Next Article in Journal
Genetic Alterations and Resectability Predict Outcome in Patients with Neuroblastoma Assigned to High-Risk Solely by MYCN Amplification
Next Article in Special Issue
Cyano- and Ketone-Containing Selenoesters as Multi-Target Compounds against Resistant Cancers
Previous Article in Journal
Skin Cancers and the Contribution of Rho GTPase Signaling Networks to Their Progression
Previous Article in Special Issue
Organoruthenium Complexes with Benzo-Fused Pyrithiones Overcome Platinum Resistance in Ovarian Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Apoptosis Deregulation and the Development of Cancer Multi-Drug Resistance

by
Christiana M. Neophytou
1,2,
Ioannis P. Trougakos
3,4,
Nuray Erin
5 and
Panagiotis Papageorgis
1,2,*
1
European University Research Center, Nicosia 2404, Cyprus
2
Tumor Microenvironment, Metastasis and Experimental Therapeutics Laboratory, Basic and Translational Cancer Research Center, Department of Life Sciences, European University Cyprus, Nicosia 2404, Cyprus
3
Department of Life Sciences, European University Cyprus, Nicosia 2404, Cyprus
4
Department of Cell Biology and Biophysics, Faculty of Biology, National and Kapodistrian University of Athens, 15784 Athens, Greece
5
Department of Medical Pharmacology, Cancer immunology and Immunotherapy Unit, Medical School, Akdeniz University, Antalya 07058, Turkey
*
Author to whom correspondence should be addressed.
Cancers 2021, 13(17), 4363; https://doi.org/10.3390/cancers13174363
Submission received: 29 July 2021 / Revised: 21 August 2021 / Accepted: 26 August 2021 / Published: 28 August 2021

Abstract

:

Simple Summary

Despite recent therapeutic advances against cancer, many patients do not respond well or respond poorly, to treatment and develop resistance to more than one anti-cancer drug, a term called multi-drug resistance (MDR). One of the main factors that contribute to MDR is the deregulation of apoptosis or programmed cell death. Herein, we describe the major apoptotic pathways and discuss how pro-apoptotic and anti-apoptotic proteins are modified in cancer cells to convey drug resistance. We also focus on our current understanding related to the interactions between survival and cell death pathways, as well as on mechanisms underlying the balance shift towards cancer cell growth and drug resistance. Moreover, we highlight the role of the tumor microenvironment components in blocking apoptosis in MDR tumors, and we discuss the significance and potential exploitation of epigenetic modifications for cancer treatment. Finally, we summarize the current and future therapeutic approaches for overcoming MDR.

Abstract

The ability of tumor cells to evade apoptosis is established as one of the hallmarks of cancer. The deregulation of apoptotic pathways conveys a survival advantage enabling cancer cells to develop multi-drug resistance (MDR), a complex tumor phenotype referring to concurrent resistance toward agents with different function and/or structure. Proteins implicated in the intrinsic pathway of apoptosis, including the Bcl-2 superfamily and Inhibitors of Apoptosis (IAP) family members, as well as their regulator, tumor suppressor p53, have been implicated in the development of MDR in many cancer types. The PI3K/AKT pathway is pivotal in promoting survival and proliferation and is often overactive in MDR tumors. In addition, the tumor microenvironment, particularly factors secreted by cancer-associated fibroblasts, can inhibit apoptosis in cancer cells and reduce the effectiveness of different anti-cancer drugs. In this review, we describe the main alterations that occur in apoptosis-and related pathways to promote MDR. We also summarize the main therapeutic approaches against resistant tumors, including agents targeting Bcl-2 family members, small molecule inhibitors against IAPs or AKT and agents of natural origin that may be used as monotherapy or in combination with conventional therapeutics. Finally, we highlight the potential of therapeutic exploitation of epigenetic modifications to reverse the MDR phenotype.

1. Introduction

Novel diagnostic and cancer therapeutic technologies have improved patient response to treatment and have lowered mortality rates. However, in several cases, the 5-year survival rate remains low, mostly due to the intrinsic resistance or to the development of acquired resistance to anticancer drugs. Multi-drug resistance (MDR) refers to the state in which cancer cells become resistant to two or more drugs that have entirely different mechanisms of action and/or chemical structures. Thus, different cancer types become difficult to treat because of MDR. These mainly include breast, lung, colorectal and prostate cancer that represent the most frequently occurring malignancies with the highest mortality rates [1,2,3,4,5].
A variety of factors and mechanisms promote the development of MDR in cancer cells, including drug inactivation, detoxification mechanisms, increased drug efflux, mutations in genes encoding drug targets, epigenetic changes, deregulation of DNA damage/repair processes, contribution of cancer stem cells, increased tumor heterogeneity, involvement of the tumor microenvironment (TME), epithelial to mesenchymal transition (EMT), modulation of reactive oxygen species (ROS) and inhibition of cell death pathways [6]. Evasion of apoptosis or programmed cell death (PCD) has been well established as a one of the major hallmarks of cancer [7]. Deregulation of apoptotic pathways can lead to tumorigenesis, autoimmune and degenerative diseases [8]. In this review, we focus on the deregulation of apoptotic pathways and the development of multi-drug resistance in a variety of tumor types. We also discuss the role of the TME in regulating apoptosis in MDR tumors, as well as current and future therapeutic approaches targeting apoptosis that are being developed to overcome cancer MDR. Finally, we discuss the potential exploitation of epigenetic modifications for new therapeutic advances.

2. Overview of Apoptotic Pathways

Apoptosis is a tightly controlled physiological process, necessary for normal embryonic development, preservation of genome integrity, proper function of the immune system and maintenance of tissue homeostasis [9]. Apoptosis may be induced by a variety of agents, including low doses of radiation, hypoxia, heat, cytotoxic drugs or more specialized anti-cancer molecules. The apoptotic process does not induce inflammation as the cell contents are eventually absorbed by phagocytic cells [10,11].
Two main pathways contribute to apoptosis, i.e., (a) the intrinsic or mitochondrial pathway that is mostly activated by intracellular stress signals, including oxidative stress and (b) the extrinsic or death receptor pathway that is engaged following extracellular signals. The latter is induced following binding of death ligands to the extracellular domain of death receptors which (among others) include Receptor 1/Tumor Necrosis Factor-α (TNFR1/TNF-α) and Fas Receptor/Fas Ligand (FasR/FasL) [12,13,14,15,16]. Downstream to this binding, death receptors establish homotrimer structures followed by self-assembly of their intracellular parts that contain death domains (DDs) [17,18]. The intracellular domains of the TNF or FAS receptors then recruit adaptor proteins, including the TNFR1-associated death domain protein (TRADD) and FAS-associated death domain protein (FADD), respectively [19,20]; adaptor proteins contain the DED protein interaction component, that by recruiting inactive initiator caspase-8 forms the Death Inducing Signaling Complex (DISC) [21] which then triggers caspase-8 activation via oligomerization [22].
Caspases are cysteinyl, aspartate-specific proteases that play critical roles in apoptosis [23]. They are expressed as inactive pro-enzymes and contain an N-terminal pro-domain and a C-terminal catalytic domain. Their C-terminal domain comprises a p20 large subunit and a p10 small subunit [24]. The so-called initiator caspases (casp-2, -8, -9, -10) are activated in the early stages of apoptosis and induce a cascade of reactions to kickstart the apoptotic process. Executioner caspases (casp-3, -6, -7) are activated during the later stages of the process and are responsible for cleaving cellular components [25]. Structurally, the pro-domain of executioner caspases is very short compared to initiator caspases. Initiator caspases contain in their pro-domain either a death effector domain (DED) (caspases-8 and -10) or a caspase-recruitment domain (CARD) (e.g., caspases-2 and -9). DED is responsible for the interaction of caspases with molecules that regulate their activity [24]. Inactive caspases are activated via cleavage at aspartate residues, while the N-terminal domain is removed by cleavage between the large and small subunits [26]. Initiator caspases then cleave downstream caspases triggering a proteolytic cascade that amplifies the apoptotic signaling pathway.
The intrinsic apoptotic pathway is initiated by different signals, including ultra-violet (UV) or gamma irradiation, hypoxia, growth factors, hormone/cytokine deprivation, viral virulence factors, heat, DNA-damaging agents and the activation of oncogenic factors (Figure 1). Apoptotic signals induce a process called mitochondrial outer membrane permeabilization (MOMP) which involves opening of the mitochondrial permeability transition (MPT) pore and release of pro-apoptotic proteins from the inter-membrane space (IMS) into the cytosol [27]. Once the outer mitochondrial membrane becomes permeable, pro-apoptotic factors located in the inter-membrane mitochondrial space exit towards the cytosol. For example, cytochrome c and Apoptotic protease activating factor 1 (Apaf-1) are released, interact with caspase-9 and activate the latter, while forming a structure known as the apoptosome [28,29]. Further, Smac/DIABLO and HtrA2/Omi are released and induce apoptosis by inhibiting a group of proteins called Inhibitors of Apoptosis Proteins (IAPs) [30,31].
MOMP is controlled by Bcl-2 family members. Bcl-2 proteins are categorized into three different groups according to their function and number of BH domains present in their structure: 1. anti-apoptotic members including Bcl-2, Bcl-xL and Mcl-1 that contain three or four BH domains, 2. pro-apoptotic members such as Bax and Bak that contain BH1, BH2 and BH3 and 3. pro-apoptotic BH3-only members including Bad, Bid, Noxa, Puma and BNIP3 [32,33]. The anti-apoptotic members can bind to BH3-only members through hydrophobic grooves formed by their BH domains [34,35]. This interaction determines the activation status of this class of proteins. BH3-only members induce apoptosis by blocking the function of anti-apoptotic family members and/or by interacting with and activating pro-apoptotic proteins such as Bax and Bak [36]. Anti-apoptotic Bcl-2 family proteins, block apoptosis by inhibiting the activity of pro-apoptotic proteins and preventing MOMP [34].
During apoptosis, anti-apoptotic protein levels decrease, while the levels of pro-apoptotic members rise; in fact, a decrease in the Bcl-2/Bax ratio is considered a reliable indicator of apoptosis. Furthermore, since the Bcl-2 family regulates mitochondrial permeability, their subcellular localization changes during apoptosis. For example, following apoptotic stimuli, Bax translocates from the cytosol to mitochondria [34]. The extrinsic pathway can also induce mitochondrial apoptotic pathways since caspase-8 can cleave Bid (a Bcl-2 pro-apoptotic protein) to its active form namely tBid which promotes MOMP [37].
Caspases -3, -6 and -7 are considered “effector” caspases and cleave cellular products during the later stages of apoptosis [38]. Caspase-9 and -8 can cleave caspase-3, while caspase-7 is a downstream target of caspase-9. Caspase-3 can also activate caspase-6 [39]. Active effector caspases break down many substrates ultimately causing DNA cleavage as well as nuclear and cytoskeletal protein degradation. The endonuclease Caspase-Activated DNase (CAD) is physiologically bound to ICAD (Inhibitor of Caspase-Activated DNase), a substrate of caspase-3. Upon its activation, CAD fragments DNA at ~180-bp pieces [40]. PARP-1 is another caspase substate; normally, it participates in DNA repair mechanisms, but it is also involved in DNA replication and transcription, cellular repair, cytoskeletal organization and protein degradation [41]. PARP-1 breakdown by caspases is crucial during apoptosis. The degradation of these targets ultimately leads to the biochemical and morphological changes observed in apoptotic cells including cell shrinkage, cytoplasmic condensation and generation of apoptotic bodies [42,43]. At the final phase of apoptosis, phagocytic cell receptors recognize ligands presented on the surface of apoptotic bodies causing their destruction by professional phagocytic cells [9]. The deregulation of apoptotic pathways that ultimately leads to MDR is described below.

3. Deregulation of the Intrinsic Apoptotic Pathway in MDR Tumors

A particularly important mechanism which promotes cancer cell resistance to chemotherapy is inhibition of apoptosis [44]. Proteins involved in the intrinsic pathway of apoptosis, including Bcl-2 family members and the tumor-suppressor p53, are commonly deregulated in MDR cancers, whereas IAPs, which control caspase activation, are often overexpressed. In addition, related survival pathways, such as PI3K/AKT, often contribute to the development of resistance.

3.1. Bcl-2 Family Deregulation in MDR

Several proteins involved in the intrinsic pathway of apoptosis have been identified as important cellular oncogenes that not only promote tumorigenesis but also contribute to anti-cancer drug resistance. Inactivating mutations or deletions of pro-apoptotic Bax or Bak are rare, but many cancers, especially those being refractory to therapy, such as colon, gastric and leukemia, overexpress one or more pro-survival family members, including Bcl-2, Bcl-xL and Mcl-1 [45,46,47,48,49]. Initial studies in Bcl-2 transgenic mice revealed accumulation of lymphocytes resistant to diverse cytotoxic agents, including chemotherapeutic drugs [50,51,52]. Multiple subsequent studies indicated that high levels of Bcl-2 gene expression correlate with severity of malignancy in cancer patients, including melanoma, breast, prostate, small cell lung, colorectal and bladder cancer, while increased Bcl-2 expression is associated with resistance to chemotherapy and radiation [53].
Another gene implicated in chemoresistance, is the tumor suppressor p53 which controls the transcription of numerous genes involved in DNA repair, metabolism, cell cycle arrest, apoptosis and senescence [54]. One of the first physiological p53 functions described was its ability to induce apoptosis in transformed cells [55]. p53 transcriptionally upregulates the expression of apoptotic-related proteins, such Puma, Noxa, Bid and Bax and can also physically interact with and neutralize the anti-apoptotic activity of Bcl-2 and Bcl-xL [56]. Thus, p53 has a dual role both as a sensitizer, as well as an activator of apoptosis. p53 mutations, mainly missense mutations, repress apoptosis thus causing therapeutic resistance [57]. Notably, mutant p53 can also inhibit apoptosis through the caspase-dependent apoptotic singling cascade [58]. Overexpression of mutant p53 has been correlated with resistance to conventional drugs including cisplatin, antimetabolites (gemcitabine), anthracyclines, (doxorubicin), alkylating agents (temozolomide) and drugs with specific targets such as EGFR-inhibitors (cetuximab) and antiestrogens (tamoxifen). In addition to mutations in TP53 gene causing protein conformational changes, p53 activity may also be impaired due to alterations in p53-regulating proteins, such as MDM2 [59]. Therefore, several small molecules, that accelerate mutant p53 protein turnover or convert it into the wild-type conformation, have been developed and applied in clinical therapy [54]. However, targeting p53 in tumor cells often leads to several side effects and drug cytotoxicity in normal tissues [60]. Importantly, ROS regulate p53 activity by oxidizing the cysteine residues present in its structure. This modification leads to p53 inability to bind to DNA and activate specific genes [61].

3.2. Inhibitors of Apoptosis Proteins (IAPs) and Their Role in MDR

IAPs are a class of proteins that are frequently overexpressed in human cancers conveying resistance to apoptosis and therapy. IAP family members include Survivin, X-linked inhibitor of apoptosis (XIAP), inhibitors of apoptosis 1 and 2 (c-IAP1 and c-IAP2), BIR-repeat-containing ubiquitin-conjugating enzyme (BRUCE/Apollon), neuronal apoptosis inhibitor protein (NAIP), IAP-like protein 2 (ILP-2) and melanoma IAP (ML-IAP/Livin) [62]. IAPs contain one or more baculovirus inhibitor repeat (BIR) domains, an amino-terminal, 70-residue structure with distinct functions. In XIAP, the region between BIR1 and BIR2 specifically targets caspases -3 and -7 while BIR3 inhibits the activity of caspase-9 [63]. The RING domain, located in XIAP, Livin, ILP2, c-IAP-1 and c-IAP-2 protein structures, catalyzes the ubiquitination and proteasomal degradation of target proteins. c-IAP and c-IAP-2 are critical regulators of the noncanonical NFkB pathway and promote malignancy by inducing the degradation of NFkB-inducing kinase (NIK) [64].
IAPs were originally thought to physically bind and block caspase activity, inhibiting both the extrinsic and intrinsic pathway of apoptosis [65]. Some family members, including c-IAP1 and c-IAP2, have a caspase recruitment domain in their structure. However, with the notable exception of XIAP, they cannot directly bind and inhibit caspases [66,67]. Under physiological conditions, IAP activity is controlled by Smac/DIABLO and Omi/HtrA2 that are released by the mitochondria and diminish their caspase-inhibitory effects. Overexpression of IAPs has been reported in many human cancers and has been correlated with resistance to therapy and worsening disease [67,68]. High levels of c-IAP1, c-IAP2, XIAP, Survivin and NAIP have been reported in breast cancer [69]. Furthermore, during early stages of pancreatic cancer, elevated levels of c-IAP2 contribute to malignant progression [70]. In esophageal cancer, increased XIAP levels inhibit caspase-3 activation and lead to apoptosis resistance [71]. Targeting IAP family members with agents that act as “SMAC mimetics” is widely investigated as a promising anti-cancer approach against MDR cancers [72].
Survivin is the smallest IAP protein. It is physiologically expressed during embryonic development to inhibit apoptosis and promote proliferation in developing tissues [73,74]. Survivin is expressed at very low levels in differentiated tissues. However, it is overexpressed in most primary tumors and has been correlated with resistance to chemotherapy and radiotherapy-induced cell death as well as poor prognosis [75,76,77,78,79,80]. Increased Survivin expression in cancer cells is partially attributed to aberrant activation of upstream survival pathways, such as NFkB, which transcriptionally upregulate Survivin [81]. The anti-apoptotic mechanism of Survivin involves formation of a complex with XIAP, that protects XIAP from ubiquitin-dependent degradation and increases its caspase-inhibiting function [82]. Furthermore, Survivin may sequester Smac/DIABLO away from XIAP or inhibit Smac/DIABLO translocation from the mitochondria to prevent XIAP inactivation [83,84]. Therapeutic exploitation of Survivin is pivotal, as it represents a cancer cell-specific drug target. However, Survivin-targeting agents have performed poorly in clinical studies, highlighting the need for developing novel approaches against this protein [85].

