk among these different mechanisms and how they are regulated are largely unknown.

Recently, numerous studies have demonstrated that microRNAs (miRNAs) play key regulatory roles in MDR through modulating many of the biological processes mentioned above. Therefore, miRNAs could be potential biomarkers and (or) targets for circumventing MDR in cancer chemotherapy. This review summarizes the various MDR mechanisms and focuses on the roles of miRNAs in regulating MDR in cancer treatment.

2. Mechanisms of multidrug resistance

2.1. Overexpression of MDR transporters

Overexpression of MDR transporters is one of the most important causes of chemoresistance4, 5. The ABC transporter family members are the most widely studied MDR transporters6, 7. These ABC transporter proteins have similar trans-membrane domains that can pump the chemotherapeutic drugs out of cancer cells against a concentration gradient in an ATP energy-dependent manner, thus reducing intracellular accumulation of chemotherapeutic agents, and protecting cancer cells from toxicity. To date, 48 ABC transporters have been detected in the human body8, among which the most extensively characterized MDR transporters include P-glycoprotein (P-gp/ABCB1), multidrug resistance associated protein-1 (MRP1/ABCC1), and breast cancer resistant proteins (BCRP/ABCG2)9, 10, 11.

The overexpression of ABCB1 has been shown to be associated with a large variety of chemotherapeutic drugs including anthracyclines, epipodophyllotoxins, vinca alkaloids, and taxanes4, 7, 12, 13. Overexpression of the ABCC1 transporter also confers resistance to a wide range of anticancer drugs, such as anthracyclines, vinca alkaloids, epipodophyllotoxins, camptothecins, methotrexate, and mitoxantrone12, 14, 15. The substrates of ABCG2 include tyrosine kinase inhibitors (TKIs), anthracyclines, camptothecin-derived topoiso-merase I inhibitors, methotrexate and flavopiridols16, 17 (Fig. 1).

Fig. 1

The mechanisms underlying the development of multidrug resistance in cancers. Anti-drug resistance can occur at many levels, including dysregulation of drugs transporters which might lead to increased drug efflux and (or) decreased drug intake, defects in cell cycle and the apoptotic machinery, induction of autophagy (see Fig. 2), alteration of drug metabolism and target, as well as disruption of redox homeostasis.

2.2. Defects in cell-cycle and the apoptotic machinery

After DNA damage is induced by anti-cancer drugs, the injured cancer cells can react in two ways: either by cell cycle arrest and damage repair, or by apoptosis and cell death if the DNA damage is too extensive to repair. The tumor suppressor protein (P53) plays a vital role during this process18. The effect of P53 on drug resistance has been studied extensively. Mutant P53, which often causes the loss of P53 function and MDR has been reported in many cancers including acute lymphoblastic leukemia, melanoma, osteosarcoma, breast, ovarian, and testicular cancers19, 20, 21.

Apoptosis is the major type of cell death triggered by chemotherapy drugs. There are two established pathways of apoptosis: the intrinsic mitochondrial pathway and the extrinsic transmembrane pathway. The intrinsic pathway is mainly under the control of the BCL-2 family, which includes both pro-apoptotic proteins (BAX, BAK, BID, BIM, BAD, and PUMA) and anti-apoptotic proteins (BCL-2, BCL-XL, and MCL-1)22, whereas the extrinsic pathway is regulated mainly by “death receptors” of the tumor necrosis factor (TNF) receptor family. Various anti-cancer drugs, such as antimetabolites, DNA cross-linking and intercalating agents, alkylating agents, topoisomerase I/II inhibitors and TKIs, have been reported to induce the intrinsic or/and the extrinsic apoptotic responses in tumor cells, resulting in caspsase activation23. Defects in these apoptotic machineries have been described to play an important role in cancer cell drug resistance24. Cancer cells can escape apoptosis either by overexpression of anti-apoptotic proteins or underexpression of pro-apoptotic proteins. Anti-apoptotic protein BCL-2 overexpression is the most common mechanism of apoptosis evasion, which has been reported be involved in the resistance of cells in a variety of drugs, including doxorubicin, paclitaxel, etoposide, camptothecin, mitoxantrone and cisplatin25, 26, 27. Several other factors, including aberrant activation of protein kinase B, nuclear factor kappa B (NF-κB), phosphatase and tensin homolog (PTEN) have also been demonstrated to play an important role in developing drug resistance in various cancer types through interfering apoptosis machinery28, 29, 30 (Fig. 1).

2.3. Induction of autophagy

Autophagy is a relative new mechanism of anti-cancer drugs resistance reported in many recent studies. It is an evolutionarily conserved catabolic process characterized by cellular self-digestion and the removal of excessive, long-lived or dysfunctional organelles and proteins via endosome and lysosome fusion, which results in the formation of autophagosomes31 (Fig. 2). Three main subsets of autophagy with different cellular functions and means by which targets are delivered to lysosomes have been identified: macroautophagy, microautophagy, and chaperone-mediated autophagy. Among the three forms, macroautophagy is the most commonly studied32.

Fig. 2

Key phases involved in the process of autophagy. Cellular stress such as chemotherapy can activate the autophagy pathway through several phases, including induction (formation of a pre-autophagosomal structure leading to an isolation membrane), vesicle nucleation (capturing and delivering cytoplasmic material to lysosomes for digestion), elongation/completion (elongating of the lipid membrane to enclose the target cargo, and completing the formation of an autophagosome), docking/fusing with the lysosome (forming a mature autolysosome), and cargo degradation (undergoing hydrolysis to degrade the vesicle׳s contents and completing macroautophagy).

Autophagy can occur as a physiological process in normal cells to eliminate damaged organelles and recycle macromolecules, thus assuring cellular homeostasis and protecting against cancer. In established tumor cells, autophagy can serve as a means of temporary survival in response to metabolic stress, such as anticancer drugs, that might mediate resistance to anticancer therapies. On the other hand, once the cellular stress is continuous and evolves to progressive autophagy, cell death ensues. This kind of autophagic cell death is a form of physiological cell death which is contradictory to type I programmed cell death (apoptosis). The double sided functions of autophagy implicate its paradoxical roles in anticancer treatments, increasing or diminishing their anticancer activity. However, an increasing amount of evidence suggests that autophagy׳s pro-survival function plays a significant role in chemoresistance in a many different cancer types33, 34, 35, 36, 37, 38.

Chemotherapeutic drugs can induce both apoptosis and autophagy. Autophagy helps cancer cells evade apoptosis and therefore contributes to chemoresistance. For example, in response to 5-fluorouracil (5-FU) and cisplatin, chemosensitive cell lines exhibited apoptosis, whereas chemoresistant populations exhibited autophagy. Generally, cancer cells that respond to drugs by inducing autophagy are more drug-resistant39. Therefore, targeting autophagy would probably be a promising therapeutic strategy to overcome antidrug resistance37.

A number of molecular mechanisms have been shown to be implicated in autophagy-mediated chemoresistance. These include the EGFR signaling pathway40, the aberrant expression of phosphatidylinositol 3-kinase/mammalian target of rapamycin (PI3K/mTOR) pathway41, vascular endothelial growth factor (VEGF)42, mitogen activated protein kinase 14 (MAPK14)/p38a signaling43, 44, as well as the tumor-suppressor gene P53 pathway43.

2.4. Alternation of anti-cancer drug metabolism

Cancer cells can acquire resistance to a specific drug by altering drug metabolism. The super family of cytochrome P450 (CYP) enzymes play a critical role in this process.

The CYP enzymes are most expressed in human liver, intestine, and kidney. These enzymes are involved in the metabolism of a variety of chemotherapy drugs, including taxanes45, 46, vinblastine45, 46, vincristine46, doxorubicin46, etoposide46, irinotecan47, cyclophosphamide48, ifosphamide48. Many factors, such as genetic polymorphisms, alterations in physiological conditions, disease status, intake of certain drugs or foods, or smoking can affect CYP activities. Such changes can alter pharmacokinetic profiles, and therefore the efficacy or toxicity of anticancer drugs. Genetic polymorphisms in CYPs sometimes result in reduced enzyme activity causing low metabolic clearance of drugs or low production of active metabolites46. The well-known example is the influence of CYP2D6 polymorphism on tamoxifen efficacy through the formation of endoxifen, which is an active metabolite of tamoxifen49 (Fig. 1).

2.5. Alteration in drug targets and DNA repair

Chemoresistance can be caused by either quantitative or qualitative alterations of the drug targets. For example, expression levels of thymidylate synthase (TS), a key enzyme and target of 5-FU, and dihydropyrimidine dehydrogenase (DPD), the rate-limiting enzyme in metabolism of 5-FU, can predict 5-FU sensitivity50. Another example is ribonucleotide reductase subunit 2 (RRM2) which is an important cellular target of gemcitabine, plays an important role in gemcitabine resistance51.

DNA topoisomerase II (Top II) is an essential nuclear enzyme that plays a critical role in DNA replication. Chemotherapy drugs, such as doxorubicin, idarubicin, mitoxantrone and etoposide, exert their anticancer function by targeting DNA-Topo II complexes, thereby leading to the DNA breakage and cancer cells death. Topo II–induced resistance to these drugs has been documented in many studies52, 53, 54.

Enhanced DNA damage repair efficiency also plays a role in the development of MDR in cancer cells. This is specifically evident in the case of platinum agents and alkylating compounds, which exert their action via directly damaging of DNA55. There are three fundamental pathways to repair damaged DNA: nuclear excision repair (NER), base excision repair (BER) and DNA mismatch repair (MMR). Dysregulation of theses repair systems may be involved in chemoresistance56 (Fig. 3). For example, hereditary nonpolyposis colorectal cancer (HNPCC) is strongly associated with specific mutations in the MMR pathway, which have been associated with reduced or absent benefit from 5-FU adjuvant chemotherapy57. Since these MMR alterations reduce the incorporation of the 5-FU metabolites into DNA, they decrease 5-FU-induced G2/M arrest apoptosis of cancer cells. In contrast, BRCA1/2 mutated breast and ovarian cancer, which exhibit homologous recombination–mediated repair deficiency, show increased sensitive to DNA-damaging chemotherapy drugs, such as platinum (Fig. 1).

Fig. 3

Biogenesis of microRNA and their functions. RNA polymerase II/III transcribes miRNA gene and generates a long primary transcript (pri-miRNA) ranging from 100 to 1000 nucleotides in length. The pri-miRNA consists of a hairpin stem, a terminal loop, and two single-stranded regions upstream and one downstream of the stem. The pri-miRNA is then processed by a RNase III endonuclease called Drosha into the precursor miRNA (pre-miRNA) which contains a hairpin structure of close to 70 nucleotides. The precise site of miRNA cleavage is determined by the DiGeorge Critical Region gene 8 protein (DGCR8) which forms a complex with Drosha. Pre-miRNA then leaves the nucleus by means of Exportin-5, a Ran-GTP dependant cytoplasmic transporter which can recognize a two-nucleotide overhang at the 3ʹ end of the RNA, and transports it to the cytoplasm. In the cytoplasm, the RNA is further processed by a second RNase III endonuclease, Dicer, into a miRNA:miRNA/duplex of approximately 19–24 nucleotides in length. One strand is selected to function as a mature miRNA and loaded into the RNA-induced silencing complex (RISC). Whereas the other miRNA/strand is degraded. The mature miRNA leads to translational repression or target mRNA.

