Skip to main content

MINI REVIEW article

Front. Oncol., 02 June 2020
Sec. Pharmacology of Anti-Cancer Drugs
This article is part of the Research Topic FDA-approved Drug Repositioning for P-glycoprotein Overexpressing Resistant Cancer View all 8 articles

Cancer Stem Cells as a Potential Target to Overcome Multidrug Resistance

  • Research Institute of Pharmaceutical Sciences, College of Pharmacy, Sookmyung Women's University, Seoul, South Korea

Multidrug resistance (MDR), which is a significant impediment to the success of cancer chemotherapy, is attributable to various defensive mechanisms in cancer. Initially, overexpression of ATP-binding cassette (ABC) transporters such as P-glycoprotein (P-gp) was considered the most important mechanism for drug resistance; hence, many investigators for a long time focused on the development of specific ABC transporter inhibitors. However, to date their efforts have failed to develop a clinically applicable drug, leaving only a number of problems. The concept of cancer stem cells (CSCs) has provided new directions for both cancer and MDR research. MDR is known to be one of the most important features of CSCs and thus plays a crucial role in cancer recurrence and exacerbation. Therefore, in recent years, research targeting CSCs has been increasing rapidly in search of an effective cancer treatment. Here, we review the drugs that have been studied and developed to overcome MDR and CSCs, and discuss the limitations and future perspectives.

Introduction

Chemotherapy is one of the most effective treatments for cancer; however, its success has been challenged by the acquisition of multidrug resistance (MDR) (1). MDR is caused by sustained (dose-dependent or time-dependent) administration of chemotherapeutic drugs, resulting in cross-resistance to a broad spectrum of structurally- and mechanically-distinct chemotherapeutic drugs (2). There are several mechanisms underlying MDR (3, 4): (i) increased pumping out of the drug through efflux pumps such as P-glycoprotein (P-gp) encoded by ABCB1 (5); (ii) decreased uptake of the drug through transporters; (iii) activation of drug-metabolizing enzymes such as cytochrome P450 and glutathione S-transferase; (iv) activation of DNA repair systems; (v) evasion of apoptosis. The first three of these processes are conducive to the development of resistance by preventing the drug from reaching an effective concentration, while the remaining two mechanisms achieve resistance by detoxifying the action of the drug. Based on the mechanisms described, we have been trying to establish strategies for overcoming MDR. Before discussing them further, we need to understand the existence of cancer stem cells (CSCs) and their roles in cancer biology. CSCs, also known as tumor-initiating cells (TICs), are a small population of cancer cells that have the ability to self-renew and differentiate similar to normal stem cells (NSCs). However, as CSCs are tumorigenic, they can contribute to the aggravation and recurrence of cancer (6). According to a CSC model that explains the relationship between CSC and MDR, increased expression of ATP-binding cassette (ABC) transporters and other genes contributes to the intrinsic resistance of CSCs to chemotherapy (7, 8). CSCs have relatively slow cell-cycle kinetics and are therefore targeted less by chemotherapeutic drugs compared to rapidly dividing cells (9). In addition to the well-known ABC transporters (2, 10), many other drug resistance mechanisms of CSCs have been identified, for example, aldehyde dehydrogenases (ALDHs) (11), epithelial-mesenchymal transition (EMT) (12), epigenetic modifications (13, 14), factors affecting tumor microenvironment, such as hypoxia (15), and signaling pathways (1618). In this review, we provide a brief outline of MDR mechanisms and focus on the investigated drugs.

Current Strategies to Overcome MDR

Targeting ABC Transporters

ABC transporters including ABCB1, ABCC1, and ABCG2 are expressed in cancer stem/progenitor cells. These transporters have broad drug specificity and pump out a wide range of structurally- and mechanically-unrelated compounds, thereby lowering the intracellular accumulation of these compounds and consequently diminishing their biological efficacies (19). Several chemotherapeutic agents in clinical use are susceptible to ABC transporter-mediated efflux, such as microtubule-targeting taxanes (e.g., docetaxel and paclitaxel) and vinca alkaloids (vinblastine and vincristine), DNA-damaging anthracyclines (daunorubicin and doxorubicin), topoisomerase inhibitors (etoposide and topotecan), and tyrosine kinase inhibitors (dasatinib and gefitinib) (20). Therefore, developing strategies to target ABC transporters is an important area of cancer research, and many studies have been conducted accordingly (21). There are three approaches: (i) regulating the function of ABC transporters using competitive or allosteric inhibitors (Table 1) as well as the antibodies that target ABC transporters, such as UIC2 and MRK16 (86); (ii) regulating gene expression of ABC transporters at the transcriptional or translational level because, as with trabectedin, it is an attractive strategy to control ABC transporters at the transcriptional level by affecting the MDR enhanceosome (8789); (iii) using anticancer drugs that are poor substrates of P-gp, such as ixabepilone (Table 1). Until ixabepilone was launched, efforts to develop drugs targeting ABC transporters had been a driving force in the development of first-, second-, and third-generation P-gp inhibitors (90, 91). However, it has been reported that first-generation inhibitors have low potency and high toxicity, and second-generation inhibitors have frequent drug-drug interactions (92). With the third-generation inhibitors, there have been many improvements with regard to the drawbacks of the previous generations, but clinical trial data are still insufficient. In effect, most clinical trials have been discontinued. Because NSCs, including hematopoietic stem cells, unrestricted somatic stem cells, and mesenchymal stem cells, also express ABC transporters to protect themselves from cytotoxic agents (93), inhibiting ABC transporters may cause serious side effects such as hematopoietic disorders due to bone marrow dysfunction. Therefore, the emergence of ixabepilone was inevitable and has been well-received. Like taxanes, ixabepilone leads to G2/M phase arrest by stabilizing microtubules and promoting tubulin polymerization. However, ixabepilone has a very important feature (not found in taxanes) effective against cancer cells that acquire MDR following repeated chemotherapy, as this drug is not pumped out through P-gp. Now, developing drugs that are not substrates of P-gp has become a trend for overcoming MDR cancer. In light of ixabepilone, chemical modifications of paclitaxel and vinblastine have also been attempted in succession, producing cabazitaxel and ortataxel, and vinflunine, respectively (Table 1). After these modifications, increased cytotoxic effects were observed in P-gp-overexpressing cell lines (43).

TABLE 1
www.frontiersin.org

Table 1. Drugs that reverse chemoresistance via various mechanisms.

Targeting Aldehyde Dehydrogenases

ALDH plays an important role in the differentiation of stem cells by converting retinol into retinoic acid, as well as detoxifying the cells by converting aldehyde into carboxylic acid, and thus ALDH is considered a biomarker for stem cells (11, 94). Several studies have shown that ALDH-1 is correlated with CSCs. Gefitinib, an EGFR inhibitor, is used for breast-, lung-, and other cancers. However, it has been confirmed that ALDH1A1-positive CSCs are more resistant to gefitinib than ALDH1A1-negative CSCs (95). ALDH-1 expression is mediated by high expression of Snail, which regulates metastasis as a transcription factor and subsequently causes CSCs to develop resistance to chemotherapy (96). To solve this problem, ALDH inhibitors, such as diethylaminobenzaldehyde, disulfiram, and tretinoin, have been proposed (Table 1). Bromodomain and extra-terminal (BET) inhibitors such as JQ1 have also been proposed due to their ability to suppress ALDH1A1 expression (97).

Targeting Epithelial-Mesenchymal Transition

EMT is a process by which epithelial cells become mesenchymal stem cells. In this process, epithelial cells lose cell polarity and cell-cell adhesion function but gain migratory and invasive functions (98, 99). In other words, EMT causes cancer cells to exhibit stem-like features such as tumorigenicity. EMT also promotes metastasis that occurs mainly due to reduced expression of E-cadherin, which itself is directly repressed by Snail (100). Because EMT plays a crucial role in chemoresistance of CSCs, strategies to inhibit this process can be effective. Some studies have shown that AMP-activated kinase (AMPK) induces apoptosis of cancer cells, inhibits TGF-β-induced EMT and, consequently, can reverse drug resistance (101). AMPK activators that have been proposed include metformin and thalidezine (Table 1). Although the data are still insufficient, these activators seem to have enough potential to overcome chemoresistance. It has also been proposed that histone deacetylase (HDAC) inhibitors suppress EMT and attenuate chemoresistance (102).