3.3. PI3K/AKT Pathway in Multi-Drug Resistance

The synergy between apoptosis resistance and increased survival signaling is critically important in cancer development. The PI3K/AKT pathway responds to a variety of external signals and is involved in the regulation of different cellular functions, including cell cycle progression, survival, metabolism, gene transcription and maintenance of DNA integrity [86]. Deregulation of this pathway has been implicated in MDR of many cancers, including leukemia, hepatocellular carcinoma, breast cancer, ovarian cancer, lung cancer and melanoma [87,88,89,90,91,92].
Binding of growth factors to receptor tyrosine kinases (RTKs) stimulate PI3K by autophosphorylation which leads to the phosphorylation and activation of serine/threonine kinase AKT (Protein Kinase B, PKB). RTK-PI3K complexes localize at the cell membrane where the PI3K subunit, p110, catalyzes the conversion of Phosphatidylinositol 4,5-bisphosphate PtdIns(4,5)P2 (PIP2) to Phosphatidylinositol (3,4,5)-triphosphate PtdIns(3,4,5)P3 (PIP3) [93]. Subsequently, AKT travels to the plasma membrane where it becomes phosphorylated. The tumor suppressor phosphatase and tensin homology deleted on chromosome 10 (PTEN) indirectly hinders AKT activity by converting PIP3 to PIP2 [94]. The phosphorylation of PTEN preserves its stability [95]. Phosphoinositide-dependent kinase-1 and -2 (PDK1 and PDK2) are responsible for activating AKT via phosphorylation on residues Thr308 and Ser473, respectively [96,97,98]. Phosphorylation of Thr308 partially activates AKT, while phosphorylation of both sites is required for its full activation [99]. Once activated, p-AKT translocates to the cytosol or the nucleus where it phosphorylates and therefore modulates the function of downstream substrates, including several targets being implicated in cancer initiation and progression. AKT enhances cell survival by negatively regulating the function or expression of pro-apoptotic proteins that inactivate Bcl-2 family members. Additionally, AKT promotes survival by eliciting p53 degradation [100].
A large percentage of tumors carry alterations in PI3K, AKT or PTEN leading to MDR. The sustained expression of pro-survival signals renders cancer cells resistant to anticancer agents. Phosphorylated AKT enhances cell survival by phosphorylating many proteins implicated in apoptotic pathways, including glycogen synthase kinase-3 (GSK-3), forkhead transcription factors (FOXO), caspases and proteins implicated in NFkB signaling [101]. AKT activates anti-apoptotic members of the Bcl-2 family, such as Bcl-2 and Bcl-xL, and IAPs including Survivin and XIAP via phosphorylation of the transcription factor cyclic AMP response element-binding protein (CREB) and IkB kinase (IKK), a positive regulator of NFkB [102,103,104,105]. Increased activity of the PI3K/AKT pathway attenuates chemotherapy-induced apoptosis by diminishing the levels of pro-apoptotic Bax and increasing the levels of anti-apoptotic Bcl-2 and XIAP [106]. In addition, AKT phosphorylates the pro-apoptotic protein Bad at Ser136, thus hindering its interaction with Bcl-xL and allowing the anti-apoptotic function of the latter. GSK-3, in response to insulin, regulates glycogen synthesis which has been shown to regulate cyclin D1 proteolysis and subcellular localization. GSK-3 activity is inhibited by AKT-mediated phosphorylation Ser21 [107,108,109]. AKT also activates the MEK-MAP kinase pathway in cancer cells promoting survival and proliferation, by phosphorylating cRAF at multiple amino acid residues, controlling its activity. [110]. Importantly, AKT can directly inhibit the caspase cascade; pro-caspase-9 is an AKT substrate and can be phosphorylated on Ser196 [111]. Moreover, the phosphorylation of FOXO transcription factors by AKT causes their degradation in the cytoplasm and enhances cell survival by blocking the transcription of death receptor ligands TRAIL and Fas, and of pro-apoptotic Bcl-2 members Bim and BNIP3 [112]. The implication of FOXOs in the development of MDR is highlighted by their key role in regulating drug efflux pump ABCD1 in leukemic and breast cancers and by eliciting resistance to agents that act via the accumulation of ROS [113]. A deeper understanding of the role of FOXOs in these two processes will enable the development of effective therapeutics.
Often, upstream and downstream proteins should also be deregulated to achieve MDR in cancer cells. Upstream regulators of PI3K, RTKs HER-2 and EGFR were found to be amplified in human cancers. HER-2 is overexpressed in 20–30% of primary breast cancers that also exhibit constitutive AKT activity, while EGFR overexpression has been reported in breast, lung and colorectal carcinoma and glioblastoma [114,115,116]. Notably, almost 70% of endometrial and ovarian cancers harbor activating mutations of PIK3CA, the gene that encodes the p110α catalytic subunit of PI3K [106]. PIK3CA mutation has been associated with elevated PI3K and AKT activity [117,118]. A mutated form of the PIK3CA protein was also found to selectively phosphorylate AKT and FOXO promoting cellular growth and cancer cell invasion [119]. The AKT2 gene is often amplified in human cancers, including lung and ovarian cancers while both AKT1 and AKT2 gene amplification has been reported in breast and colorectal cancers [120,121,122,123,124]. As previously mentioned, PTEN is responsible for the indirect inactivation of AKT by converting PIP3 to PIP2, thus acting as a tumor suppressor. Loss of PTEN can occur either via gene mutation, deletion or promoter hypermethylation leading to elevated concentrations of the PIP3 substrate. Consequently, downstream components of the PI3K pathway, including AKT and mTOR, are constitutively active [94,125]. Based on its implication in cancer cell progression, the PI3K/AKT pathway has been extensively studied as a promising drug target against malignant progression [126].

4. Implications of the TME in Apoptosis and MDR

The TME plays a crucial role in tumor growth, metastasis and development of MDR. The TME consists of immune cells, fibroblasts and endothelial cells that communicate with cancer cells through paracrine signaling [127]. Immune cells present in the TME can activate or inhibit apoptotic pathways and affect response to therapy [128]. Cytotoxic lymphocytes induce the activation of effector mechanisms, such as release of death ligands such as FasL and TRAIL [129,130], as well as activation of granule exocytosis pathway [131,132]. Neutrophils and monocytes express TRAIL and target TRAIL receptor-expressing tumor cells [133]. Macrophages, the major phagocytic cells of the innate immune system, can also induce TRAIL-mediated apoptosis of cancer cells [134]. Cytokines, such IFNs, CD137 and IL-24, secreted by tumor-associated immune cells promote apoptosis in cancer cells [133,135,136,137]. In addition, non-cellular TME constituents are important mediators of cancer cell behavior, such as excess extracellular matrix (ECM) deposition, which compresses blood vessels and reduces perfusion, as well as acidic and hypoxic milieu, which collectively impair drug delivery [138,139,140]. Here, we mainly focus on the cellular components of the TME that regulate apoptosis of cancer cells and, more specifically, cancer associated fibroblasts (CAFs) which alter the apoptotic responses of cancer cells to cytotoxic drugs (Figure 2).

4.1. Cancer Associated Fibroblasts in Apoptosis and MDR

During carcinogenesis, CAFs are activated as a major component of the tumor stroma [144,145]. CAFs secrete various extracellular matrix proteins, chemokines, cytokines, as well as growth factors and extensively contribute to tumor progression, invasion and metastasis [145,146]. CAFs are also linked to poor survival in most cancers and are considered potential therapeutic targets [147]. Factors released by CAFs increase tumor cell survival via the activation of anti-apoptotic pathways or by induction of the epithelial to mesenchymal transition (EMT) and cancer stem cell (CSC) phenotype, as demonstrated in melanoma, non-small cell lung cancer (NSCLC) and colorectal cancer [148,149,150,151,152,153].

4.1.1. CAF-Derived Extracellular Vesicles

The significance of extracellular vesicles (EVs) derived from CAFs (CAF-EVs) in the progression of carcinomas and resistance to apoptosis has been increasingly recognized. CAFs secrete exosomes that are loaded with proteins, lipids and RNAs to functionally control the properties of the TME [154]. CAF-EVs exert tumor-promoting functions and microRNAs (miRNAs) found in CAF-EVs participate in the interaction between cancer cells and CAFs [154,155,156]. Apoptosis of breast cancer cells is inhibited by CAF-EVs that reduce miR-30e expression to upregulate collagen triple helix repeat containing 1 (CTHRC1); this, in turn, activates the Wnt/β-catenin pathway to facilitate breast cancer development and progression. Overexpression of miR-30e or silencing of CTHRC1 suppresses proliferation, migration/invasion of breast cancer cells and promotes apoptosis [154]. CAF-derived exosomes could also transfer miR-181d-5p to enhance breast cancer aggressiveness. CAFs antagonize apoptosis in MCF-7 cells via transfer of miR-181d-5p which downregulates homeobox A5 (HOXA5) and caudal-related homeobox 2 (CDX2) [157]. Long non-coding RNAs (lncRNAs) are a heterogeneous class of transcripts longer than 200 nucleotides with limited protein-coding potential [158,159]. CAFs were shown to transfer lncRNA H19 to neighboring colorectal cancer cells [151,160,161,162] and promote the stemness via activating Wnt/β-catenin signaling causing resistance to oxaliplatin-mediated apoptosis [163]. Exosomal lncRNA enhance invasion, migration, proliferation and inhibit apoptosis in cervical cancer cells and in NSCLC [164,165]. The importance of lncRNAs in regulating drug resistance in cancer cells has recently been described in a detailed review [166].

4.1.2. Interleukins Secreted by CAFs

Lung adenocarcinoma is one of the most CAF-rich cancers. The role of CAFs in resistance to chemotherapy of lung adenocarcinoma is well appreciated [167,168]. Cisplatin treatment increases Interleukin-11 (IL-11) levels in CAFs which protects lung adenocarcinoma cells from apoptosis [169]. Cisplatin induces DNA damage and subsequently activation of apoptosis [170]. IL-11, a member of IL-6 family, binds to IL-11Ra2 to activate signaling. [171]. Patients with high IL-11Rα levels show poor response to cisplatin-based chemotherapy [169]. IL-11 induces Signal transducer and activator of transcription 3 (STAT3) phosphorylation and increases the expression of anti-apoptotic protein Bcl-2 and Survivin in cancer cells. As mentioned above, Bcl-2 and Survivin have been extensively implicated in the development of chemoresistance in cancer [172,173]. The anti-apoptotic effects of IL-11 can be prevented by suppressing STAT3 phosphorylation or silencing IL-11Rα expression in lung adenocarcinoma [169].
IL-6 is another key cytokine, secreted by cancer cells, immune cells and CAFs, which inhibits apoptosis of cancer cells through STAT3 activation [174,175,176]. IL-6 binds to the cell surface receptor glycoprotein 130 (gp130) and activates several cell survival-related pathways promoting chemotherapeutic resistance in breast, ovarian and endometrial cancers [177,178,179]. Significant amount of IL-6 in the TME originates from CAFs and is involved in carcinogenesis and metastasis [180,181]. IL-6 derived from CAFs prevents chemotherapy-induced apoptosis by increasing the phosphorylation of Jak1 and STAT3, and the expression of the anti-apoptotic proteins Bcl-2 and Survivin in gastric carcinoma [182]. Studies conducted in CAFs derived from human gastric carcinomas further demonstrated the role of CAFs in prevention of early apoptosis of gastric cancer cells treated with 5-fluorouracil (5-FU). Factors secreted from CAFs not only inhibit apoptosis but also induce an EMT phenotype in gastric carcinoma [183]. Treatment with conditioned medium from activated gastric carcinoma CAFs decreases response of gastric cancer cells to 5-FU by suppression of apoptosis-related proteins, such as Bak, Bax, cleaved caspase 3 and cleaved PARP [183]. Similarly, chemotherapy-induced apoptosis of pancreatic cancer cells significantly decreases in the presence of CAFs [184,185]. Overall, the role of IL-6, IL-11 as well as of other interleukins, has been observed in MDR cancer cells; the clinical attempts to block their effects for therapeutic intervention have been described in a recent review [186].

4.1.3. Regulation of Sex Determining Region Y-box 2 by CAFs

Sex determining region Y -box 2 (Sox2), an essential embryonal stem cell transcription factor, may also play an important role in CAF-induced drug resistance. Sox2 is linked to the formation and maintenance of CSC phenotype and is implicated in drug resistance and poor patient prognosis [187,188,189,190,191,192]. In ER+ breast cancer cells, a mixture of CAF-secreted factors strongly induced Sox2 expression. In addition, Sox2 blocked apoptosis, enabled cellular growth and shielded cells against the anti-estrogen fulvestrant [193]. CAFs also minimized the effectiveness of tamoxifen in breast cancer cells [194]. Whereas Sox2 is an attractive therapeutic target, direct targeting of Sox2 via siRNA has shown poor outcomes due to inefficient delivery and efficacy; novel approaches include the design of artificial transcription factors (ATFs), that bind to proximal SOX2 promoters and reduce its expression [195].

4.1.4. Growth Promoting Proteins Released by CAFs

CAFs secrete hepatocyte growth factor (HGF) that mediates resistance to cancer cell apoptosis [196]. HGF also induces cell proliferation, cancer cell motility and migration. Many cancer types, including ovarian, gastric, colorectal and pancreatic, overexpress c-Met, which functions as a specific HGF receptor [197,198]. HGF-induced c-Met activation triggers downstream the PI3K/Akt pathway, enabling cancer progression [199,200,201]. In lung cancer, HGF derived from CAFs, attenuates the apoptotic effects of paclitaxel (PAC) by upregulating glucose-regulated protein 78 (GRP78) [167]. GRP78 acts as a chaperon protein in the endoplasmic reticulum (ER) where it regulates protein folding; it is highly expressed on the surface of cancer cells [202] and enables malignant growth, motility, migration and resistance to therapy [203,204]. In ovarian cancer, HGF secreted by CAFs attenuated paclitaxel-induced apoptosis by activating the c-Met/PI3K/Akt pathway and signaling involving GRP78 [196]. In addition, GRP78 inhibits apoptosis by interacting with caspase-7 or p53 [205,206] and also by binding to Bax and Bik to prevent mitochondrial release of cyt-c [207,208]. This evidence suggests that the activation of PI3K/Akt and GRP78 may be implicated in cancer progression and resistance to therapy. Activation of PI3K/Akt pathway by other CAF-originated factors can also lead to resistance to apoptosis induced by cytotoxic drugs. More specifically, the CAF-derived chemokine CCL5, promoted cisplatin resistance in ovarian cancer cells by affecting the PI3K/Akt signaling pathway [209,210]. Netrin-1 is a multifunctional secreted glycoprotein upregulated in various cancers, such as gastric and lung, and may inhibit apoptosis induced by the dependence receptors DCC and UNC5H [211,212]. Netrin-1 and its receptor (UNC5B) are upregulated in CAFs of lung and colon tumors [213]. Inhibition of netrin-1 abrogates CAF-mediated increase in cancer stemness [213]. Hence, Netrin-1 secreted by CAFs may play an important role in inhibition of apoptosis and drug resistance.
TP53-regulated inhibitor of apoptosis 1 (TRIAP1) is a small, 76-amino acid long, evolutionary conserved protein which inhibits apoptosis and promotes DNA repair [214,215]. Loss of caveolin-1 in CAFs augmented the secretion of TRIAP1 from CAFs causing radiation resistance of prostate cancer cells by hindering apoptosis [216]. Similarly, knockdown of TRIAP1, using microRNA miR-320b, induced mitochondrial apoptosis [214,217].
Folicular lymphoma-associated CAFs, isolated from malignant lymphoma patients, were shown to protect tumor cells from apoptosis in response to cytotoxic drugs [218]. These CAFs do not alter proliferation rate of cancer cells but markedly upregulate the expression of the anti-apoptotic BCL2L1 gene in folicular lymphoma cells [219,220,221]. Midkine can also mediate CAF-induced inhibition of apoptosis and chemoresistance. Midkine is a heparin-binding growth factor and induces tumor progression by enhancing carcinoma cell growth, survival [222,223], invasiveness, migration, and chemotherapy resistance [224,225]. Primary CAFs from oral squamous cell carcinoma secrete high levels of midkine, which abrogate cisplatin-induced cell death [226]. Midkine also enables glioma cells to become resistant to tetrahydrocannabinol by obstructing the ALK receptor and inhibiting autophagy-mediated cell death via the Akt/mTORC1 pathway [227]. Midkine induces the expression of lncRNA ANRIL in cancer cells while lncRNA ANRIL knockdown blocked proliferation and promoted apoptosis to augment cisplatin cytotoxicity via impairment of the drug transporters MRP1 and ABCC2 [226]. Furthermore, knockdown of lncRNA ANRIL increased the activation of caspase-3 and inhibited Bcl-2 expression [226].
Conclusively, CAFs not only enhance the aggressiveness of cancer cells but also render them resistant to therapy-induced apoptotic effects by secreting various factors including miRNAs, lncRNAs, cytokines and chemokines. All these CAF-secreted factors induce activation or upregulation of factors implicated in apoptosis inhibition and overactivation of survival pathways in cancer cells in response to chemotherapy. Given the crucial roles of CAFs in carcinogenesis and drug resistance, better understanding of the underlying mechanisms will uncover novel targets to overcome drug resistance mediated via deregulation of cell death pathways.