2.6. Disruption of redox homeostasis

Disruption of redox homeostasis is another important resistance mechanism of anti-cancer drugs. Normal cells are capable of maintaining a balance between cellular oxidants and antioxidants, which is called redox homeostasis; whereas cancer cells usually exhibit higher levels of reactive oxygen species (ROS), which can promote tumor progression and development. However, extremely high ROS will lead to cancer cell death. A variety of drugs exert their anti-cancer function (at least partly) through increasing ROS production, such as cisplatin58, alkylating agents (Adriamycin and temozolomide)59, 60 and paclitaxel61. Nonetheless, some tumor cells can overcome drug-induced oxidative stress by enhancing their antioxidant systems, including heme oxygenase 1 (HMOX1), superoxide dismutase 1 (SOD1) and glutathione (GSH)62. Therefore, a new redox balance at a higher level of ROS accumulation and stronger antioxidant systems is established, which is called ‘Redox Resetting’. Redox resetting has been shown to be implicated in drug resistance by interfering with other mechanisms, including elevated drug efflux, dysregulated apoptosis and autophagy, and altered drug metabolism drug targets63.

Among antioxidant systems, GSH is the widely antioxidant agent reported to be involved in anti-cancer drugs resistance. Increased levels of GSH lead to chemotherapeutic drug resistance in numerous cancers64. GSH-dependent enzymes also have been implicated in MDR of cancer cells. One example is γ-glutamyltransferase (GGT), a key enzyme of GSH metabolism, which is able to sustain the ‘GSH cycling’ and maintain a high intracellular GSH level by metabolizing extracellular GSH and continuously supplying cysteine for the intracellular GSH re-synthesis. High GGT expression and/or increased GGT activity results in increased resistance to a number of anti-cancer drugs65. Another enzyme is glutathione S-transferases (GST) which can catalyze the conjugation of glutathione to chemical toxins. GST functions as the major detoxification mechanism in human body that can suppress oxidative stress and maintain normal cellular redox homeostasis. Many studies have shown that GST is relevant to the development of resistance to chemotherapy agents in different cancers66, 67, 68, 69. Many drug-resistant cancers express high levels of GST. GST polymorphisms may affect drug metabolism and influence chemotherapy and cancer survival70, 71, 72. Some novel anticancer agents targeting GSTs are in development both in preclinical and clinical stages73, 74.

3. miRNAs

miRNAs are a family of small single-stranded non-coding RNAs of 20–25 nucleotide length that are broadly conserved across species. Most miRNA loci are found in non-coding intronic transcription regions, but some are located in exonic regions75. Therefore, miRNAs do not encode any proteins. The main function of miRNA is regulating protein-coding gene expression post-transcriptionally by directly base pairing between the 5ʹ seed region of a miRNA and the 3ʹ untranslated region (3ʹUTR) of multiple target messenger RNA (mRNA), resulting in translational repression or mRNA degradation and lacking the ability to encode proteins75. It is of note that miRNA is not always associated with inhibitory or down regulatory effects. In rare circumstances, dependent on cell cycle and co-factors, miRNA can activate mRNA translation, thus up regulating protein levels76.

The first miRNA molecule was identified in 1993 by Lee and collaborators77. To date, around 2600 unique mature human miRNAs have been discovered and there are certainly more to come (miRBase version 20)78. miRNAs have been shown to have vast effects on gene translation. More than 50% of all human gene translation is regulated by miRNA. Moreover, each miRNA can regulate numerous target genes, and, vice versa, the same target gene can be regulated by several types of miRNAs, creating a complex network79, 80, 81. The inherent complexity of this regulatory system allows miRNAs to control the global activity of the cell, including cell differentiation, proliferation, stress response, metabolism, cell cycle, apoptosis, and angiogenesis. Therefore, miRNAs may be involved in a broad range of human diseases, including cancer82.

Deregulated miRNAs may cause up- or down-regulation of the miRNAs of interest, thus affecting the function of multiple target mRNAs, altering the expression of multiple proteins that are involved in cancer development, metastasis, angiogenesis and drug resistance83, 84, 85, 86. miRNAs have also been shown to be potential biomarkers for early diagnosis and prognosis prediction in various cancers86. The present review summarizes the current knowledge on the role of miRNAs in anticancer drugs resistance.

4. Aberrant expression of miRNAs and cancer drug resistance

Different miRNA expression profiles between cancerous cells and paired normal tissues from the same organ and cancer types have been documented in a number of recent studies87, 88. Furthermore, significant changes in miRNA expression profiles were observed in drug-resistant cancer cells in comparison with parental drug-sensitive cancer cells89. The dysregulation of miRNAs expression profiles in cancer cells can lead to anti-cancer drugs resistance by abnormally modulating the expression of genes involved MDR mechanisms of action, such as ABC transporters genes, apoptosis and autophagy relates genes, drug metabolism genes, and redox systems related genes. These miRNAs could regulate MDR through targeting a specific gene or cellular signaling pathway, or simultaneously targeting several genes or cellular signaling pathways. Evidence pointing to the role of miRNAs in determining drug sensitivity and MDR is emerging90, 91. Below, we focus on the MDR regulatory role of miRNA stratified by different MDR mechanisms.

4.1. miRNAs regulate MDR transporters

4.1.1. ABCB1/MDR1

Numerous studies have shown that miRNAs can modulate chemotherapy drug resistance through regulating the expression of ABC membrane transporters. P-gp, one of the most important MDR transporters, is responsible for the resistance to a large range of chemotherapy drugs. Overexpression of P-gp results from activation of the ABCB1/MDR1 gene. Our laboratory demonstrated for the first time that both miR-451 and miR-27a were up-regulated in MDR cancer cell lines and caused a high level of P-gp92. Similar results were observed by Li et al.93 in drug-resistant ovarian cancer cells. In contrast, other studies showed conflicting results. Kocalchuk et al.94 reported the negative regulating role of miR-451 on P-gp expression in resistance of the MCF-7 breast cancer cells. Subsequently, similar phenomena were observed separately both in leukemia and hepatocellular carcinoma cell lines95, 96, suggesting the cell lines and environment may regulate miRNA functions.

A number of other miRNAs have been found to have a MDR modulation role by regulation ABCB1 expression. Zhao et al.97 reported that up-regulation of miR-138 could significantly down-regulate the expression of P-gp and reverse adriamycin resistance on the MDR leukemia (HL-60/VCR) cell line. Bao et al.98 found that miR-298 could decrease P-gp expression in a dose-dependent manner by directing bound to P-gp 3ʹUTR and reverse doxorubicinresistance in breast cancer cells. miR-381 and miR-495 were also shown to be inversely associated with the expression of the MDR1 gene and development of MDR99. Yang et al.100 reported that miR-223 could down-regulate ABCB1 at both mRNA and protein levels and increase the HCC cell sensitivity to anti-cancer drugs. Other ABCB1-modulating miRNAs reported include miR-9101, miR-122102, miR-873103.

Recently, high-throughput functional screening was used to identify additional MDR-related miRNAs. One miRNA found by this method is miR-508-5p. Overexpression of miR-508-5p was sufficient to reverse gastric cancer cell resistance to multiple chemotherapeutics in vitro and sensitize tumors to chemotherapy in vivo104. The most recent update study showed gastric cancer cells with up-regulated both miR-27b and miR-508-5p were more sensitive to chemotherapy. The two miRNAs have synergic effect and can form the miR-27b/CCNG1/P53/miR-508-5p axis which plays an important role in GC-associated MDR105 (Table 1).

Table 1

Roles of miRNA on regulation of drug resistance in Cancers.

MiRNA functionmiR(s)Target of miR(s)Effect(s)Ref.
Regulation of MDR transportersmiR-451ABCB1/MDR1Downregulates P-gp in cancer cells94, 95, 96
miR-27amiR-451ABCB1/MDR1Upregulate P-gp in MDR cancer cells92, 93
miR-138ABCB1/MDR1Down regulates P-gp and reverses adriamycinresistance on the MDR cell line in leukemia.97
miR-298ABCB1/MDR1Decreases P-gp expression and reversedoxorubicin resistance in breast cancer cells98
miR-381miR-495ABCB1/MDR1Negatively regulate MDR1 gene99
miR-223ABCB1/MDR1Down-regulates ABCB1 mRNA and protein levels andincreases the HCC cell sensitivity to anti-cancer drugs100
miR-9miR-122miR-122miR-508-5pABCB1/MDR1Mediate MDR in cancer cells by targeting ABCB1.101, 102, 103, 104, 105
miR-328ABCG2/BCRPDown-regulates BCRP and increases the sensitivity tomitoxantrone in breast cancer cells107, 108
miR-519miR-520(h)miR-212miR-181amiR-487aABCG2/BCRPNegatively regulate ABCG2 expression109, 110, 111, 112
miR-326ABCC1/MRP1Down-regulates MRP-1 expression and sensitizes cancercells to VP-16 and doxorubicin113
Hsa-MiR-1291ABCC1/MRP1Down-regulates ABCC1 expression and sensitizes cellsto doxorubicin.114
miR-125bmiR-140P53Suppress p53-dependent apoptosis and inducechemoresistance119, 120
miR-122P53Increases p53 protein stability and contribute tochemosensitivity.121
miR-34aCDK6Induces apoptosis and inhibits cell growth123
miR-15b miR-16miR-21BCL2Upregulate of BCL-2 protein expression126, 127, 128
miR-497miR-200bc/429miR-1915miR-214miR-195miR-205BCL2Directly target BCL-2129, 130, 131, 132, 133, 134
miR-574-3pBCL-XLModulates the anti-apoptotic protein BCL-XL135
miR-101MCL-1Sensitizes hepatocellular carcinoma cells todoxorubicin-induced apoptosis136
miR-494BIMDown-regulates the BIM137
miR-365BAXDown-regulates BAX expression and inducesgemcitabine resistance138
miR-30 b/cmiR-21Caspase-3 PDCD4Impair TRAIL dependent apoptosis140
miR-21miR-22miR-221miR-214miR-19a/bmiRNA-17-5pmiR-222PTENTarget PTEN and/or its downstream kinase141, 142, 143, 144, 145, 146, 147, 148


Induction of autophagymiR-30aBeclin 1 and ATG5Activates beclin 1–related autophagy and confersanti-cancer drugs resistance149, 150, 151, 152, 153
miR-30dBeclinInhibits beclin 1–mediated autophagy154
miR-155miR-15amiR-16Induce autophagy and enhance chemosensitivity158, 160
miR-200bmiR-181aATG12ATG5Suppress autophagy159, 161


Modulation anti-cancer drug metabolismmiR-27bCYP1B1Negatively regulates CYP1B1 expression162, 163
miR-892aCYP1A1Sensitizes cancer cells to a broad spectrum of anticancerdrugs.165
let-7bCYP2J2166
miR-148aCYP3A4Downregulates the expression of CYP3A4.167


Modulation of drug targetsmiR-192miR-215TS enzymeInfluences 5-Fu sensitivity168
miR-27amiR-27bmiR-134miR-582-5pDPD enzymeModulate the sensitivity of 5-Fu–based basedchemotherapy169
miR-211let-7RRM2Regulate RRM2 expression and sensitize PDAC cells togemcitabine170, 171
miR-21MMR proteinsDownregulates hMSH2, hMSH6 expression and reduces5-Fu sensitivity172
miR-155MMR proteinsNegatively regulates MSH2, MSH6 and MLH1-PMS2expression173
miR-182miR-9BRCA1Down-regulate BRCA1 expression and increased thesensitivity of cancer cells to cisplatin and PARPinhibitors174, 175


Regulation GSH and GSH-depended enzymesmiRNA-27aGSHModulates GSH biosynthesis178
miR-513a-3pGSTNegatively regulates GSTP1 gene expression179
miR-133bGSTReduces GST expression, and inverses chemotherapyresistance180

—Not known.