Targeting Epigenetic Modifications

Histone acetylation, one of the most common post-translational modifications, is closely associated with CSC chemoresistance. This process, in which lysine residues are acetylated, is tightly regulated by histone acetyltransferases (HATs) and HDACs. Bromodomains (BRD) of the BET family proteins read acetyl-lysine (Kac) residues on histones and regulate gene expression. BET family proteins recruit the positive transcriptional elongation factor (P-TEFb), resulting in a transcriptional cascade of oncogenes (54). Thus, inhibition of BET binding to acetylated histones suppresses cell proliferation and induces apoptosis. JQ1 is a potent, highly specific, and Kac competitive inhibitor for BET family proteins (97). Because JQ1 has a short half-life, its derivatives and structurally similar forms are undergoing clinical trials (Table 1). On the other hand, HDACs, which are epigenetic erasers, induce stem-like features by promoting EMT. HDACs also affect hypoxia-inducible factors (HIFs) and NF-κB related to apoptosis (103, 104), which are components of the tumor microenvironment. HDAC inhibitors, including vorinostat (SAHA) and panobinostat, suppress EGFR expression and reverse EMT (Table 1). In addition, these drugs have been successfully used in combination with BET inhibitors. Such epigenetic combination therapy improves clinical efficacy by reducing Myc expression. Lysine-specific demethylase 1 (KDM1, also known as LSD1) modulates histone methylation, and demethylation of H3K9me2, H3K4me3, and H3K36me3 contributes to KDM1-mediated chemoresistance (105). Inhibition of KDM1 activity was initially observed in monoamine oxidase (MAO) inhibitors such as tranylcypromine, which was accompanied by suppression of the stem cell properties of CSCs in vivo (106). Based on their structure, several selective KDM1 inhibitors have been developed (Table 1). In addition, the activity of KDM1 can be regulated by HDAC inhibitors, as crosstalk exists between KDM1 and HDAC. KDM1 and HDAC1/2 form the CoREST complex, which is associated with silencing gene expression. As a result, combination therapy with KDM1 inhibitors and HDAC inhibitors has been expected to have synergistic effects and has often been evaluated. Recently, KDM1-HDAC dual inhibitors such as corin have been reported (107). BMI1 and EZH2, which induce epigenetic silencing as polycomb group (PcG) members, have been reported to be associated with chemoresistance (108, 109). Besides, tumor suppressor genes are silenced by hypermethylation of promoter regions of DNA, and thus, chemotherapy loses its efficacy against many cancers (110).

Targeting Microenvironment

The cellular microenvironment plays an important role in determining cellular behavior. NSCs and CSCs are generated, maintained, and regulated within this microenvironment. The tumor microenvironment (TME) creates a niche for itself that influences not only the proliferation and differentiation of CSCs but also the response to drugs (111). Although cancer-associated fibroblasts (CAFs) as well as inflammation and immune cells are components of the TME, we have to discuss the crucial role of hypoxia in the TME. Hypoxia signaling contributes to chemoresistance of CSCs by increasing the expression of ABC transporters and ALDH (8, 112, 113). In solid tumors, hypoxic regions are necessarily present and lead to angiogenesis through HIF1A and VEGF (114, 115). However, tumor angiogenesis is sloppy; hence, drugs do not reach effective concentrations in hypoxic cells. Thus, VEGF inhibitors such as bevacizumab and receptor tyrosine kinase (RTK) inhibitors such as sorafenib can enhance chemosensitivity. Vascular disrupting agents (VDAs), including tubulin-binding agents such as combretastatin A4 and plocabulin, also enhance chemosensitivity by increasing vessel permeability (Table 1). Of course, we should be careful not to inject these agents prior to chemotherapy because it could impair the delivery of chemotherapeutic drugs by reducing blood supply.

Targeting Signaling Pathways

Signaling pathways control cell responses, particularly the self-renewal, differentiation, and survival of CSCs. Among them, the Notch, Hedgehog (Hh), and Wnt/β-catenin signaling pathways are responsible for drug resistance (116). Notch induces the expression of survivin, which is an anti-apoptotic gene that inhibits apoptosis. This signaling is suppressed by γ-secretase inhibitors (GSIs) such as DAPT and RO4929097, which block the second cleavage of Notch receptors and release of the Notch-IC fragment from the cell membrane (Table 1). Hh signaling is critical in embryogenesis and has been found in many cancers (75, 117). This signaling induces the activation of Smo and Gli1, which are involved in the drug resistance caused by overexpression of P-gp and BCRP (118). Thus, Smo inhibitors such as cyclopamine, vismodegib, and sonidegib can suppress chemoresistance (Table 1). Wnt/β-catenin signaling activated by Frizzleds (FZDs) is also associated with drug resistance due to overexpression of P-gp, BCRP, and MRP (119). Besides, this pathway promotes cell cycling, inhibits apoptosis, and mediates DNA repair processes. Although it has been suggested that the Wnt/β-catenin signaling pathway is undruggable, strategies to target it have been explored because of their various therapeutic potential. Indeed, there has been an effort to develop drugs that inhibit Wnt/β-catenin signaling (Table 1).

Future Perspectives

To date, many chemotherapeutic drugs have been developed and many studies conducted to overcome MDR. The newer strategies that we have covered above have limitations, however, and so more work is needed with different approaches being explored to find effective and lasting treatment. Any strategy for overcoming MDR should not affect NCSs, but only CSCs. There are a few things to pay attention to for selective targeting of CSCs: (i) encapsulation of anticancer drugs in liposomes, micelles, and nanoparticles, i.e., nanotechnology-based drug delivery (120). The improvement of pharmacokinetic properties, tumor-specific delivery due to the enhanced permeability and retention (EPR) effect, and resistance to efflux pumps because of the size-exclusion effect are representative advantages of nanomaterials (NMs) (121123). Thus, NMs are expected to increase the therapeutic effect and reduce unwanted side effects of anticancer drugs. Moreover, dual drug-loaded NMs can lead to better therapeutic effects (124). For example, pluronics, which are amphiphilic polymers and form nano-sized micellar structures, have been increasingly regarded as CSC modulators (125, 126). Pluronics are not just drug delivery carriers; rather, they inhibit metastasis and activate apoptosis by mediating the release of cytochrome c and apoptosis inducing factor (AIF). In addition, they can alter the microenvironment and suppress CSCs effectively. At present, only SP1049C (doxorubicin with pluronics L61 and F127 micelles) is in phase 3 trials (127), but there are reasons to be optimistic. (ii) Targeting microRNAs (miRNAs), which mediate translational repression and mRNA degradation mainly by binding to the 3′ UTR. Previous studies have shown that miRNAs regulate ABC transporters in CSCs (128, 129); for instance, miR-212, miR-328, miR-451, and so on. Some miRNAs, including miR-21 and miR-222, are oncogenic and up-regulated in cancer cells, while other miRNAs, including miR-15 and miR-181, are tumor suppressive and down-regulated in cancer cells (29, 130). (iii) Targeting MDR mRNA (131, 132), through which antisense oligonucleotides (aODNs) are used to down-regulate the expression of specific genes. Small interfering RNAs (siRNAs) also silence specific genes because they are artificial double-stranded RNA (dsRNA) molecules with functions similar to those of miRNA. If they selectively target tumorigenic genes, cancer cells will become chemosensitive. There are no clinical data yet to suggest that MDR can be completely reversed using this RNA interference (RNAi) technology, but it is likely that drugs related to RNAi will be developed in the near future. (iv) Transferring MDR genes into NSCs, especially bone marrow stem cells (133135). Most chemotherapeutic drugs suppress the bone marrow and, as a result, blood cells such as erythrocytes, leukocytes, and thrombocytes do not function properly, and the ensuing loss of oxygen transport, immune response, and bleeding control functions causes serious adverse effects. To prevent these adverse effects, gene transfer technology has emerged, making it possible to administer high doses of chemotherapeutic drugs. In the transplantation model of CD34-positive peripheral blood stem cell (PBSC) infection with retroviral vectors containing MDR genes, long-term myeloprotection was achieved, demonstrating the safety of transplantation (136).