5. Therapeutic Approaches to Induce Apoptosis in MDR Cancers

Potent apoptosis-inducing approaches can prevent tumor initiation and progression. Many proteins involved in apoptosis have been targeted with small molecule inhibitors, epigenetic drugs and natural or synthetic compounds (Table 1). These agents may be used as monotherapy, but they have been often evaluated in combination with other targeted or conventional anti-tumor therapeutics. The Bcl-2 family of proteins represent an attractive target for therapy as it is often deregulated and confers resistance in cancer. Consequently, small molecule inhibitors that can interact with BH3 domains and antisense oligonucleotides have been developed [228,229]. These small molecules against Bcl-2 proteins can be categorized as BH3 mimetics (i.e., ABT-737, ABT-263) and small molecules with BH3 putative mimetic action (i.e., gossypol, obatoclax etc.) [230].
Table 1. Types of anti-cancer treatments against multi-drug resistance involving apoptotic pathways.
Table 1. Types of anti-cancer treatments against multi-drug resistance involving apoptotic pathways.
Therapeutic ClassCompoundObserved EffectModelRef
Small molecule inhibitorsVenetoclax
(Bcl-2 inhibitor)
Directly blocked the wild-type ABCG2 efflux function and inhibited the ATPase activity of ABCG2.Human embryonic kidney cell line HEK293 overexpressing ABCG2 in vitro.[231]
ABT-737
(BH3-mimetic)
In combination with Fenretinide, synergistically induced cyt-c release, activation of caspases, Bax, t-Bid and Bak.MDR neuroblastoma cell lines in vitro.[232]
Nutlin5
(MDM2-p53 antagonist)
Reversed MDR-1-mediated multidrug resistance in a p53-independent manner.High MDR-1-expressing p53 mutant neuroblastoma cell lines in vitro.[233]
MI-219
(MDM2 inhibitor)
Sensitized cells to androgen ablation and radiotherapy by inducing DNA damage and apoptosis.Prostate Cancer Cells in vitro.[234]
ThiosemicarbazoneInhibited cell cycle progression at the G1 phase.MCF7 and MCF7/ADR cells in vitro.[235]
LY294002
(PI3K inhibitor)
Inhibited the expression of p-Akt and P-gp.Leukemia cell line K562/DNR and gastric cancer cell line SGC7901/ADR in vitro.[236]
Metformin
(Metabolic inhibitor)
In combination with 2-deoxyglucose selectively enhanced cytotoxicity of Doxorubicin leading to G2/M arrest and apoptosis.MCF-7/Dox breast cancer cells in vitro.[237]
BEZ235
(PI3K/mTOR inhibitor)
Caused a dose-dependent decrease in cell viability in combination with Dox, associated with an increase in cleaved PARP.Ovarian A2780 and pancreatic MiaPaca2 cancer cells in vitro.[238]
AZ D8055
(mTORC1/2 inhibitor)
Inhibition of mTOR and caspase-3 cleavage in platinum-resistant cells.Advanced-stage ovarian clear cell carcinoma patient-derived xenograft models.[239]
Rapamycine
(mTOR inhibitor)
Inhibited PI3K/AKT pathway, blocked proliferation, sensitized cells to Tamoxifen and Fulvestrant.Breast cancer cells resistant to endocrine therapy in vitro.[240]
YM155
(Survivin inhibitor)
Survivin depletion and p53 activation.Neuroblastoma cell lines and their sublines with acquired resistance to clinically relevant drugs in vitro.[241]
Natural agents
and derivatives
Wagonin Promoted TRAIL-induced apoptosis in vitro and downregulated the expression of anti-apoptotic XIAP, cFLIPL, cIAP-1 and cIAP-2.Non-small cell lung cancer in vivo.[242]
LuteolinGenerated ROS leading to DNA damage and activated the ATR/Chk2/p53 pathway independently of the P-gp efflux pump.MDR breast cancer cells in vitro.[243]
Fisetin Concurrent treatment with chemotherapeutic drugs activated caspases -8 and -3, release of cyt-c and inhibited survival pathways IGF1R and AKT.Colon cancer cells resistant to both Irinotecan and Oxaliplatin in vitro/in vivo.[244]
Genistein Pre-treatment inhibited NFkB activity and led to increased growth inhibition and apoptosis in combination with Cisplatin and Docetaxel.Prostate and lung cancer cells in vitro/in vivo.[245]
Resveratrol - Induced apoptosis by upregulating miR-34c and p53.- Platinum-resistant colorectal cancer cells, in vivo.[246]
- Reversed MDR by targeting Survivin and activating caspase-3.
- Non-small cell lung MDR cancer cells, in vivo.
[247]
Curcumin - Sensitized cells to capecitabine by inhibiting NFkB, reduced Bcl-2, IAP-1, Survivin, COX-2, MMP-2, ICAM-1, CXCR4 and VEGF- Colorectal cancer to capecitabine in vivo[248]
- Difluorinated Curcumin downregulated PTEN inhibitor, miR-21.- Colorectal cancer cells resistant to 5-FU and oxaliplatin in vitro.[249]
- In combination with EGCG led to synergistic effects through activation of the caspase-dependent signaling pathway, and downregulation of Bcl-2 and Survivin.- Resistant breast cancer cells in vitro.[250]
Ellagic acid In combination with 5-FU increased the Bax/Bcl-2 ratio, caused changes in mitochondrial membrane potential, activated caspase-3 and induced apoptosis.Colorectal cancer cells in vitro.[251]
O-methylated coumarinInhibited the PI3K/Akt signaling pathway.Myelogenous leukemia K562/ADM cells in vitro.[252]
Vitamin E and derivativesTPGS induced cell cycle arrest and apoptosis selectively in Survivin-overexpressing breast cancer cells.Breast cancer cells in vitro.[253]
TME/Immune regulationPirfenidoneInduced apoptosis in CAFs at high concentration; at low concentrations induced apoptosis and decreased tumor progression synergistically with Cisplatin.NSCLC cells in vitro and in vivo.[254]
Combination of anti–CTL-4 plus anti–PD1 therapyMediated a switch from expansion of phenotypically exhausted CD8+ T cells to expansion of activated effector CD8+ T cells.Melanoma patients.[255]
Combination of EGFR-TKIs and anti-PD-1/PD-L1 antibodiesPD-L1 mediated by EGFR activation could induce the apoptosis of T cells through PD-L1/PD-1 axis in tumor cells.EGFR-TKIs-resistant NSCLC cells with EGFR mutation in vitro.[256]
MEDI9447Antibody targeting ectoenzyme CD73, increased CD8+ effector cells and activated macrophages.Mouse syngeneic colorectal tumor growth in vivo.[257]
Epigenetic drugsHydralazine
(DNMTi)
In combination with Magnesium Valproate LP improved progression-free survival.Metastatic Recurrent or Persistent Cervical Cancer patients.[258]
Parthenolide
(HDACi)
NFkB and HIF1-α Inhibition.Brain, breast, colon cancer cell lines in vitro.[259]
Decitabine (DNMTi) and Panobinostat (HDACi)In combination with alkylating agent temozolomide showed great improvements in disease stabilization and remission.Resistant metastatic melanoma patients[260]
Azacitidine and Valproic acidIn combination with carboplatin demonstrates decreased DR4 methylation and shows modest evidence of antitumor activity Patients with heavily treated advanced ovarian cancer.[261]
BRD4i
(BRD4 inhibitor)
Induced homologous recombination deficiency and sensitized cells to PARP inhibition.Multiple tumor lineages regardless of BRCA1/2, TP53, RAS or BRAF mutation status in vitro and in vivo.[262]
Inhibiting AKT has for long been a major focus as a promising therapeutic approach in cancer. To date, there are two classes of AKT inhibitors, namely, ATP-competitive and allosteric inhibitors which either block ATP binding or prevent AKT phosphorylation and activation [263]. However, despite the development of many compounds with promising results for targeting AKT, none of these inhibitors has been approved yet for clinical use. This is, at least in part, attributed to the complex and pleiotropic functions that AKT exerts in cells. Therefore, combinational therapy approaches seem to represent a major research direction for the successful clinical utilization of AKT inhibitors [264].
Conventional drugs, such as cisplatin and doxorubicin, exert their anti-cancer effects via the accumulation of ROS and DNA damage. However, reduction of ROS generation leads to resistance [265]. Moreover, synthetic agents tested against MDR cancers, can sometimes cause toxic side effects and lack specificity. For these reasons, efforts have also been focused on natural agents and their derivatives, to take advantage of their beneficial properties. Polyphenols represent a large family of organic, naturally occurring compounds that are characterized by the presence of many phenol groups in their structure. Phenolic compounds, including flavones, ellagitannins and curcumin, are known to act as chemopreventive agents due to their antioxidant properties and their ability to inactivate pro-carcinogens. Certain natural compounds, rely on the upregulation of ROS to induce DNA damage. However, low oxygen levels within the tumor leads to limited generation of ROS and allows cancer cells to escape death [266]. Reduced ROS levels in cancer cells have been reported to increase the levels of P-gp efflux pump through the JNK pathway, further promoting drug resistance [267]. Importantly, natural compounds also exert chemotherapeutic properties because they can regulate signaling pathways to inhibit the proliferation of cancer cells, block angiogenesis and metastasis, and induce immune and inflammatory responses [268]. Importantly, phenolic compounds can induce apoptosis in cancer cells by activating various pro-apoptotic machineries and, interestingly, several have been reported to be effective against MDR tumors [269]. The TME is also being targeted to improve drug efficacy in difficult to treat cancers; immune checkpoint inhibitors (ICIs) that have shown promising clinical efficacy, are now being tested in combination with other agents to overcome intrinsic or acquired tumor resistance [270].

6. Future Perspectives: The Implication and Therapeutic Exploitation of Epigenetics in MDR

For decades, genetic mutations during cancer progression and acquired MDR were considered a major cause of treatment failure in relapsed cases, ignoring the non-genetic basis of tumor heterogeneity [271]. A breakthrough in anti-cancer therapy was achieved when it was realized that the high predominance of MDR is attributed not only to DNA mutations but also to a variety of epigenetic alterations. Moreover, it became increasingly evident that the probability of an irreversible mutation to appear, increases as more critical modifications emerge in the epigenome of tumor cells [272]. Among the observations that led to this conclusion were the reversal of drug resistance observed upon drug-free periods, the frequent absence of mutations in drug targets or activated pathways, as well as the heterogeneity in acquired MDR and in relapsed cases [273,274]. Recent data highlight the major role of epigenetic changes in tumorigenesis and in the development of MDR [275]. Cancer cells can escape from a poised drug-tolerant condition and enter into an epigenetically fixed acquired-resistant state via poorly understood mechanisms. During carcinogenesis, environmental pressure upon tumor cells results in an array of epigenetic aberrations, such as DNA and RNA methylation, alterations in the miRNAs expression and histone modifications, which eventually lead to epigenetically-induced transcriptional adaptation [272,275].
Epigenetic changes are generally reversible and susceptible to external factors; these characteristics make them appealing targets either for monotherapy or in combination with other anti-cancer agents to treat MDR [276,277]. Therefore, multiple generations of drugs that target epigenetic regulators, called epi-drugs, have been designed during the last 40 years, demonstrating valuable effects on cancer therapy in clinical trials [278,279]. Epigenetic modifications such as DNA hypermethylation of gene promoters could partially explain the acquired resistance after prolonged treatment [280,281]. Recent studies have shown that epi-drugs, such as the DNA methylation inhibitor (iDNMT), 5-aza-20-deoxycytidine (decitabine; DAC), can be effective against resistant cancers (including lung cancer and AML), in combination with conventional chemotherapeutics by reversing DNA methylation which sensitizes cancer cells to other chemotherapeutics, including carboplatin, cisplatin and 5-FU [282,283,284].
Several challenges remain, however, to improve the effectiveness of epi-drugs against MDR cancers. In contrast to hematological malignancies, solid tumors do not respond well to epi-drugs possibly due to the contribution of the TME [285]. To overcome these obstacles, single-cell sequencing technologies (i.e., scRNA-seq, scATAC-seq, sc-Hi-C and scChIP-seq) using patients’ biopsies immediately before and/or after epi-drug administration could provide more detailed information about the drug resistance landscape derived from genome/epigenome interactions. In addition, many epi-drugs, such as HDACs inhibitors which were shown to be effective against hematological malignancies and MDR, result in numerous off-target effects, since they act as pan-HDAC inhibitors [286]. A novel therapeutic approach, called proteolysis-targeting chimera (PROTAC) has been developed to improve the specificity of drugs against targets at low concentrations. The PROTACs technology promotes proteasomal protein degradation through E3 ubiquitin ligase activity. Importantly, side effects are significantly reduced [57]. PROTAC drugs targeting the epigenetic reader BRD4 are under preclinical evaluation in various cancer types showing promising therapeutic effects in reversing drug resistance phenotypes [57]. In addition, the replacement of preclinical models, such as two-dimensional (2D) in vitro cancer cell lines, with more clinically relevant 3D in vitro or mouse models that recapitulate the TME of resistant cells in relapsed patients should be considered [287,288]. Importantly, epigenetic interpatient and intratumor heterogeneity, a hallmark of human cancers that plays crucial roles in developing MDR need to be more thoroughly investigated [271,277]. Taken together, current evidence suggests that to improve efficacy of epi-drugs, new generations of more selective agents should be developed with optimized drug dosage, pharmacodynamics and pharmacokinetics properties, along with low toxicity levels in normal cells and tissues.

7. Conclusions

Deregulation of the major apoptotic pathways, and the related survival pathways that control the expression and/or activation of apoptotic proteins, may lead to MDR. Imbalance in Bcl-2 family levels, overexpression of IAPs and p53 inactivation have been widely reported in various types of MDR tumors. The TME appears to be pivotal in tumor progression and is known to impair the effectiveness of many therapeutics. A better understanding of how epigenetic alterations control cancer development may lead improved drug efficacy and contribute to the improvement of several agents already employed against MDR tumors. Recent studies using in vitro models have shown that different MDR mechanisms can be derived from a single ancestor cell [289]. Subsequently, in these cases, detection and characterization of residual tumor cells using single-cell sequencing technologies, will elucidate the contribution of genetic and epigenetic variability in developing drug resistance phenotypes [290,291]. It has been suggested that one of the better strategies in the fight against acquired MDR could be a therapeutic scheme that targets cancer cells prior to the acquisition of drug resistance, i.e., before they express a pro-survival program [280]. At the dawn of personalized medicine, the study of these alterations could provide novel and promising predictive biomarkers with great clinical significance against the evolution of acquired MDR.

Author Contributions

Conceptualization, C.M.N.; investigation, C.M.N., I.P.T. and N.E.; writing—original draft preparation, C.M.N., I.P.T. and N.E.; writing—review and editing, C.M.N., I.P.T., N.E. and P.P.; visualization, C.M.N.; supervision, P.P.; funding acquisition, C.M.N. All authors have read and agreed to the published version of the manuscript.