ATG12, autophagy-associated gene 12; ATG5, autophagy-associated gene 5; CDK6, cyclin-dependent kinase 6; DDP, dihydropyrimidine dehydrogenase; GSH, glutathione; GST, glutathione S-transferases; MDR, multidrug resistance; MMR, mismatch repair; PDCD4, pro-apoptotic factors programmed cell death 4; PTEN, phosphatase and tensin homolog; RRM2, ribonucleotide reductase subunit 2; TS, thymidylate synthase.

4.1.2. ABCG2/BCRP

ABCG2/BCRP is the first MDR transporter reported to be regulated by miRNA106. To date, several miRNAs have been identified to regulate ABCG2 expression. Overexpression of miR-328 down-regulated BCRP in breast cancer cells, thus increasing their sensitivity to mitoxantrone107, 108. Other miRNAs showed similarly negative regulatory role on ABCG2 expression, including miR-519, miR-520(h), miR-212, MiR-181a, and MiR-487a109, 110, 111, 112.

4.1.3. ABCC1/MRP1

ABCC1/MRP1 is up-regulated in VP-16-resistant breast cancer cells (MCF-7/VP). Two miRNAs are reported to down-regulate ABCC1 expression and to reverse ABCC1-related MDR. Liang et al.113 found MiR-326 was significantly down-regulated in a MCF-7/VP cell line compared to its parental cell line, while up-regulating miR-326 level in the mimics-transfected VP-16-resistant cell line could down-regulate MRP-1 expression and sensitize these cells to VP-16 and doxorubicin. Pan et al.114 found hsa-miR-1291 could directly down-regulate ABCC1 expression and sensitize the cancer cells to doxorubicin.

miRNAs could also regulate MDR by targeting other members of the ABC transporter family. For example, miR-23a enhances 5-FU resistance in microsatellite instability (MSI) CRC cells through targeting ABCF1115. miR-let-7g/i (let-7g/i) inhibits ABCC10 expression and enhances cellular sensitivity to DDP in human esophageal carcinoma (EC) cell lines116.

Sometimes miRNAs exert their MDR modulating function by directly targeting several MDR related proteins at the same time. For example, over-expressed miR-129-5p can reverse chemo-resistance by simultaneously targeting three members of ABC transporters (ABCB1, ABCC5 and ABCG1)117. Down-regulating miR-106a reversed MDR in human glioma cells by decreasing the expression of P-gp, MDR1, MRP1, as well as the expression of other apoptosis, survival, and inflammatory related proteins118.

4.2. miRNAs regulate cell cycle and the apoptotic machinery

The tumor suppressor protein P53 is a critical mediator of cell cycle and apoptosis in response to different chemotherapeutic drugs. Several miRNAs are involved in the regulation of P53. Iida et al.119 reported that upregulating miR-125b could suppress P53-dependent apoptosis and induce chemoresistance to doxorubicine, vincristine, etoposide and mafosfamide in Ewing sarcoma/primitive neuroectodermal tumor (EWS) cells. Liang and colleagues120 demonstrated that overexpression of miR-140 could interfere with the growth and invasion of pancreatic duct adenocarcinoma cells by directly targeting inhibitor of apoptosis-stimulating protein of P53 (iASPP). MiR-122/cyclin G1 interaction was demonstrated to positively regulated P53 protein stability and transcriptional activity and affected doxorubicin sensitivity of human hepatocarcinoma cells121. On the other hand, the P53 protein itself could regulate certain miRNAs to induce cell cycle arrest and apoptosis. MiR-34a is one of P53 effector genes. The expression of miR-34a shows a strong linear correlation with wild-type P53 expression122. Overexpression of miR-34a can inhibit cell growth, induce apoptosis by targeting cyclin-dependent kinase 6 (CDK6)123.

Several recent studies have showed that introduction of synthetic miR-34a mimics was able to induce cell death in P53-mutated medulloblastoma and glioblastoma cell lines124. “Restoration of P53/miR-34a regulatory axis decreases survival advantage and ensures BAX-dependent apoptosis of non-small cell lung carcinoma cells”125.

BCL-2 is the most important anti-apoptosis protein. Quite a number of miRNAs have been shown to modulate MDR by targeting BCL-2. Xia and colleagues126 found that in MDR gastric cancer cells significant downregulation of miR-15b and miR-16 was concurrent with the upregulation of BCL-2 protein expression, whereas upregulation of miR-15b or miR-16 dramatically reduced BCL-2 protein level and sensitized the cells to anti-cancer drugs. Another study by Cittelly et al.127 found that downregulation of miR-15a/16 mediated BCL-2 activation and promoted tamoxifen resistance in breast cancer. Dong et al.׳s study128 found that miR-21 was involved in gemcitabine resistance by directly upregulating BCL-2 expression in pancreatic cancer cells. Other miRNAs found to directly target BCL-2 using Western blot and luciferase activity assays include miR-497129, miR-200bc/429 cluster130, miR-1915131, miR-214132 miR-195133, and miR-205134.

Furthermore, miRNAs can exert the apoptosis-regulating function by target other members of BCL-2 family proteins. For example, the anti-apoptotic protein BCL-XL can be modulated by miR-574-3p135. miRNA-101 can directly targeting MCL-1 and sensitizes hepatocellular carcinoma cells to doxorubicin-induced apoptosis136. On the other hand, miRNAs could also target some pro-apoptotic BCL-2 family proteins to modulate apoptosis. For instance, miR-494 could induce TNF-related apoptosis-inducing ligand (TRAIL) resistance in non-small cell lung cancer (NSCLC) through the down-modulation of BIM137. Up-regulation of miR-365 can induce gemcitabine resistance by directly down-regulating apoptosis-promoting protein BAX expression138.

In addition, several miRNAs have been shown to regulate extrinsic apoptotic pathways. Quintavalle and colleague139 reported that increased levels of miR-30 b/c and miR-21 in TRAIL resistant glioma cells could impair TRAIL dependent apoptosis by inhibiting the expression of caspase-3 and TAp63. Up-regulation of miR-21 could promote the resistance of nasopharyngeal carcinoma cells to cisplatin by suppressing the pro-apoptotic factors programmed cell death 4 (PDCD4) and FAS ligand (FAS-L)140.

PTEN is another important tumor suppressor gene correlated to chemotherapeutic response. Different miRNAs have been shown to regulate tumor cell chemoresistance by targeting PTEN and/or its downstream kinase, including miR-21141, 142, miR-22143, miR-221144,miR-214145,miR-19a/b146,miRNA-17-5p147andmiR-222148 (Table 1).

4.3. miRNAs regulate autophagy

Induction of autophagy is another important mechanism of anti-cancer drug resistance which can be modulated by miRNAs. The entire process of autophagy, including autophagic induction, vesicle nucleation, vesicle elongation and completion can be modulated by different miRNAs. Our laboratory first reported the regulatory role of miRNAs on autophagy in 2009149. Since then, a growing body of evidence indicates that miRNAs can regulate autophagy related genes to modulate anti-cancer drug resistance. However, the precise roles of miRNAs in the autophagy pathways have not yet been well elucidated.

miR-30a is the first microRNAs reported by our group to suppress stress-induced autophagy through inhibition of beclin 1 expression149. Beclin is an essential protein in autophagy. miR-30a can inhibit autophagy via suppression of beclin 1 and ATG5 (another key autophagy promoting protein). Thus, upregulation of miR-30a can sensitize chronic myelogenous leukemia (CML) cells to imatinib treatment. Targeting miR-30a promotes autophagy in response to imatinib treatment and enhances imatinib activity against CML150, 151. Dysregulation of “miRNA-30a activating beclin-1 related autophagy” is also found to be contributed to chemoresistance of osteosarcoma cells152, as well as resistance to sorafenib in renal cell carcinoma cells153.

miR-30d is another member of the miR-30 family identified by our group which acts similarly to miR-30a in regulating autophagy. miR-30d can directly target the binding sequences in the 3′UTR of beclin 1, affecting the expression of this key autophagy-promoting protein. Moreover, we found that inhibition of the beclin 1–mediated autophagy by the miR-30d mimics sensitized anaplastic thyroid carcinoma cells to cisplatin both in vitro (cell culture) and in vivo (animal xenograft model)154.

It is noticeable that both autophagy and apoptosis function in cell growth, survival, development, and death. Therefore, these two pathways might have cross-talk, and be regulated by the same miRNA. For example, miR-204 shows both an anti-apoptosis effect and autophagy inhibitory effect155, 156. Up or down regulation of miR-204 may change the transition of apoptosis and autophagy; subsequently, influence chemosensitivity. The most recent study showed that combined overexpression of miR-16 and miR-17 can suppress the expression of beclin 1 and Bcl-2, and, in turn, inhibit autophagy and promote apoptosis. Thus, upregulation of miR-16 and miR-17 can dramatically sensitize paclitaxel-resistant lung cancer cells to paclitaxel treatment157. There are still a number of miRNAs reported to regulate anti-cancer drugs sensitivity by targeting autophagy. Examples include miR-155 mediated drug resistance in osteosarcoma cells via induction of autophagy158, miR-200b regulated autophagy associated with chemoresistance in human lung adenocarcinoma159. In addition, miR-15a and miR-16 induced autophagy and enhanced chemosensitivity of camptothecin160, and miR-181a suppressed autophagy and sensitized gastric cancer cells to cisplatin161 (Table 1).

4.4. miRNAs control anti-cancer drug metabolism

miRNAs may regulate the superfamily of P450 (CYP) metabolic enzymes, thereby modulating patterns of drug metabolism, including those of anti-cancer drugs. Accumulating evidence suggests that miRNAs may modulate MDA through regulation of CYP enzymes. For example, miR-27b may negatively regulate CYP1B1, a key member of the CYP family mediating metabolism of a wide range of drugs. Decreased expression of miR-27b and the subsequent high expression of CYP1B1 could be one of causes for resistance to docetaxel in cancerous cells162, 163. The most recent studies showed that miR-27b can also sensitize cancer cells to a broad spectrum of anti-cancer drugs in vitro and in vivo by activating P53-dependent apoptosis and reducing CYP1B1-mediated drug detoxification164. Other CYPs are also reported to be modulated by miRNA. For instance, CYP1A1 can be targeted by miR-892a165, CYP2J2 is inhibited by let-7b166, and CYP3A4 is downregulated by miR-148a167 (Table 1).