Conclusion

Over the past few decades there have been many strategies used to treat cancer, from conventional radiation therapy, chemotherapy to recent targeted therapy, and immunotherapy. It is no exaggeration to say that the flow of this change has been led by the discovery of various factors that cause MDR acquisition in cancer cells, especially ABC transporters. By discovering ABC transporters and identifying their functions, new possibilities have emerged for cancer treatment. Initially, many studies were focused on the direct inhibition of ABC transporters such as P-gp inhibitors. However, because they are less selective and less potent, differ in their in vitro and in vivo data, and often cause severe adverse effects, so far no drugs that directly target or inhibit P-gp have been accepted for clinical use. Although direct inhibition of ABC transporters may not be effective, the transporters remain attractive targets because enhanced drug efflux through the transporters is one of the most important causes of MDR acquisition. Therefore, rather than directly inhibiting the ABC transporters, it would be effective to devise alternative strategies to avoid drug efflux via transporters. The transporter-mediated MDR might be overcome by developing novel anticancer drugs with P-gp non-substrates. In addition, we have introduced alternative approaches for targeting CSCs that focus on ALDHs, EMT, epigenetic modifications, the microenvironment, and signaling pathways (Figure 1). We have also discussed the drugs being developed in each approach: (i) ABC transporter inhibitors and non-substrates; (ii) ALDH inhibitors; (iii) AMPK activators; (iv) BET inhibitors, HDAC inhibitors, and KDM1 inhibitors; (v) VEGF inhibitors, RTK inhibitors, and vascular disrupting agents; (vi) Notch inhibitors, Smo inhibitors, and Wnt/β-catenin inhibitors. One of the big obstacles we are facing now is whether we can selectively influence CSCs only. NSCs and CSCs share the same characteristics; therefore, solving this problem is a complex task. Although promising solutions have been proposed, more research should be conducted to support the arguments relating to nanotechnology-based drug delivery, RNAi technology, and gene transfer technology. It is worth noting that these approaches are directly or indirectly related to ABC transporters. It is hoped that all of the approaches reviewed here will help devise new strategies to overcome MDR and to eradicate MDR cancer.

FIGURE 1
www.frontiersin.org

Figure 1. Current strategies to overcome multidrug resistance. In addition to the well-known ABC transporters, many drug resistance mechanisms of CSCs have been identified, including ALDHs, EMT, epigenetic changes, tumor microenvironment, and stemness-related signaling pathways. The drugs that inhibit each pathway are described: (i) ABC transporter inhibitors and non-substrates; (ii) ALDH inhibitors; (iii) AMPK activators; (iv) BET inhibitors, HDAC inhibitors, and KDM1 inhibitors; (v) VEGF inhibitors, RTK inhibitors, and vascular disrupting agents; (VI) Notch inhibitors, Smo inhibitors, and Wnt/β-catenin inhibitors. ABC, ATP binding cassette; ALDHs, aldehyde dehydrogenases; AMPK, AMP-activated kinase; BET, bromodomain and extra-terminal; CSCs, cancer stem cells; EMT, epithelial to mesenchymal transition; HDAC, histone deacetylase; KDM1, lysine-specific demethylase 1; MDR, multidrug resistance; RTK, receptor tyrosine kinase; VEGF, vascular endothelial growth factor.

Author Contributions

YC and YK participated in manuscript drafting and revision of this article.

Funding

This work was supported by a National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIP) [NRF-2018R1A2B2005646 (to YK)].

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We would like to thank Bokyeong Kim for providing the illustration used herein.

References

1. Thomas H, Coley HM. Overcoming multidrug resistance in cancer: an update on the clinical strategy of inhibiting p-glycoprotein. Cancer Control. (2003) 10:159–65. doi: 10.1177/107327480301000207

PubMed Abstract | CrossRef Full Text | Google Scholar

2. Ambudkar SV, Kimchi-Sarfaty C, Sauna ZE, Gottesman MM. P-glycoprotein: from genomics to mechanism. Oncogene. (2003) 22:7468–85. doi: 10.1038/sj.onc.1206948

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Moitra K. Overcoming multidrug resistance in cancer stem cells. BioMed Res Int. (2015) 2015:635745–8. doi: 10.1155/2015/635745

PubMed Abstract | CrossRef Full Text | Google Scholar

4. Pan ST, Li ZL, He ZX, Qiu JX, Zhou SF. Molecular mechanisms for tumour resistance to chemotherapy. Clin Exp Pharmacol Physiol. (2016) 43:723–37. doi: 10.1111/1440-1681.12581

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Ambudkar SV, Dey S, Hrycyna CA, Ramachandra M, Pastan I, Gottesman MM. Biochemical, cellular, and pharmacological aspects of the multidrug transporter. Annu Rev Pharmacol Toxicol. (1999) 39:361–98. doi: 10.1146/annurev.pharmtox.39.1.361

PubMed Abstract | CrossRef Full Text | Google Scholar

6. Hirschmann-Jax C, Foster AE, Wulf GG, Nuchtern JG, Jax TW, Gobel U, et al. A distinct “side population” of cells with high drug efflux capacity in human tumor cells. Proc Natl Acad Sci USA. (2004) 101:14228–33. doi: 10.1073/pnas.0400067101

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Jung Y, Kim W. Cancer stem cell targeting: are we there yet? Arch Pharm Res. (2015) 38:414–22. doi: 10.1007/s12272-015-0570-2

PubMed Abstract | CrossRef Full Text | Google Scholar

8. Ma L, Lai D, Liu T, Cheng W, Guo L. Cancer stem-like cells can be isolated with drug selection in human ovarian cancer cell line SKOV3. Acta Biochim Biophys Sin (Shanghai). (2010) 42:593–602. doi: 10.1093/abbs/gmq067

PubMed Abstract | CrossRef Full Text | Google Scholar

9. Moore N, Lyle S. Quiescent, slow-cycling stem cell populations in cancer: a review of the evidence and discussion of significance. J Oncol. (2011) 2011:396076. doi: 10.1155/2011/396076

PubMed Abstract | CrossRef Full Text | Google Scholar

10. Gottesman MM, Fojo T, Bates SE. Multidrug resistance in cancer: role of ATP-dependent transporters. Nat Rev Cancer. (2002) 2:48–58. doi: 10.1038/nrc706

PubMed Abstract | CrossRef Full Text | Google Scholar

11. Raha D, Wilson TR, Peng J, Peterson D, Yue P, Evangelista M, et al. The cancer stem cell marker aldehyde dehydrogenase is required to maintain a drug-tolerant tumor cell subpopulation. Cancer Res. (2014) 74:3579–90. doi: 10.1158/0008-5472.CAN-13-3456

PubMed Abstract | CrossRef Full Text | Google Scholar

12. Shibue T, Weinberg RA. EMT, CSCs, and drug resistance: the mechanistic link and clinical implications. Nat Rev Clin Oncol. (2017) 14:611–29. doi: 10.1038/nrclinonc.2017.44

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Wainwright EN, Scaffidi P. Epigenetics and cancer stem cells: unleashing, hijacking, and restricting cellular plasticity. Trends Cancer. (2017) 3:372–86. doi: 10.1016/j.trecan.2017.04.004

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Munoz P, Iliou MS, Esteller M. Epigenetic alterations involved in cancer stem cell reprogramming. Mol Oncol. (2012) 6:620–36. doi: 10.1016/j.molonc.2012.10.006

PubMed Abstract | CrossRef Full Text | Google Scholar

15. Bao B, Azmi AS, Ali S, Ahmad A, Li Y, Banerjee S, et al. The biological kinship of hypoxia with CSC and EMT and their relationship with deregulated expression of miRNAs and tumor aggressiveness. Biochim Biophys Acta. (2012) 1826:272–96. doi: 10.1016/j.bbcan.2012.04.008

PubMed Abstract | CrossRef Full Text | Google Scholar

16. Dontu G, Jackson KW, McNicholas E, Kawamura MJ, Abdallah WM, Wicha MS. Role of Notch signaling in cell fate determination of human mammary stemprogenitor cells. Breast Cancer Res. (2004) 6:R605–15. doi: 10.1186/bcr920