Funding

This article/publication is based upon work from COST Action CA17104 STRATAGEM, supported by COST (European Cooperation in Science and Technology), www.cost.eu, accessed on 13 April 2018.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Nuray Erin was supported by TUBITAK grant no:118S378 and Ioannis Trougakos from the Hellenic GRST project PlantUP-GR (MIS 5002803). This work was co-funded by the European Regional Development Fund and the Republic of Cyprus through the Research and Innovation Foundation (Project: POST-DOC/0916/0044).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2020. CA Cancer J. Clin. 2020, 70, 7–30. [Google Scholar] [CrossRef]
  2. Yersal, O.; Barutca, S. Biological subtypes of breast cancer: Prognostic and therapeutic implications. World J. Clin. Oncol. 2014, 5, 412–424. [Google Scholar] [CrossRef] [PubMed]
  3. Tsaur, I.; Heidegger, I.; Kretschmer, A.; Borgmann, H.; Gandaglia, G.; Briganti, A.; de Visschere, P.; Mathieu, R.; Valerio, M.; van den Bergh, R.; et al. Aggressive variants of prostate cancer—Are we ready to apply specific treatment right now? Cancer Treat. Rev. 2019, 75, 20–26. [Google Scholar] [CrossRef]
  4. Nguyen, K.S.; Neal, J.W.; Wakelee, H. Review of the current targeted therapies for non-small-cell lung cancer. World J. Clin. Oncol. 2014, 5, 576–587. [Google Scholar] [CrossRef] [PubMed]
  5. Kozovska, Z.; Gabrisova, V.; Kucerova, L. Colon cancer: Cancer stem cells markers, drug resistance and treatment. Biomed. Pharmacother. Biomed. Pharmacother. 2014, 68, 911–916. [Google Scholar] [CrossRef] [PubMed]
  6. Assaraf, Y.G.; Brozovic, A.; Goncalves, A.C.; Jurkovicova, D.; Line, A.; Machuqueiro, M.; Saponara, S.; Sarmento-Ribeiro, A.B.; Xavier, C.P.R.; Vasconcelos, M.H. The multi-factorial nature of clinical multidrug resistance in cancer. Drug Resist. Updates Rev. Comment. Antimicrob. Anticancer Chemother. 2019, 46, 100645. [Google Scholar] [CrossRef]
  7. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  8. Plati, J.; Bucur, O.; Khosravi-Far, R. Apoptotic cell signaling in cancer progression and therapy. Integr. Biol. 2011, 3, 279–296. [Google Scholar] [CrossRef] [PubMed]
  9. Elmore, S. Apoptosis: A review of programmed cell death. Toxicol. Pathol. 2007, 35, 495–516. [Google Scholar] [CrossRef]
  10. Savill, J.; Fadok, V. Corpse clearance defines the meaning of cell death. Nature 2000, 407, 784–788. [Google Scholar] [CrossRef]
  11. Kurosaka, K.; Takahashi, M.; Watanabe, N.; Kobayashi, Y. Silent cleanup of very early apoptotic cells by macrophages. J. Immunol. 2003, 171, 4672–4679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Suliman, A.; Lam, A.; Datta, R.; Srivastava, R.K. Intracellular mechanisms of TRAIL: Apoptosis through mitochondrial-dependent and -independent pathways. Oncogene 2001, 20, 2122–2133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Ashkenazi, A.; Dixit, V.M. Death receptors: Signaling and modulation. Science 1998, 281, 1305–1308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Chicheportiche, Y.; Bourdon, P.R.; Xu, H.; Hsu, Y.M.; Scott, H.; Hession, C.; Garcia, I.; Browning, J.L. TWEAK, a new secreted ligand in the tumor necrosis factor family that weakly induces apoptosis. J. Biol. Chem. 1997, 272, 32401–32410. [Google Scholar] [CrossRef] [Green Version]
  15. Peter, M.E.; Krammer, P.H. Mechanisms of CD95 (APO-1/Fas)-mediated apoptosis. Curr. Opin. Immunol. 1998, 10, 545–551. [Google Scholar] [CrossRef]
  16. Rubio-Moscardo, F.; Blesa, D.; Mestre, C.; Siebert, R.; Balasas, T.; Benito, A.; Rosenwald, A.; Climent, J.; Martinez, J.I.; Schilhabel, M.; et al. Characterization of 8p21.3 chromosomal deletions in B-cell lymphoma: TRAIL-R1 and TRAIL-R2 as candidate dosage-dependent tumor suppressor genes. Blood 2005, 106, 3214–3222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Guicciardi, M.E.; Gores, G.J. Life and death by death receptors. FASEB J. 2009, 23, 1625–1637. [Google Scholar] [CrossRef] [Green Version]
  18. Boldin, M.P.; Mett, I.L.; Varfolomeev, E.E.; Chumakov, I.; Shemer-Avni, Y.; Camonis, J.H.; Wallach, D. Self-association of the “death domains” of the p55 tumor necrosis factor (TNF) receptor and Fas/APO1 prompts signaling for TNF and Fas/APO1 effects. J. Biol. Chem. 1995, 270, 387–391. [Google Scholar] [CrossRef] [Green Version]
  19. Wajant, H. The Fas signaling pathway: More than a paradigm. Science 2002, 296, 1635–1636. [Google Scholar] [CrossRef]
  20. Hsu, H.; Xiong, J.; Goeddel, D.V. The TNF receptor 1-associated protein TRADD signals cell death and NF-kappa B activation. Cell 1995, 81, 495–504. [Google Scholar] [CrossRef] [Green Version]
  21. Jin, Z.; El-Deiry, W.S. Overview of cell death signaling pathways. Cancer Biol. Ther. 2005, 4, 139–163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Kischkel, F.C.; Hellbardt, S.; Behrmann, I.; Germer, M.; Pawlita, M.; Krammer, P.H.; Peter, M.E. Cytotoxicity-dependent APO-1 (Fas/CD95)-associated proteins form a death-inducing signaling complex (DISC) with the receptor. EMBO J. 1995, 14, 5579–5588. [Google Scholar] [CrossRef]
  23. Li, J.; Yuan, J. Caspases in apoptosis and beyond. Oncogene 2008, 27, 6194–6206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Pop, C.; Salvesen, G.S. Human caspases: Activation, specificity, and regulation. J. Biol. Chem. 2009, 284, 21777–21781. [Google Scholar] [CrossRef] [Green Version]
  25. Cohen, G.M. Caspases: The executioners of apoptosis. Biochem. J. 1997, 326 Pt 1, 1–16. [Google Scholar] [CrossRef] [Green Version]
  26. Degterev, A.; Boyce, M.; Yuan, J. A decade of caspases. Oncogene 2003, 22, 8543–8567. [Google Scholar] [CrossRef] [Green Version]
  27. Saelens, X.; Festjens, N.; Vande Walle, L.; van Gurp, M.; van Loo, G.; Vandenabeele, P. Toxic proteins released from mitochondria in cell death. Oncogene 2004, 23, 2861–2874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Chinnaiyan, A.M. The apoptosome: Heart and soul of the cell death machine. Neoplasia 1999, 1, 5–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Hill, M.M.; Adrain, C.; Duriez, P.J.; Creagh, E.M.; Martin, S.J. Analysis of the composition, assembly kinetics and activity of native Apaf-1 apoptosomes. EMBO J. 2004, 23, 2134–2145. [Google Scholar] [CrossRef] [Green Version]
  30. Van Loo, G.; van Gurp, M.; Depuydt, B.; Srinivasula, S.M.; Rodriguez, I.; Alnemri, E.S.; Gevaert, K.; Vandekerckhove, J.; Declercq, W.; Vandenabeele, P. The serine protease Omi/HtrA2 is released from mitochondria during apoptosis. Omi interacts with caspase-inhibitor XIAP and induces enhanced caspase activity. Cell Death Differ. 2002, 9, 20–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Schimmer, A.D. Inhibitor of apoptosis proteins: Translating basic knowledge into clinical practice. Cancer Res. 2004, 64, 7183–7190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Giam, M.; Huang, D.C.; Bouillet, P. BH3-only proteins and their roles in programmed cell death. Oncogene 2008, 27 (Suppl. S1), S128–136. [Google Scholar] [CrossRef] [PubMed]
  33. Danial, N.N. BCL-2 family proteins: Critical checkpoints of apoptotic cell death. Clin. Cancer Res. 2007, 13, 7254–7263. [Google Scholar] [CrossRef] [Green Version]
  34. 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. [Google Scholar] [CrossRef] [PubMed]
  35. Hinds, M.G.; Day, C.L. Regulation of apoptosis: Uncovering the binding determinants. Curr. Opin. Struct. Biol. 2005, 15, 690–699. [Google Scholar] [CrossRef] [PubMed]
  36. Ghiotto, F.; Fais, F.; Bruno, S. BH3-only proteins: The death-puppeteer’s wires. Cytom. A 2010, 77, 11–21. [Google Scholar] [CrossRef] [PubMed]
  37. Esposti, M.D. The roles of Bid. Apoptosis 2002, 7, 433–440. [Google Scholar] [CrossRef]
  38. Igney, F.H.; Krammer, P.H. Death and anti-death: Tumour resistance to apoptosis. Nat. Rev. Cancer 2002, 2, 277–288. [Google Scholar] [CrossRef]
  39. Cullen, S.P.; Martin, S.J. Caspase activation pathways: Some recent progress. Cell Death Differ. 2009, 16, 935–938. [Google Scholar] [CrossRef]
  40. Enari, M.; Sakahira, H.; Yokoyama, H.; Okawa, K.; Iwamatsu, A.; Nagata, S. A caspase-activated DNase that degrades DNA during apoptosis, and its inhibitor ICAD. Nature 1998, 391, 43–50. [Google Scholar] [CrossRef]
  41. Hong, S.J.; Dawson, T.M.; Dawson, V.L. Nuclear and mitochondrial conversations in cell death: PARP-1 and AIF signaling. Trends Pharm. Sci. 2004, 25, 259–264. [Google Scholar] [CrossRef]
  42. Luthi, A.U.; Martin, S.J. The CASBAH: A searchable database of caspase substrates. Cell Death Differ. 2007, 14, 641–650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Khosravi-Far, R.; Esposti, M.D. Death receptor signals to mitochondria. Cancer Biol. Ther. 2004, 3, 1051–1057. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Housman, G.; Byler, S.; Heerboth, S.; Lapinska, K.; Longacre, M.; Snyder, N.; Sarkar, S. Drug resistance in cancer: An overview. Cancers 2014, 6, 1769–1792. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Campana, D.; Coustan-Smith, E.; Manabe, A.; Buschle, M.; Raimondi, S.C.; Behm, F.G.; Ashmun, R.; Arico, M.; Biondi, A.; Pui, C.H. Prolonged survival of B-lineage acute lymphoblastic leukemia cells is accompanied by overexpression of bcl-2 protein. Blood 1993, 81, 1025–1031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Mason, K.D.; Vandenberg, C.J.; Scott, C.L.; Wei, A.H.; Cory, S.; Huang, D.C.; Roberts, A.W. In vivo efficacy of the Bcl-2 antagonist ABT-737 against aggressive Myc-driven lymphomas. Proc. Natl. Acad. Sci. USA 2008, 105, 17961–17966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Othman, R.T.; Kimishi, I.; Bradshaw, T.D.; Storer, L.C.; Korshunov, A.; Pfister, S.M.; Grundy, R.G.; Kerr, I.D.; Coyle, B. Overcoming multiple drug resistance mechanisms in medulloblastoma. Acta Neuropathol. Commun. 2014, 2, 57. [Google Scholar] [CrossRef] [PubMed]
  48. Beroukhim, R.; Mermel, C.H.; Porter, D.; Wei, G.; Raychaudhuri, S.; Donovan, J.; Barretina, J.; Boehm, J.S.; Dobson, J.; Urashima, M.; et al. The landscape of somatic copy-number alteration across human cancers. Nature 2010, 463, 899–905. [Google Scholar] [CrossRef]
  49. Cory, S.; Roberts, A.W.; Colman, P.M.; Adams, J.M. Targeting BCL-2-like Proteins to Kill Cancer Cells. Trends Cancer 2016, 2, 443–460. [Google Scholar] [CrossRef] [PubMed]
  50. McDonnell, T.J.; Deane, N.; Platt, F.M.; Nunez, G.; Jaeger, U.; McKearn, J.P.; Korsmeyer, S.J. bcl-2-immunoglobulin transgenic mice demonstrate extended B cell survival and follicular lymphoproliferation. Cell 1989, 57, 79–88. [Google Scholar] [CrossRef]
  51. Strasser, A.; Whittingham, S.; Vaux, D.L.; Bath, M.L.; Adams, J.M.; Cory, S.; Harris, A.W. Enforced BCL2 expression in B-lymphoid cells prolongs antibody responses and elicits autoimmune disease. Proc. Natl. Acad. Sci. USA 1991, 88, 8661–8665. [Google Scholar] [CrossRef] [Green Version]
  52. Sentman, C.L.; Shutter, J.R.; Hockenbery, D.; Kanagawa, O.; Korsmeyer, S.J. bcl-2 inhibits multiple forms of apoptosis but not negative selection in thymocytes. Cell 1991, 67, 879–888. [Google Scholar] [CrossRef]
  53. Thomas, S.; Quinn, B.A.; Das, S.K.; Dash, R.; Emdad, L.; Dasgupta, S.; Wang, X.Y.; Dent, P.; Reed, J.C.; Pellecchia, M.; et al. Targeting the Bcl-2 family for cancer therapy. Expert Opin. Ther. Targets 2013, 17, 61–75. [Google Scholar] [CrossRef] [Green Version]
  54. Khoo, K.H.; Verma, C.S.; Lane, D.P. Drugging the p53 pathway: Understanding the route to clinical efficacy. Nat. Rev. Drug Discov. 2014, 13, 217–236. [Google Scholar] [CrossRef]
  55. Kruiswijk, F.; Labuschagne, C.F.; Vousden, K.H. p53 in survival, death and metabolic health: A lifeguard with a licence to kill. Nat. Rev. Mol. Cell Biol. 2015, 16, 393–405. [Google Scholar] [CrossRef] [PubMed]
  56. Green, D.R.; Kroemer, G. Cytoplasmic functions of the tumour suppressor p53. Nature 2009, 458, 1127–1130. [Google Scholar] [CrossRef] [PubMed]
  57. Zou, Y.; Ma, D.; Wang, Y. The PROTAC technology in drug development. Cell Biochem. Funct. 2019, 37, 21–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Marino, G.; Niso-Santano, M.; Baehrecke, E.H.; Kroemer, G. Self-consumption: The interplay of autophagy and apoptosis. Nat. Rev. Mol. Cell Biol. 2014, 15, 81–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Hientz, K.; Mohr, A.; Bhakta-Guha, D.; Efferth, T. The role of p53 in cancer drug resistance and targeted chemotherapy. Oncotarget 2017, 8, 8921–8946. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Sanz, G.; Singh, M.; Peuget, S.; Selivanova, G. Inhibition of p53 inhibitors: Progress, challenges and perspectives. J. Mol. Cell Biol. 2019, 11, 586–599. [Google Scholar] [CrossRef] [Green Version]
  61. Liu, B.; Chen, Y.; St Clair, D.K. ROS and p53: A versatile partnership. Free Radic. Biol. Med. 2008, 44, 1529–1535. [Google Scholar] [CrossRef] [Green Version]
  62. Mita, A.C.; Mita, M.M.; Nawrocki, S.T.; Giles, F.J. Survivin: Key regulator of mitosis and apoptosis and novel target for cancer therapeutics. Clin. Cancer Res. 2008, 14, 5000–5005. [Google Scholar] [CrossRef] [Green Version]
  63. Vaux, D.L.; Silke, J. IAPs, RINGs and ubiquitylation. Nat. Rev. Mol. Cell Biol. 2005, 6, 287–297. [Google Scholar] [CrossRef]
  64. Varfolomeev, E.; Blankenship, J.W.; Wayson, S.M.; Fedorova, A.V.; Kayagaki, N.; Garg, P.; Zobel, K.; Dynek, J.N.; Elliott, L.O.; Wallweber, H.J.; et al. IAP antagonists induce autoubiquitination of c-IAPs, NF-kappaB activation, and TNFalpha-dependent apoptosis. Cell 2007, 131, 669–681. [Google Scholar] [CrossRef] [Green Version]
  65. Salvesen, G.S.; Duckett, C.S. IAP proteins: Blocking the road to death’s door. Nat. Rev. Mol. Cell Biol. 2002, 3, 401–410. [Google Scholar] [CrossRef]
  66. Eckelman, B.P.; Salvesen, G.S.; Scott, F.L. Human inhibitor of apoptosis proteins: Why XIAP is the black sheep of the family. EMBO Rep. 2006, 7, 988–994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. LaCasse, E.C.; Mahoney, D.J.; Cheung, H.H.; Plenchette, S.; Baird, S.; Korneluk, R.G. IAP-targeted therapies for cancer. Oncogene 2008, 27, 6252–6275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Hunter, A.M.; LaCasse, E.C.; Korneluk, R.G. The inhibitors of apoptosis (IAPs) as cancer targets. Apoptosis 2007, 12, 1543–1568. [Google Scholar] [CrossRef] [PubMed]
  69. Peng, X.H.; Karna, P.; O’Regan, R.M.; Liu, X.; Naithani, R.; Moriarty, R.M.; Wood, W.C.; Lee, H.Y.; Yang, L. Down-regulation of inhibitor of apoptosis proteins by deguelin selectively induces apoptosis in breast cancer cells. Mol. Pharm. 2007, 71, 101–111. [Google Scholar] [CrossRef] [Green Version]
  70. Esposito, I.; Kleeff, J.; Abiatari, I.; Shi, X.; Giese, N.; Bergmann, F.; Roth, W.; Friess, H.; Schirmacher, P. Overexpression of cellular inhibitor of apoptosis protein 2 is an early event in the progression of pancreatic cancer. J. Clin. Pathol. 2007, 60, 885–895. [Google Scholar] [CrossRef] [Green Version]
  71. Zhang, S.; Ding, F.; Luo, A.; Chen, A.; Yu, Z.; Ren, S.; Liu, Z.; Zhang, L. XIAP is highly expressed in esophageal cancer and its downregulation by RNAi sensitizes esophageal carcinoma cell lines to chemotherapeutics. Cancer Biol. Ther. 2007, 6, 973–980. [Google Scholar] [CrossRef] [Green Version]
  72. Rathore, R.; McCallum, J.E.; Varghese, E.; Florea, A.M.; Busselberg, D. Overcoming chemotherapy drug resistance by targeting inhibitors of apoptosis proteins (IAPs). Apoptosis 2017, 22, 898–919. [Google Scholar] [CrossRef]
  73. Shiozaki, A.; Kataoka, K.; Fujimura, M.; Yuki, H.; Sakai, M.; Saito, S. Survivin inhibits apoptosis in cytotrophoblasts. Placenta 2003, 24, 65–76. [Google Scholar] [CrossRef] [PubMed]
  74. Lehner, R.; Bobak, J.; Kim, N.W.; Shroyer, A.L.; Shroyer, K.R. Localization of telomerase hTERT protein and survivin in placenta: Relation to placental development and hydatidiform mole. Obstet. Gynecol. 2001, 97, 965–970. [Google Scholar] [CrossRef] [PubMed]
  75. O’Connor, D.S.; Grossman, D.; Plescia, J.; Li, F.; Zhang, H.; Villa, A.; Tognin, S.; Marchisio, P.C.; Altieri, D.C. Regulation of apoptosis at cell division by p34cdc2 phosphorylation of survivin. Proc. Natl. Acad. Sci. USA 2000, 97, 13103–13107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Altieri, D.C. Survivin and IAP proteins in cell-death mechanisms. Biochem. J. 2010, 430, 199–205. [Google Scholar] [CrossRef] [Green Version]