4.5. miRNAs modulate drug targets and DNA repair

miRNAs seem to impact anti-cancer drugs sensitivity by modulating the expression of drug targets. For example, miR-192 and miR-215 may influence 5-FU sensitivity by targeting TS enzyme in colorectal cancer cells168. Furthermore miR-27a, miR-27b, miR-134, and miR-582-5p are able to post-transcriptionally regulate DPD protein expression, which is also involved in sensitivity to 5-FU–based chemotherapy169. miR-211 can reduce the expression of RRM2, the important cellular target of gemcitabine, and increase the sensitivity of pancreatic cancer cells to gemcitabine170. miRNA let-7 was also found to negatively regulate RRM2 expression and sensitize PDAC cells to gemcitabine171.

Several studies have demonstrated that miRNAs can influence the chemosensitivity of cancer cells through interfering with DNA-repair pathways in cancer cells. Valeri and collaborators172 showed that in colorectal cancer cells, overexpression of miR-21 dramatically downregulated the expression of MMR proteins (hMSH2 and hMSH6) and reduced the therapeutic efficacy of 5-FU. MMR proteins, MSH2, MSH6 and MLH1-PMS2, were also reported to be negatively regulated by miR-155173. In breast cancer cell lines, miR-182 was able to downregulate BRCA1 protein expression, impair homologous recombination–mediated repair, thereby increasing cellular sensitivity to poly (ADP-ribose) polymerase (PARP) 1 inhibitor174. Sun et al׳s study175 showed that miR-9 could downregulate BRCA1 and impede DNA damage repair in ovarian cancer, subsequently increasing the sensitivity of cancer cells to cisplatin and PARP inhibitors (Table 1).

4.6. miRNAs regulate GSH and GSH-depended enzymes

Several recent studies have shown the regulatory role of miRNA on redox systems176, 177. However, such a role in the field of cancer MDR has not been fully studied. One report found that miRNA-27a contributed to cisplatin resistance through modulation of GSH biosynthesis178. Several other studies demonstrated that miRNA was able to target GST mediated drug metabolism to regulate MDR. Zhang et al.179 reported that miRNA-513a-3p could negatively regulate GSTP1 gene expression. Overexpression of miR-513a-3p resensitized cisplatin-resistant A549 cells to cisplatin179. Another study found that increased miR-133b expression could reduce GST-π expression, and reverse chemotherapy drug resistance180.

4.7. Master miRNAs modulate multiple targets

The most recent studies have suggested the existence of master miRNAs which could target multiple essential drug resistance pathways, therefore, are capable of improving the sensitivity to a broad spectrum of anticancer drugs. For example, miR-1271 can regulate cisplatin resistance of human gastric cancer cell lines by targeting IGF1R, IRS1, mTOR, and BCL-2181. miRNA-127 reversed adriamycin resistance via modulating ABC transporters MDR1 and MRP1, apoptosis related proteins (RUNX2, P53, BCL-2, survivin), as well as the AKT signal pathway182. miR-214 behaved as a key hub by coordinating fundamental signaling networks, such as PTEN/AKT, β-catenin, and tyrosine kinase receptor pathways, and also regulated the levels of crucial gene expression modulators, such as epigenetic repressor EZH2, P53, transcription factors TFAP2, and another miRNA (miR-148b)183. Discovery of more master miRNAs may be a powerful tool to overcome MDR.

5. Conclusions

MDR in cancer treatment is a highly complex process encompassing many different mechanisms. miRNAs, due to their extensive gene regulatory roles, are able to regulate nearly all the mechanisms of MDR. Therefore, miRNA, especially the master miRNAs, could be ideal biomarkers to predict chemotherapeutic response, as well as potential targets to overcome MDR in the future.

Acknowledgments

The Zhu׳s laboratory is supported by U. S. National Institute of Health Grants R01 #HL124122, #AR067766 and American Heart Association Grant 12SDG12070174. The Tan׳s laboratory is supported by the National Natural Science Foundation of China (Grant No. 81401155).

Footnotes

Peer review under responsibility of Institute of Materia Medica, Chinese Academy of Medical Sciences and Chinese Pharmaceutical Association.