PubMed Abstract | CrossRef Full Text | Google Scholar

17. Yang L, Xie G, Fan Q, Xie J. Activation of the hedgehog-signaling pathway in human cancer and the clinical implications. Oncogene. (2010) 29:469–81. doi: 10.1038/onc.2009.392

PubMed Abstract | CrossRef Full Text | Google Scholar

18. Yang W, Yan HX, Chen L, Liu Q, He YQ, Yu LX, et al. Wnt/beta-catenin signaling contributes to activation of normal and tumorigenic liver progenitor cells. Cancer Res. (2008) 68:4287–95. doi: 10.1158/0008-5472.CAN-07-6691

PubMed Abstract | CrossRef Full Text | Google Scholar

19. Wu CP, Calcagno AM, Ambudkar SV. Reversal of ABC drug transporter-mediated multidrug resistance in cancer cells: evaluation of current strategies. Curr Mol Pharmacol. (2008) 1:93–105. doi: 10.2174/1874467210801020093

PubMed Abstract | CrossRef Full Text | Google Scholar

20. Krishna R, Mayer LD. Multidrug resistance (MDR) in cancer. Mechanisms, reversal using modulators of MDR and the role of MDR modulators in influencing the pharmacokinetics of anticancer drugs. Eur J Pharm Sci. (2000) 11:265–83. doi: 10.1016/s0928-0987(00)00114-7

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Chung FS, Santiago JS, Jesus MF, Trinidad CV, See MF. Disrupting P-glycoprotein function in clinical settings: what can we learn from the fundamental aspects of this transporter? Am J Cancer Res. (2016) 6:1583–98.

PubMed Abstract | Google Scholar

22. Chao NJ, Aihara M, Blume KG, Sikic BI. Modulation of etoposide (VP-16) cytotoxicity by verapamil or cyclosporine in multidrug-resistant human leukemic cell lines and normal bone marrow. Exp Hematol. (1990) 18:1193–8.

PubMed Abstract | Google Scholar

23. Jin Y, Bin ZQ, Qiang H, Liang C, Hua C, Jun D, 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) 135:1369–76. doi: 10.1007/s00432-009-0578-4

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Solary E, Velay I, Chauffert B, Bidan J, Caillot D, Dumas M, et al. Sufficient levels of quinine in the serum circumvent the multidrug resistance of the human leukemic cell line K562/ADM. Cancer. (1991) 68:1714–9. doi: 10.1002/1097-0142(19911015)68:8<1714::AID-CNCR2820680811>3.0.CO;2-2

PubMed Abstract | CrossRef Full Text | Google Scholar

25. Li C, Kim M, Choi H, Choi J. Effects of baicalein on the pharmacokinetics of tamoxifen and its main metabolite, 4-hydroxytamoxifen, in rats: possible role of cytochrome P450 3A4 and P-glycoprotein inhibition by baicalein. Arch Pharm Res. (2011) 34:1965–72. doi: 10.1007/s12272-011-1117-9

PubMed Abstract | CrossRef Full Text | Google Scholar

26. Spoelstra EC, Westerhoff HV, Pinedo HM, Dekker H, Lankelma J. The multidrug-resistance-reverser verapamil interferes with cellular P-glycoprotein-mediated pumping of daunorubicin as a non-competing substrate. Eur J Biochem. (1994) 221:363–73. doi: 10.1111/j.1432-1033.1994.tb18748.x

PubMed Abstract | CrossRef Full Text | Google Scholar

27. Solary E, Mannone L, Moreau D, Caillot D, Casasnovas RO, Guy H, et al. Phase I study of cinchonine, a multidrug resistance reversing agent, combined with the CHVP regimen in relapsed and refractory lymphoproliferative syndromes. Leukemia. (2000) 14:2085–94. doi: 10.1038/sj.leu.2401945

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Wilson WH, Jamis-Dow C, Bryant G, Balis FM, Klecker RW, Bates SE, et al. Phase I and pharmacokinetic study of the multidrug resistance modulator dexverapamil with EPOCH chemotherapy. J Clin Oncol. (1995) 13:1985–94. doi: 10.1200/JCO.1995.13.8.1985

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Mubashar M, Harrington KJ, Chaudhary KS, Lalani E, Stamp GW, Peters AM. Differential effects of toremifene on doxorubicin, vinblastine and Tc-99m-sestamibi in P-glycoprotein-expressing breast and head and neck cancer cell lines. Acta Oncol. (2004) 43:443–52. doi: 10.1080/02841860410031048

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Twentyman PR, Bleehen NM. Resistance modification by PSC-833, a novel non-immunosuppressive cyclosporin [corrected]. Eur J Cancer. (1991) 27:1639–42. doi: 10.1016/0277-5379(91)90435-G

PubMed Abstract | CrossRef Full Text | Google Scholar

31. Gandhi L, Harding MW, Neubauer M, Langer CJ, Moore M, Ross HJ, et al. A phase II study of the safety and efficacy of the multidrug resistance inhibitor VX-710 combined with doxorubicin and vincristine in patients with recurrent small cell lung cancer. Cancer. (2007) 109:924–32. doi: 10.1002/cncr.22492

PubMed Abstract | CrossRef Full Text | Google Scholar

32. Nakanishi O, Baba M, Saito A, Yamashita T, Sato W, Abe H, et al. Potentiation of the antitumor activity by a novel quinoline compound, MS-209, in multidrug-resistant solid tumor cell lines. Oncol Res. (1997) 9:61–9.

PubMed Abstract | Google Scholar

33. van Zuylen L, Sparreboom A, van der Gaast A, Nooter K, Eskens FA, Brouwer E, et al. Disposition of docetaxel in the presence of P-glycoprotein inhibition by intravenous administration of R101933. Eur J Cancer. (2002) 38:1090–9. doi: 10.1016/S0959-8049(02)00035-7

PubMed Abstract | CrossRef Full Text | Google Scholar

34. Shi B, Yaremko B, Hajian G, Terracina G, Bishop WR, Liu M, et al. The farnesyl protein transferase inhibitor SCH66336 synergizes with taxanes in vitro and enhances their antitumor activity in vivo. Cancer Chemother Pharmacol. (2000) 46:387–93. doi: 10.1007/s002800000170

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Green LJ, Marder P, Slapak CA. Modulation by LY335979 of P-glycoprotein function in multidrug-resistant cell lines and human natural killer cells. Biochem Pharmacol. (2001) 61:1393–9. doi: 10.1016/S0006-2952(01)00599-8

PubMed Abstract | CrossRef Full Text | Google Scholar

36. de Bruin M, Miyake K, Litman T, Robey R, Bates SE. Reversal of resistance by GF120918 in cell lines expressing the ABC half-transporter, MXR. Cancer Lett. (1999) 146:117–26. doi: 10.1016/S0304-3835(99)00182-2

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Lusvarghi S, Ambudkar SV. ATP-dependent thermostabilization of human P-glycoprotein (ABCB1) is blocked by modulators. Biochem J. (2019) 476:3737–50. doi: 10.1042/BCJ20190736

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Rabindran SK, He H, Singh M, Brown E, Collins KI, Annable T, et al. Reversal of a novel multidrug resistance mechanism in human colon carcinoma cells by fumitremorgin C. Cancer Res. (1998) 58:5850–8.

PubMed Abstract | Google Scholar

39. Ghosh P, Moitra K, Maki N, Dey S. Allosteric modulation of the human P-glycoprotein involves conformational changes mimicking catalytic transition intermediates. Arch Biochem Biophys. (2006) 450:100–12. doi: 10.1016/j.abb.2006.02.025

PubMed Abstract | CrossRef Full Text | Google Scholar

40. Fornier MN. Ixabepilone, first in a new class of antineoplastic agents: the natural epothilones and their analogues. Clin Breast Cancer. (2007) 7:9–15. doi: 10.3816/CBC.2007.n.036

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Di Nunno V, Mollica V, Massari F. Cabazitaxel in metastatic prostate cancer. N Engl J Med. (2020) 382:1286. doi: 10.1056/NEJMc2000990

PubMed Abstract | CrossRef Full Text | Google Scholar

42. Brooks TA, Minderman H, O'Loughlin KL, Pera P, Ojima I, Baer MR, et al. Taxane-based reversal agents modulate drug resistance mediated by P-glycoprotein, multidrug resistance protein, and breast cancer resistance protein. Mol Cancer Ther. (2003) 2:1195–205.

PubMed Abstract | Google Scholar

43. Waghray D, Zhang Q. Inhibit or evade multidrug resistance P-glycoprotein in cancer treatment. J Med Chem. (2018) 61:5108–21. doi: 10.1021/acs.jmedchem.7b01457

PubMed Abstract | CrossRef Full Text | Google Scholar

44. MacDonagh L, Gallagher MF, Ffrench B, Gasch C, Breen E, Gray SG, et al. Targeting the cancer stem cell marker, aldehyde dehydrogenase 1, to circumvent cisplatin resistance in NSCLC. Oncotarget. (2017) 8:72544–63. doi: 10.18632/oncotarget.19881

PubMed Abstract | CrossRef Full Text | Google Scholar

45. Toledo-Guzman ME, Hernandez MI, Gomez-Gallegos AA, Ortiz-Sanchez E. ALDH as a stem cell marker in solid tumors. Curr Stem Cell Res Ther. (2019) 14:375–88. doi: 10.2174/1574888X13666180810120012

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Moreb JS, Ucar-Bilyeu DA, Khan A. Use of retinoic acid/aldehyde dehydrogenase pathway as potential targeted therapy against cancer stem cells. Cancer Chemother Pharmacol. (2017) 79:295–301. doi: 10.1007/s00280-016-3213-5

PubMed Abstract | CrossRef Full Text | Google Scholar

47. Choi YK, Park KG. Metabolic roles of AMPK and metformin in cancer cells. Mol Cells. (2013) 36:279–87. doi: 10.1007/s10059-013-0169-8

PubMed Abstract | CrossRef Full Text | Google Scholar

48. Qu C, Qu C, Zhang W, Zhang W, Zheng G, Zheng G, et al. Metformin reverses multidrug resistance and epithelial–mesenchymal transition (EMT) via activating AMP-activated protein kinase (AMPK) in human breast cancer cells. Mol Cell Biochem. (2014) 386:63–71. doi: 10.1007/s11010-013-1845-x

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Law BYK, Gordillo-Martínez F, Qu YQ, Zhang N, Xu SW, Coghi PS, et al. Thalidezine, a novel AMPK activator, eliminates apoptosis-resistant cancer cells through energy-mediated autophagic cell death. Oncotarget. (2017) 8:30077–91. doi: 10.18632/oncotarget.15616

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Heijmen L, Ter Voert EG, Punt CJ, Heerschap A, Oyen WJ, Bussink J, et al. Monitoring hypoxia and vasculature during bevacizumab treatment in a murine colorectal cancer model. Contrast Media Mol Imaging. (2014) 9:237–45. doi: 10.1002/cmmi.1564

PubMed Abstract | CrossRef Full Text | Google Scholar

51. Qiu Y, Shan W, Yang Y, Jin M, Dai Y, Yang H, et al. Reversal of sorafenib resistance in hepatocellular carcinoma: epigenetically regulated disruption of 14-3-3η/hypoxia-inducible factor-1α. Cell Death Discov. (2019) 5:120. doi: 10.1038/s41420-019-0200-8

PubMed Abstract | CrossRef Full Text | Google Scholar

52. Close A. Antiangiogenesis and vascular disrupting agents in cancer: circumventing resistance and augmenting their therapeutic utility. Future Med Chem. (2016) 8:443–62. doi: 10.4155/fmc.16.6

PubMed Abstract | CrossRef Full Text | Google Scholar

53. Horsman MR, Siemann DW. Pathophysiologic effects of vascular-targeting agents and the implications for combination with conventional therapies. Cancer Res. (2006) 66:11520–39. doi: 10.1158/0008-5472.CAN-06-2848

PubMed Abstract | CrossRef Full Text | Google Scholar

54. Alqahtani A, Choucair K, Ashraf M, Hammouda DM, Alloghbi A, Khan T, et al. Bromodomain and extra-terminal motif inhibitors: a review of preclinical and clinical advances in cancer therapy. Future Sci OA. (2019) 5:FSO372. doi: 10.4155/fsoa-2018-0115

PubMed Abstract | CrossRef Full Text | Google Scholar

55. Bolden JE, Tasdemir N, Dow LE, van Es JH, Wilkinson JE, Zhao Z, et al. Inducible in vivo silencing of Brd4 identifies potential toxicities of sustained BET protein inhibition. Cell Rep. (2014) 8:1919–29. doi: 10.1016/j.celrep.2014.08.025

PubMed Abstract | CrossRef Full Text | Google Scholar

56. Siu KT, Ramachandran J, Yee AJ, Eda H, Santo L, Panaroni C, et al. Preclinical activity of CPI-0610, a novel small-molecule bromodomain and extra-terminal protein inhibitor in the therapy of multiple myeloma. Leukemia. (2017) 31:1760–9. doi: 10.1038/leu.2016.355

PubMed Abstract | CrossRef Full Text | Google Scholar

57. Long J, Li B, Rodriguez-Blanco J, Pastori C, Volmar C, Wahlestedt C, et al. The BET bromodomain inhibitor I-BET151 acts downstream of smoothened protein to abrogate the growth of hedgehog protein-driven cancers. J Biol Chem. (2014) 289:35494–502. doi: 10.1074/jbc.M114.595348

PubMed Abstract | CrossRef Full Text | Google Scholar

58. Coude MM, Braun T, Berrou J, Dupont M, Bertrand S, Masse A, et al. BET inhibitor OTX015 targets BRD2 and BRD4 and decreases c-MYC in acute leukemia cells. Oncotarget. (2015) 6:17698–712. doi: 10.18632/oncotarget.4131

PubMed Abstract | CrossRef Full Text | Google Scholar

59. Boi M, Gaudio E, Bonetti P, Kwee I, Bernasconi E, Tarantelli C, et al. The BET bromodomain inhibitor OTX015 affects pathogenetic pathways in preclinical B-cell tumor models and synergizes with targeted drugs. Clin Cancer Res. (2015) 21:1628–38. doi: 10.1158/1078-0432.CCR-14-1561

PubMed Abstract | CrossRef Full Text | Google Scholar

60. Bui MH, Lin X, Albert DH, Li L, Lam LT, Faivre EJ, et al. Preclinical characterization of BET family bromodomain inhibitor ABBV-075 suggests combination therapeutic strategies. Cancer Res. (2017) 77:2976–89. doi: 10.1158/0008-5472.CAN-16-1793

PubMed Abstract | CrossRef Full Text | Google Scholar

61. Zhang D, Leal AS, Carapellucci S, Zydeck K, Sporn MB, Liby KT. Chemoprevention of preclinical breast and lung cancer with the bromodomain inhibitor I-BET 762. Cancer Prev Res. (2018) 11:143–56. doi: 10.1158/1940-6207.CAPR-17-0264

PubMed Abstract | CrossRef Full Text | Google Scholar

62. Bruzzese F, Leone A, Rocco M, Carbone C, Piro G, Caraglia M, et al. HDAC inhibitor vorinostat enhances the antitumor effect of gefitinib in squamous cell carcinoma of head and neck by modulating ErbB receptor expression and reverting EMT. J Cell Physiol. (2011) 226:2378–90. doi: 10.1002/jcp.22574

PubMed Abstract | CrossRef Full Text | Google Scholar

63. Di Costanzo A, Del Gaudio N, Migliaccio A, Altucci L. Epigenetic drugs against cancer: an evolving landscape. Arch Toxicol. (2014) 88:1651–68. doi: 10.1007/s00204-014-1315-6

PubMed Abstract | CrossRef Full Text | Google Scholar

64. Ling Y, Liu J, Qian J, Meng C, Guo J, Gao W, et al. Recent advances in multi-target drugs targeting protein kinases and histone deacetylases in cancer therapy. Curr Med Chem. (2020). doi: 10.2174/0929867327666200102115720. [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

65. Binda C, Valente S, Romanenghi M, Pilotto S, Cirilli R, Karytinos A, et al. Biochemical, structural, and biological evaluation of tranylcypromine derivatives as inhibitors of histone demethylases LSD1 and LSD2. J Am Chem Soc. (2010) 132:6827–33. doi: 10.1021/ja101557k

PubMed Abstract | CrossRef Full Text | Google Scholar

66. Mohammad HP, Smitheman KN, Kamat CD, Soong D, Federowicz KE, Van Aller GS, et al. A DNA hypomethylation signature predicts antitumor activity of LSD1 inhibitors in SCLC. Cancer Cell. (2015) 28:57–69. doi: 10.1016/j.ccell.2015.06.002

PubMed Abstract | CrossRef Full Text | Google Scholar

67. Fu X, Zhang P, Yu B. Advances toward LSD1 inhibitors for cancer therapy. Future Med Chem. (2017) 9:1227–42. doi: 10.4155/fmc-2017-0068

PubMed Abstract | CrossRef Full Text | Google Scholar

68. Shaw HV, Koval A, Katanaev VL. Targeting the Wnt signalling pathway in cancer: prospects and perils. Swiss Med Wkly. (2019) 149:w20129. doi: 10.4414/smw.2019.20129

PubMed Abstract | CrossRef Full Text | Google Scholar

69. Maes T, Mascaro C, Tirapu I, Estiarte A, Ciceri F, Lunardi S, et al. ORY-1001, a potent and selective covalent KDM1A inhibitor, for the treatment of acute leukemia. Cancer Cell. (2018) 33:495–511.e12. doi: 10.1016/j.ccell.2018.02.002

PubMed Abstract | CrossRef Full Text | Google Scholar

70. Hollebecque A, de Bono JS, Plummer R, Isambert N, Martin-Romano P, Baudin E, et al. Phase I study of CC-90011 in patients with advanced solid tumors and relapsed/refractory non-Hodgkin lymphoma (R/R NHL). Ann Oncol. (2019) 29(Suppl. 8):viii649–69. doi: 10.1093/annonc/mdy303

PubMed Abstract | CrossRef Full Text | Google Scholar

71. Raffaella Soldi, Tithi, Ghosh Halder, Alexis Weston, Trason Thode, Kevin Drenner, Rhonda Lewis, et al. The novel reversible LSD1 inhibitor SP-2577 promotes anti-tumor immunity in SWItch/Sucrose-NonFermentable (SWI/SNF) complex mutated ovarian cancer. bioRxiv. doi: 10.1101/2020.01.10.902528

PubMed Abstract | CrossRef Full Text | Google Scholar

72. Young MJ, Wu YH, Chiu WT, Weng TY, Huang YF, Chou CY. All-trans retinoic acid downregulates ALDH1-mediated stemness and inhibits tumour formation in ovarian cancer cells. Carcinogenesis. (2015) 36:498–507. doi: 10.1093/carcin/bgv018

PubMed Abstract | CrossRef Full Text | Google Scholar

73. Meng RD, Shelton CC, Li Y, Qin L, Notterman D, Paty PB, et al. gamma-secretase inhibitors abrogate oxaliplatin-induced activation of the notch-1 signaling pathway in colon cancer cells resulting in enhanced chemosensitivity. Cancer Res. (2009) 69:573–82. doi: 10.1158/0008-5472.CAN-08-2088

PubMed Abstract | CrossRef Full Text | Google Scholar

74. Lee SM, Moon J, Redman BG, Chidiac T, Flaherty LE, Zha Y, et al. Phase 2 study of RO4929097, a gamma-secretase inhibitor, in metastatic melanoma: SWOG 0933. Cancer. (2015) 121:432–40. doi: 10.1002/cncr.29055

PubMed Abstract | CrossRef Full Text | Google Scholar

75. Song Z, Yue W, Wei B, Wang N, Li T, Guan L, et al. Sonic hedgehog pathway is essential for maintenance of cancer stem-like cells in human gastric cancer. PLoS ONE. (2011) 6:e17687. doi: 10.1371/journal.pone.0017687

PubMed Abstract | CrossRef Full Text | Google Scholar

76. Dirix L. Discovery and exploitation of novel targets by approved drugs. J Clin Oncol. (2014) 32:720–1. doi: 10.1200/JCO.2013.53.7118

PubMed Abstract | CrossRef Full Text | Google Scholar

77. Dummer R, Ascierto PA, Basset-Seguin N, Dreno B, Garbe C, Gutzmer R, et al. Sonidegib and vismodegib in the treatment of patients with locally advanced basal cell carcinoma: a joint expert opinion. J Eur Acad Dermatol Venereol. (2020). doi: 10.1111/jdv.16230. [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

78. Jimeno A, Gordon M, Chugh R, Messersmith W, Mendelson D, Dupont J, et al. A first-in-human phase I study of the anticancer stem cell agent Ipafricept (OMP-54F28), a decoy receptor for Wnt ligands, in patients with advanced solid tumors. Clin Cancer Res. (2017) 23:7490–7. doi: 10.1158/1078-0432.CCR-17-2157

PubMed Abstract | CrossRef Full Text | Google Scholar

79. Katoh M, Katoh M. Molecular genetics and targeted therapy of WNT-related human diseases (Review). Int J Mol Med. (2017) 40:587–606. doi: 10.3892/ijmm.2017.3071

PubMed Abstract | CrossRef Full Text | Google Scholar

80. Katoh M. Canonical and non-canonical WNT signaling in cancer stem cells and their niches: cellular heterogeneity, omics reprogramming, targeted therapy and tumor plasticity (Review). Int J Oncol. (2017) 51:1357–69. doi: 10.3892/ijo.2017.4129

PubMed Abstract | CrossRef Full Text | Google Scholar

81. Safholm A, Leandersson K, Dejmek J, Nielsen CK, Villoutreix BO, Andersson T. A formylated hexapeptide ligand mimics the ability of Wnt-5a to impair migration of human breast epithelial cells. J Biol Chem. (2006) 281:2740–9. doi: 10.1074/jbc.M508386200

PubMed Abstract | CrossRef Full Text | Google Scholar

82. Madan B, Ke Z, Harmston N, Ho SY, Frois AO, Alam J, et al. Wnt addiction of genetically defined cancers reversed by PORCN inhibition. Oncogene. (2016) 35:2197–207. doi: 10.1038/onc.2015.280

PubMed Abstract | CrossRef Full Text | Google Scholar

83. Liu J, Pan S, Hsieh MH, Ng N, Sun F, Wang T, et al. Targeting Wnt-driven cancer through the inhibition of porcupine by LGK974. Proc Natl Acad Sci USA. (2013) 110:20224–9. doi: 10.1073/pnas.1314239110

PubMed Abstract | CrossRef Full Text | Google Scholar

84. Pak S, Park S, Kim Y, Park JH, Park CH, Lee KJ, et al. Correction to: the small molecule WNT/beta-catenin inhibitor CWP232291 blocks the growth of castration-resistant prostate cancer by activating the endoplasmic reticulum stress pathway. J Exp Clin Cancer Res. (2019) 38:440–1. doi: 10.1186/s13046-019-1451-1

PubMed Abstract | CrossRef Full Text | Google Scholar

85. Gang EJ, Hsieh YT, Pham J, Zhao Y, Nguyen C, Huantes S, et al. Small-molecule inhibition of CBP/catenin interactions eliminates drug-resistant clones in acute lymphoblastic leukemia. Oncogene. (2014) 33:2169–78. doi: 10.1038/onc.2013.169

PubMed Abstract | CrossRef Full Text | Google Scholar

86. Ritchie TK, Kwon H, Atkins WM. Conformational analysis of human ATP-binding cassette transporter ABCB1 in lipid nanodiscs and inhibition by the antibodies MRK16 and UIC2. J Biol Chem. (2011) 286:39489–96. doi: 10.1074/jbc.M111.284554

PubMed Abstract | CrossRef Full Text | Google Scholar

87. Scotto KW, Johnson RA. Transcription of the multidrug resistance gene MDR1: a therapeutic target. Mol Interv. (2001) 1:117–25.

PubMed Abstract | Google Scholar

88. Scotto KW. Transcriptional regulation of ABC drug transporters. Oncogene. (2003) 22:7496–511. doi: 10.1038/sj.onc.1206950

PubMed Abstract | CrossRef Full Text | Google Scholar

89. Souid S, Aissaoui D, Srairi-Abid N, Essafi-Benkhadir K. Trabectedin (yondelis®) as a therapeutic option in gynecological cancers: a focus on its mechanisms of action, clinical activity, and genomic predictors of drug response. Curr Drug Targets. (2020) 21. doi: 10.2174/1389450121666200128161733. [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

90. Szakacs G, Paterson JK, Ludwig JA, Booth-Genthe C, Gottesman MM. Targeting multidrug resistance in cancer. Nat Rev Drug Discov. (2006) 5:219–34. doi: 10.1038/nrd1984

PubMed Abstract | CrossRef Full Text | Google Scholar

91. Palmeira A, Sousa E, Vasconcelos MH, Pinto MM. Three decades of P-gp inhibitors: skimming through several generations and scaffolds. Curr Med Chem. (2012) 19:1946–2025. doi: 10.2174/092986712800167392

PubMed Abstract | CrossRef Full Text | Google Scholar

92. Amin ML. P-glycoprotein inhibition for optimal drug delivery. Drug Target Insights. (2013) 7:27–34. doi: 10.4137/DTI.S12519

PubMed Abstract | CrossRef Full Text | Google Scholar

93. Shackleton M. Normal stem cells and cancer stem cells: similar and different. Semin Cancer Biol. (2010) 20:85–92. doi: 10.1016/j.semcancer.2010.04.002

PubMed Abstract | CrossRef Full Text | Google Scholar

94. Chute JP, Muramoto GG, Whitesides J, Colvin M, Safi R, Chao NJ, et al. Inhibition of aldehyde dehydrogenase and retinoid signaling induces the expansion of human hematopoietic stem cells. Proc Natl Acad Sci USA. (2006) 103:11707–12. doi: 10.1073/pnas.0603806103

PubMed Abstract | CrossRef Full Text | Google Scholar

95. Huang CP, Tsai MF, Chang TH, Tang WC, Chen SY, Lai HH, et al. ALDH-positive lung cancer stem cells confer resistance to epidermal growth factor receptor tyrosine kinase inhibitors. Cancer Lett. (2013) 328:144–51. doi: 10.1016/j.canlet.2012.08.021

PubMed Abstract | CrossRef Full Text | Google Scholar

96. Zhou W, Lv R, Qi W, Wu D, Xu Y, Liu W. Snail contributes to the maintenance of stem cell like phenotype cells in human pancreatic cancer. PLoS ONE. (2014) 9:e87409. doi: 10.1371/journal.pone.0087409

PubMed Abstract | CrossRef Full Text | Google Scholar

97. Yokoyama Y, Zhu H, Lee JH, Kossenkov AV, Wu SY, Wickramasinghe JM, et al. BET inhibitors suppress ALDH activity by targeting ALDH1A1 super-enhancer in ovarian cancer. Cancer Res. (2016) 76:6320–30. doi: 10.1158/0008-5472.CAN-16-0854

PubMed Abstract | CrossRef Full Text | Google Scholar

98. Celia-Terrassa T, Jolly MK. Cancer stem cells and epithelial-to-mesenchymal transition in cancer metastasis. Cold Spring Harb Perspect Med. (2019) a036905. doi: 10.1101/cshperspect.a036905. [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

99. Oskarsson T, Batlle E, Massague J. Metastatic stem cells: sources, niches, and vital pathways. Cell Stem Cell. (2014) 14:306–21. doi: 10.1016/j.stem.2014.02.002

PubMed Abstract | CrossRef Full Text | Google Scholar

100. Puisieux A, Brabletz T, Caramel J. Oncogenic roles of EMT-inducing transcription factors. Nat Cell Biol. (2014) 16:488–94. doi: 10.1038/ncb2976

PubMed Abstract | CrossRef Full Text | Google Scholar

101. Lin H, Li N, He H, Ying Y, Sunkara S, Luo L, et al. AMPK inhibits the stimulatory effects of TGF-beta on Smad2/3 activity, cell migration, and epithelial-to-mesenchymal transition. Mol Pharmacol. (2015) 88:1062–71. doi: 10.1124/mol.115.099549

PubMed Abstract | CrossRef Full Text | Google Scholar

102. Sakamoto T, Kobayashi S, Yamada D, Nagano H, Tomokuni A, Tomimaru Y, et al. A histone deacetylase inhibitor suppresses epithelial-mesenchymal transition and attenuates chemoresistance in biliary tract cancer. PLoS ONE. (2016) 11:e0145985. doi: 10.1371/journal.pone.0145985

PubMed Abstract | CrossRef Full Text | Google Scholar

103. Hwang JW, Cho H, Lee JY, Jeon Y, Kim SN, Lee SJ, et al. The synthetic ajoene analog SPA3015 induces apoptotic cell death through crosstalk between NF-kappaB and PPARgamma in multidrug-resistant cancer cells. Food Chem Toxicol. (2016) 96:35–42. doi: 10.1016/j.fct.2016.07.020

PubMed Abstract | CrossRef Full Text | Google Scholar

104. Baud V, Karin M. Is NF-kappaB a good target for cancer therapy? Hopes and pitfalls. Nat Rev Drug Discov. (2009) 8:33–40. doi: 10.1038/nrd2781

PubMed Abstract | CrossRef Full Text | Google Scholar

105. McDonald OG, Wu H, Timp W, Doi A, Feinberg AP. Genome-scale epigenetic reprogramming during epithelial-to-mesenchymal transition. Nat Struct Mol Biol. (2011) 18:867–74. doi: 10.1038/nsmb.2084

PubMed Abstract | CrossRef Full Text | Google Scholar

106. Thinnes CC, England KS, Kawamura A, Chowdhury R, Schofield CJ, Hopkinson RJ. Targeting histone lysine demethylases - progress, challenges, and the future. Biochim Biophys Acta. (2014) 1839:1416–32. doi: 10.1016/j.bbagrm.2014.05.009

PubMed Abstract | CrossRef Full Text | Google Scholar

107. Kalin JH, Wu M, Gomez AV, Song Y, Das J, Hayward D, et al. Targeting the CoREST complex with dual histone deacetylase and demethylase inhibitors. Nat Commun. (2018) 9:53–4. doi: 10.1038/s41467-017-02242-4

PubMed Abstract | CrossRef Full Text | Google Scholar

108. Crea F, Danesi R, Farrar WL. Cancer stem cell epigenetics and chemoresistance. Epigenomics. (2009) 1:63–79. doi: 10.2217/epi.09.4

PubMed Abstract | CrossRef Full Text | Google Scholar

109. Pietersen AM, Horlings HM, Hauptmann M, Langerod A, Ajouaou A, Cornelissen-Steijger P, et al. EZH2 and BMI1 inversely correlate with prognosis and TP53 mutation in breast cancer. Breast Cancer Res. (2008) 10:R109. doi: 10.1186/bcr2214

PubMed Abstract | CrossRef Full Text | Google Scholar

110. Ibanez de Caceres I, Cortes-Sempere M, Moratilla C, Machado-Pinilla R, Rodriguez-Fanjul V, Manguan-Garcia C, et al. IGFBP-3 hypermethylation-derived deficiency mediates cisplatin resistance in non-small-cell lung cancer. Oncogene. (2010) 29:1681–90. doi: 10.1038/onc.2009.454

PubMed Abstract | CrossRef Full Text | Google Scholar

111. Lau EY, Ho NP, Lee TK. Cancer stem cells and their microenvironment: biology and therapeutic implications. Stem Cells Int. (2017) 2017:3714190. doi: 10.1155/2017/3714190

PubMed Abstract | CrossRef Full Text | Google Scholar

112. Seo EJ, Kim DK, Jang IH, Choi EJ, Shin SH, Lee SI, et al. Hypoxia-NOTCH1-SOX2 signaling is important for maintaining cancer stem cells in ovarian cancer. Oncotarget. (2016) 7:55624–38. doi: 10.18632/oncotarget.10954

PubMed Abstract | CrossRef Full Text | Google Scholar

113. Garson K, Vanderhyden BC. Epithelial ovarian cancer stem cells: underlying complexity of a simple paradigm. Reproduction. (2015) 149:59. doi: 10.1530/REP-14-0234

PubMed Abstract | CrossRef Full Text | Google Scholar

114. Hajizadeh F, Okoye I, Esmaily M, Ghasemi Chaleshtari M, Masjedi A, Azizi G, et al. Hypoxia inducible factors in the tumor microenvironment as therapeutic targets of cancer stem cells. Life Sci. (2019) 237:116952. doi: 10.1016/j.lfs.2019.116952

PubMed Abstract | CrossRef Full Text | Google Scholar

115. Jing X, Yang F, Shao C, Wei K, Xie M, Shen H, et al. Role of hypoxia in cancer therapy by regulating the tumor microenvironment. Mol Cancer. (2019) 18:157–9. doi: 10.1186/s12943-019-1089-9

PubMed Abstract | CrossRef Full Text | Google Scholar

116. Takebe N, Miele L, Harris PJ, Jeong W, Bando H, Kahn M, et al. Targeting notch, hedgehog, and Wnt pathways in cancer stem cells: clinical update. Nat Rev Clin Oncol. (2015) 12:445–64. doi: 10.1038/nrclinonc.2015.61

PubMed Abstract | CrossRef Full Text | Google Scholar

117. Tang SN, Fu J, Nall D, Rodova M, Shankar S, Srivastava RK. Inhibition of sonic hedgehog pathway and pluripotency maintaining factors regulate human pancreatic cancer stem cell characteristics. Int J Cancer. (2012) 131:30–40. doi: 10.1002/ijc.26323

PubMed Abstract | CrossRef Full Text | Google Scholar

118. Sims-Mourtada J, Izzo JG, Ajani J, Chao KS. Sonic hedgehog promotes multiple drug resistance by regulation of drug transport. Oncogene. (2007) 26:5674–9. doi: 10.1038/sj.onc.1210356

PubMed Abstract | CrossRef Full Text | Google Scholar

119. Flahaut M, Meier R, Coulon A, Nardou KA, Niggli FK, Martinet D, et al. The Wnt receptor FZD1 mediates chemoresistance in neuroblastoma through activation of the Wnt/beta-catenin pathway. Oncogene. (2009) 28:2245–56. doi: 10.1038/onc.2009.80

PubMed Abstract | CrossRef Full Text | Google Scholar

120. Zhang M, Liu E, Cui Y, Huang Y. Nanotechnology-based combination therapy for overcoming multidrug-resistant cancer. Cancer Biol Med. (2017) 14:212–27. doi: 10.20892/j.issn.2095-3941.2017.0054

PubMed Abstract | CrossRef Full Text | Google Scholar

121. Gao Z, Zhang L, Sun Y. Nanotechnology applied to overcome tumor drug resistance. J Control Release. (2012) 162:45–55. doi: 10.1016/j.jconrel.2012.05.051

PubMed Abstract | CrossRef Full Text | Google Scholar

122. Steichen SD, Caldorera-Moore M, Peppas NA. A review of current nanoparticle and targeting moieties for the delivery of cancer therapeutics. Eur J Pharm Sci. (2013) 48:416–27. doi: 10.1016/j.ejps.2012.12.006

PubMed Abstract | CrossRef Full Text | Google Scholar

123. Rempe R, Cramer S, Qiao R, Galla HJ. Strategies to overcome the barrier: use of nanoparticles as carriers and modulators of barrier properties. Cell Tissue Res. (2014) 355:717–26. doi: 10.1007/s00441-014-1819-7

PubMed Abstract | CrossRef Full Text | Google Scholar

124. Ma L, Kohli M, Smith A. Nanoparticles for combination drug therapy. ACS Nano. (2013) 7:9518–25. doi: 10.1021/nn405674m

PubMed Abstract | CrossRef Full Text | Google Scholar

125. Kabanov AV, Batrakova EV, Alakhov VY. Pluronic block copolymers for overcoming drug resistance in cancer. Adv Drug Deliv Rev. (2002) 54:759–79. doi: 10.1016/S0169-409X(02)00047-9

PubMed Abstract | CrossRef Full Text | Google Scholar

126. Dehghankelishadi P, Dorkoosh F. Pluronic based nano-delivery systems; Prospective warrior in war against cancer. Nanomed Res J. (2016) 1:1–7. doi: 10.7508/nmrj.2016.01.001

PubMed Abstract | CrossRef Full Text | Google Scholar

127. Alakhova DY, Zhao Y, Li S, Kabanov AV. Effect of doxorubicin/pluronic SP1049C on tumorigenicity, aggressiveness, DNA methylation and stem cell markers in murine leukemia. PLoS ONE. (2013) 8:e72238. doi: 10.1371/journal.pone.0072238

PubMed Abstract | CrossRef Full Text | Google Scholar

128. DeSano JT, Xu L. MicroRNA regulation of cancer stem cells and therapeutic implications. AAPS J. (2009) 11:682–92. doi: 10.1208/s12248-009-9147-7

PubMed Abstract | CrossRef Full Text | Google Scholar

129. Xing F, Wu K, Watabe K. MicroRNAs in cancer stem cells: new regulators of stemness. Curr Pharm Des. (2014) 20:5319–27. doi: 10.2174/1381612820666140128210912

PubMed Abstract | CrossRef Full Text | Google Scholar

130. Liang Z, Wu H, Xia J, Li Y, Zhang Y, Huang K, et al. Involvement of miR-326 in chemotherapy resistance of breast cancer through modulating expression of multidrug resistance-associated protein 1. Biochem Pharmacol. (2010) 79:817–24. doi: 10.1016/j.bcp.2009.10.017

PubMed Abstract | CrossRef Full Text | Google Scholar

131. Zou S, Cao N, Cheng D, Zheng R, Wang J, Zhu K, et al. Enhanced apoptosis of ovarian cancer cells via nanocarrier-mediated codelivery of siRNA and doxorubicin. Int J Nanomed. (2012) 7:3823–35. doi: 10.2147/IJN.S29328

PubMed Abstract | CrossRef Full Text | Google Scholar

132. Xiong X, Lavasanifar A. Traceable multifunctional micellar nanocarriers for cancer-targeted co-delivery of MDR-1 siRNA and doxorubicin. ACS Nano. (2011) 5:5202–13. doi: 10.1021/nn2013707

PubMed Abstract | CrossRef Full Text | Google Scholar

133. Budak-Alpdogan T, Banerjee D, Bertino JR. Hematopoietic stem cell gene therapy with drug resistance genes: an update. Cancer Gene Ther. (2005) 12:849–63. doi: 10.1038/sj.cgt.7700866

PubMed Abstract | CrossRef Full Text | Google Scholar

134. Zaboikin M, Srinivasakumar N, Schuening F. Gene therapy with drug resistance genes. Cancer Gene Ther. (2006) 13:335–45. doi: 10.1038/sj.cgt.7700912

PubMed Abstract | CrossRef Full Text | Google Scholar

135. Bertino JR. Transfer of drug resistance genes into hematopoietic stem cells for marrow protection. Oncologist. (2008) 13:1036–42. doi: 10.1634/theoncologist.2008-0173

PubMed Abstract | CrossRef Full Text | Google Scholar

136. O'Shaughnessy JA, Cowan KH, Nienhuis AW, McDonagh KT, Sorrentino BP, Dunbar CE, et al. Retroviral mediated transfer of the human multidrug resistance gene (MDR-1) into hematopoietic stem cells during autologous transplantation after intensive chemotherapy for metastatic breast cancer. Hum Gene Ther. (1994) 5:891–911. doi: 10.1089/hum.1994.5.7-891

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: multidrug resistance, ABC transporters, P-glycoprotein, cancer stem cells, epigenetics

Citation: Cho Y and Kim YK (2020) Cancer Stem Cells as a Potential Target to Overcome Multidrug Resistance. Front. Oncol. 10:764. doi: 10.3389/fonc.2020.00764

Received: 10 March 2020; Accepted: 21 April 2020;
Published: 02 June 2020.

Edited by:

Sungpil Yoon, Sungkyunkwan University, South Korea

Reviewed by:

Hyung Sik Kim, Sungkyunkwan University, South Korea
Xiaoju Wang, University of Michigan, United States

Copyright © 2020 Cho and Kim. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Yong Kee Kim, yksnbk@sookmyung.ac.kr

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.