  77. Waligorska-Stachura, J.; Jankowska, A.; Wasko, R.; Liebert, W.; Biczysko, M.; Czarnywojtek, A.; Baszko-Blaszyk, D.; Shimek, V.; Ruchala, M. Survivin--prognostic tumor biomarker in human neoplasms—Review. Ginekol. Pol. 2012, 83, 537–540. [Google Scholar]
  78. Xu, C.; Yamamoto-Ibusuki, M.; Yamamoto, Y.; Yamamoto, S.; Fujiwara, S.; Murakami, K.; Okumura, Y.; Yamaguchi, L.; Fujiki, Y.; Iwase, H. High survivin mRNA expression is a predictor of poor prognosis in breast cancer: A comparative study at the mRNA and protein level. Breast Cancer 2012. [Google Scholar] [CrossRef]
  79. Kelly, R.J.; Lopez-Chavez, A.; Citrin, D.; Janik, J.E.; Morris, J.C. Impacting tumor cell-fate by targeting the inhibitor of apoptosis protein survivin. Mol. Cancer 2011, 10, 35. [Google Scholar] [CrossRef] [Green Version]
  80. Neophytou, C.M.; Constantinou, A.C. C. Survivin: Transcriptional Regulation and Protein Function in Cancer. J. Immunol. 2017. [Google Scholar]
  81. Richmond, A. Nf-kappa B, chemokine gene transcription and tumour growth. Nat. Rev. Immunol. 2002, 2, 664–674. [Google Scholar] [CrossRef]
  82. Dohi, T.; Okada, K.; Xia, F.; Wilford, C.E.; Samuel, T.; Welsh, K.; Marusawa, H.; Zou, H.; Armstrong, R.; Matsuzawa, S.; et al. An IAP-IAP complex inhibits apoptosis. J. Biol. Chem. 2004, 279, 34087–34090. [Google Scholar] [CrossRef] [Green Version]
  83. Song, Z.; Yao, X.; Wu, M. Direct interaction between survivin and Smac/DIABLO is essential for the anti-apoptotic activity of survivin during taxol-induced apoptosis. J. Biol. Chem. 2003, 278, 23130–23140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ceballos-Cancino, G.; Espinosa, M.; Maldonado, V.; Melendez-Zajgla, J. Regulation of mitochondrial Smac/DIABLO-selective release by survivin. Oncogene 2007, 26, 7569–7575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Li, F.; Aljahdali, I.; Ling, X. Cancer therapeutics using survivin BIRC5 as a target: What can we do after over two decades of study? J. Exp. Clin. Cancer Res. 2019, 38, 368. [Google Scholar] [CrossRef] [Green Version]
  86. Hay, N. The Akt-mTOR tango and its relevance to cancer. Cancer Cell 2005, 8, 179–183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Chen, Y.; Wang, T.; Du, J.; Li, Y.; Wang, X.; Zhou, Y.; Yu, X.; Fan, W.; Zhu, Q.; Tong, X.; et al. The Critical Role of PTEN/PI3K/AKT Signaling Pathway in Shikonin-Induced Apoptosis and Proliferation Inhibition of Chronic Myeloid Leukemia. Cell. Physiol. Biochem. Int. J. Exp. Cell. Physiol. Biochem. Pharmacol. 2018, 47, 981–993. [Google Scholar] [CrossRef]
  88. Rahmani, F.; Ziaeemehr, A.; Shahidsales, S.; Gharib, M.; Khazaei, M.; Ferns, G.A.; Ryzhikov, M.; Avan, A.; Hassanian, S.M. Role of regulatory miRNAs of the PI3K/AKT/mTOR signaling in the pathogenesis of hepatocellular carcinoma. J. Cell. Physiol. 2020, 235, 4146–4152. [Google Scholar] [CrossRef]
  89. Soltani, A.; Torki, S.; Ghahfarokhi, M.S.; Jami, M.S.; Ghatrehsamani, M. Targeting the phosphoinositide 3-kinase/AKT pathways by small molecules and natural compounds as a therapeutic approach for breast cancer cells. Mol. Biol. Rep. 2019, 46, 4809–4816. [Google Scholar] [CrossRef] [PubMed]
  90. Ediriweera, M.K.; Tennekoon, K.H.; Samarakoon, S.R. Role of the PI3K/AKT/mTOR signaling pathway in ovarian cancer: Biological and therapeutic significance. Semin. Cancer Biol. 2019, 59, 147–160. [Google Scholar] [CrossRef] [PubMed]
  91. Wu, D.M.; Zhang, T.; Liu, Y.B.; Deng, S.H.; Han, R.; Liu, T.; Li, J.; Xu, Y. The PAX6-ZEB2 axis promotes metastasis and cisplatin resistance in non-small cell lung cancer through PI3K/AKT signaling. Cell Death Dis. 2019, 10, 349. [Google Scholar] [CrossRef]
  92. Rittler, D.; Baranyi, M.; Molnar, E.; Garay, T.; Jalsovszky, I.; Varga, I.K.; Hegedus, L.; Aigner, C.; Tovari, J.; Timar, J.; et al. The Antitumor Effect of Lipophilic Bisphosphonate BPH1222 in Melanoma Models: The Role of the PI3K/Akt Pathway and the Small G Protein Rheb. Int. J. Mol. Sci. 2019, 20, 4917. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Fayard, E.; Tintignac, L.A.; Baudry, A.; Hemmings, B.A. Protein kinase B/Akt at a glance. J. Cell Sci. 2005, 118, 5675–5678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Stambolic, V.; Suzuki, A.; de la Pompa, J.L.; Brothers, G.M.; Mirtsos, C.; Sasaki, T.; Ruland, J.; Penninger, J.M.; Siderovski, D.P.; Mak, T.W. Negative regulation of PKB/Akt-dependent cell survival by the tumor suppressor PTEN. Cell 1998, 95, 29–39. [Google Scholar] [CrossRef] [Green Version]
  95. Vazquez, F.; Ramaswamy, S.; Nakamura, N.; Sellers, W.R. Phosphorylation of the PTEN tail regulates protein stability and function. Mol. Cell Biol. 2000, 20, 5010–5018. [Google Scholar] [CrossRef] [Green Version]
  96. Stephens, L.; Anderson, K.; Stokoe, D.; Erdjument-Bromage, H.; Painter, G.F.; Holmes, A.B.; Gaffney, P.R.; Reese, C.B.; McCormick, F.; Tempst, P.; et al. Protein kinase B kinases that mediate phosphatidylinositol 3,4,5-trisphosphate-dependent activation of protein kinase B. Science 1998, 279, 710–714. [Google Scholar] [CrossRef]
  97. Sarbassov, D.D.; Guertin, D.A.; Ali, S.M.; Sabatini, D.M. Phosphorylation and regulation of Akt/PKB by the rictor-mTOR complex. Science 2005, 307, 1098–1101. [Google Scholar] [CrossRef] [Green Version]
  98. Jacinto, E.; Facchinetti, V.; Liu, D.; Soto, N.; Wei, S.; Jung, S.Y.; Huang, Q.; Qin, J.; Su, B. SIN1/MIP1 maintains rictor-mTOR complex integrity and regulates Akt phosphorylation and substrate specificity. Cell 2006, 127, 125–137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Alessi, D.R.; Andjelkovic, M.; Caudwell, B.; Cron, P.; Morrice, N.; Cohen, P.; Hemmings, B.A. Mechanism of activation of protein kinase B by insulin and IGF-1. EMBO J. 1996, 15, 6541–6551. [Google Scholar] [CrossRef] [Green Version]
  100. Carnero, A.; Blanco-Aparicio, C.; Renner, O.; Link, W.; Leal, J.F. The PTEN/PI3K/AKT signalling pathway in cancer, therapeutic implications. Curr. Cancer Drug Targets 2008, 8, 187–198. [Google Scholar] [CrossRef]
  101. Sale, E.M.; Sale, G.J. Protein kinase B: Signalling roles and therapeutic targeting. Cell Mol. Life Sci. 2008, 65, 113–127. [Google Scholar] [CrossRef]
  102. Asanuma, H.; Torigoe, T.; Kamiguchi, K.; Hirohashi, Y.; Ohmura, T.; Hirata, K.; Sato, M.; Sato, N. Survivin expression is regulated by coexpression of human epidermal growth factor receptor 2 and epidermal growth factor receptor via phosphatidylinositol 3-kinase/AKT signaling pathway in breast cancer cells. Cancer Res. 2005, 65, 11018–11025. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Crawford, A.; Nahta, R. Targeting Bcl-2 in Herceptin-Resistant Breast Cancer Cell Lines. Curr. Pharm. Pers. Med. 2011, 9, 184–190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Du, K.; Montminy, M. CREB is a regulatory target for the protein kinase Akt/PKB. J. Biol. Chem. 1998, 273, 32377–32379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Song, G.; Ouyang, G.; Bao, S. The activation of Akt/PKB signaling pathway and cell survival. J. Cell Mol. Med. 2005, 9, 59–71. [Google Scholar] [CrossRef] [PubMed]
  106. Liu, R.; Chen, Y.; Liu, G.; Li, C.; Song, Y.; Cao, Z.; Li, W.; Hu, J.; Lu, C.; Liu, Y. PI3K/AKT pathway as a key link modulates the multidrug resistance of cancers. Cell Death Dis. 2020, 11, 797. [Google Scholar] [CrossRef]
  107. Welsh, G.I.; Wilson, C.; Proud, C.G. GSK3: A SHAGGY frog story. Trends Cell Biol. 1996, 6, 274–279. [Google Scholar] [CrossRef]
  108. Srivastava, A.K.; Pandey, S.K. Potential mechanism(s) involved in the regulation of glycogen synthesis by insulin. Mol. Cell Biochem. 1998, 182, 135–141. [Google Scholar] [CrossRef]
  109. Diehl, J.A.; Cheng, M.; Roussel, M.F.; Sherr, C.J. Glycogen synthase kinase-3beta regulates cyclin D1 proteolysis and subcellular localization. Genes Dev. 1998, 12, 3499–3511. [Google Scholar] [CrossRef] [Green Version]
  110. Somanath, P.R.; Vijai, J.; Kichina, J.V.; Byzova, T.; Kandel, E.S. The role of PAK-1 in activation of MAP kinase cascade and oncogenic transformation by Akt. Oncogene 2009, 28, 2365–2369. [Google Scholar] [CrossRef] [Green Version]
  111. Manning, B.D.; Cantley, L.C. AKT/PKB signaling: Navigating downstream. Cell 2007, 129, 1261–1274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Zhang, X.; Tang, N.; Hadden, T.J.; Rishi, A.K. Akt, FoxO and regulation of apoptosis. Biochim. Biophys. Acta 2011, 1813, 1978–1986. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Farhan, M.; Silva, M.; Xingan, X.; Huang, Y.; Zheng, W. Role of FOXO Transcription Factors in Cancer Metabolism and Angiogenesis. Cells 2020, 9, 1586. [Google Scholar] [CrossRef]
  114. Slamon, D.J.; Clark, G.M.; Wong, S.G.; Levin, W.J.; Ullrich, A.; McGuire, W.L. Human breast cancer: Correlation of relapse and survival with amplification of the HER-2/neu oncogene. Science 1987, 235, 177–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Zhou, B.P.; Hu, M.C.; Miller, S.A.; Yu, Z.; Xia, W.; Lin, S.Y.; Hung, M.C. HER-2/neu blocks tumor necrosis factor-induced apoptosis via the Akt/NF-kappaB pathway. J. Biol. Chem. 2000, 275, 8027–8031. [Google Scholar] [CrossRef] [Green Version]
  116. Salomon, D.S.; Brandt, R.; Ciardiello, F.; Normanno, N. Epidermal growth factor-related peptides and their receptors in human malignancies. Crit. Rev. Oncol. Hematol. 1995, 19, 183–232. [Google Scholar] [CrossRef]
  117. Shayesteh, L.; Lu, Y.; Kuo, W.L.; Baldocchi, R.; Godfrey, T.; Collins, C.; Pinkel, D.; Powell, B.; Mills, G.B.; Gray, J.W. PIK3CA is implicated as an oncogene in ovarian cancer. Nat. Genet. 1999, 21, 99–102. [Google Scholar] [CrossRef]
  118. Byun, D.S.; Cho, K.; Ryu, B.K.; Lee, M.G.; Park, J.I.; Chae, K.S.; Kim, H.J.; Chi, S.G. Frequent monoallelic deletion of PTEN and its reciprocal associatioin with PIK3CA amplification in gastric carcinoma. Int. J. Cancer 2003, 104, 318–327. [Google Scholar] [CrossRef]
  119. Samuels, Y.; Diaz, L.A., Jr.; Schmidt-Kittler, O.; Cummins, J.M.; Delong, L.; Cheong, I.; Rago, C.; Huso, D.L.; Lengauer, C.; Kinzler, K.W.; et al. Mutant PIK3CA promotes cell growth and invasion of human cancer cells. Cancer Cell 2005, 7, 561–573. [Google Scholar] [CrossRef] [Green Version]
  120. Stal, O.; Perez-Tenorio, G.; Akerberg, L.; Olsson, B.; Nordenskjold, B.; Skoog, L.; Rutqvist, L.E. Akt kinases in breast cancer and the results of adjuvant therapy. Breast Cancer Res. 2003, 5, R37–R44. [Google Scholar] [CrossRef] [Green Version]
  121. Roy, H.K.; Olusola, B.F.; Clemens, D.L.; Karolski, W.J.; Ratashak, A.; Lynch, H.T.; Smyrk, T.C. AKT proto-oncogene overexpression is an early event during sporadic colon carcinogenesis. Carcinogenesis 2002, 23, 201–205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Liu, T.; Zhu, J.; Du, W.; Ning, W.; Zhang, Y.; Zeng, Y.; Liu, Z.; Huang, J.A. AKT2 drives cancer progression and is negatively modulated by miR-124 in human lung adenocarcinoma. Respir. Res. 2020, 21, 227. [Google Scholar] [CrossRef] [PubMed]
  123. Rychahou, P.G.; Kang, J.; Gulhati, P.; Doan, H.Q.; Chen, L.A.; Xiao, S.Y.; Chung, D.H.; Evers, B.M. Akt2 overexpression plays a critical role in the establishment of colorectal cancer metastasis. Proc. Natl. Acad. Sci. USA 2008, 105, 20315–20320. [Google Scholar] [CrossRef] [Green Version]
  124. Arboleda, M.J.; Lyons, J.F.; Kabbinavar, F.F.; Bray, M.R.; Snow, B.E.; Ayala, R.; Danino, M.; Karlan, B.Y.; Slamon, D.J. Overexpression of AKT2/protein kinase Bbeta leads to up-regulation of beta1 integrins, increased invasion, and metastasis of human breast and ovarian cancer cells. Cancer Res. 2003, 63, 196–206. [Google Scholar] [PubMed]
  125. Ozes, O.N.; Akca, H.; Mayo, L.D.; Gustin, J.A.; Maehama, T.; Dixon, J.E.; Donner, D.B. A phosphatidylinositol 3-kinase/Akt/mTOR pathway mediates and PTEN antagonizes tumor necrosis factor inhibition of insulin signaling through insulin receptor substrate-1. Proc. Natl. Acad. Sci. USA 2001, 98, 4640–4645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Osaki, M.; Oshimura, M.; Ito, H. PI3K-Akt pathway: Its functions and alterations in human cancer. Apoptosis 2004, 9, 667–676. [Google Scholar] [CrossRef] [PubMed]
  127. Neophytou, C.M.; Panagi, M.; Stylianopoulos, T.; Papageorgis, P. The Role of Tumor Microenvironment in Cancer Metastasis: Molecular Mechanisms and Therapeutic Opportunities. Cancers 2021, 13, 2053. [Google Scholar] [CrossRef] [PubMed]
  128. Neophytou, C.M.; Pierides, C.; Christodoulou, M.I.; Costeas, P.; Kyriakou, T.C.; Papageorgis, P. The Role of Tumor-Associated Myeloid Cells in Modulating Cancer Therapy. Front. Oncol. 2020, 10, 899. [Google Scholar] [CrossRef] [PubMed]
  129. Martinez-Lostao, L.; Anel, A.; Pardo, J. How Do Cytotoxic Lymphocytes Kill Cancer Cells? Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2015, 21, 5047–5056. [Google Scholar] [CrossRef] [Green Version]
  130. Anel, A.; Buferne, M.; Boyer, C.; Schmitt-Verhulst, A.M.; Golstein, P. T cell receptor-induced Fas ligand expression in cytotoxic T lymphocyte clones is blocked by protein tyrosine kinase inhibitors and cyclosporin A. Eur. J. Immunol. 1994, 24, 2469–2476. [Google Scholar] [CrossRef] [PubMed]
  131. Bossi, G.; Griffiths, G.M. CTL secretory lysosomes: Biogenesis and secretion of a harmful organelle. Semin. Immunol. 2005, 17, 87–94. [Google Scholar] [CrossRef]
  132. De Saint Basile, G.; Menasche, G.; Fischer, A. Molecular mechanisms of biogenesis and exocytosis of cytotoxic granules. Nat. Rev. Immunol. 2010, 10, 568–579. [Google Scholar] [CrossRef] [PubMed]
  133. Tecchio, C.; Huber, V.; Scapini, P.; Calzetti, F.; Margotto, D.; Todeschini, G.; Pilla, L.; Martinelli, G.; Pizzolo, G.; Rivoltini, L.; et al. IFNalpha-stimulated neutrophils and monocytes release a soluble form of TNF-related apoptosis-inducing ligand (TRAIL/Apo-2 ligand) displaying apoptotic activity on leukemic cells. Blood 2004, 103, 3837–3844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Lee, G.T.; Kim, J.H.; Kwon, S.J.; Stein, M.N.; Hong, J.H.; Nagaya, N.; Billakanti, S.; Kim, M.M.; Kim, W.J.; Kim, I.Y. Dihydrotestosterone Increases Cytotoxic Activity of Macrophages on Prostate Cancer Cells via TRAIL. Endocrinology 2019, 160, 2049–2060. [Google Scholar] [CrossRef]
  135. Freeman, Z.T.; Nirschl, T.R.; Hovelson, D.H.; Johnston, R.J.; Engelhardt, J.J.; Selby, M.J.; Kochel, C.M.; Lan, R.Y.; Zhai, J.; Ghasemzadeh, A.; et al. A conserved intratumoral regulatory T cell signature identifies 4-1BB as a pan-cancer target. J. Clin. Investig. 2020, 130, 1405–1416. [Google Scholar] [CrossRef]
  136. Hu, B.S.; Tang, T.; Jia, J.L.; Xie, B.C.; Wu, T.L.; Sheng, Y.Y.; Xue, Y.Z.; Tang, H.M. CD137 agonist induces gastric cancer cell apoptosis by enhancing the functions of CD8(+) T cells via NF-kappaB signaling. Cancer Cell Int. 2020, 20, 513. [Google Scholar] [CrossRef] [PubMed]
  137. Yu, D.; Zhong, Y.; Li, X.; Li, Y.; Li, X.; Cao, J.; Fan, H.; Yuan, Y.; Ji, Z.; Qiao, B.; et al. ILs-3, 6 and 11 increase, but ILs-10 and 24 decrease stemness of human prostate cancer cells in vitro. Oncotarget 2015, 6, 42687–42703. [Google Scholar] [CrossRef]
  138. Quail, D.F.; Joyce, J.A. Microenvironmental regulation of tumor progression and metastasis. Nat. Med. 2013, 19, 1423–1437. [Google Scholar] [CrossRef]
  139. Bejarano, L.; Jordao, M.J.C.; Joyce, J.A. Therapeutic Targeting of the Tumor Microenvironment. Cancer Discov. 2021, 11, 933–959. [Google Scholar] [CrossRef] [PubMed]
  140. Khalaf, K.; Hana, D.; Chou, J.T.; Singh, C.; Mackiewicz, A.; Kaczmarek, M. Aspects of the Tumor Microenvironment Involved in Immune Resistance and Drug Resistance. Front. Immunol. 2021, 12, 656364. [Google Scholar] [CrossRef]
  141. Jain, R.K.; Martin, J.D.; Stylianopoulos, T. The role of mechanical forces in tumor growth and therapy. Annu. Rev. Biomed. Eng. 2014, 16, 321–346. [Google Scholar] [CrossRef] [Green Version]
  142. Muz, B.; de la Puente, P.; Azab, F.; Azab, A.K. The role of hypoxia in cancer progression, angiogenesis, metastasis, and resistance to therapy. Hypoxia 2015, 3, 83–92. [Google Scholar] [CrossRef] [Green Version]
  143. Gerweck, L.E.; Vijayappa, S.; Kozin, S. Tumor pH controls the in vivo efficacy of weak acid and base chemotherapeutics. Mol. Cancer Ther. 2006, 5, 1275–1279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Yamaguchi, H.; Sakai, R. Direct Interaction between Carcinoma Cells and Cancer Associated Fibroblasts for the Regulation of Cancer Invasion. Cancers 2015, 7, 2054–2062. [Google Scholar] [CrossRef]