References

1. Fojo A., Hamilton T.C., Young R.C., Ozols R.F. Multidrug resistance in ovarian cancer. Cancer. 1987;60:S2075–S2080. [PubMed] [Google Scholar]
2. Gong J., Jaiswal R., Mathys J.M., Combes V., Grau G.E., Bebawy M. Microparticles and their emerging role in cancer multidrug resistance. Cancer Treat Rev. 2012;38:226–234. [PubMed] [Google Scholar]
3. Baguley B.C. Multidrug resistance in cancer. Methods Mol Biol. 2010;596:1–14. [PubMed] [Google Scholar]
4. Gottesman M.M., Fojo T., Bates S.E. Multidrug resistance in cancer: role of ATP-dependent transporters. Nat Rev Cancer. 2002;2:48–58. [PubMed] [Google Scholar]
5. Szakács G., Paterson J.K., Ludwig J.A., Booth-Genthe C., Gottesman M.M. Targeting multidrug resistance in cancer. Nat Rev Drug Discov. 2006;5:219–234. [PubMed] [Google Scholar]
6. Gottesman M.M., Ling V. The molecular basis of multidrug resistance in cancer: the early years of P-glycoprotein research. FEBS Lett. 2006;580:998–1009. [PubMed] [Google Scholar]
7. Schinkel A.H., Jonker J.W. Mammalian drug efflux transporters of the ATP binding cassette (ABC) family: an overview. Adv Drug Deliv Rev. 2003;55:3–29. [PubMed] [Google Scholar]
8. Gillet J.P., Efferth T., Remacle J. Chemotherapy-induced resistance by ATP-binding cassette transporter genes. Biochim Biophys Acta. 2007;1775:237–262. [PubMed] [Google Scholar]
9. Consortium International Transporter, Giacomini K.M., Huang S.M., Tweedie D.J., Benet L.Z., Brouwer K.L. Membrane transporters in drug development. Nat Rev Drug Discov. 2010;9:215–236. [PMC free article] [PubMed] [Google Scholar]
10. Piecuch A., Obłąk E., Yeast A.B.C. proteins involved in multidrug resistance. Cell Mol Biol Lett. 2014;19:1–22. [PMC free article] [PubMed] [Google Scholar]
11. Huang S., Ye J., Yu J., Chen L., Zhou L., Wang H. The accumulation and efflux of lead partly depend on ATP-dependent efflux pump-multidrug resistance protein 1 and glutathione in testis Sertoli cells. Toxicol Lett. 2014;226:277–284. [PubMed] [Google Scholar]
12. Sodani K., Patel A., Kathawala R.J., Chen Z.S. Multidrug resistance associated proteins in multidrug resistance. Chin J Cancer. 2012;31:58–72. [PMC free article] [PubMed] [Google Scholar]
13. Tiwari A.K., Sodani K., Dai C.L., Ashby C.R. Jr, Chen Z.S. Revisiting the ABCs of multidrug resistance in cancer chemotherapy. Curr Pharm Biotechnol. 2011;12:570–594. [PubMed] [Google Scholar]
14. Assaraf Y.G., Rothem L., Hooijberg J.H., Stark M., Ifergan I., Kathmann I. Loss of multidrug resistance protein 1 (MRP1) expression and folate efflux activity results in a highly concentrative folate transport in human leukemia cells. J Biol Chem. 2003;278:6680–6686. [PubMed] [Google Scholar]
15. Wang D.S., Patel A., Shukla S., Zhang Y.K., Wang Y.J., Kathawala R.J. Icotinib antagonizes ABCG2-mediated multidrug resistance, but not the pemetrexed resistance mediated by thymidylate synthase and ABCG2. Oncotarget. 2014;5:4529–4542. [PMC free article] [PubMed] [Google Scholar]
16. Mao Q., Unadkat J.D. Role of the breast cancer resistance protein (ABCG2) in drug transport. AAPS J. 2005;7:E118–E133. [PMC free article] [PubMed] [Google Scholar]
17. Sun Y.L., Kathawala R.J., Singh S., Zheng K., Talele T.T., Jiang W.Q. Zafirlukast antagonizes ATP-binding cassette subfamily G member 2-mediated multidrug resistance. Anticancer Drugs. 2012;23:865–873. [PubMed] [Google Scholar]
18. Kastan M.B., Canman C.E., Leonard C.J. P53, cell cycle control and apoptosis: implications for cancer. Cancer Metastasis Rev. 1995;14:3–15. [PubMed] [Google Scholar]
19. Turzanski J., Zhu Y.M., Pallis M.J., Russell N.H. Comments on: multidrug resistance-associated protein (MRP) expression is correlated with expression of aberrant p53 protein in colorectal cancer, Fukushima Y, Oshika Y, Tokunaga T, et al., Eur J Cancer 1999, 35, 935-938. Mutant p53 and high expression of MRP are associated in acute myeloblastic leukaemia. Eur J Cancer. 2000;36:270–271. [PubMed] [Google Scholar]
20. Milicevic Z., Kasapovic J., Gavrilovic L., Milovanovic Z., Bajic V., Spremo-Potparevic B. Mutant p53 protein expression and antioxidant status deficiency in breast cancer. EXCLI J. 2014;13:691–708. [PMC free article] [PubMed] [Google Scholar]
21. Oshika Y., Nakamura M., Tokunaga T., Fukushima Y., Abe Y., Ozeki Y. Multidrug resistance-associated protein and mutant p53 protein expression in non-small cell lung cancer. Mod Pathol. 1998;11:1059–1063. [PubMed] [Google Scholar]
22. Pommier Y., Sordet O., Antony S., Hayward R.L., Kohn K.W. Apoptosis defects and chemotherapy resistance: molecular interaction maps and networks. Oncogene. 2004;23:2934–2949. [PubMed] [Google Scholar]
23. Kaufmann S.H., Earnshaw W.C. Induction of apoptosis by cancer chemotherapy. Exp Cell Res. 2000;256:42–49. [PubMed] [Google Scholar]
24. Wilson T.R., Johnston P.G., Longley D.B. Anti-apoptotic mechanisms of drug resistance in cancer. Curr Cancer Drug Targets. 2009;9:307–319. [PubMed] [Google Scholar]
25. Dole M., Nuñez G., Merchant A.K., Maybaum J., Rode C.K., Bloch C.A. Bcl-2 inhibits chemotherapy-induced apoptosis in neuroblastoma. Cancer Res. 1994;54:3253–3259. [PubMed] [Google Scholar]
26. Youle R.J., Strasser A. The BCL-2 protein family: opposing activities that mediate cell death. Nat Rev Mol Cell Biol. 2008;9:47–59. [PubMed] [Google Scholar]
27. Rong Y., Distelhorst C.W. Bcl-2 protein family members: versatile regulators of calcium signaling in cell survival and apoptosis. Annu Rev Physiol. 2008;70:73–91. [PubMed] [Google Scholar]
28. Ying H., Qu D., Liu C., Ying T., Lv J., Jin S. Chemoresistance is associated with Beclin-1 and PTEN expression in epithelial ovarian cancers. Oncol Lett. 2015;9:1759–1763. [PMC free article] [PubMed] [Google Scholar]
29. Bentires-Alj M., Barbu V., Fillet M., Chariot A., Relic B., Jacobs N. NFB transcription factor induces drug resistance through MDR1 expression in cancer cells. Oncogene. 2003;22:90–97. [PubMed] [Google Scholar]
30. Huang W.C., Hung M.C. Induction of Akt activity by chemotherapy confers acquired resistance. J Formos Med Assoc. 2009;108:180–194. [PubMed] [Google Scholar]
31. Levine B., Kroemer G. Autophagy in the pathogenesis of disease. Cell. 2008;132:27–42. [PMC free article] [PubMed] [Google Scholar]
32. Yang Z., Klionsky D.J. Eaten alive: a history of macroautophagy. Nat Cell Biol. 2010;12:814–822. [PMC free article] [PubMed] [Google Scholar]
33. Wen J., Yeo S., Wang C., Chen S., Sun S., Haas M.A. Autophagy inhibition re-sensitizes pulse stimulation-selected paclitaxel-resistant triple negative breast cancer cells to chemotherapy-induced apoptosis. Breast Cancer Res Treat. 2015;149:619–629. [PMC free article] [PubMed] [Google Scholar]
34. Yu L., Gu C., Zhong D., Shi L., Kong Y., Zhou Z. Induction of autophagy counteracts the anticancer effect of cisplatin in human esophageal cancer cells with acquired drug resistance. Cancer Lett. 2014;355:34–45. [PubMed] [Google Scholar]
35. Chen M., He M., Song Y., Chen L., Xiao P., Wan X. The cytoprotective role of gemcitabine-induced autophagy associated with apoptosis inhibition in triple-negative MDA-MB-231 breast cancer cells. Int J Mol Med. 2014;34:276–282. [PubMed] [Google Scholar]
36. Pan Y.Z., Wang X., Bai H., Wang C.B., Zhang Q., Xi R. Autophagy in drug resistance of the multiple myeloma cell line RPMI8226 to doxorubicin. Genet Mol Res. 2015;14:5621–5629. [PubMed] [Google Scholar]
37. Kumar A., Singh U.K., Chaudhary A. Targeting autophagy to overcome drug resistance in cancer therapy. Future Med Chem. 2015;7:1535–1542. [PubMed] [Google Scholar]
38. Chen S., Zhu X., Qiao H., Ye M., Lai X., Yu S. Protective autophagy promotes the resistance of HER2-positive breast cancer cells to lapatinib. Tumour Biol. 2016;37:2321–2331. [PubMed] [Google Scholar]
39. O׳Donovan T.R., O׳Sullivan G.C., McKenna S.L. Induction of autophagy by drug-resistant esophageal cancer cells promotes their survival and recovery following treatment with chemotherapeutics. Autophagy. 2011;7:509–524. [PMC free article] [PubMed] [Google Scholar]
40. Henson E.S., Gibson S.B. Surviving cell death through epidermal growth factor (EGF) signal transduction pathways: implications for cancer therapy. Cell Signal. 2006;18:2089–2097. [PubMed] [Google Scholar]
41. Ghadimi M.P., Lopez G., Torres K.E., Belousov R., Young E.D., Liu J. Targeting the PI3K/mTOR axis, alone and in combination with autophagy blockade, for the treatment of malignant peripheral nerve sheath tumors. Mol Cancer Ther. 2012;11:1758–1769. [PMC free article] [PubMed] [Google Scholar]
42. Stanton M.J., Dutta S., Zhang H., Polavaram N.S., Leontovich A.A., Hönscheid P. Autophagy control by the VEGF-C/NRP-2 axis in cancer and its implication for treatment resistance. Cancer Res. 2013;73:160–171. [PMC free article] [PubMed] [Google Scholar]
43. Paillas S., Causse A., Marzi L., de Medina P., Poirot M., Denis V. MAPK14/p38α confers irinotecan resistance to TP53-defective cells by inducing survival autophagy. Autophagy. 2012;8:1098–1112. [PMC free article] [PubMed] [Google Scholar]
44. Hennigan R.F., Moon C.A., Parysek L.M., Monk K.R., Morfini G., Berth S. The NF2 tumor suppressor regulates microtubule-based vesicle trafficking via a novel Rac, MLK and p38SAPK pathway. Oncogene. 2013;32:1135–1143. [PMC free article] [PubMed] [Google Scholar]
45. Crommentuyn K.M., Schellens J.H., van den Berg J.D., Beijnen J.H. In-vitro metabolism of anti-cancer drugs, methods and applications: paclitaxel, docetaxel, tamoxifen and ifosfamide. Cancer Treat Rev. 1998;24:345–366. [PubMed] [Google Scholar]
46. Kivisto K.T., Kroemer H.K., Eichelbaum M. The role of human cytochrome P450 enzymes in the metabolism of anticancer agents: implications for drug interactions. Br J Clin Pharmacol. 1995;40:523–530. [PMC free article] [PubMed] [Google Scholar]
47. Vassal G., Pondarré C., Boland I., Cappelli C., Santos A., Thomas C. Preclinical development of camptothecin derivatives and clinical trials in pediatric oncology. Biochimie. 1998;80:271–280. [PubMed] [Google Scholar]
48. Patterson L.H., Murray G.I. Tumour cytochrome P450 and drug activation. Curr Pharm Des. 2002;8:1335–1347. [PubMed] [Google Scholar]
49. Mwinyi J., Vokinger K., Jetter A., Breitenstein U., Hiller C., Kullak-Ublick G.A. Impact of variable CYP genotypes on breast cancer relapse in patients undergoing adjuvant tamoxifen therapy. Cancer Chemother Pharmacol. 2014;73:1181–1188. [PubMed] [Google Scholar]