  145. Joshi, R.S.; Kanugula, S.S.; Sudhir, S.; Pereira, M.P.; Jain, S.; Aghi, M.K. The Role of Cancer-Associated Fibroblasts in Tumor Progression. Cancers 2021, 13, 1399. [Google Scholar] [CrossRef]
  146. Augsten, M. Cancer-associated fibroblasts as another polarized cell type of the tumor microenvironment. Front. Oncol. 2014, 4, 62. [Google Scholar] [CrossRef]
  147. Guo, L.; Li, B.; Yang, J.; Shen, J.; Ji, J.; Miao, M. Fibroblastderived exosomal microRNA369 potentiates migration and invasion of lung squamous cell carcinoma cells via NF1mediated MAPK signaling pathway. Int. J. Mol. Med. 2020, 46, 595–608. [Google Scholar] [CrossRef] [PubMed]
  148. Valcz, G.; Sipos, F.; Tulassay, Z.; Molnar, B.; Yagi, Y. Importance of carcinoma-associated fibroblast-derived proteins in clinical oncology. J. Clin. Pathol. 2014, 67, 1026–1031. [Google Scholar] [CrossRef] [Green Version]
  149. Paraiso, K.H.; Smalley, K.S. Fibroblast-mediated drug resistance in cancer. Biochem. Pharmacol. 2013, 85, 1033–1041. [Google Scholar] [CrossRef] [PubMed]
  150. Choe, C.; Shin, Y.S.; Kim, S.H.; Jeon, M.J.; Choi, S.J.; Lee, J.; Kim, J. Tumor-stromal interactions with direct cell contacts enhance motility of non-small cell lung cancer cells through the hedgehog signaling pathway. Anticancer Res. 2013, 33, 3715–3723. [Google Scholar]
  151. Ren, J.; Ding, L.; Zhang, D.; Shi, G.; Xu, Q.; Shen, S.; Wang, Y.; Wang, T.; Hou, Y. Carcinoma-associated fibroblasts promote the stemness and chemoresistance of colorectal cancer by transferring exosomal lncRNA H19. Theranostics 2018, 8, 3932–3948. [Google Scholar] [CrossRef]
  152. Chen, W.J.; Ho, C.C.; Chang, Y.L.; Chen, H.Y.; Lin, C.A.; Ling, T.Y.; Yu, S.L.; Yuan, S.S.; Chen, Y.J.; Lin, C.Y.; et al. Cancer-associated fibroblasts regulate the plasticity of lung cancer stemness via paracrine signalling. Nat. Commun. 2014, 5, 3472. [Google Scholar] [CrossRef] [PubMed]
  153. Lacina, L.; Plzak, J.; Kodet, O.; Szabo, P.; Chovanec, M.; Dvorankova, B.; Smetana, K., Jr. Cancer Microenvironment: What Can We Learn from the Stem Cell Niche. Int. J. Mol. Sci. 2015, 16, 24094–24110. [Google Scholar] [CrossRef]
  154. Xi, C.; Wang, J.; Sun, H.; Zhang, X.; Kang, H. Loss of microRNA-30e induced by extracellular vesicles from cancer-associated fibroblasts promotes breast cancer progression by binding to CTHRC1. Exp. Mol. Pathol. 2021, 118, 104586. [Google Scholar] [CrossRef]
  155. Eichelmann, A.K.; Matuszcak, C.; Hummel, R.; Haier, J. Role of miRNAs in cell signaling of cancer associated fibroblasts. Int. J. Biochem. Cell Biol. 2018, 101, 94–102. [Google Scholar] [CrossRef]
  156. Boomgarden, A.C.; Sheehan, C.; D’Souza-Schorey, C. Extracellular Vesicles in the Tumor Microenvironment: Various Implications in Tumor Progression. Adv. Exp. Med. Biol. 2020, 1259, 155–170. [Google Scholar] [CrossRef] [PubMed]
  157. Wang, H.; Wei, H.; Wang, J.; Li, L.; Chen, A.; Li, Z. MicroRNA-181d-5p-Containing Exosomes Derived from CAFs Promote EMT by Regulating CDX2/HOXA5 in Breast Cancer. Mol. Ther. Nucleic Acids 2020, 19, 654–667. [Google Scholar] [CrossRef]
  158. Beermann, J.; Piccoli, M.T.; Viereck, J.; Thum, T. Non-coding RNAs in Development and Disease: Background, Mechanisms, and Therapeutic Approaches. Physiol. Rev. 2016, 96, 1297–1325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Di Martino, M.T.; Riillo, C.; Scionti, F.; Grillone, K.; Polera, N.; Caracciolo, D.; Arbitrio, M.; Tagliaferri, P.; Tassone, P. miRNAs and lncRNAs as Novel Therapeutic Targets to Improve Cancer Immunotherapy. Cancers 2021, 13, 1587. [Google Scholar] [CrossRef] [PubMed]
  160. Cui, H.; Onyango, P.; Brandenburg, S.; Wu, Y.; Hsieh, C.L.; Feinberg, A.P. Loss of imprinting in colorectal cancer linked to hypomethylation of H19 and IGF2. Cancer Res. 2002, 62, 6442–6446. [Google Scholar]
  161. Berteaux, N.; Lottin, S.; Monte, D.; Pinte, S.; Quatannens, B.; Coll, J.; Hondermarck, H.; Curgy, J.J.; Dugimont, T.; Adriaenssens, E. H19 mRNA-like noncoding RNA promotes breast cancer cell proliferation through positive control by E2F1. J. Biol. Chem. 2005, 280, 29625–29636. [Google Scholar] [CrossRef] [Green Version]
  162. Lecerf, C.; Peperstraete, E.; Le Bourhis, X.; Adriaenssens, E. Propagation and Maintenance of Cancer Stem Cells: A Major Influence of the Long Non-Coding RNA H19. Cells 2020, 9, 2613. [Google Scholar] [CrossRef] [PubMed]
  163. Vermeulen, L.; De Sousa, E.M.F.; van der Heijden, M.; Cameron, K.; de Jong, J.H.; Borovski, T.; Tuynman, J.B.; Todaro, M.; Merz, C.; Rodermond, H.; et al. Wnt activity defines colon cancer stem cells and is regulated by the microenvironment. Nat. Cell Biol. 2010, 12, 468–476. [Google Scholar] [CrossRef]
  164. Gao, Z.; Wang, Q.; Ji, M.; Guo, X.; Li, L.; Su, X. Exosomal lncRNA UCA1 modulates cervical cancer stem cell self-renewal and differentiation through microRNA-122-5p/SOX2 axis. J. Transl. Med. 2021, 19, 229. [Google Scholar] [CrossRef]
  165. Lv, X.; Lian, Y.; Liu, Z.; Xiao, J.; Zhang, D.; Yin, X. Exosomal long non-coding RNA LINC00662 promotes non-small cell lung cancer progression by miR-320d/E2F1 axis. Aging 2021, 13, 6010–6024. [Google Scholar] [CrossRef]
  166. Liu, K.; Gao, L.; Ma, X.; Huang, J.J.; Chen, J.; Zeng, L.; Ashby, C.R., Jr.; Zou, C.; Chen, Z.S. Long non-coding RNAs regulate drug resistance in cancer. Mol. Cancer 2020, 19, 54. [Google Scholar] [CrossRef] [PubMed]
  167. Ying, L.; Zhu, Z.; Xu, Z.; He, T.; Li, E.; Guo, Z.; Liu, F.; Jiang, C.; Wang, Q. Cancer Associated Fibroblast-Derived Hepatocyte Growth Factor Inhibits the Paclitaxel-Induced Apoptosis of Lung Cancer A549 Cells by Up-Regulating the PI3K/Akt and GRP78 Signaling on a Microfluidic Platform. PLoS ONE 2015, 10, e0129593. [Google Scholar] [CrossRef] [Green Version]
  168. Chen, C.; Hou, J.; Yu, S.; Li, W.; Wang, X.; Sun, H.; Qin, T.; Claret, F.X.; Guo, H.; Liu, Z. Role of cancer-associated fibroblasts in the resistance to antitumor therapy, and their potential therapeutic mechanisms in non-small cell lung cancer. Oncol. Lett. 2021, 21, 413. [Google Scholar] [CrossRef] [PubMed]
  169. Tao, L.; Huang, G.; Wang, R.; Pan, Y.; He, Z.; Chu, X.; Song, H.; Chen, L. Cancer-associated fibroblasts treated with cisplatin facilitates chemoresistance of lung adenocarcinoma through IL-11/IL-11R/STAT3 signaling pathway. Sci. Rep. 2016, 6, 38408. [Google Scholar] [CrossRef]
  170. Bockhorn, J.; Dalton, R.; Nwachukwu, C.; Huang, S.; Prat, A.; Yee, K.; Chang, Y.F.; Huo, D.; Wen, Y.; Swanson, K.E.; et al. MicroRNA-30c inhibits human breast tumour chemotherapy resistance by regulating TWF1 and IL-11. Nat. Commun. 2013, 4, 1393. [Google Scholar] [CrossRef] [Green Version]
  171. Lokau, J.; Agthe, M.; Garbers, C. Generation of Soluble Interleukin-11 and Interleukin-6 Receptors: A Crucial Function for Proteases during Inflammation. Mediat. Inflamm. 2016, 2016, 1785021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Delbridge, A.R.; Grabow, S.; Strasser, A.; Vaux, D.L. Thirty years of BCL-2: Translating cell death discoveries into novel cancer therapies. Nat. Rev. Cancer 2016, 16, 99–109. [Google Scholar] [CrossRef]
  173. Real, P.J.; Sierra, A.; De Juan, A.; Segovia, J.C.; Lopez-Vega, J.M.; Fernandez-Luna, J.L. Resistance to chemotherapy via Stat3-dependent overexpression of Bcl-2 in metastatic breast cancer cells. Oncogene 2002, 21, 7611–7618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Borsellino, N.; Belldegrun, A.; Bonavida, B. Endogenous interleukin 6 is a resistance factor for cis-diamminedichloroplatinum and etoposide-mediated cytotoxicity of human prostate carcinoma cell lines. Cancer Res. 1995, 55, 4633–4639. [Google Scholar]
  175. Garcia-Tunon, I.; Ricote, M.; Ruiz, A.; Fraile, B.; Paniagua, R.; Royuela, M. IL-6, its receptors and its relationship with bcl-2 and bax proteins in infiltrating and in situ human breast carcinoma. Histopathology 2005, 47, 82–89. [Google Scholar] [CrossRef] [PubMed]
  176. Gritsko, T.; Williams, A.; Turkson, J.; Kaneko, S.; Bowman, T.; Huang, M.; Nam, S.; Eweis, I.; Diaz, N.; Sullivan, D.; et al. Persistent activation of stat3 signaling induces survivin gene expression and confers resistance to apoptosis in human breast cancer cells. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2006, 12, 11–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Kumari, N.; Dwarakanath, B.S.; Das, A.; Bhatt, A.N. Role of interleukin-6 in cancer progression and therapeutic resistance. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2016, 37, 11553–11572. [Google Scholar] [CrossRef] [PubMed]
  178. Jiang, X.P.; Yang, D.C.; Elliott, R.L.; Head, J.F. Down-regulation of expression of interleukin-6 and its receptor results in growth inhibition of MCF-7 breast cancer cells. Anticancer Res. 2011, 31, 2899–2906. [Google Scholar] [PubMed]
  179. So, K.A.; Min, K.J.; Hong, J.H.; Lee, J.K. Interleukin-6 expression by interactions between gynecologic cancer cells and human mesenchymal stem cells promotes epithelial-mesenchymal transition. Int. J. Oncol. 2015, 47, 1451–1459. [Google Scholar] [CrossRef] [PubMed]
  180. Kinoshita, H.; Hirata, Y.; Nakagawa, H.; Sakamoto, K.; Hayakawa, Y.; Takahashi, R.; Nakata, W.; Sakitani, K.; Serizawa, T.; Hikiba, Y.; et al. Interleukin-6 mediates epithelial-stromal interactions and promotes gastric tumorigenesis. PLoS ONE 2013, 8, e60914. [Google Scholar] [CrossRef] [PubMed]
  181. Wu, X.; Tao, P.; Zhou, Q.; Li, J.; Yu, Z.; Wang, X.; Li, J.; Li, C.; Yan, M.; Zhu, Z.; et al. IL-6 secreted by cancer-associated fibroblasts promotes epithelial-mesenchymal transition and metastasis of gastric cancer via JAK2/STAT3 signaling pathway. Oncotarget 2017, 8, 20741–20750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Ham, I.H.; Oh, H.J.; Jin, H.; Bae, C.A.; Jeon, S.M.; Choi, K.S.; Son, S.Y.; Han, S.U.; Brekken, R.A.; Lee, D.; et al. Targeting interleukin-6 as a strategy to overcome stroma-induced resistance to chemotherapy in gastric cancer. Mol. Cancer 2019, 18, 68. [Google Scholar] [CrossRef]
  183. Ma, Y.; Zhu, J.; Chen, S.; Li, T.; Ma, J.; Guo, S.; Hu, J.; Yue, T.; Zhang, J.; Wang, P.; et al. Activated gastric cancer-associated fibroblasts contribute to the malignant phenotype and 5-FU resistance via paracrine action in gastric cancer. Cancer Cell Int. 2018, 18, 104. [Google Scholar] [CrossRef] [PubMed]
  184. Hwang, R.F.; Moore, T.; Arumugam, T.; Ramachandran, V.; Amos, K.D.; Rivera, A.; Ji, B.; Evans, D.B.; Logsdon, C.D. Cancer-associated stromal fibroblasts promote pancreatic tumor progression. Cancer Res. 2008, 68, 918–926. [Google Scholar] [CrossRef] [Green Version]
  185. Muerkoster, S.; Wegehenkel, K.; Arlt, A.; Witt, M.; Sipos, B.; Kruse, M.L.; Sebens, T.; Kloppel, G.; Kalthoff, H.; Folsch, U.R.; et al. Tumor stroma interactions induce chemoresistance in pancreatic ductal carcinoma cells involving increased secretion and paracrine effects of nitric oxide and interleukin-1beta. Cancer Res. 2004, 64, 1331–1337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Setrerrahmane, S.; Xu, H. Tumor-related interleukins: Old validated targets for new anti-cancer drug development. Mol. Cancer 2017, 16, 153. [Google Scholar] [CrossRef] [PubMed]
  187. Tripathi, S.C.; Fahrmann, J.F.; Celiktas, M.; Aguilar, M.; Marini, K.D.; Jolly, M.K.; Katayama, H.; Wang, H.; Murage, E.N.; Dennison, J.B.; et al. MCAM Mediates Chemoresistance in Small-Cell Lung Cancer via the PI3K/AKT/SOX2 Signaling Pathway. Cancer Res. 2017, 77, 4414–4425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Jung, K.; Wang, P.; Gupta, N.; Gopal, K.; Wu, F.; Ye, X.; Alshareef, A.; Bigras, G.; McMullen, T.P.; Abdulkarim, B.S.; et al. Profiling gene promoter occupancy of Sox2 in two phenotypically distinct breast cancer cell subsets using chromatin immunoprecipitation and genome-wide promoter microarrays. Breast Cancer Res. 2014, 16, 470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Srinivasan, D.; Senbanjo, L.; Majumdar, S.; Franklin, R.B.; Chellaiah, M.A. Androgen receptor expression reduces stemness characteristics of prostate cancer cells (PC3) by repression of CD44 and SOX2. J. Cell. Biochem. 2018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Gao, H.; Teng, C.; Huang, W.; Peng, J.; Wang, C. SOX2 Promotes the Epithelial to Mesenchymal Transition of Esophageal Squamous Cells by Modulating Slug Expression through the Activation of STAT3/HIF-alpha Signaling. Int. J. Mol. Sci. 2015, 16, 21643–21657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Piva, M.; Domenici, G.; Iriondo, O.; Rabano, M.; Simoes, B.M.; Comaills, V.; Barredo, I.; Lopez-Ruiz, J.A.; Zabalza, I.; Kypta, R.; et al. Sox2 promotes tamoxifen resistance in breast cancer cells. EMBO Mol. Med. 2014, 6, 66–79. [Google Scholar] [CrossRef] [PubMed]
  192. Finicelli, M.; Benedetti, G.; Squillaro, T.; Pistilli, B.; Marcellusi, A.; Mariani, P.; Santinelli, A.; Latini, L.; Galderisi, U.; Giordano, A. Expression of stemness genes in primary breast cancer tissues: The role of SOX2 as a prognostic marker for detection of early recurrence. Oncotarget 2014, 5, 9678–9688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Dittmer, A.; Dittmer, J. Carcinoma-Associated Fibroblasts Promote Growth of Sox2-Expressing Breast Cancer Cells. Cancers 2020, 12, 3435. [Google Scholar] [CrossRef] [PubMed]
  194. Shekhar, M.P.; Santner, S.; Carolin, K.A.; Tait, L. Direct involvement of breast tumor fibroblasts in the modulation of tamoxifen sensitivity. Am. J. Pathol. 2007, 170, 1546–1560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Zhang, S.; Sun, Y. Targeting oncogenic SOX2 in human cancer cells: Therapeutic application. Protein Cell 2020, 11, 82–84. [Google Scholar] [CrossRef] [Green Version]
  196. Deying, W.; Feng, G.; Shumei, L.; Hui, Z.; Ming, L.; Hongqing, W. CAF-derived HGF promotes cell proliferation and drug resistance by up-regulating the c-Met/PI3K/Akt and GRP78 signalling in ovarian cancer cells. Biosci. Rep. 2017, 37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Moran-Jones, K. The Therapeutic Potential of Targeting the HGF/cMET Axis in Ovarian Cancer. Mol. Diagn. Ther. 2016, 20, 199–212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Otte, J.M.; Schmitz, F.; Kiehne, K.; Stechele, H.U.; Banasiewicz, T.; Krokowicz, P.; Nakamura, T.; Folsch, U.R.; Herzig, K. Functional expression of HGF and its receptor in human colorectal cancer. Digestion 2000, 61, 237–246. [Google Scholar] [CrossRef] [PubMed]
  199. Yamada, T.; Matsumoto, K.; Wang, W.; Li, Q.; Nishioka, Y.; Sekido, Y.; Sone, S.; Yano, S. Hepatocyte growth factor reduces susceptibility to an irreversible epidermal growth factor receptor inhibitor in EGFR-T790M mutant lung cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2010, 16, 174–183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Meng, Q.; Xia, C.; Fang, J.; Rojanasakul, Y.; Jiang, B.H. Role of PI3K and AKT specific isoforms in ovarian cancer cell migration, invasion and proliferation through the p70S6K1 pathway. Cell. Signal. 2006, 18, 2262–2271. [Google Scholar] [CrossRef] [PubMed]
  201. Yao, R.; Cooper, G.M. Requirement for phosphatidylinositol-3 kinase in the prevention of apoptosis by nerve growth factor. Science 1995, 267, 2003–2006. [Google Scholar] [CrossRef]
  202. Shin, B.K.; Wang, H.; Yim, A.M.; Le Naour, F.; Brichory, F.; Jang, J.H.; Zhao, R.; Puravs, E.; Tra, J.; Michael, C.W.; et al. Global profiling of the cell surface proteome of cancer cells uncovers an abundance of proteins with chaperone function. J. Biol. Chem. 2003, 278, 7607–7616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Fu, Y.; Lee, A.S. Glucose regulated proteins in cancer progression, drug resistance and immunotherapy. Cancer Biol. Ther. 2006, 5, 741–744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Zhang, J.; Jiang, Y.; Jia, Z.; Li, Q.; Gong, W.; Wang, L.; Wei, D.; Yao, J.; Fang, S.; Xie, K. Association of elevated GRP78 expression with increased lymph node metastasis and poor prognosis in patients with gastric cancer. Clin. Exp. Metastasis 2006, 23, 401–410. [Google Scholar] [CrossRef]
  205. Davidson, D.J.; Haskell, C.; Majest, S.; Kherzai, A.; Egan, D.A.; Walter, K.A.; Schneider, A.; Gubbins, E.F.; Solomon, L.; Chen, Z.; et al. Kringle 5 of human plasminogen induces apoptosis of endothelial and tumor cells through surface-expressed glucose-regulated protein 78. Cancer Res. 2005, 65, 4663–4672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Arnaudeau, S.; Arboit, P.; Bischof, P.; Shin-ya, K.; Tomida, A.; Tsuruo, T.; Irion, O.; Cohen, M. Glucose-regulated protein 78: A new partner of p53 in trophoblast. Proteomics 2009, 9, 5316–5327. [Google Scholar] [CrossRef] [PubMed]