50. Peters G.J., Backus H.H., Freemantle S., van Triest B., Codacci-Pisanelli G., van der Wilt C.L. Induction of thymidylate synthase as a 5-fluorouracil resistance mechanism. Biochim Biophys Acta. 2002;1587:194–205. [PubMed] [Google Scholar]
51. Nakamura J., Kohya N., Kai K., Ohtaka K., Hashiguchi K., Hiraki M. Ribonucleotide reductase subunit M1 assessed by quantitative double-fluorescence immunohistochemistry predicts the efficacy of gemcitabine in biliary tract carcinoma. Int J Oncol. 2010;37:845–852. [PubMed] [Google Scholar]
52. Nitiss J.L. Targeting DNA topoisomerase II in cancer chemotherapy. Nat Rev Cancer. 2009;9:338–350. [PMC free article] [PubMed] [Google Scholar]
53. Ganapathi R.N., Ganapathi M.K. Mechanisms regulating resistance to inhibitors of topoisomerase II. Front Pharmacol. 2013;4:89. [PMC free article] [PubMed] [Google Scholar]
54. Geng M., Wang L., Chen X., Cao R., Li P. The association between chemosensitivity and Pgp, GST and Topo II expression in gastric cancer. Diagn Pathol. 2013;8:198. [PMC free article] [PubMed] [Google Scholar]
55. Wilson T.R., Longley D.B., Johnston P.G. Chemoresistance in solid tumours. Ann Oncol. 2006;17 Suppl 10:x315–x324. [PubMed] [Google Scholar]
56. Allen K.A. Implementation of new technologies in cytotechnology education. Cancer. 1998;84:324–327. [PubMed] [Google Scholar]
57. Ribic C.M., Sargent D.J., Moore M.J., Thibodeau S.N., French A.J., Goldberg R.M. Tumor microsatellite-instability status as a predictor of benefit from fluorouracil-based adjuvant chemotherapy for colon cancer. N Engl J Med. 2003;349:247–257. [PMC free article] [PubMed] [Google Scholar]
58. Marullo R., Werner E., Degtyareva N., Moore B., Altavilla G., Ramalingam S.S. Cisplatin induces a mitochondrial-ROS response that contributes to cytotoxicity depending on mitochondrial redox status and bioenergetic functions. PLoS One. 2013;8:e81162. [PMC free article] [PubMed] [Google Scholar]
59. Fan C., Zheng W., Fu X., Li X., Wong Y.S., Chen T. Strategy to enhance the therapeutic effect of doxorubicin in human hepatocellular carcinoma by selenocystine, a synergistic agent that regulates the ROS-mediated signaling. Oncotarget. 2014;5:2853–2863. [PMC free article] [PubMed] [Google Scholar]
60. Lin C.J., Lee C.C., Shih Y.L., Lin C.H., Wang S.H., Chen T.H. Inhibition of mitochondria- and endoplasmic reticulum stress-mediated autophagy augments temozolomide-induced apoptosis in glioma cells. PLoS One. 2012;7:e38706. [PMC free article] [PubMed] [Google Scholar]
61. Liu W., Gu J., Qi J., Zeng X.N., Ji J., Chen Z.Z. Lentinan exerts synergistic apoptotic effects with paclitaxel in A549 cells via activating ROS-TXNIP-NLRP3 inflammasome. J Cell Mol Med. 2015;19:1949–1955. [PMC free article] [PubMed] [Google Scholar]
62. Acharya A., Das I., Chandhok D., Saha T. Redox regulation in cancer: a double-edged sword with therapeutic potential. Oxid Med Cell Longev. 2010;3:23–34. [PMC free article] [PubMed] [Google Scholar]
63. Liu Y., Li Q., Zhou L., Xie N., Nice E.C., Zhang H. Cancer drug resistance: redox resetting renders a way. Oncotarget. 2016 Available from: http://dx.doi.org/10.18632/oncotarget.8600. [PMC free article] [PubMed] [Google Scholar]
64. Traverso N., Ricciarelli R., Nitti M., Marengo B., Furfaro A.L., Pronzato M.A. Role of glutathione in cancer progression and chemoresistance. Oxid Med Cell Longev. 2013;2013:972913. [PMC free article] [PubMed] [Google Scholar]
65. Pompella A., Corti A., Paolicchi A., Giommarelli C., Zunino F. γ-glutamyltransferase, redox regulation and cancer drug resistance. Curr Opin Pharmacol. 2007;7:360–366. [PubMed] [Google Scholar]
66. Schnekenburger M., Karius T., Diederich M. Regulation of epigenetic traits of the glutathione S-transferase P1 gene: from detoxification toward cancer prevention and diagnosis. Front Pharmacol. 2014;5:170. [PMC free article] [PubMed] [Google Scholar]
67. Jardim B.V., Moschetta M.G., Leonel C., Gelaleti G.B., Regiani V.R., Ferreira L.C. Glutathione and glutathione peroxidase expression in breast cancer: an immunohistochemical and molecular study. Oncol Rep. 2013;30:1119–1128. [PubMed] [Google Scholar]
68. Backos D.S., Franklin C.C., Reigan P. The role of glutathione in brain tumor drug resistance. Biochem Pharmacol. 2012;83:1005–1012. [PubMed] [Google Scholar]
69. Singh S. Cytoprotective and regulatory functions of glutathione S-transferases in cancer cell proliferation and cell death. Cancer Chemother Pharmacol. 2015;75:1–15. [PubMed] [Google Scholar]
70. Beeghly A., Katsaros D., Chen H., Fracchioli S., Zhang Y., Massobrio M. Glutathione S-transferase polymorphisms and ovarian cancer treatment and survival. Gynecol Oncol. 2006;100:330–337. [PubMed] [Google Scholar]
71. Ge J., Tian A.X., Wang Q.S., Kong P.Z., Yu Y., Li X.Q. The GSTP1 105Val allele increases breast cancer risk and aggressiveness but enhances response to cyclophosphamide chemotherapy in North China. PLoS One. 2013;8:e67589. [PMC free article] [PubMed] [Google Scholar]
72. Kap E.J., Richter S., Rudolph A., Jansen L., Ulrich A., Hoffmeister M. Genetic variants in the glutathione S-transferase genes and survival in colorectal cancer patients after chemotherapy and differences according to treatment with oxaliplatin. Pharmacogenet Genom. 2014;24:340–347. [PubMed] [Google Scholar]
73. Townsend D.M., Tew K.D. The role of glutathione-S-transferase in anti-cancer drug resistance. Oncogene. 2003;22:7369–7375. [PMC free article] [PubMed] [Google Scholar]
74. Ramsay E.E., Dilda P.J., Glutathione S-conjugates as prodrugs to target drug-resistant tumors. Front Pharmacol. 2014;5:181. [PMC free article] [PubMed] [Google Scholar]
75. Rutnam Z.J., Wight T.N., Yang B.B. miRNAs regulate expression and function of extracellular matrix molecules. Matrix Biol. 2013;32:74–85. [PMC free article] [PubMed] [Google Scholar]
76. Vasudevan S., Tong Y., Steitz J.A. Switching from repression to activation: microRNAs can up-regulate translation. Science. 2007;318:1931–1934. [PubMed] [Google Scholar]
77. Lee R.C., Feinbaum R.L., Ambros V. The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell. 1993;75:843–854. [PubMed] [Google Scholar]
78. Kozomara A., Griffiths-Jones S. miRBase: integrating microRNA annotation and deep-sequencing data. Nucleic Acids Res. 2011;39:D152–D157. [PMC free article] [PubMed] [Google Scholar]
79. Spizzo R., Almeida M.I., Colombatti A., Calin G.A. Long non-coding RNAs and cancer: a new frontier of translational research. Oncogene. 2012;31:4577–4587. [PMC free article] [PubMed] [Google Scholar]
80. Mendell J.T., Olson E.N. MicroRNAs in stress signaling and human disease. Cell. 2012;148:1172–1187. [PMC free article] [PubMed] [Google Scholar]
81. Esteller M. Non-coding RNAs in human disease. Nat Rev Genet. 2011;12:861–874. [PubMed] [Google Scholar]
82. Jiang Q., Wang Y., Hao Y., Juan L., Teng M., Zhang X. miR2Disease: a manually curated database for microRNA deregulation in human disease. Nucleic Acids Res. 2009;37:D98–104. [PMC free article] [PubMed] [Google Scholar]
83. Ambros V. MicroRNA pathways in flies and worms: growth, death, fat, stress, and timing. Cell. 2003;113:673–676. [PubMed] [Google Scholar]
84. Croce C.M. Causes and consequences of microRNA dysregulation in cancer. Nat Rev Genet. 2009;10:704–714. [PMC free article] [PubMed] [Google Scholar]
85. Liang L.H., He X.H. Macro-management of microRNAs in cell cycle progression of tumor cells and its implications in anti-cancer therapy. Acta Pharmacol Sin. 2011;32:1311–1320. [PMC free article] [PubMed] [Google Scholar]
86. Mitchell P.S., Parkin R.K., Kroh E.M., Fritz B.R., Wyman S.K., Pogosova-Agadjanyan E.L. Circulating microRNAs as stable blood-based markers for cancer detection. Proc Natl Acad Sci USA. 2008;105:10513–10518. [PMC free article] [PubMed] [Google Scholar]
87. Lu J., Getz G., Miska E.A., Alvarez-Saavedra E., Lamb J., Peck D. MicroRNA expression profiles classify human cancers. Nature. 2005;435:834–838. [PubMed] [Google Scholar]
88. Zhu J., Zheng Z., Wang J., Sun J., Wang P., Cheng X. Different miRNA expression profiles between human breast cancer tumors and serum. Front Genet. 2014;5:149. [PMC free article] [PubMed] [Google Scholar]
89. Fojo T. Multiple paths to a drug resistance phenotype: mutations, translocations, deletions and amplification of coding genes or promoter regions, epigenetic changes and microRNAs. Drug Resist Updat. 2007;10:59–67. [PubMed] [Google Scholar]
90. Blower P.E., Chung J.H., Verducci J.S., Lin S., Park J.K., Dai Z. MicroRNAs modulate the chemosensitivity of tumor cells. Mol Cancer Ther. 2008;7:1–9. [PubMed] [Google Scholar]
91. Wang J., Yang M., Li Y., Han B. The role of microRNAs in the chemoresistance of breast cancer. Drug Dev Res. 2015;76:368–374. [PubMed] [Google Scholar]
92. Zhu H., Wu H., Liu X., Evans B.R., Medina D.J., Liu C.G. Role of microRNA miR-27a and miR-451 in the regulation of MDR1/P-glycoprotein expression in human cancer cells. Biochem Pharmacol. 2008;76:582–588. [PMC free article] [PubMed] [Google Scholar]
93. Li Z., Hu S., Wang J., Cai J., Xiao L., Yu L. miR-27a modulates MDR1/P-glycoprotein expression by targeting HIPK2 in human ovarian cancer cells. Gynecol Oncol. 2010;119:125–130. [PubMed] [Google Scholar]
94. Kovalchuk O., Filkowski J., Meservy J., Ilnytskyy Y., Tryndyak V.P., Chekhun V.F. Involvement of microRNA-451 in resistance of the MCF-7 breast cancer cells to chemotherapeutic drug doxorubicin. Mol Cancer Ther. 2008;7:2152–2159. [PubMed] [Google Scholar]
95. Feng D.D., Zhang H., Zhang P., Zheng Y.S., Zhang X.J., Han B.W. Down-regulated miR-331-5p and miR-27a are associated with chemotherapy resistance and relapse in leukaemia. J Cell Mol Med. 2011;15:2164–2175. [PMC free article] [PubMed] [Google Scholar]
96. Chen Z., Ma T., Huang C., Zhang L., Lv X., Xu T. MiR-27a modulates the MDR1/P-glycoprotein expression by inhibiting FZD7-catenin pathway in hepatocellular carcinoma cells. Cell Signal. 2013;25:2693–2701. [PubMed] [Google Scholar]
97. Zhao X., Yang L., Hu J., Ruan J. miR-138 might reverse multidrug resistance of leukemia cells. Leuk Res. 2010;34:1078–1082. [PubMed] [Google Scholar]
98. Bao L., Hazari S., Mehra S., Kaushal D., Moroz K., Dash S. Increased expression of P-glycoprotein and doxorubicin chemoresistance of metastatic breast cancer is regulated by miR-298. Am J Pathol. 2012;180:2490–2503. [PMC free article] [PubMed] [Google Scholar]
99. Xu Y., Ohms S.J., Li Z., Wang Q., Gong G., Hu Y. Changes in the expression of miR-381 and miR-495 are inversely associated with the expression of the MDR1 gene and development of multi-drug resistance. PLoS One. 2013;8:e82062. [PMC free article] [PubMed] [Google Scholar]
100. Yang T., Zheng Z.M., Li X.N., Li Z.F., Wang Y., Geng Y.F. miR-223 modulates multidrug resistance via downregulation of ABCB1 in hepatocellular carcinoma cells. Exp Biol Med. 2013;238:1024–1032. [PubMed] [Google Scholar]
101. Munoz J.L., Bliss S.A., Greco S.J., Ramkissoon S.H., Ligon K.L., Rameshwar P. Delivery of functional anti-miR-9 by mesenchymal stem cell-derived exosomes to glioblastoma multiforme cells conferred chemosensitivity. Mol Ther Nucleic Acids. 2013;2:e126. [PMC free article] [PubMed] [Google Scholar]
102. Lin C.J., Gong H.Y., Tseng H.C., Wang W.L., Wu J.L. miR-122 targets an anti-apoptotic gene, Bcl-w, in human hepatocellular carcinoma cell lines. Biochem Biophys Res Commun. 2008;375:315–320. [PubMed] [Google Scholar]
103. Wu D.D., Li X.S., Meng X.N., Yan J., Zong Z.H. MicroRNA-873 mediates multidrug resistance in ovarian cancer cells by targeting ABCB1. Tumour Biol. 2016;37:10499–10506. [PubMed] [Google Scholar]
104. Shang Y., Zhang Z., Liu Z., Feng B., Ren G., Li K. miR-508-5p regulates multidrug resistance of gastric cancer by targeting ABCB1 and ZNRD1. Oncogene. 2014;33:3267–3276. [PubMed] [Google Scholar]
105. Shang Y., Feng B., Zhou L., Ren G., Zhang Z., Fan X. The miR27b-CCNG1-P53-miR-508-5p axis regulates multidrug resistance of gastric cancer. Oncotarget. 2016;7:538–549. [PMC free article] [PubMed] [Google Scholar]
106. To K.K., Zhan Z., Litman T., Bates S.E. Regulation of ABCG2 expression at the 3′ untranslated region of its mRNA through modulation of transcript stability and protein translation by a putative microRNA in the S1 colon cancer cell line. Mol Cell Biol. 2008;28:5147–5161. [PMC free article] [PubMed] [Google Scholar]
107. Pan Y.Z., Morris M.E., Yu A.M. MicroRNA-328 negatively regulates the expression of breast cancer resistance protein (BCRP/ABCG2) in human cancer cells. Mol Pharmacol. 2009;75:1374–1379. [PMC free article] [PubMed] [Google Scholar]
108. Li W.Q., Li Y.M., Tao B.B., Lu Y.C., Hu G.H., Liu H.M. Downregulation of ABCG2 expression in glioblastoma cancer stem cells with miRNA-328 may decrease their chemoresistance. Med Sci Monit. 2010;16:HY27–HY30. [PubMed] [Google Scholar]
109. Li X., Pan Y.Z., Seigel G.M., Hu Z.H., Huang M., Yu A.M. Breast cancer resistance protein BCRP/ABCG2 regulatory microRNAs (hsa-miR-328, -519c and -520h) and their differential expression in stem-like ABCG2+ cancer cells. Biochem Pharmacol. 2011;81:783–792. [PMC free article] [PubMed] [Google Scholar]
110. Turrini E., Haenisch S., Laechelt S., Diewock T., Bruhn O., Cascorbi I. MicroRNA profiling in K-562 cells under imatinib treatment: influence of miR-212 and miR-328 on ABCG2 expression. Pharmacogenet Genom. 2012;22:198–205. [PubMed] [Google Scholar]
111. Jiao X., Zhao L., Ma M., Bai X., He M., Yan Y. miR-181a enhances drug sensitivity in mitoxantone-resistant breast cancer cells by targeting breast cancer resistance protein (BCRP/ABCG2) Breast Cancer Res Treat. 2013;139:717–730. [PubMed] [Google Scholar]
112. Ma M.T., He M., Wang Y., Jiao X.Y., Zhao L., Bai X.F. miR-487a resensitizes mitoxantrone (MX)-resistant breast cancer cells (MCF-7/MX) to MX by targeting breast cancer resistance protein (BCRP/ABCG2) Cancer Lett. 2013;339:107–115. [PubMed] [Google Scholar]
113. Liang Z., Wu H., Xia J., Li Y., Zhang Y., Huang K. Involvement of miR-326 in chemotherapy resistance of breast cancer through modulating expression of multidrug resistance-associated protein 1. Biochem Pharmacol. 2010;79:817–824. [PubMed] [Google Scholar]
114. Pan Y.Z., Zhou A., Hu Z., Yu A.M. Small nucleolar RNA-derived microRNA hsa-miR-1291 modulates cellular drug disposition through direct targeting of ABC transporter ABCC1. Drug Metab Dispos. 2013;41:1744–1751. [PMC free article] [PubMed] [Google Scholar]
115. Li X., Li X., Liao D., Wang X., Wu Z., Nie J. Elevated microRNA-23a expression enhances the chemoresistance of colorectal cancer cells with microsatellite instability to 5-fluorouracil by directly targeting ABCF1. Curr Protein Pept Sci. 2015;16:301–309. [PubMed] [Google Scholar]
116. Wu K., Yang Y., Zhao J., Zhao S. BAG3-mediated miRNA let-7g and let-7i inhibit proliferation and enhance apoptosis of human esophageal carcinoma cells by targeting the drug transporter ABCC10. Cancer Lett. 2016;371:125–133. [PubMed] [Google Scholar]
117. Wu Q., Yang Z., Xia L., Nie Y., Wu K., Shi Y. Methylation of miR-129-5p CpG island modulates multi-drug resistance in gastric cancer by targeting ABC transporters. Oncotarget. 2014;5:11552–11563. [PMC free article] [PubMed] [Google Scholar]
118. Wang Q., Wang Z., Chu L., Li X., Kan P., Xin X. The effects and molecular mechanisms of MiR-106a in multidrug resistance reversal in human glioma U87/DDP and U251/G cell lines. PLoS One. 2015;10:e0125473. [PMC free article] [PubMed] [Google Scholar]
119. Iida K., Fukushi J., Matsumoto Y., Oda Y., Takahashi Y., Fujiwara T. miR-125b develops chemoresistance in Ewing sarcoma/primitive neuroectodermal tumor. Cancer Cell Int. 2013;13:21. [PMC free article] [PubMed] [Google Scholar]
120. Liang S., Gong X., Zhang G., Huang G., Lu Y., Li Y. MicroRNA-140 regulates cell growth and invasion in pancreatic duct adenocarcinoma by targeting iASPP. Acta Biochim Biophys Sin. 2016;48:174–181. [PubMed] [Google Scholar]
121. Fornari F., Gramantieri L., Giovannini C., Veronese A., Ferracin M., Sabbioni S. miR-122/cyclin G1 interaction modulates p53 activity and affects doxorubicin sensitivity of human hepatocarcinoma cells. Cancer Res. 2009;69:5761–5767. [PubMed] [Google Scholar]
122. Fujita Y., Kojima K., Hamada N., Ohhashi R., Akao Y., Nozawa Y. Effects of miR-34a on cell growth and chemoresistance in prostate cancer PC3 cells. Biochem Biophys Res Commun. 2008;377:114–119. [PubMed] [Google Scholar]
123. Lodygin D., Tarasov V., Epanchintsev A., Berking C., Knyazeva T., Körner H. Inactivation of miR-34a by aberrant CpG methylation in multiple types of cancer. Cell Cycle. 2008;7:2591–2600. [PubMed] [Google Scholar]
124. Fan Y.N., Meley D., Pizer B., Sée V. miR-34a mimics are potential therapeutic agents for p53-mutated and chemo-resistant brain tumour cells. PLoS One. 2014;9:e108514. [PMC free article] [PubMed] [Google Scholar]
125. Chakraborty S., Mazumdar M., Mukherjee S., Bhattacharjee P., Adhikary A., Manna A. Restoration of p53/miR-34a regulatory axis decreases survival advantage and ensures Bax-dependent apoptosis of non-small cell lung carcinoma cells. FEBS Lett. 2014;588:549–559. [PubMed] [Google Scholar]
126. Xia L., Zhang D., Du R., Pan Y., Zhao L., Sun S. miR-15b and miR-16 modulate multidrug resistance by targeting BCL2 in human gastric cancer cells. Int J Cancer. 2008;123:372–379. [PubMed] [Google Scholar]
127. Cittelly D.M., Das P.M., Salvo V.A., Fonseca J.P., Burow M.E., Jones F.E. Oncogenic HER2Δ16 suppresses miR-15a/16 and deregulates BCL-2 to promote endocrine resistance of breast tumors. Carcinogenesis. 2010;31:2049–2057. [PMC free article] [PubMed] [Google Scholar]
128. Dong J., Zhao Y.P., Zhou L., Zhang T.P., Chen G. Bcl-2 upregulation induced by miR-21 via a direct interaction is associated with apoptosis and chemoresistance in MIA PaCa-2 pancreatic cancer cells. Arch Med Res. 2011;42:8–14. [PubMed] [Google Scholar]
129. Zhu W., Zhu D., Lu S., Wang T., Wang J., Jiang B. miR-497 modulates multidrug resistance of human cancer cell lines by targeting BCL2. Med Oncol. 2012;29:384–391. [PubMed] [Google Scholar]
130. Zhu W., Xu H., Zhu D., Zhi H., Wang T., Wang J. miR-200bc/429 cluster modulates multidrug resistance of human cancer cell lines by targeting BCL2 and XIAP. Cancer Chemother Pharmacol. 2012;69:723–731. [PubMed] [Google Scholar]
131. Xu K., Liang X., Cui D., Wu Y., Shi W., Liu J. miR-1915 inhibits Bcl-2 to modulate multidrug resistance by increasing drug-sensitivity in human colorectal carcinoma cells. Mol Carcinog. 2013;52:70–78. [PubMed] [Google Scholar]
132. Wang F., Liu M., Li X., Tang H. miR-214 reduces cell survival and enhances cisplatin-induced cytotoxicity via down-regulation of Bcl2l2 in cervical cancer cells. FEBS Lett. 2013;587:488–495. [PubMed] [Google Scholar]
133. Qu J., Zhao L., Zhang P., Wang J., Xu N., Mi W. MicroRNA-195 chemosensitizes colon cancer cells to the chemotherapeutic drug doxorubicin by targeting the first binding site of BCL2L2 mRNA. J Cell Physiol. 2015;230:535–545. [PubMed] [Google Scholar]
134. Verdoodt B., Neid M., Vogt M., Kuhn V., Liffers S.T., Palisaar R.J. MicroRNA-205, a novel regulator of the anti-apoptotic protein Bcl2, is downregulated in prostate cancer. Int J Oncol. 2013;43:307–314. [PubMed] [Google Scholar]
135. Chiyomaru T., Yamamura S., Fukuhara S., Hidaka H., Majid S., Saini S. Genistein up-regulates tumor suppressor microRNA-574-3p in prostate cancer. PLoS One. 2013;8:e58929. [PMC free article] [PubMed] [Google Scholar]
136. He H., Tian W., Chen H., Deng Y. MicroRNA-101 sensitizes hepatocellular carcinoma cells to doxorubicin-induced apoptosis via targeting Mcl-1. Mol Med Rep. 2016;13:1923–1929. [PubMed] [Google Scholar]
137. Romano G., Acunzo M., Garofalo M., Di Leva G., Cascione L., Zanca C. miR-494 is regulated by ERK1/2 and modulates TRAIL-induced apoptosis in non-small-cell lung cancer through BIM down-regulation. Proc Natl Acad Sci U S A. 2012;109:16570–16575. [PMC free article] [PubMed] [Google Scholar]
138. Hamada S., Masamune A., Miura S., Satoh K., Shimosegawa T. miR-365 induces gemcitabine resistance in pancreatic cancer cells by targeting the adaptor protein SHC1 and pro-apoptotic regulator BAX. Cell Signal. 2014;26:179–185. [PubMed] [Google Scholar]
139. Quintavalle C., Donnarumma E., Iaboni M., Roscigno G., Garofalo M., Romano G. Effect of miR-21 and miR-30b/c on TRAIL-induced apoptosis in glioma cells. Oncogene. 2013;32:4001–4008. [PubMed] [Google Scholar]
140. Yang G.D., Huang T.J., Peng L.X., Yang C.F., Liu R.Y., Huang H.B. Epstein-barr virus_encoded LMP1 upregulates microRNA-21 to promote the resistance of nasopharyngeal carcinoma cells to cisplatin-induced Apoptosis by suppressing PDCD4 and Fas-L. PLoS One. 2013;8:e78355. [PMC free article] [PubMed] [Google Scholar]
141. Meng F., Henson R., Wehbe-Janek H., Ghoshal K., Jacob S.T., Patel T. MicroRNA-21 regulates expression of the PTEN tumor suppressor gene in human hepatocellular cancer. Gastroenterology. 2007;133:647–658. [PMC free article] [PubMed] [Google Scholar]
142. Yang S.M., Huang C., Li X.F., Yu M.Z., He Y., Li J. miR-21 confers cisplatin resistance in gastric cancer cells by regulating PTEN. Toxicology. 2013;306:162–168. [PubMed] [Google Scholar]
143. Li J., Zhang Y., Zhao J., Kong F., Chen Y. Overexpression of miR-22 reverses paclitaxel-induced chemoresistance through activation of PTEN signaling in p53-mutated colon cancer cells. Mol Cell Biochem. 2011;357:31–38. [PubMed] [Google Scholar]
144. Zhao G., Cai C., Yang T., Qiu X., Liao B., Li W. MicroRNA-221 induces cell survival and cisplatin resistance through PI3K/Akt pathway in human osteosarcoma. PLoS One. 2013;8:e53906. [PMC free article] [PubMed] [Google Scholar]
145. Wang Y.S., Wang Y.H., Xia H.P., Zhou S.W., Schmid-Bindert G., Zhou C.C. MicroRNA-214 regulates the acquired resistance to gefitinib via the PTEN/AKT pathway in EGFR-mutant cell lines. Asian Pac J Cancer Prev. 2012;13:255–260. [PubMed] [Google Scholar]
146. Wang F., Li T., Zhang B., Li H., Wu Q., Yang L. MicroRNA-19a/b regulates multidrug resistance in human gastric cancer cells by targeting PTEN. Biochem Biophys Res Commun. 2013;434:688–694. [PubMed] [Google Scholar]
147. Fang L., Li H., Wang L., Hu J., Jin T., Wang J. MicroRNA-17-5p promotes chemotherapeutic drug resistance and tumour metastasis of colorectal cancer by repressing PTEN expression. Oncotarget. 2014;5:2974–2987. [PMC free article] [PubMed] [Google Scholar]
148. Zeng L.P., Hu Z.M., Li K., Xia K. miR-222 attenuates cisplatin-induced cell death by targeting the PPP2R2A/Akt/mTOR Axis in bladder cancer cells. J Cell Mol Med. 2016;20:559–567. [PMC free article] [PubMed] [Google Scholar]
149. Zhu H., Wu H., Liu X., Li B., Chen Y., Ren X. Regulation of autophagy by a beclin 1–targeted microRNA, miR–30a, in cancer cells. Autophagy. 2009;5:816–823. [PMC free article] [PubMed] [Google Scholar]
150. Yu Y., Cao L., Yang L., Kang R., Lotze M., Tang D. microRNA 30A promotes autophagy in response to cancer therapy. Autophagy. 2012;8:853–855. [PMC free article] [PubMed] [Google Scholar]
151. Yu Y., Yang L., Zhao M., Zhu S., Kang R., Vernon P. Targeting microRNA-30a–mediated autophagy enhances imatinib activity against human chronic myeloid leukemia cells. Leukemia. 2012;26:1752–1760. [PubMed] [Google Scholar]
152. Xu R., Liu S., Chen H., Lao L. MicroRNA-30a downregulation contributes to chemoresistance of osteosarcoma cells through activating Beclin-1–mediated autophagy. Oncol Rep. 2016;35:1757–1763. [PubMed] [Google Scholar]
153. Zheng B., Zhu H., Gu D., Pan X., Qian L., Xue B. miRNA-30a-mediated autophagy inhibition sensitizes renal cell carcinoma cells to sorafenib. Biochem Biophys Res Commun. 2015;459:234–239. [PubMed] [Google Scholar]
154. Zhang Y., Yang W.Q., Zhu H., Qian Y.Y., Zhou L., Ren Y.J. Regulation of autophagy by miR-30d impacts sensitivity of anaplastic thyroid carcinoma to cisplatin. Biochem Pharmacol. 2014;87:562–570. [PMC free article] [PubMed] [Google Scholar]
155. Sümbül A.T., Gög˘ebakan B., Ergün S., Yengil E., Batmacı C.Y., Tonyalı Ö. miR-204-5p expression in colorectal cancer: an autophagy-associated gene. Tumour Biol. 2014;35:12713–12719. [PubMed] [Google Scholar]
156. Xiao J., Zhu X., He B., Zhang Y., Kang B., Wang Z. miR-204 regulates cardiomyocyte autophagy induced by ischemia–reperfusion through LC3-II. J Biomed Sci. 2011;18:35. [PMC free article] [PubMed] [Google Scholar]
157. Chatterjee A., Chattopadhyay D., Chakrabarti G. miR-16 targets Bcl-2 in paclitaxel-resistant lung cancer cells and overexpression of miR-16 along with miR-17 causes unprecedented sensitivity by simultaneously modulating autophagy and apoptosis. Cell Signal. 2015;27:189–203. [PubMed] [Google Scholar]
158. Chen L., Jiang K., Jiang H., Wei P. miR-155 mediates drug resistance in osteosarcoma cells via inducing autophagy. Exp Ther Med. 2014;8:527–532. [PMC free article] [PubMed] [Google Scholar]
159. Pan B., Feng B., Chen Y., Huang G., Wang R., Chen L. miR-200b regulates autophagy associated with chemoresistance in human lung adenocarcinoma. Oncotarget. 2015;6:32805–32820. [PMC free article] [PubMed] [Google Scholar]
160. Huang N., Wu J., Qiu W., Lyu Q., He J., Xie W. miR-15a and miR-16 induce autophagy and enhance chemosensitivity of Camptothecin. Cancer Biol Ther. 2015;16:941–948. [PMC free article] [PubMed] [Google Scholar]
161. Zhao J., Nie Y., Wang H., Lin Y. miR-181a suppresses autophagy and sensitizes gastric cancer cells to cisplatin. Gene. 2016;576:828–833. [PubMed] [Google Scholar]
162. Tsuchiya Y., Nakajima M., Takagi S., Taniya T., Yokoi T. MicroRNA regulates the expression of human cytochrome P450 1B1. Cancer Res. 2006;66:9090–9098. [PubMed] [Google Scholar]
163. Martinez V.G., O׳Connor R., Liang Y., Clynes M. CYP1B1 expression is induced by docetaxel: effect on cell viability and drug resistance. Br J Cancer. 2008;98:564–570. [PMC free article] [PubMed] [Google Scholar]
164. Mu W., Hu C., Zhang H., Qu Z., Cen J., Qiu Z. miR-27b synergizes with anticancer drugs via p53 activation and CYP1B1 suppression. Cell Res. 2015;25:477–495. [PMC free article] [PubMed] [Google Scholar]
165. Choi Y.M., An S., Lee E.M., Kim K., Choi S.J., Kim J.S. CYP1A1 is a target of miR-892a-mediated post-transcriptional repression. Int J Oncol. 2012;41:331–336. [PubMed] [Google Scholar]
166. Chen F., Chen C., Yang S., Gong W., Wang Y., Cianflone K. Let-7b inhibits human cancer phenotype by targeting cytochrome P450 epoxygenase 2J2. PLoS One. 2012;7:e39197. [PMC free article] [PubMed] [Google Scholar]
167. Takagi S., Nakajima M., Mohri T., Yokoi T. Post-transcriptional regulation of human pregnane X receptor by micro-RNA affects the expression of cytochrome P450 3A4. J Biol Chem. 2008;283:9674–9680. [PubMed] [Google Scholar]
168. Boni V., Bitarte N., Cristobal I., Zarate R., Rodriguez J., Maiello E. miR-192/miR-215 influence 5-fluorouracil resistance through cell cycle-mediated mechanisms complementary to its post-transcriptional thymidilate synthase regulation. Mol Cancer Ther. 2010;9:2265–2275. [PubMed] [Google Scholar]
169. Hirota T., Date Y., Nishibatake Y., Takane H., Fukuoka Y., Taniguchi Y. Dihydropyrimidine dehydrogenase (DPD) expression is negatively regulated by certain microRNAs in human lung tissues. Lung Cancer. 2012;77:16–23. [PubMed] [Google Scholar]
170. Maftouh M., Avan A., Funel N., Frampton A.E., Fiuji H., Pelliccioni S. miR-211 modulates gemcitabine activity through downregulation of ribonucleotide reductase and inhibits the invasive behavior of pancreatic cancer cells. Nucleosides Nucleotides Nucleic Acids. 2014;33:384–393. [PubMed] [Google Scholar]
171. Bhutia Y.D., Hung S.W., Krentz M., Patel D., Lovin D., Manoharan R. Differential processing of let-7a precursors influences RRM2 expression and chemosensitivity in pancreatic cancer: role of LIN-28 and SET oncoprotein. PLoS One. 2013;8:e53436. [PMC free article] [PubMed] [Google Scholar]
172. Valeri N., Gasparini P., Braconi C., Paone A., Lovat F., Fabbri M. MicroRNA-21 induces resistance to 5-fluorouracil by down-regulating human DNA MutS homolog 2 (hMSH2) Proc Natl Acad Sci U S A. 2010;107:21098–21103. [PMC free article] [PubMed] [Google Scholar]
173. Valeri N., Gasparini P., Fabbri M., Braconi C., Veronese A., Lovat F. Modulation of mismatch repair and genomic stability by miR-155. Proc Natl Acad Sci U S A. 2010;107:6982–6987. [PMC free article] [PubMed] [Google Scholar]
174. Moskwa P., Buffa F.M., Pan Y., Panchakshari R., Gottipati P., Muschel R.J. miR-182-mediated downregulation of BRCA1 impacts DNA repair and sensitivity to PARP inhibitors. Mol Cell. 2011;41:210–220. [PMC free article] [PubMed] [Google Scholar]
175. Sun C., Li N., Yang Z., Zhou B., He Y., Weng D. miR-9 regulation of BRCA1 and ovarian cancer sensitivity to cisplatin and PARP inhibition. J Natl Cancer Inst. 2013;105:1750–1758. [PubMed] [Google Scholar]
176. Mikhed Y., Görlach A., Knaus U.G., Daiber A. Redox regulation of genome stability by effects on gene expression, epigenetic pathways and DNA damage/repair. Redox Biol. 2015;5:275–289. [PMC free article] [PubMed] [Google Scholar]
177. Ouyang Y.B., Stary C.M., White R.E., Giffard R.G. The use of microRNAs to modulate redox and immune response to stroke. Antioxid Redox Signal. 2015;22:187–202. [PMC free article] [PubMed] [Google Scholar]
178. Drayton R.M., Dudziec E., Peter S., Bertz S., Hartmann A., Bryant H.E. Reduced expression of miRNA-27a modulates cisplatin resistance in bladder cancer by targeting the cystine/glutamate exchanger SLC7A11. Clin Cancer Res. 2014;20:1990–2000. [PMC free article] [PubMed] [Google Scholar]
179. Zhang X., Zhu J., Xing R., Tie Y., Fu H., Zheng X. miR-513a-3p sensitizes human lung adenocarcinoma cells to chemotherapy by targeting GSTP1. Lung Cancer. 2012;77:488–494. [PubMed] [Google Scholar]
180. Chen S., Jiao J.W., Sun K.X., Zong Z.H., Zhao Y. MicroRNA-133b targets glutathione S-transferase π expression to increase ovarian cancer cell sensitivity to chemotherapy drugs. Drug Des Devel Ther. 2015;9:5225–5235. [PMC free article] [PubMed] [Google Scholar]
181. Yang M., Shan X., Zhou X., Qiu T., Zhu W., Ding Y. miR-1271 regulates cisplatin resistance of human gastric cancer cell lines by targeting IGF1R, IRS1, mTOR, and BCL2. Anticancer Agents Med Chem. 2014;14:884–891. [PubMed] [Google Scholar]
182. Feng R., Dong L. Knockdown of microRNA-127 reverses adriamycin resistance via cell cycle arrest and apoptosis sensitization in adriamycin-resistant human glioma cells. Int J Clin Exp Pathol. 2015;8:6107–6116. [PMC free article] [PubMed] [Google Scholar]
183. Penna E., Orso F., Taverna D. miR-214 as a key hub that controls cancer networks: small player, multiple functions. J Investig Dermatol. 2015;135:960–969. [PubMed] [Google Scholar]

Articles from Acta Pharmaceutica Sinica. B are provided here courtesy of Elsevier

-