  207. Fu, Y.; Li, J.; Lee, A.S. GRP78/BiP inhibits endoplasmic reticulum BIK and protects human breast cancer cells against estrogen starvation-induced apoptosis. Cancer Res. 2007, 67, 3734–3740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Zhou, H.; Zhang, Y.; Fu, Y.; Chan, L.; Lee, A.S. Novel mechanism of anti-apoptotic function of 78-kDa glucose-regulated protein (GRP78): Endocrine resistance factor in breast cancer, through release of B-cell lymphoma 2 (BCL-2) from BCL-2-interacting killer (BIK). J. Biol. Chem. 2011, 286, 25687–25696. [Google Scholar] [CrossRef] [Green Version]
  209. Hao, Y.; Zhang, L.; He, J.; Guo, Z.; Ying, L.; Xu, Z.; Zhang, J.; Lu, J.; Wang, Q. Functional investigation of NCI-H460-inducible myofibroblasts on the chemoresistance to VP-16 with a microfluidic 3D co-culture device. PLoS ONE 2013, 8, e61754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Zhou, B.; Sun, C.; Li, N.; Shan, W.; Lu, H.; Guo, L.; Guo, E.; Xia, M.; Weng, D.; Meng, L.; et al. Cisplatin-induced CCL5 secretion from CAFs promotes cisplatin-resistance in ovarian cancer via regulation of the STAT3 and PI3K/Akt signaling pathways. Int. J. Oncol. 2016, 48, 2087–2097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  211. Mehlen, P.; Delloye-Bourgeois, C.; Chedotal, A. Novel roles for Slits and netrins: Axon guidance cues as anticancer targets? Nat. Rev. Cancer 2011, 11, 188–197. [Google Scholar] [CrossRef]
  212. Kefeli, U.; Ucuncu Kefeli, A.; Cabuk, D.; Isik, U.; Sonkaya, A.; Acikgoz, O.; Ozden, E.; Uygun, K. Netrin-1 in cancer: Potential biomarker and therapeutic target? Tumour Biol. 2017, 39, 1010428317698388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Sung, P.J.; Rama, N.; Imbach, J.; Fiore, S.; Ducarouge, B.; Neves, D.; Chen, H.W.; Bernard, D.; Yang, P.C.; Bernet, A.; et al. Cancer-Associated Fibroblasts Produce Netrin-1 to Control Cancer Cell Plasticity. Cancer Res. 2019, 79, 3651–3661. [Google Scholar] [CrossRef]
  214. Adams, C.; Cazzanelli, G.; Rasul, S.; Hitchinson, B.; Hu, Y.; Coombes, R.C.; Raguz, S.; Yague, E. Apoptosis inhibitor TRIAP1 is a novel effector of drug resistance. Oncol. Rep. 2015, 34, 415–422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Park, W.R.; Nakamura, Y. p53CSV, a novel p53-inducible gene involved in the p53-dependent cell-survival pathway. Cancer Res. 2005, 65, 1197–1206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Ketteler, J.; Panic, A.; Reis, H.; Wittka, A.; Maier, P.; Herskind, C.; Yague, E.; Jendrossek, V.; Klein, D. Progression-Related Loss of Stromal Caveolin 1 Levels Mediates Radiation Resistance in Prostate Carcinoma via the Apoptosis Inhibitor TRIAP1. J. Clin. Med. 2019, 8, 348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Li, Y.; Tang, X.; He, Q.; Yang, X.; Ren, X.; Wen, X.; Zhang, J.; Wang, Y.; Liu, N.; Ma, J. Overexpression of Mitochondria Mediator Gene TRIAP1 by miR-320b Loss Is Associated with Progression in Nasopharyngeal Carcinoma. PLoS Genet. 2016, 12, e1006183. [Google Scholar] [CrossRef] [PubMed]
  218. Kunou, S.; Shimada, K.; Takai, M.; Sakamoto, A.; Aoki, T.; Hikita, T.; Kagaya, Y.; Iwamoto, E.; Sanada, M.; Shimada, S.; et al. Exosomes secreted from cancer-associated fibroblasts elicit anti-pyrimidine drug resistance through modulation of its transporter in malignant lymphoma. Oncogene 2021, 40, 3989–4003. [Google Scholar] [CrossRef]
  219. Staiger, A.M.; Duppel, J.; Dengler, M.A.; van der Kuip, H.; Vohringer, M.C.; Aulitzky, W.E.; Rosenwald, A.; Ott, G.; Horn, H. An analysis of the role of follicular lymphoma-associated fibroblasts to promote tumor cell viability following drug-induced apoptosis. Leuk. Lymphoma 2017, 58, 1922–1930. [Google Scholar] [CrossRef] [PubMed]
  220. Brien, G.; Trescol-Biemont, M.C.; Bonnefoy-Berard, N. Downregulation of Bfl-1 protein expression sensitizes malignant B cells to apoptosis. Oncogene 2007, 26, 5828–5832. [Google Scholar] [CrossRef] [Green Version]
  221. Scarfo, L.; Ghia, P. Reprogramming cell death: BCL2 family inhibition in hematological malignancies. Immunol. Lett. 2013, 155, 36–39. [Google Scholar] [CrossRef]
  222. Zhang, H.; Okamoto, M.; Panzhinskiy, E.; Zawada, W.M.; Das, M. PKCdelta/midkine pathway drives hypoxia-induced proliferation and differentiation of human lung epithelial cells. Am. J. Physiol. Cell Physiol. 2014, 306, C648–C658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Wang, Q.L.; Wang, H.; Zhao, S.L.; Huang, Y.H.; Hou, Y.Y. Over-expressed and truncated midkines promote proliferation of BGC823 cells in vitro and tumor growth in vivo. World J. Gastroenterol. 2008, 14, 1858–1865. [Google Scholar] [CrossRef]
  224. Huang, Y.; Hoque, M.O.; Wu, F.; Trink, B.; Sidransky, D.; Ratovitski, E.A. Midkine induces epithelial-mesenchymal transition through Notch2/Jak2-Stat3 signaling in human keratinocytes. Cell Cycle 2008, 7, 1613–1622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Zhao, G.; Nie, Y.; Lv, M.; He, L.; Wang, T.; Hou, Y. ERbeta-mediated estradiol enhances epithelial mesenchymal transition of lung adenocarcinoma through increasing transcription of midkine. Mol. Endocrinol. 2012, 26, 1304–1315. [Google Scholar] [CrossRef]
  226. Zhang, D.; Ding, L.; Li, Y.; Ren, J.; Shi, G.; Wang, Y.; Zhao, S.; Ni, Y.; Hou, Y. Midkine derived from cancer-associated fibroblasts promotes cisplatin-resistance via up-regulation of the expression of lncRNA ANRIL in tumour cells. Sci. Rep. 2017, 7, 16231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Lorente, M.; Torres, S.; Salazar, M.; Carracedo, A.; Hernandez-Tiedra, S.; Rodriguez-Fornes, F.; Garcia-Taboada, E.; Melendez, B.; Mollejo, M.; Campos-Martin, Y.; et al. Stimulation of the midkine/ALK axis renders glioma cells resistant to cannabinoid antitumoral action. Cell Death Differ. 2011, 18, 959–973. [Google Scholar] [CrossRef] [PubMed]
  228. O’Brien, S.M.; Cunningham, C.C.; Golenkov, A.K.; Turkina, A.G.; Novick, S.C.; Rai, K.R. Phase I to II multicenter study of oblimersen sodium, a Bcl-2 antisense oligonucleotide, in patients with advanced chronic lymphocytic leukemia. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2005, 23, 7697–7702. [Google Scholar] [CrossRef]
  229. O’Brien, S.; Moore, J.O.; Boyd, T.E.; Larratt, L.M.; Skotnicki, A.; Koziner, B.; Chanan-Khan, A.A.; Seymour, J.F.; Bociek, R.G.; Pavletic, S.; et al. Randomized phase III trial of fludarabine plus cyclophosphamide with or without oblimersen sodium (Bcl-2 antisense) in patients with relapsed or refractory chronic lymphocytic leukemia. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2007, 25, 1114–1120. [Google Scholar] [CrossRef]
  230. Soderquist, R.S.; Eastman, A. BCL2 Inhibitors as Anticancer Drugs: A Plethora of Misleading BH3 Mimetics. Mol. Cancer Ther. 2016, 15, 2011–2017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Wang, J.Q.; Li, J.Y.; Teng, Q.X.; Lei, Z.N.; Ji, N.; Cui, Q.; Zeng, L.; Pan, Y.; Yang, D.H.; Chen, Z.S. Venetoclax, a BCL-2 Inhibitor, Enhances the Efficacy of Chemotherapeutic Agents in Wild-Type ABCG2-Overexpression-Mediated MDR Cancer Cells. Cancers 2020, 12, 466. [Google Scholar] [CrossRef] [Green Version]
  232. Fang, H.; Harned, T.M.; Kalous, O.; Maldonado, V.; DeClerck, Y.A.; Reynolds, C.P. Synergistic activity of fenretinide and the Bcl-2 family protein inhibitor ABT-737 against human neuroblastoma. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2011, 17, 7093–7104. [Google Scholar] [CrossRef] [Green Version]
  233. Chen, L.; Zhao, Y.; Halliday, G.C.; Berry, P.; Rousseau, R.F.; Middleton, S.A.; Nichols, G.L.; Del Bello, F.; Piergentili, A.; Newell, D.R.; et al. Structurally diverse MDM2-p53 antagonists act as modulators of MDR-1 function in neuroblastoma. Br. J. Cancer 2014, 111, 716–725. [Google Scholar] [CrossRef] [Green Version]
  234. Feng, F.Y.; Zhang, Y.; Kothari, V.; Evans, J.R.; Jackson, W.C.; Chen, W.; Johnson, S.B.; Luczak, C.; Wang, S.; Hamstra, D.A. MDM2 Inhibition Sensitizes Prostate Cancer Cells to Androgen Ablation and Radiotherapy in a p53-Dependent Manner. Neoplasia 2016, 18, 213–222. [Google Scholar] [CrossRef] [Green Version]
  235. Bai, J.; Wang, R.-H.; Qiao, Y.; Wang, A.; Fang, C.-J. Schiff base derived from thiosemicarbazone and anthracene showed high potential in overcoming multidrug resistance in vitro with low drug resistance index. Drug Des. Devel. Ther. 2017, 11, 2227–2237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Zhang, Y.; Qu, X.J.; Liu, Y.P.; Yang, X.H.; Hou, K.Z.; Teng, Y.E.; Zhang, J.D. Reversal effect of PI3-K inhibitor LY294002 on P-glycoprotein-mediated multidrug resistance of human leukemia cell line K562/DNR and gastric cancer cell line SGC7901/ADR. Ai Zheng Aizheng Chin. J. Cancer 2009, 28, 97–99. [Google Scholar]
  237. Xue, C.; Wang, C.; Sun, Y.; Meng, Q.; Liu, Z.; Huo, X.; Sun, P.; Sun, H.; Ma, X.; Ma, X.; et al. Targeting P-glycoprotein function, p53 and energy metabolism: Combination of metformin and 2-deoxyglucose reverses the multidrug resistance of MCF-7/Dox cells to doxorubicin. Oncotarget 2017, 8, 8622–8632. [Google Scholar] [CrossRef] [Green Version]
  238. Durrant, D.E.; Das, A.; Dyer, S.; Kukreja, R.C. A dual PI3 kinase/mTOR inhibitor BEZ235 reverses doxorubicin resistance in ABCB1 overexpressing ovarian and pancreatic cancer cell lines. Biochim. Biophys. Acta Gen. Subj. 2020, 1864, 129556. [Google Scholar] [CrossRef]
  239. Caumanns, J.J.; Berns, K.; Wisman, G.B.A.; Fehrmann, R.S.N.; Tomar, T.; Klip, H.; Meersma, G.J.; Hijmans, E.M.; Gennissen, A.M.C.; Duiker, E.W.; et al. Integrative Kinome Profiling Identifies mTORC1/2 Inhibition as Treatment Strategy in Ovarian Clear Cell Carcinoma. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2018, 24, 3928–3940. [Google Scholar] [CrossRef] [Green Version]
  240. Ghayad, S.E.; Bieche, I.; Vendrell, J.A.; Keime, C.; Lidereau, R.; Dumontet, C.; Cohen, P.A. mTOR inhibition reverses acquired endocrine therapy resistance of breast cancer cells at the cell proliferation and gene-expression levels. Cancer Sci. 2008, 99, 1992–2003. [Google Scholar] [CrossRef]
  241. Voges, Y.; Michaelis, M.; Rothweiler, F.; Schaller, T.; Schneider, C.; Politt, K.; Mernberger, M.; Nist, A.; Stiewe, T.; Wass, M.N.; et al. Effects of YM155 on survivin levels and viability in neuroblastoma cells with acquired drug resistance. Cell Death Dis. 2016, 7, e2410. [Google Scholar] [CrossRef]
  242. Yang, L.; Wang, Q.; Li, D.; Zhou, Y.; Zheng, X.; Sun, H.; Yan, J.; Zhang, L.; Lin, Y.; Wang, X. Wogonin enhances antitumor activity of tumor necrosis factor-related apoptosis-inducing ligand in vivo through ROS-mediated downregulation of cFLIPL and IAP proteins. Apoptosis 2013, 18, 618–626. [Google Scholar] [CrossRef] [PubMed]
  243. Rao, P.S.; Satelli, A.; Moridani, M.; Jenkins, M.; Rao, U.S. Luteolin induces apoptosis in multidrug resistant cancer cells without affecting the drug transporter function: Involvement of cell line-specific apoptotic mechanisms. Int. J. Cancer 2012, 130, 2703–2714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Jeng, L.B.; Kumar Velmurugan, B.; Chen, M.C.; Hsu, H.H.; Ho, T.J.; Day, C.H.; Lin, Y.M.; Padma, V.V.; Tu, C.C.; Huang, C.Y. Fisetin mediated apoptotic cell death in parental and Oxaliplatin/irinotecan resistant colorectal cancer cells in vitro and in vivo. J. Cell. Physiol. 2018, 233, 7134–7142. [Google Scholar] [CrossRef]
  245. Li, Y.; Ahmed, F.; Ali, S.; Philip, P.A.; Kucuk, O.; Sarkar, F.H. Inactivation of nuclear factor kappaB by soy isoflavone genistein contributes to increased apoptosis induced by chemotherapeutic agents in human cancer cells. Cancer Res. 2005, 65, 6934–6942. [Google Scholar] [CrossRef] [Green Version]
  246. Yang, S.; Li, W.; Sun, H.; Wu, B.; Ji, F.; Sun, T.; Chang, H.; Shen, P.; Wang, Y.; Zhou, D. Resveratrol elicits anti-colorectal cancer effect by activating miR-34c-KITLG in vitro and in vivo. BMC Cancer 2015, 15, 969. [Google Scholar] [CrossRef] [Green Version]
  247. Zhao, W.; Bao, P.; Qi, H.; You, H. Resveratrol down-regulates survivin and induces apoptosis in human multidrug-resistant SPC-A-1/CDDP cells. Oncol. Rep. 2010, 23, 279–286. [Google Scholar] [CrossRef] [Green Version]
  248. Kunnumakkara, A.B.; Diagaradjane, P.; Anand, P.; Harikumar, K.B.; Deorukhkar, A.; Gelovani, J.; Guha, S.; Krishnan, S.; Aggarwal, B.B. Curcumin sensitizes human colorectal cancer to capecitabine by modulation of cyclin D1, COX-2, MMP-9, VEGF and CXCR4 expression in an orthotopic mouse model. Int. J. Cancer 2009, 125, 2187–2197. [Google Scholar] [CrossRef]
  249. Roy, S.; Yu, Y.; Padhye, S.B.; Sarkar, F.H.; Majumdar, A.P. Difluorinated-curcumin (CDF) restores PTEN expression in colon cancer cells by down-regulating miR-21. PLoS ONE 2013, 8, e68543. [Google Scholar] [CrossRef]
  250. Wang, S.; Chen, R.; Zhong, Z.; Shi, Z.; Chen, M.; Wang, Y. Epigallocatechin-3-gallate potentiates the effect of curcumin in inducing growth inhibition and apoptosis of resistant breast cancer cells. Am. J. Chin. Med. 2014, 42, 1279–1300. [Google Scholar] [CrossRef]
  251. Kao, T.Y.; Chung, Y.C.; Hou, Y.C.; Tsai, Y.W.; Chen, C.H.; Chang, H.P.; Chou, J.L.; Hsu, C.P. Effects of ellagic acid on chemosensitivity to 5-fluorouracil in colorectal carcinoma cells. Anticancer Res. 2012, 32, 4413–4418. [Google Scholar]
  252. Wang, H.; Jia, X.H.; Chen, J.R.; Wang, J.Y.; Li, Y.J. Osthole shows the potential to overcome P-glycoproteinmediated multidrug resistance in human myelogenous leukemia K562/ADM cells by inhibiting the PI3K/Akt signaling pathway. Oncol. Rep. 2016, 35, 3659–3668. [Google Scholar] [CrossRef] [Green Version]
  253. Neophytou, C.M.; Constantinou, C.; Papageorgis, P.; Constantinou, A.I. D-alpha-tocopheryl polyethylene glycol succinate (TPGS) induces cell cycle arrest and apoptosis selectively in Survivin-overexpressing breast cancer cells. Biochem. Pharm. 2014, 89, 31–42. [Google Scholar] [CrossRef]
  254. Mediavilla-Varela, M.; Boateng, K.; Noyes, D.; Antonia, S.J. The anti-fibrotic agent pirfenidone synergizes with cisplatin in killing tumor cells and cancer-associated fibroblasts. BMC Cancer 2016, 16, 176. [Google Scholar] [CrossRef] [Green Version]
  255. Wei, S.C.; Anang, N.A.S.; Sharma, R.; Andrews, M.C.; Reuben, A.; Levine, J.H.; Cogdill, A.P.; Mancuso, J.J.; Wargo, J.A.; Pe’er, D.; et al. Combination anti-CTLA-4 plus anti-PD-1 checkpoint blockade utilizes cellular mechanisms partially distinct from monotherapies. Proc. Natl. Acad. Sci. USA 2019, 116, 22699–22709. [Google Scholar] [CrossRef] [Green Version]
  256. Chen, N.; Fang, W.; Zhan, J.; Hong, S.; Tang, Y.; Kang, S.; Zhang, Y.; He, X.; Zhou, T.; Qin, T.; et al. Upregulation of PD-L1 by EGFR Activation Mediates the Immune Escape in EGFR-Driven NSCLC: Implication for Optional Immune Targeted Therapy for NSCLC Patients with EGFR Mutation. J. Thorac. Oncol. Off. Publ. Int. Assoc. Study Lung Cancer 2015, 10, 910–923. [Google Scholar] [CrossRef] [Green Version]
  257. Hay, C.M.; Sult, E.; Huang, Q.; Mulgrew, K.; Fuhrmann, S.R.; McGlinchey, K.A.; Hammond, S.A.; Rothstein, R.; Rios-Doria, J.; Poon, E.; et al. Targeting CD73 in the tumor microenvironment with MEDI9447. Oncoimmunology 2016, 5, e1208875. [Google Scholar] [CrossRef]
  258. Soto, H.; Sanchez, K.; Escobar, J.Y.; Constanzo, A.; Fernandez, Z.; Melendez, C. Cost-Effectiveness Analysis of Hydralazine and Magnesium Valproate LP Associated with Treatment for Adult Patients with Metastatic Recurrent or Persistent Cervical Cancer in Mexico. Value Health J. Int. Soc. Pharm. Outcomes Res. 2014, 17, A639. [Google Scholar] [CrossRef]
  259. Dawood, M.; Ooko, E.; Efferth, T. Collateral Sensitivity of Parthenolide via NF-kappaB and HIF-alpha Inhibition and Epigenetic Changes in Drug-Resistant Cancer Cell Lines. Front. Pharm. 2019, 10, 542. [Google Scholar] [CrossRef] [Green Version]
  260. Xia, C.; Leon-Ferre, R.; Laux, D.; Deutsch, J.; Smith, B.J.; Frees, M.; Milhem, M. Treatment of resistant metastatic melanoma using sequential epigenetic therapy (decitabine and panobinostat) combined with chemotherapy (temozolomide). Cancer Chemother. Pharmacol. 2014, 74, 691–697. [Google Scholar] [CrossRef] [Green Version]
  261. Falchook, G.S.; Fu, S.; Naing, A.; Hong, D.S.; Hu, W.; Moulder, S.; Wheler, J.J.; Sood, A.K.; Bustinza-Linares, E.; Parkhurst, K.L.; et al. Methylation and histone deacetylase inhibition in combination with platinum treatment in patients with advanced malignancies. Invest. New Drugs 2013, 31, 1192–1200. [Google Scholar] [CrossRef] [Green Version]
  262. Sun, C.; Yin, J.; Fang, Y.; Chen, J.; Jeong, K.J.; Chen, X.; Vellano, C.P.; Ju, Z.; Zhao, W.; Zhang, D.; et al. BRD4 Inhibition Is Synthetic Lethal with PARP Inhibitors through the Induction of Homologous Recombination Deficiency. Cancer Cell 2018, 33, 401–416.e408. [Google Scholar] [CrossRef] [Green Version]
  263. Shariati, M.; Meric-Bernstam, F. Targeting AKT for cancer therapy. Expert Opin. Investig. Drugs 2019, 28, 977–988. [Google Scholar] [CrossRef]
  264. Nitulescu, G.M.; Margina, D.; Juzenas, P.; Peng, Q.; Olaru, O.T.; Saloustros, E.; Fenga, C.; Spandidos, D.; Libra, M.; Tsatsakis, A.M. Akt inhibitors in cancer treatment: The long journey from drug discovery to clinical use (Review). Int. J. Oncol. 2016, 48, 869–885. [Google Scholar] [CrossRef] [Green Version]
  265. Yang, H.; Villani, R.M.; Wang, H.; Simpson, M.J.; Roberts, M.S.; Tang, M.; Liang, X. The role of cellular reactive oxygen species in cancer chemotherapy. J. Exp. Clin. Cancer Res. 2018, 37, 266. [Google Scholar] [CrossRef] [PubMed]
  266. Minassian, L.M.; Cotechini, T.; Huitema, E.; Graham, C.H. Hypoxia-Induced Resistance to Chemotherapy in Cancer. Adv. Exp. Med. Biol. 2019, 1136, 123–139. [Google Scholar] [CrossRef] [PubMed]
  267. Li, L.; Xu, J.; Min, T.; Huang, W. Up-regulation of P-glycoprotein expression by catalase via JNK activation in HepG2 cells. Redox Rep. Commun. Free Radic. Res. 2006, 11, 173–178. [Google Scholar] [CrossRef]
  268. Kotecha, R.; Takami, A.; Espinoza, J.L. Dietary phytochemicals and cancer chemoprevention: A review of the clinical evidence. Oncotarget 2016, 7, 52517–52529. [Google Scholar] [CrossRef] [Green Version]
  269. Neophytou, C.M.; Constantinou, G.Y. Pro-apoptotic properties of chemopreventive agents. In Natural Products for Cancer Chemoprevention: Single Compounds and Combinations; Pezzuto, J.M., Vang, O., Eds.; Springer International Publishing: Berlin/Heidelberg, Germany, 2020; pp. 517–559. [Google Scholar]
  270. Hu-Lieskovan, S.; Malouf, G.G.; Jacobs, I.; Chou, J.; Liu, L.; Johnson, M.L. Addressing resistance to immune checkpoint inhibitor therapy: An urgent unmet need. Future Oncol. 2021, 17, 1401–1439. [Google Scholar] [CrossRef]
  271. Wilting, R.H.; Dannenberg, J.H. Epigenetic mechanisms in tumorigenesis, tumor cell heterogeneity and drug resistance. Drug Resist. Updates Rev. Comment. Antimicrob. Anticancer Chemother. 2012, 15, 21–38. [Google Scholar] [CrossRef] [Green Version]
  272. Talukdar, S.; Bhoopathi, P.; Emdad, L.; Das, S.; Sarkar, D.; Fisher, P.B. Dormancy and cancer stem cells: An enigma for cancer therapeutic targeting. Adv. Cancer Res. 2019, 141, 43–84. [Google Scholar] [CrossRef]
  273. Shipitsin, M.; Campbell, L.L.; Argani, P.; Weremowicz, S.; Bloushtain-Qimron, N.; Yao, J.; Nikolskaya, T.; Serebryiskaya, T.; Beroukhim, R.; Hu, M.; et al. Molecular definition of breast tumor heterogeneity. Cancer Cell 2007, 11, 259–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Issa, M.E.; Takhsha, F.S.; Chirumamilla, C.S.; Perez-Novo, C.; Vanden Berghe, W.; Cuendet, M. Epigenetic strategies to reverse drug resistance in heterogeneous multiple myeloma. Clin. Epigenet. 2017, 9, 17. [Google Scholar] [CrossRef] [Green Version]
  275. Lu, Y.; Chan, Y.T.; Tan, H.Y.; Li, S.; Wang, N.; Feng, Y. Epigenetic regulation in human cancer: The potential role of epi-drug in cancer therapy. Mol. Cancer 2020, 19, 79. [Google Scholar] [CrossRef] [PubMed]
  276. Ganesan, A.; Arimondo, P.B.; Rots, M.G.; Jeronimo, C.; Berdasco, M. The timeline of epigenetic drug discovery: From reality to dreams. Clin. Epigenet. 2019, 11, 174. [Google Scholar] [CrossRef] [Green Version]
  277. Guo, M.; Peng, Y.; Gao, A.; Du, C.; Herman, J.G. Epigenetic heterogeneity in cancer. Biomark. Res. 2019, 7, 23. [Google Scholar] [CrossRef]
  278. Dawson, M.A. The cancer epigenome: Concepts, challenges, and therapeutic opportunities. Science 2017, 355, 1147–1152. [Google Scholar] [CrossRef]
  279. Mohammad, H.P.; Barbash, O.; Creasy, C.L. Targeting epigenetic modifications in cancer therapy: Erasing the roadmap to cancer. Nat. Med. 2019, 25, 403–418. [Google Scholar] [CrossRef]
  280. Guo, L.; Lee, Y.T.; Zhou, Y.; Huang, Y. Targeting epigenetic regulatory machinery to overcome cancer therapy resistance. Semin. Cancer Biol. 2021. [Google Scholar] [CrossRef]
  281. Brown, R.; Curry, E.; Magnani, L.; Wilhelm-Benartzi, C.S.; Borley, J. Poised epigenetic states and acquired drug resistance in cancer. Nat. Rev. Cancer 2014, 14, 747–753. [Google Scholar] [CrossRef]
  282. Bukowski, K.; Kciuk, M.; Kontek, R. Mechanisms of Multidrug Resistance in Cancer Chemotherapy. Int. J. Mol. Sci. 2020, 21, 3233. [Google Scholar] [CrossRef] [PubMed]
  283. Juergens, R.A.; Wrangle, J.; Vendetti, F.P.; Murphy, S.C.; Zhao, M.; Coleman, B.; Sebree, R.; Rodgers, K.; Hooker, C.M.; Franco, N.; et al. Combination epigenetic therapy has efficacy in patients with refractory advanced non-small cell lung cancer. Cancer Discov. 2011, 1, 598–607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Fuller, M.; Klein, M.; Schmidt, E.; Rohde, C.; Gollner, S.; Schulze, I.; Qianli, J.; Berdel, W.E.; Edemir, B.; Muller-Tidow, C.; et al. 5-azacytidine enhances efficacy of multiple chemotherapy drugs in AML and lung cancer with modulation of CpG methylation. Int. J. Oncol. 2015, 46, 1192–1204. [Google Scholar] [CrossRef] [Green Version]
  285. Hammerlindl, H.; Schaider, H. Tumor cell-intrinsic phenotypic plasticity facilitates adaptive cellular reprogramming driving acquired drug resistance. J. Cell Commun. Signal. 2018, 12, 133–141. [Google Scholar] [CrossRef] [Green Version]
  286. Guha, M. HDAC inhibitors still need a home run, despite recent approval. Nat. Rev. Drug Discov. 2015, 14, 225–226. [Google Scholar] [CrossRef]
  287. Imamura, Y.; Mukohara, T.; Shimono, Y.; Funakoshi, Y.; Chayahara, N.; Toyoda, M.; Kiyota, N.; Takao, S.; Kono, S.; Nakatsura, T.; et al. Comparison of 2D- and 3D-culture models as drug-testing platforms in breast cancer. Oncol. Rep. 2015, 33, 1837–1843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  288. Gengenbacher, N.; Singhal, M.; Augustin, H.G. Preclinical mouse solid tumour models: Status quo, challenges and perspectives. Nat. Rev. Cancer 2017, 17, 751–765. [Google Scholar] [CrossRef]
  289. Ramirez, M.; Rajaram, S.; Steininger, R.J.; Osipchuk, D.; Roth, M.A.; Morinishi, L.S.; Evans, L.; Ji, W.; Hsu, C.H.; Thurley, K.; et al. Diverse drug-resistance mechanisms can emerge from drug-tolerant cancer persister cells. Nat. Commun. 2016, 7, 10690. [Google Scholar] [CrossRef]
  290. Boumahdi, S.; de Sauvage, F.J. The great escape: Tumour cell plasticity in resistance to targeted therapy. Nat. Rev. Drug Discov. 2020, 19, 39–56. [Google Scholar] [CrossRef]
  291. Morel, D.; Jeffery, D.; Aspeslagh, S.; Almouzni, G.; Postel-Vinay, S. Combining epigenetic drugs with other therapies for solid tumours—Past lessons and future promise. Nat. Rev. Clin. Oncol. 2020, 17, 91–107. [Google Scholar] [CrossRef]
Figure 1. The extrinsic and intrinsic pathways of apoptosis. The activation of the extrinsic pathway involves binding of an external ligand to a transmembrane Death Receptor which then induces its conformation as a homotrimer. In the internal part of the receptor, the exposed Death Domain (DD) recruit adaptor proteins such as TRADD. Through its Death Effector Domain (DED), TRADD can then recruit pro-Caspase-8 which is activated through self-proteolysis. Subsequently the death ligand, receptor, TRADD molecule and caspase-8 form the DISC complex. Active caspase-8 can cleave downstream effector caspases 3, 6 and 7 which degrade nuclear lamins and other cellular components. Caspase-8 connects the extrinsic pathway with intrinsic or mitochondrial apoptotic signaling, though the cleavage of Bid to truncated Bid (t-Bid). Bid, a member of the Bcl-2 family, facilitates the opening of mitochondrial pores in a process called MOMP by inducing the polymerization of Bax on the outer mitochondrial membrane. MOMP is also induced by internal signals, such as extensive DNA damage. Cytochrome-c (cyt-c) is released from mitochondria and along with Apaf-1 and pro-caspase-9 form the apoptosome heptamer structure. Following proteolysis, caspase-9 cleaves and activates effector caspases, further amplifying the apoptotic process. Apoptotic Protease Activating Factor 1 (APAF-1); Mitochondrial Outer Membrane Permeabilization (MOMP); TNFR1-Associated Death Domain protein, (TRADD).
Figure 1. The extrinsic and intrinsic pathways of apoptosis. The activation of the extrinsic pathway involves binding of an external ligand to a transmembrane Death Receptor which then induces its conformation as a homotrimer. In the internal part of the receptor, the exposed Death Domain (DD) recruit adaptor proteins such as TRADD. Through its Death Effector Domain (DED), TRADD can then recruit pro-Caspase-8 which is activated through self-proteolysis. Subsequently the death ligand, receptor, TRADD molecule and caspase-8 form the DISC complex. Active caspase-8 can cleave downstream effector caspases 3, 6 and 7 which degrade nuclear lamins and other cellular components. Caspase-8 connects the extrinsic pathway with intrinsic or mitochondrial apoptotic signaling, though the cleavage of Bid to truncated Bid (t-Bid). Bid, a member of the Bcl-2 family, facilitates the opening of mitochondrial pores in a process called MOMP by inducing the polymerization of Bax on the outer mitochondrial membrane. MOMP is also induced by internal signals, such as extensive DNA damage. Cytochrome-c (cyt-c) is released from mitochondria and along with Apaf-1 and pro-caspase-9 form the apoptosome heptamer structure. Following proteolysis, caspase-9 cleaves and activates effector caspases, further amplifying the apoptotic process. Apoptotic Protease Activating Factor 1 (APAF-1); Mitochondrial Outer Membrane Permeabilization (MOMP); TNFR1-Associated Death Domain protein, (TRADD).
Cancers 13 04363 g001
Figure 2. Deregulation of apoptotic pathways during the MDR development in cancer. The PI3K/AKT survival pathway is frequently overactive in MDR cancers, partly due to increased levels of growth ligands and receptors. Binding of growth factors to receptor tyrosine kinases (RTKs) stimulates PI3K by autophosphorylation; PI3K then catalyzes the conversion of PIP2 to PIP3 while tumor suppressor PTEN has an opposing function. PTEN is often mutated and thus inactive in MDR cancers. PI3K mediates the activation of AKT via phosphorylation on Thr308 and Ser473 by PDK1 and PDK2, respectively. AKT can then activate MDM2 which blocks the function of p53. P53 regulates (among others) the levels and activation status of the Bcl-2 family of proteins. The balance between the levels of the pro-apoptotic and anti-apoptotic Bcl-2 family proteins, controls the release of pro-apoptotic factors from mitochondria. Once released, cytochrome-c activates firstly caspase-9 and then the executioner caspases-3, -6 and -7. Furthermore, in MDR cancers, members of the IAP family are overexpressed, blocking caspase function. SMAC/Diablo are also released from the mitochondria during apoptosis and can inhibit the function of IAPs. NFkB is indirectly activated by growth factors via the PI3K/AKT pathway and up-regulates IAPs, e.g., survivin via regulation at the transcriptional level. Factors released by CAFs increase tumor cells survival via the activation of the PI3K/AKT pathway and inhibit apoptotic pathways. The cellular and non-cellular components of the TME as well as modifications in metabolic pathways and mechanical stress have been also implicated in resistance to cancer-cells targeting pro-apoptotic therapeutic agents. The composition and structure of stromal components in tumors increase interstitial fluid pressure (IFP) hindering the penetration of macromolecules through tissue and influence the sensitivity of tumor cells to therapy [141]. The hypoxic TME favours cells that have lost sensitivity to p53-mediated apoptosis and that are deficient in DNA mismatch repair leading to resistance to platinum-based chemotherapeutic agents [142]. Moreover, the low extracellular pH in tumors decreases the cellular uptake of weakly basic drugs such as doxorubicin, mitoxantrone, vincristine and vinblastine [143]. Multi Drug Resistant (MDR); Receptor Tyrosine Kinases (RTKs); Phosphatidylinositol 4,5-bisphosphate PtdIns(4,5)P2, (PIP2); Phosphatidylinositol (3,4,5)-triphosphate PtdIns(3,4,5)P3, (PIP3); Phosphoinositide-dependent kinase, (PDK); Tumor Microenvironment (TME); Inhibitors of Apoptosis Proteins, (IAPs); Cancer Associated Fibroblasts, (CAFs).
Figure 2. Deregulation of apoptotic pathways during the MDR development in cancer. The PI3K/AKT survival pathway is frequently overactive in MDR cancers, partly due to increased levels of growth ligands and receptors. Binding of growth factors to receptor tyrosine kinases (RTKs) stimulates PI3K by autophosphorylation; PI3K then catalyzes the conversion of PIP2 to PIP3 while tumor suppressor PTEN has an opposing function. PTEN is often mutated and thus inactive in MDR cancers. PI3K mediates the activation of AKT via phosphorylation on Thr308 and Ser473 by PDK1 and PDK2, respectively. AKT can then activate MDM2 which blocks the function of p53. P53 regulates (among others) the levels and activation status of the Bcl-2 family of proteins. The balance between the levels of the pro-apoptotic and anti-apoptotic Bcl-2 family proteins, controls the release of pro-apoptotic factors from mitochondria. Once released, cytochrome-c activates firstly caspase-9 and then the executioner caspases-3, -6 and -7. Furthermore, in MDR cancers, members of the IAP family are overexpressed, blocking caspase function. SMAC/Diablo are also released from the mitochondria during apoptosis and can inhibit the function of IAPs. NFkB is indirectly activated by growth factors via the PI3K/AKT pathway and up-regulates IAPs, e.g., survivin via regulation at the transcriptional level. Factors released by CAFs increase tumor cells survival via the activation of the PI3K/AKT pathway and inhibit apoptotic pathways. The cellular and non-cellular components of the TME as well as modifications in metabolic pathways and mechanical stress have been also implicated in resistance to cancer-cells targeting pro-apoptotic therapeutic agents. The composition and structure of stromal components in tumors increase interstitial fluid pressure (IFP) hindering the penetration of macromolecules through tissue and influence the sensitivity of tumor cells to therapy [141]. The hypoxic TME favours cells that have lost sensitivity to p53-mediated apoptosis and that are deficient in DNA mismatch repair leading to resistance to platinum-based chemotherapeutic agents [142]. Moreover, the low extracellular pH in tumors decreases the cellular uptake of weakly basic drugs such as doxorubicin, mitoxantrone, vincristine and vinblastine [143]. Multi Drug Resistant (MDR); Receptor Tyrosine Kinases (RTKs); Phosphatidylinositol 4,5-bisphosphate PtdIns(4,5)P2, (PIP2); Phosphatidylinositol (3,4,5)-triphosphate PtdIns(3,4,5)P3, (PIP3); Phosphoinositide-dependent kinase, (PDK); Tumor Microenvironment (TME); Inhibitors of Apoptosis Proteins, (IAPs); Cancer Associated Fibroblasts, (CAFs).
Cancers 13 04363 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Neophytou, C.M.; Trougakos, I.P.; Erin, N.; Papageorgis, P. Apoptosis Deregulation and the Development of Cancer Multi-Drug Resistance. Cancers 2021, 13, 4363. https://doi.org/10.3390/cancers13174363

AMA Style

Neophytou CM, Trougakos IP, Erin N, Papageorgis P. Apoptosis Deregulation and the Development of Cancer Multi-Drug Resistance. Cancers. 2021; 13(17):4363. https://doi.org/10.3390/cancers13174363

Chicago/Turabian Style

Neophytou, Christiana M., Ioannis P. Trougakos, Nuray Erin, and Panagiotis Papageorgis. 2021. "Apoptosis Deregulation and the Development of Cancer Multi-Drug Resistance" Cancers 13, no. 17: 4363. https://doi.org/10.3390/cancers13174363

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop