Skip to main content
Translational Oncology logoLink to Translational Oncology
. 2021 Mar 6;14(5):101056. doi: 10.1016/j.tranon.2021.101056

Chemotherapeutic drugs: Cell death- and resistance-related signaling pathways. Are they really as smart as the tumor cells?

Mojtaba Mollaei a,, Zuhair Mohammad Hassan a, Fatemeh Khorshidi a,b, Ladan Langroudi c
PMCID: PMC7938256  PMID: 33684837

Highlight

  • Chemotherapeutic drugs are classified into alkylating substrates, antimetabolite agents, anti-tumor antibiotics, inhibitors of topoisomerase I and II, mitotic inhibitors, and corticosteroids.

  • Most of the chemotherapeutic drugs act in a dose-dependent manner; they either stimulate apoptosis or chemoresistance.

  • Two major complications during chemotherapy are chemoresistance and tumor relapse. We have discussed underlying intracellular signaling cascades responsible for these events.

  • Combination chemotherapies, targeted therapies, and immunotherapies have represented more promising responses than single chemotherapy applications.

  • The role of omics, next-generation sequencing (NGS), whole-exome sequencing (WES), and machine learning technologies in precision medicine are currently under investigations for overcoming chemoresistance and tumor relapse.

Key words: Chemotherapeutic drugs, Intracellular signaling, Chemoresistance, Death-related intracellular signaling, Resistance-related intracellular signaling, Combination chemotherapy, Precision medicine

Abstract

Chemotherapeutic drugs kill cancer cells or control their progression all over the patient's body, while radiation- and surgery-based treatments perform in a particular site. Based on their mechanisms of action, they are classified into different groups, including alkylating substrates, antimetabolite agents, anti-tumor antibiotics, inhibitors of topoisomerase I and II, mitotic inhibitors, and finally, corticosteroids. Although chemotherapeutic drugs have brought about more life expectancy, two major and severe complications during chemotherapy are chemoresistance and tumor relapse. Therefore, we aimed to review the underlying intracellular signaling pathways involved in cell death and resistance in different chemotherapeutic drug families to clarify the shortcomings in the conventional single chemotherapy applications. Moreover, we have summarized the current combination chemotherapy applications, including numerous combined-, and encapsulated-combined-chemotherapeutic drugs. We further discussed the possibilities and applications of precision medicine, machine learning, next-generation sequencing (NGS), and whole-exome sequencing (WES) in promoting cancer immunotherapies. Finally, some of the recent clinical trials concerning the application of immunotherapies and combination chemotherapies were included as well, in order to provide a practical perspective toward the future of therapies in cancer cases.

Graphical abstract

Image, graphical abstract

Introduction

Cancer is characterized by the uncontrolled cell proliferation, invasion, and check-point evasion of abnormal cells that are mostly nonfunctional. Cancer can arise due to diet insufficiencies, inherited mutations, and tobacco consumption, to say the least [1, 2]. Cancer's incident is increasing due to the sedentary lifestyle, overpopulated, polluted megacities, and individuals’ growing desire for consuming processed foods containing preservatives additives [3], [4], [5]. Since cancers might not manifest symptoms in their early onset, it would be difficult or even improbable to treat them when they are diagnosed in their late stage. By and large, tumors are composed of two main parts, including the proliferating cells and stroma, which contains connective tissue and blood supply [6]. Chemotherapy has been among our best options against malignancies.

Chemotherapy is defined by the administration of numerous drugs and chemicals either alone or in combination to kill the cancer cells. Chemotherapeutic drugs kill cancer cells or control their progression all over the patient's body, while radiation- and surgery-based treatments are directed to a particular site. Cure, control, and palliation are the three objectives of chemotherapies. Killing cancer cells by implementing chemotherapy drugs means “Cure”, whereas “Control” defines the situation that full remission seems far-fetched; therefore, the objective of the therapy would be to decrease the tumor size or to diminish the growth rate and angiogenesis. Palliation aims to alleviate the pain, symptoms, and medical conditions arisen due to cancer. It is mostly accomplished when cancer is in the advanced stages and cannot be eradicated; therefore, our aim would be to improve the quality of life.

The chemotherapy prescription approaches rely on various elements, including the cancer's type, the cancer's stage, the patient's age, the patient's general health status, the other concurrent health issues, and the history of receiving chemotherapies. Since chemotherapeutic drugs cannot distinguish normal cells against cancerous cells, the prescribed dosage is the other crucial aspect toward achieving the best possible response. The administration dosage depends on the patient's weight, body surface area, age, nutrition status, history of radiation therapy, and blood cell count. Besides, a suitable drug administration schedule might help obtain the most efficient anti-cancer activity and minimum side effects [7, 8].

Chemotherapeutic agents target different stages of the cell cycle. They are classified into different groups based on their mechanisms of action, including alkylating substrates that damage the DNA and interfering with cell reproduction in various cell cycle phases. These elements leave impacts on both bone marrow stem cells and somatic cells, making them a preferred option for solid tumors and leukemia. On the other hand, antimetabolite agents replace the typical structures of RNA and DNA and are known as RNA and DNA blockers. The so-called process blocks the multiplying of chromosomes that makes these drugs suitable against leukemia, breast, and ovary cancers. While, anti-tumor antibiotics, unlike traditional antibiotics, hinder the activity of DNA duplication-related enzymes. These are known as Anthracyclines, which have been prescribed for numerous malignancies. However, since permanent heart failure is one of their serious side effects, non-Anthracyclines have been developed to mitigate these side effects. Moreover, topoisomerase I and II Inhibitors impede the activity of topoisomerases, therefore hindering the replication of DNA. Nevertheless, topoisomerase II inhibitors enhance the risk factors for secondary cancers, such as acute myeloid leukemia (AML). Besides, mitotic inhibitors are compounds driven from plants that interfere with different phases cell cycle phases resulting in cell proliferation blockage; nonetheless incidental damages to the nerve system are among their disadvantages. Finally, corticosteroids have been prescribed to palliate the chemotherapy drug's side effects, such as nausea, vomiting, and allergic reactions [7, 9] Table 1.

Table 1.

The category of chemotherapy drugs.

Category Drugs Mechanisms of action
Alkylating agents Altertamine Damage the DNA
Busulfan
Carboplatin
Carmustine
Chlorambucil
Cisplatin
Cyclophosphamide
Dacarbazine
Lomustine
Melphalan
Oxaliplatin
Temozolomide
Thiotepa
Antimetabolite 5-fluorouracil (5-FU) Substitute the RNA and DNA blocks
6-mercaptopurine (6-MP)
Capecitabine (Xeloda)
Cytarabine (Ara-C)
Floxuridine
Fludarabine
Gemcitabine (Gemzar)
Hydroxyurea
Methotrexate
Pemetrexed (Alimta)
Anti-tumor Antibiotics Anthracyclines Epirubicin Interfere with the activity of DNA replication enzymes
Idarubicin
Daunorubicin
Doxorubicin (Adriamycin)
non-Anthracyclines Actinomycin-D
Bleomycin
Mitomycin-C
Mitoxantrone
Topoisomerase inhibitors Topoisomerase inhibitor I Topotecan Interfere with the topoisomerase enzymes and incorporate the unwinding DNA in replication and transcription
Irinotecan (CPT-11)
Topoisomerase inhibitor II Etoposide (VP-16)
Teniposide
Mitoxantrone
Mitotic inhibitors Docetaxel Hinder the cell proliferation and division
Estramustine
Ixabepilone
Paclitaxel
Vinblastine
Vincristine
Vinorelbine
Corticosteroids Prednisone Palliate the chemotherapy side effects
Methylprednisone (Solumedrol)
Dexamethasone (Decadron)
EGFR inhibitors Tarceva (Erlotinib) Blocks the epidermal growth factor receptors on tumor cells
Erbitux (Cetuximab)
Iressa (Gefitinib)

Having said that, two significant and severe complications during chemotherapy are chemoresistance and tumor relapse. Resistance is either intrinsic or extrinsic (acquired); in one way, cancer cells become resistant to different chemotherapeutic drugs through common resistance mechanisms, which is called multi-drug resistance. In other ways, cancer cells develop resistance through the increased drug efflux, dampened apoptosis, enhanced drug detoxification, altered expression in the drug target, and improved DNA repair mechanisms [10].

In this regard, we have selected representatives in each chemotherapeutic drug family to discuss their molecular and cellular mechanisms toward cell death and resistance. Besides, we have summarized the current combination chemotherapy applications, including numerous combined-, and encapsulated-combined-chemotherapeutic drugs. We further discussed the possibilities and applications of precision medicine, machine learning, next-generation sequencing (NGS), and whole-exome sequencing (WES) in promoting cancer immunotherapies. Finally, some of the recent clinical trials concerning the application of immunotherapies and combination chemotherapies were included as well, in order to provide a practical perspective toward the future of therapies in cancer cases.

Current chemotherapeutic drugs and their mechanisms of action

Cisplatin

Cisplatin mechanism of action

Cis-diamminedichloroplatinum, known as cisplatin, is one of the most beneficial chemotherapeutic drugs which is prescribed for half of the malignancies [11]. Cisplatin is administered intravenously, where it binds to plasma proteins like albumin and is conveyed through the bloodstream, and ultimately enters the target cells via copper transporters [12, 13]. Interestingly, just about 1 percent of the administered drug impacts DNA [14]. Cisplatin contains platinum-based combinations that construct inter-, and intra-strand adducts within the DNA strands between guanine and adenine located in the opposite strands [15], resulting in cell cycle arrest and ignition of apoptosis [16, 17]. To illustrate this, cisplatin forms these adducts in a two-steps manner. Initially, it constructs a bond with N7 guanine; after that, it forms the second link with guanine or adenine in the opposite or the same strand. Since N7 atoms in adenine and guanine are easily accessible, cisplatin forms numerous adducts and cross-links within the DNA structure, leading to a general distortion in the DNA framework [15].

Mechanisms of cisplatin in cell death through signaling cascades

As cisplatin binds to DNA, it causes a bend within the DNA structure. Various proteins, such as the high mobility group (HMG), which are a kind of non-histone- and chromosome-structural protein, identify this bend and selectively bind to cisplatin adducts in DNA [18]. HMG proteins cover the affected DNA from being recognized by nucleotide excision repair (NER) and mismatch repair (MMR) [19]; therefore, p53, the genome security guard, is activated, which increases the activity of waf1, p21, and MDM2, leading to cell cycle arrest [20]. More on that, cisplatin triggers the oxidative stress within the cell. Reactive oxygen species (ROS) cripple DNA, proteins, and cell lipids [21]. Providing that the damage is severe or irreparable, the intrinsic pathway of apoptosis is ignited through BCL2 inhibition and caspase activation [22, 23]. The inhibition of BCL2 impacts the mitochondrial membrane integrity as well. Besides, cisplatin decreases the mitochondrial glutathione leading to hydroxyl radicals and oxidative stress, which damage the mitochondrial membrane integrity [21, 24]. Consequently, cytochrome C is released, where it further interplays with Apaf-1. This interaction activates pro-caspase 9, which activates caspase 3 and 7, known as the executioner caspases, leading to the cleavage and activation of poly ADP-ribose polymerase (PARP). PARP commences cell death due to the loss of function and degradation of numerous vital proteins and DNA fragmentation [25], [26], [27].

In addition, cisplatin is considered a mitochondrial-DNA-targeting element. Since mitochondrial DNA (mtDNA) has abundant guanine-rich stretches, cisplatin forms more adducts within the mtDNA structure than cellular DNA. Therefore, it can be conjectured that mitochondrial functions would be impaired as well [28].

Moreover, studies have thoroughly appreciated that MAPK family is impacted by cisplatin. MAPK includes serine/threonine kinases, such as JNK, p38, and ERK, which are incredibly crucial for cell proliferation and survival [29], [30], [31], [32]. To illustrate this, cisplatin activates ERK, which further allows the phosphorylation and activation of p53, which induces the transcription of BAX [33], overexpression of p21, and cell cycle arrest [34, 35]. Besides, DNA damages caused by platinum-based combinations in cisplatin activate JNK which further stabilizes and activates p73, a pro-apoptotic protein. P73 promotes cisplatin-mediated apoptosis in correlation with JNK [36]. Furthermore, P18 interacts with p53, which stabilizes and increases the correlation of p53 with a pro-apoptotic gene called NOXA; hence, cisplatin has the potential to stimulate the cell death pathways through p18-p38 as well [37]. Interestingly, p38 makes another contribution to cisplatin-induced cell death, which is accomplished through inducing the internalization of epidermal growth factor receptors (EGFRs) via the activation of p38 [38].

Besides, it has been reported that cisplatin arrests the cell cycle at the G2 phase through the phosphorylation of Chk1 and Chk2. It also stimulates the activation of Cdc25C and its trafficking to the cytoplasm, which blocks the cell cycle transition to the M phase [39].

Mechanisms of cisplatin resistance through signaling cascades and how to be tackled

Chemoresistance is an unwelcome phenomenon in cancer cells. It has limited the application of different drugs against various cancers. To be precise, chemoresistance is subcategorized into two main categories: 1) innate resistance, in which the administered drug has no impact in the first place, 2) acquired resistance, in which although the chemotherapeutic drug was responsive at the beginning, it loses its beneficiary impacts consequently. The so-called condition arises due to alterations in the cellular drug absorption pattern, changes in drug influx and efflux pattern, conjugation to glutathione or metallothionein, increased drug detoxification, stimulation of DNA repair mechanisms, and inhibition of apoptosis pathways [21, 40].

There have been arguments about cisplatin resistance in patients suffering from colorectal, lung, and ovarian cancer [41, 42]. Numerous mechanisms have been demonstrated in the context of cisplatin resistance; including 1) tumor cells increase drug efflux, therefore decreasing the intracellular accumulation of cisplatin, 2) escalated intracellular level of glutathione and metallothioneins, that are intracellular scavengers, increases drug detoxification, 3) stimulation of DNA repair machinery such as, nucleotide excision repair (NER), inter-strand bound repair and mismatch repair (MMR) which attenuates the impact of cisplatin, and finally 4) adjustment in the approaches of apoptosis-based cell death [43], [44], [45], [46]. The aforementioned events give rise to cisplatin inactivation and prevent the formation of cisplatin-DNA adducts.

Three copper transporters (CTRs) play a distinct role in cisplatin transportation [21, 47]. CTR1 improves cisplatin uptake; however studies on human ovarian carcinoma have represented a decrease in CTR1 expression level [48]. Therefore, loss of CTR1 has been reported to decrease the intracellular accumulation of cisplatin, which leads to cisplatin resistance [49, 50]. Besides, high amounts of ATP7A and ATP7B, which contribute to copper efflux, have been shown to enhance cisplatin resistance [51].

Moreover, many elements such as RNA, thiol peptides, and cell microfilaments bind to cisplatin subsequent to its intracellular accumulation, which results in the blockage of cisplatin [21]. To illustrate this, thiol peptides and RNA strands create adducts with cisplatin within the cell. Glutathione S transferase (GST), an enzyme with a regulatory sensor formed of cysteine residues [52], modulates glutathione and cisplatin conjugation. Therefore, resulting in the blockage of cisplatin-DNA adducts as well. According to studies, high levels of glutathione and GST is associated with cisplatin resistance [53, 54]. Furthermore, metallothionein binds to cisplatin, attenuating the administered dosage and contributing to drug inactivation and resistance [55, 56].

Various proteins like xeroderma pigmentosum A (XPA), XPG, and replication protein A (RPA), are involved in NER [57, 58], and studies have demonstrated that NER deficiencies increase the sensitivity to platinum-based chemotherapeutic drugs [59, 60]. Therefore, over-expression of NER proteins is believed to correlate with cisplatin resistance [61].

Moreover, aberrant activation of cell survival signaling pathways, including PI3K, AKT, and mammalian target of rapamycin (mTOR), is related to increased tumor mass [62], [63], [64]. RSK proteins, such as Ras-ERK mediators that play a significant role in cell survival necessities [65, 66], have been shown to activate BCL-2 and NHE1 to dampen apoptosis pathways [67, 68]. Fig. 1.

Fig. 1.

Fig. 1

Platinum-based combinations in cisplatin constructs inter- and intrastrand adducts, namely lesions, within the nuclear and mitochondrial DNA. Besides it stimulates ROS and oxidative stress. Through the activation of p53, in which p21, Waf1, and MDM2 are recruited, and through the activation of Bax, in which the pro-caspase-9 and -3 are activated, the programed cell death, apoptosis pathways are ignited. To illustrate this, theses apoptosis pathways are accomplished through RAS-MAPK, p38-p18, and JNK cascades.

Paclitaxel

Paclitaxel mechanism of action

Paclitaxel is a member of the chemotherapeutic drug family known as taxane, which agitates the polymerization of tubulins. All taxanes stabilize microtubules, except for vinblastine, which stacks up tubulin and depolymerizes the microtubules [69], [70], [71], [72]. Tubulins are bricks for the structure of microtubules. By impeding the polymerization of microtubules, taxanes can block the cell division in metaphase and anaphase. Needless to mention that, microtubules are vital frameworks in a variety of cell activities, including mitosis, transportation, cell shape sustainability, cell secretion, and phagocytosis, to say the least [73]. Paclitaxel targets these microtubules and wreaks the havoc on spindles during the G2 and M phases, especially during prophase [74, 75]. According to Horwitz, paclitaxel prevents cell division and replication through increasing the polymerization of β-tubulin microtubules and preventing their depolymerization; therefore, poisoned cells will be arrested in the G2/M phases [76] and undergo apoptosis [77]. Shreds of recent evidence have indicated that low concentration, approximately less than nanomolar, inhibits the depolymerization of microtubules, whereas high-dose administration of paclitaxel stimulates an excessive polymerization of microtubules, which renders them nonfunctional owing to their increased stability [78]. To be precise, paclitaxel couples directly and specifically to the N-terminal of the β-tubulin units in a reversible manner instead of binding to the tubulin dimers [79, 80]. Interestingly, paclitaxel has the potential to stimulate the tubulin polymerization even in 4-degree centigrade condition independent of GTP [77].

Mechanisms of paclitaxel in cell death through signaling cascades

It is assumed that weekly consumption of paclitaxel has a dose-dependent apoptotic impact on malignant cells independent of its microtubule stabilization potential. 10 nM of paclitaxel in 12 h brings about apoptosis in the S phase without mitotic arrest [78]. Besides, administration concentration of ≥9 nM stimulates the activation of Raf-1, which promotes further apoptosis mechanisms, whereas administration of ≤9 nM induces p53 and p21, furthering apoptosis independent of Raf-1 activation [78, 81, 82].

In addition, paclitaxel is capable of inducing other pro-apoptotic pathways. The toll-like receptor-4- (TLR-4) related pathway, JNK, MAPK, NF-KB, JAK, and STAT, are associated with the cell-death activity of paclitaxel. The activation of MAPK leads to the dephosphorylation of Bad and Bax as well as the phosphorylation of Bcl2, which ultimately triggers apoptosis cascade [83], [84], [85], [86].

Weekly administration of paclitaxel has displayed anti-angiogenic activities as well [87, 88]. In other words, pieces of evidence have lent support to the fact that low-dose administration of paclitaxel inhibits the expression of vascular endothelial growth factor (VEGF) [89, 90].

The other mechanism of action for paclitaxel is accomplished through the promotion of ROS generation. This process is triggered by the induction of nicotinamide adenine dinucleotide phosphate (NADPH) oxidase that results in the oxidative stress and the subsequent cell death pathways [91, 92].

Paclitaxel has impacts on the immune system as well. It suppresses the immune system in a dose-dependent manner. Providing that it is administered in a standard dose, it dampens the function of macrophages, natural killer (NK) cells, and effector T cells [93]. Whereas, prescribing low-dose paclitaxel stimulates immune responses and anti-tumor properties in immune cells [93, 94].

As it has been mentioned earlier, paclitaxel is capable of binding to TLR-4, which is expressed on innate-immunity cells. The binding of paclitaxel on macrophages, for instance, recruits MyD88, MAPK, and NF-KB, further leading to the expression and secretion of TNF-α, IL-1, IL-6, and IL-8 [95], [96], [97]. These stimulated macrophages can either directly lyse tumor cells through secreting nitric oxide (NO) and lysosomal enzymes or activate dendritic cells (DCs), NK cells, and cytotoxic T lymphocytes (CTLs) to invade tumor cells indirectly [93]. In addition, DCs can also be directly stimulated by paclitaxel through TLR-4. This leads to DC maturation through increasing the expression level of antigen-presenting related genes, costimulatory molecules, and IL-12P70 [94]. Therefore, it promotes the function of DCs to stimulate T cells toward anti-tumor responses [98].

It is now well appreciated that tumor cells can evade cytotoxic T lymphocytes by reducing the specific tumor antigens and MHC class I. However, NK cells seem to assist in eliminating tumors in this regard. It has been shown that paclitaxel induces mRNA transcription and protein expression of perforin in NK cells, therefore stimulating their cytotoxic ability against tumor cells [99]. However, one should bear in mind that the impact of paclitaxel on NK cells is dose-dependent as well [100, 101].

Paclitaxel also has other impacts on immune-related functions. To illustrate this, paclitaxel upregulates mannose-6-phosphate, which increases the cellular permeability to granzyme B, hence improving the cytotoxic function of T lymphocytes. It also promotes the secretion of type I cytokines, including IL-2 from CD4+ T cells and IFN-γ form CD8+ T cells [97, 102]. Activated CD8+ T cells can further differentiate into MHC I-CTL type 1 (Tc1) secreting IFN-γ, responsible for tumor cell lysis [103]. CD4+ CD25+ Tregs are known to favor tumor cells by suppressing immune responses and promoting cancer progression [78]. Murine studies have demonstrated that paclitaxel diminishes both the number and the size of Tregs, enhancing the anti-tumor properties in immune cells [104], [105], [106]. Paclitaxel also diminished the expression level of Bcl-2, an anti-apoptotic agent, whereas increasing the pro-apoptotic factor, namely Bax, in Tregs, rendering them more susceptible to undergo apoptosis [107].

Furthermore, myeloid-derived suppressor cells (MDSCs), which are being labored by tumor cells to expand and to suppress immune responses, are dampened in a C57Bl/6 mice study by implementing an ultra-low dose of paclitaxel. To be precise, an ultra-low dose administration of paclitaxel stimulates the differentiation of MDSCs into DCs independent of TLR-4 [102]. Owing to the data mentioned above, altogether with the suppression of Tregs, paclitaxel can promote immune responses against tumor cells and ignite the apoptosis pathways within the these cells.

Mechanisms of paclitaxel resistance through signaling cascades and how to be tackled

Multifactor approaches and numerous genes correlate with paclitaxel resistance, yet not all are well elucidated [78]. It has been proposed that paclitaxel-resistant cells mainly contain changes in their mRNA and protein synthesis level, including numerous ribosomal genes and transcriptional factors. They also represent shifts in oxidative stress-related genes such as UGT1A6, MAOA, and CYBA, as well as in glycolysis, including ADH1A, HK1, and ENO3, and glutathione metabolism pathways [108]. Findings of a study, in which tumor cells were treated with paclitaxel and other chemotherapeutic drugs, have lent support to this proposal. In other words, the expression level of 337 genes among a total of 845 genes was changed [109]. Alterations in their gene expression profiles interfere with the drug impacts, tumor microenvironment, cellular structure, and metabolism shifts that can lead tumor cells toward resistance [78].

Moreover, due to the aggressive proliferation in cancers, new malformed vessels are constructed to meet their oxygen demand; nonetheless, their growth rate leads to oxygen depletion and hypoxia [110]. Tumor cells evolve and adapt to this microenvironment condition. Therefore, they remain alive with the potential to increase their proliferation rate, dissemination, invasion, progression [111], and diminishing their chemosensitivity [112].

This does not end here; hypoxia induces chemoresistance through changing both the genomics and proteomics profiles. In other words, oxygen depletion activates P53, inhibits apoptosis, and enhances angiogenesis through increasing VEGF and angiogenin. It also promotes the production and secretion of growth factors such as platelet-derived growth factor (PDGF), transforming factor-beta (TGF-β), and insulin-like growth factor (IGF), as well as inducing glycolysis, which altogether contribute to cell survival [112]. Tumor cells avoid apoptosis, which as discussed above, is mediated by the hypoxic tumor microenvironment and hypoxia-induced factor-1 (HIF-1), that arrests cells at G0/G1 checkpoint, rendering them resistant to chemotherapeutic drugs, including paclitaxel [109, 113]. Furthermore, other transcriptional factors like NF-KB and STAT3, which stimulate pro-inflammatory responses, cell survival, angiogenesis, cell division, and metastasis, are activated under hypoxic conditions. Besides, paclitaxel phosphorylates BCL-2, an anti-apoptotic factor, through NF-KB, which dampens apoptosis and promotes chemoresistance under hypoxic conditions [114]. It has been shown that inhibiting the activation of the over-expressed STAT3 has reduced resistance against paclitaxel [115]. Moreover, the activation of lipopolysaccharide-inducible genes, MAPK, Raf-1 kinase, PI3K, as well as increased pro-inflammatory cytokines such as TNF-α, IL-6, and IL-8 contribute to paclitaxel resistance [84, 107, 116, 117].

Drug efflux is a commonplace mechanism of chemoresistance in tumor cells treated with various chemotherapeutic drugs. Drug efflux is accomplished through ATP-binding cassette (ABC) transporters such as P-glycoprotein (P-gp). P-gp is encoded by MDR-1 and has been shown to contribute to chemosensitivity to paclitaxel [83]. It has been reported that P-gp is over-expressed in paclitaxel-resistant tumor cells [109].

In addition, tubulin-related mechanisms have been introduced to play a role in paclitaxel resistance. In one way, their intracellular concentration is reduced and they are not sufficient for paclitaxel to perform its action. On the other hand, point mutations in tubulin genes and changes in the expression level of various tubulin isotypes, such as class-III β-tubulin, contribute to paclitaxel resistance [118], [119], [120]. Fig. 2.

Fig. 2.

Fig. 2

Paclitaxel targets microtubules and wreck the havoc on spindles in G2, prophase, and M phase. To be precise, it increases the polymerization of β-tubulin microtubules and prevents their depolymerization, which leads to the cell cycle arrest in G2/M phase. 10 nM of paclitaxel in 12 hrs stalls the cell cycle at the S phase, and by the administration of ≥9 nM ignites the apoptosis through the activation of Raf-1, whereas, administration of ≤9 nM recruits p53 and p21. TLR-4 is involved in two manners as well. Regarding apoptosis, JNK, NF-KB, MAPK, JAK, and STAT participate, where they lead the cell's fate to apoptosis through the dephosphorylation of Bad and Bax, and through the phosphorylation of Bcl2. Regarding the immune system, through the stimulation of TLR-4, paclitaxel recruits MyD88, MAPK, NF-KB which results in the expression and secretion of IL-1, IL-6, IL-2, IL-8, TNF-α, and IFN-γ.

Gefitinib and erlotinib

Gefitinib and erlotinib mechanism of action

Tyrosine kinase inhibitors (TKIs) are non-peptide combinations prescribed orally and are structurally identical to adenosine triphosphate (ATP). They compete for the ATP-binding domains in kinases, therefore stopping the subsequent signaling transduction, which ultimately leads to apoptosis and the prevention of cell proliferation [121, 122]. Lapatinib, gefitinib, and erlotinib are representatives for TKIs. The expression level of tyrosine kinase receptors of epidermal growth factors is increased on cancer cells; therefore, gefitinib, for instance, is administered to block them. EGFRI-gefitinib prohibits cell proliferation and survival by stopping the growth signaling through EGFRs [123], [124], [125]. Gefitinib has decreased the size of thyroid tumors [126], and has had beneficial impacts on non-small cell lung cancer as well [125, 127], yet chemoresistance is inevitable [128]. Erlotinib acts through similar approaches; however, it is expensive, and chemoresistance has limited its application [129, 130]. The half-life of erlotinib is about 36 hours, and it has a high affinity for plasma proteins such as albumin and α-1 acid glycoprotein [131]. Erlotinib is mainly metabolized by cytochrome P450 (CYP) 3A4, yet it can slightly be metabolized by CYP1A2. It has been reported that co-administration of erlotinib and a CYP3A4 inhibitor, namely ketoconazole, has enhanced the plasma level of erlotinib [132].

Mechanisms of gefitinib and erlotinib in cell death through signaling cascades

EGFRs are overexpressed in a variety of malignancies including, non-small cell lung cancer (40–80%), colorectal cancer (72–82%), head and neck cancer (95–100%), breast cancer (14–91%), and renal cell cancer (50–90%) [133]. Moreover, numerous studies proposed a cross-talk between EGFR and VEGF pathways [134].

As mentioned earlier, gefitinib and erlotinib are EGFR-tyrosine kinase inhibitors, which means that they act as an ATP antagonist and compete for their intracellular binding sites on the tyrosine kinase domain of EGFRs, therefore dampening their tyrosine kinase activity. EGFR is a 170-kDa transmembrane glycoprotein and is a member of the ErbB family, including ErbB-2/Neu/HER2, ErbB-3/HER3, and ErbB-4/HER4. These cell surface receptors contribute to cell growth regulation, differentiation, and cell survival [135]. EGFR dimerizes in either hetero or homo manner by the time their extracellular domain binds to its ligand. This dimerization allows the intracellular domains to perform their tyrosine kinase activity, which leads to the autophosphorylation of their cytoplasmic tyrosine residues. The tyrosine residues further provide a docking platform, where a variety of intracellular signaling proteins bind to and become activated, which ultimately impact the expression level of multiple genes [132, 135].

Various central downstream signaling cascades associate with EGFR activation. These cascades include PI3/Akt, Jak2/STAT3, extracellular signal-regulated kinase (ERK1,2), phospholipase C (PLCγ), Ras/Raf mitogen-activated protein, and modulation of calcium channels which inhibit apoptosis, nonetheless, promote cell proliferation, angiogenesis, metastasis, and invasion [132, 136]. Therefore, inhibiting EGFRs and their downstream signaling cascades dampen cell proliferation, metastasis, invasion, and angiogenesis, as well as the induction of apoptosis and assisting the concurrent radiotherapy treatments.

Recent somatic mutation discoveries in the ATP-binding domain of EGFR genes have shed light on achieving desired responses in EGFR-TKIs like gefitinib and erlotinib [137], [138], [139]. To be precise, frame deletion mutations in exon 19, codons 746–750, which has the frequency of 45–50% among all EGFR mutations, and missense mutations at codon 858, that lead to the substitution of leucine with arginine in exon 21 with the frequency of 35–45%, were among the most interesting mutations [139]. A population of non-small cell lung cancer that contained the aforementioned somatic mutations represented better responses subsequent to EGFR-TKIs administration [140, 141]. Recent studies have lent support that erlotinib/gefitinib responders carry at least one of these EFGR gene mutations [132].

In addition to similar functions among different TKIs, gefitinib has other impacts as well. In other words, gefitinib administration increases the level of P27, while reduces cyclin D1 and Cdk4 during sub-G1/G1 phases in the cell cycle. Moreover, gefitinib diminishes the phosphorylation of glycogen synthase kinase-3 beta (GSK-3β), which is an AKT kinase target [142, 143]. Shreds of evidence supported that gefitinib induces apoptosis partly through an increased level of P38 MAPK, dephosphorylation of FOXOs with a successive enhanced level of P27, and increasing the expression of caspase 3 and BIM [143], [144], [145]. Besides, mTOR signaling has been reported to be down-regulated following gefitinib administration [146].

Mechanisms of gefitinib and erlotinib resistance through signaling cascades and how to be tackled

Most patients develop resistance against TKIs within just a year. Although some mechanisms have been discovered so far, many others have remained elusive.

Dysregulated microRNAs are shown to contribute to TKI-chemoresistance [147]. To be precise, for instance, miR-21 is dramatically overexpressed in erlotinib-resistant non-small cell lung cancer, which gives rise to the downregulation of PTEN and PDCD4, leading to an increased cancer progression and proliferation through the activation of AKT [148, 149]. However, the tumor-suppressive miR-34 is shown to be downregulated in TKI-resistant cancers, including those which were treated with erlotinib. MiR-34 has been overexpressed alone or in concurrence with erlotinib administration to resensitize tumors and promote cell cycle arrest [147]. Interestingly, it is now well established that miR-147b can stimulate resistance against even the third generation of EGFR TKIs, namely osimertinib, through altering the TCA cycle [150]. Moreover, miR-17-5p and miR-641 are also involved in erlotinib resistance. It has been shown that the overexpression of miR-17-5p resensitizes tumor cells to erlotinib through EZH1 in non-small cell lung cancer [151]; nonetheless, the overexpression of miR-641 works in favor of erlotinib resistance via the downregulation of NF1 [147]. A.S Pal, et al. have firmly demonstrated that miR-5693, miR-3618, and miR-432-5p regulate erlotinib resistance and drug efflux. Besides, miR-5693, miR-3618, and miR-432-5p were evaluated in the presence of other EGFR TKIs, including gefitinib and afatinib. Interestingly, they have significantly promoted cell proliferation and cancer progression in the presence of these TKIs [147].

Recent studies have demonstrated that the stimulation of epithelial-mesenchymal transition (EMT) promotes resistance to EGFR TKIs [152, 153]. It is well appreciated that EMT is triggered by the activation of the transforming growth factor-beta (TGF-β) pathway, which seems to be one of the main drivers of tumor invasiveness and metastasis by inducing migration in various cancers, such as lung malignancies. To illustrate this, TGF-β ligands bind to the TGF-β receptor complex, which is constructed from TGF-βRI and TGF-βRII. After that, the receptor phosphorylates and activates Smad2 and Smad3, which further construct a transcriptional complex with Smad4, that, accompanying other transcriptional factors, can stimulate the expression of tumor-progression related genes [154]. Therefore, a high expression level of TGF-β contributes to tumor expansion, proliferation, and metastasis [155]. It has been envisaged that EGFR TKIs can dampen TGF-β-induced cell movements [156]. Therefore, Serizawa et al. have shown that administration of TGF-β inhibitors, namely LY364947, along with EGFR TKIs such as erlotinib and gefitinib, dampens the progression and invasion of EGFR TKI-resistant cancer cells [157].

In addition, integrins have been reported to be involved in erlotinib-resistant cancers. Kanda et al. have represented that all resistant cells showed an increased level of β1, α2, and α5 integrins. Furthermore, they illustrated that using siRNA, in order to knockdown integrin β1, they could resensitize malignant cells to erlotinib, in which Akt phosphorylation was thoroughly dampened after the administration of erlotinib. This notion signifies the fact that silencing integrin β1 alleviates erlotinib-chemoresistance. Nevertheless, silencing α2 and α5 integrins resulted in moderate inhibition of Akt phosphorylation in these cancers. Moreover, the expression level of integrin α5 was also downregulated when silencing integrin β1, suggesting a link between integrin β1 and α5. Altogether these findings signify the role of integrins in erlotinib-chemoresistance [158].

In addition, ras-associated binding protein-25 (Rab25) mediates EGFR TKI-resistance in non-small cell lung cancer. In this regard, Wang et al. have reported that the overexpression of Rab25 stimulated resistance against erlotinib. To be precise, Rab25 interacts with integrin β1, which facilitates its migration toward the membrane, where it induces the phosphorylation of Akt. It further triggers the activation of the Wnt/β‐catenin signaling cascade and promotes cell proliferation and cancer progression. They also represented when silencing Rab25, Wnt/β‐catenin and Akt-related signaling cascades were dampened, which resensitized cancer cells to erlotinib [159].

The activation of EGFR downstream cascades correlates with gefitinib resistance as well [160]. The hyperactivation of the MAPK pathway, especially P42-MAPK, contributes to the intrinsic and acquired resistance in breast cancer models. Therefore, inhibition of cell proliferation and induction of apoptosis were witnessed by dampening MAPK-related pathways [161, 162]. Interestingly, some studies have indicated a transcriptional role for gefitinib, which is accomplished by increasing the expression of HER-specific ligands and augmenting the importation and nuclear accumulation of an HER ligand, namely neuregulin [163]. Moreover, some results demonstrated the upregulation of PI3K, which results from a non-functioning PTEN, as another reason underlying gefitinib resistance. In other words, in studies carried out on subtypes of breast cancer overexpressing EGFR, but without PTEN activity, gefitinib terminated the phosphorylation of EGFR and MAPK, leaving AKT intact. However, regarding the cell populations subjected to gefitinib with restored-function PTEN, the phosphorylation of AKT, GSK-3β, and translation repressor protein 4EBP1 were dampened. They were also resensitized to gefitinib and were arrested at the G1 phase [164], [165], [166].

Other pathways related to the interaction between EGFRs and hepatocyte growth factor receptor (MET) also correlate with gefitinib chemoresistance. MET is a tyrosine kinase receptor that increases the phosphorylation of EGFR through the activation of c-Src kinase; therefore, it is involved in breast cancer progression and gefitinib resistance. It has been shown that, by inhibiting MET in HER-2 positive and EGFR-overexpressed breast cancers, cell proliferation was dampened, and cells were resensitized to gefitinib [167]. Furthermore, The interplay between EGFRs, G-coupled protein receptors (GCPRs) [168], K-RAS —an EGFR-downstream executioner— [169], and EGFR mutations have all been reported to be involved in erlotinib-, gefitinib-resistant lung cancer, breast cancer [170], and adenocarcinomas [171, 172]. Fig. 3.

Fig. 3.

Fig. 3

Tyrosine kinase inhibitors (TKIs) are similar to ATP, hence they compete for ATP-binding domains in kinases of EGFRs. They inhibit PI3K/Akt, JAK/STAT, PLCγ, and Ras/Raf, therefore they are able to ignite apoptosis machinery in the absence of the aforementioned pathways. Gefitinib also increases the level of p27 and induces p38-MAPK as well which give rise to apoptosis. One major signaling that cells develop resistance against EGFR TKIs is through TGF-β signaling. Recruiting Smads 2,3, and 4, it stimulates cell survival and EMT. Moreover, Rab25 which has close interplay with integrin β1 correlates with TKI resistance. To be precise it phosphorylates Akt and ultimately assists survival and proliferation signaling cascades.

Gemcitabine

Gemcitabine mechanism of action

2′,2′-difluoro-2′-deoxycytidine (dFdCTP), which is known as gemcitabine, acts as the analog of deoxycytidine during a process called “masked chain termination” [173]. Implementing the dFdCTP enables its substitution with typical nucleoside, therefore stopping the DNA polymerase movements and DNA duplication. Nonetheless, DNA repair mechanisms fail to detach gemcitabine from the DNA strands [174].

The second mechanism that gemcitabine acts through is by empowering its ability to inhibit crucial enzymes in deoxynucleotides metabolism. Deoxycytidylate deaminase (dCTD), a member of these enzymes, is inhibited directly and indirectly by dFdCTP and dFdCDP, respectively. To be precise, dFdCDP binds covalently to the active sites of ribonucleotide reductase (RR), impeding the conversion of ribonucleotides into deoxynucleotides, which results in dNTP pool shortage. This shortage consequently leads to a diminished activity in dCTD. Moreover, deoxycytidine kinase (dCK) is modulated by dCTP; therefore, diminished-dNTP pool stimulates dFdC phosphorylation through the activation of dCK, which ultimately increases the level of dFdCTP and its interference within the DNA structure [175], [176], [177], [178].

Moreover, caspase signaling cascades are the third mechanism of action leading to cell apoptosis that has been defined for gemcitabine [179, 180].

Gemcitabine is among the most energetic chemotherapeutic drugs with a wide range of prescription for numerous malignancies, including bladder cancer [181], pancreatic cancer [182], non-small cell lung cancer [183], and breast cancer [184]. There have been promising results for approval of gemcitabine for ovarian cancer. Needless to mention that, it has fewer side effects and toxicity in comparison to the other chemotherapeutic agents, hence suggesting gemcitabine as a better option for chemotherapies [173]. However, the responses are not always satisfying, which brings about the demand for developing new approaches in its application, like combination chemotherapy, or even develop new regimens.

Mechanisms of gemcitabine in cell death through signaling cascades

Interestingly, gemcitabine activates p38 mitogen-activated protein kinase (MAPK), which ignites apoptosis in cancer cells, leaving normal cells intact [185, 186]. Besides, gemcitabine activates MAPK-activated protein kinase (MK2), which induces the phosphorylation of heat-shock protein-27 (HSP-27), resulting in the suppression of tumor cell growth [187].

There have been pieces of evidence representing that the S-phase checkpoint is activated and slowed due to the breaks and lesions in DNA replication after gemcitabine treatment [188, 189]. Typically, as the replication comes to a halt, probably due to a break in DNA, the ataxia-telangiectasia mutated kinase (ATM) and checkpoint kinase 2 (Chk2) are activated, which manage the cell cycle arrest, apoptosis, and DNA repair. Besides, when there is a DNA lesion, the ataxia-telangiectasia mutated and Rad3-related kinase (ATR), checkpoint kinase 1 (Chk1), and Rad9-Hus1-Rad1 (9-1-1 complex) are activated that inhibit cell cycle progression, stabilize the replication fork, and induce DNA repair mechanisms by the assist of Rad17 and replication factor C [190, 191]. The criteria mentioned above argue that gemcitabine administration leads to cell cycle arrest initially; however, ATM facilitates the cell cycle progression independent of p53. Supporting this idea, Larry et al. showed that ATM depletion sensitizes tumors toward gemcitabine [189].

Mechanisms of gemcitabine resistance through signaling cascades and how to be tackled

Similar to any other chemotherapeutic drugs, tumor cells develop either intrinsic or acquired chemoresistance against this gemcitabine. Gemcitabine has been the first-line treatment for pancreatic cancer. However, one major obstacle toward tackling pancreatic cancer is the extensive desmoplastic reaction, which constructs approximately 90% of the tumor mass. Desmoplastic reaction accounts for poor drug delivery and innate-gemcitabine chemoresistance in pancreatic cancer cases [178, 192]. Besides, Hedgehog (Hh) signaling has been shown to play a distinct role in gemcitabine chemoresistance. Hh signaling correlates with morphogenesis [193], tumorigenesis, desmoplastic reaction, and the alteration in extracellular matrix construction [194, 195]. In this regard, genetic analyses were accomplished to locate frequent mutations in pancreatic tumors. Hh is the most genetically shifted signaling in this cancer [196]. Interestingly, Olive et al. have shown that by inhibiting the Hh signaling in mice, they could increase the intratumor drug delivery up to 60% [197].

Moreover, the activation of MK2 plays an essential role in apoptosis induction after gemcitabine administration. A study carried out by Cöpper et al. has shown that the translesion polymerase activity has inhibited the activation of MK2, which ultimately facilitated osteosarcoma survival [198].

Since a successful drug delivery is the first step toward achieving desired chemotherapy responses, any disturbances or dysfunctions in this process might facilitate chemoresistance. In this regard, studies have represented that a low level of hENT1, which is the primary transporter of gemcitabine, is associated with gemcitabine resistance and a low survival rate among pancreatic cancer patients [199, 200].

It is well appreciated that gemcitabine is a prodrug that needs to be processed by dCK. There has been evidence that loss of dCK mRNA in ovarian and pancreatic cancer cell lines was associated with gemcitabine resistance [201, 202].

Besides, ribonucleotide reductase (RR) is a holoenzyme formed of RRM1 and RRM2 and has been shown to downregulate the gemcitabine activity. There have been shreds of evidence indicating that an increased level of these subunits is associated with gemcitabine resistance [203], [204], [205].

Transcription factors are responsible for gemcitabine resistance as well. One of them is high mobility group A1 (HMGA1), which are framework-related transcription factors that modulate various genes and are escalated in various tumors [206]. Moreover, NF-KB is the other transcription factor that is related to gemcitabine resistance. Arlt et al. represented that among five different pancreatic cancer cell lines, resistant cell lines, including BxPc-3, Capan-1, and PancTu-1, have displayed an overexpressed level of NF-KB [207]. Gemcitabine resistant cells exhibiting overexpressed NF-KB might have correlations with the impact of Apurinic/apyrimidinic endonuclease 1/redox factor-1 (APE1/Ref1) on the activation of transcription factors such as NF-KB [208]. Moreover, NF-KB has been shown to assist gemcitabine resistance by inhibiting the expression of hCNT1, a nucleoside transporter, in pancreatic cancer cell lines [209]. Excessive cellular accumulation of NF-KB targets hypoxia-inducible factor 1-α (HIF-1α) that contributes to gemcitabine resistance. Firstly, HIF-1, a transcription factor that plays a role in angiogenesis, invasiveness, metastasis, and chemoresistance in tumors [210]; secondly, there has been related evidence that HIF-1α down-regulates hENT1 and hENT2 [211, 212]. There are also other mechanisms that both NF-KB and HIF-1α might act through, ultimately resulting in gemcitabine resistance. To illustrate this, CXCL12/CXCR4 downstream pathway stimulates cell survival and proliferation factors, such as NF-KB [213]. Solid witnesses claim that CXCL12/CXCR4 interaction gives rise to gemcitabine resistance by promoting PI3K/Akt, extracellular-signal-regulated kinase (ERK), and focal adhesion kinase (FAK) in pancreatic cancer cells [214]. What is more, gemcitabine induces the expression of CXCR4 through NF-KB and HIF-1α, which leads to increased invasiveness in pancreatic cancer cells [215]. This is in accordance with other experiments reporting that low doses of gemcitabine stimulate the expression of NF-KB in non-small cell lung cancer and pancreatic cancer [207, 216]. Besides, reports show that CXCL12/CXCR4 downstream pathway promotes Hh signaling through NF-KB [217]. Altogether creates a cycle that facilitates cancer cell survival via NF-KB signaling, perhaps the key player in gemcitabine resistance.

Besides, Nagano, et al. have represented that miRNA-29a induces gemcitabine resistance through Wnt/β-catenin in pancreatic cell lines [218]. Wnt/β-catenin signaling pathway contributes to cell differentiation, cell proliferation, the onset of numerous malignancies, and chemoresistance [218], [219], [220]. Studies have demonstrated that roughly 65% of pancreatic cancer patients display activated Wnt/β-catenin signaling, which helps generate chemoresistant cancer stem cells [221], [222], [223]. Lately, Kapinas et al. has shown that Wnt/β-catenin is activated through Dikkopf-1 (Dkk1), Kremen2, and secreted frizzled-related protein 2 (sFRP2) by miRNA-29a [224].

As mentioned earlier, gemcitabine is the best option for pancreatic cancer treatment [182]. However, due to the central-tumor hypoxic condition, chemotherapy responses are not satisfactory. Hypoxia is associated with poor prognosis, chemo- and radio-resistance, metastasis, and apoptosis inhibition [225]. PI3K, Akt, and MAPK as vital downstream signaling pathways associated with cell proliferation, cell migration (metastasis in this regard), inhibition of apoptosis, and responding to the growth factors, are phosphorylated and activated under hypoxic condition [226, 227]. Yokoi et al. have shown an enhancement in the DNA-binding activity of NF-KB and in the activation of PI3K/Akt and MAPK under hypoxia, which leads to gemcitabine resistance in L3.6pl cells, nonetheless, by inhibiting PI3K using PKI166, they were able to induce apoptosis in this cell population [228].

Another factor associated with gemcitabine resistance is Annexin II. There has been evidence arguing that the overexpression of Annexin II is related to tumor relapse in pancreatic cancer patients receiving gemcitabine; nevertheless, by inhibiting Annexin II, the toxicity of gemcitabine has increased [229]. Besides, Kagawa et al. have demonstrated that Akt/mTOR signaling is involved in Annexin-II-related gemcitabine resistance, to say the least [230]. Fig. 4.

Fig. 4.

Fig. 4

By the time gemcitabine is administered, it activates p38-MAPK that ignites apoptosis through MK2 and phosphorylation of HSP-27. Moreover, due to the DNA damage caused by gemcitabine, S phase checkpoint is activated, ATM/Chk2 and ATR/Chk1 are recruited which altogether with the company of 9-1-1-complex phosphorylate and activate p53. Regarding resistance, desmoplastic reaction seems to be accomplished through Hh signaling which accounts for poor drug delivery. RR also has been shown tot attenuate the activity of gemcitabine. Furthermore, CXCL12/CXCR4 stimulates PI3K/Akt, ERK, and NF-KB which along with Wnt/β catenin decreases the efficiency of gemcitabine therapy.

Doxorubicin

Doxorubicin mechanism of action

Doxorubicin is a non-selective class-I anthracycline chemotherapeutic agent that contains aglyconic and sugar components. It binds to the plasma protein carriers and enters its target cell through passive diffusion subsequent to its administration, where it further accumulates intracellularly. After that, it mainly traffics into the nucleus excelling its cytosolic concentration [6]. It is mainly prescribed for breast, bladder, stomach, lung, ovaries, thyroid cancers, and soft tissue sarcoma [231].

Although its mechanism of action is elaborate, it is proposed to act as an anti-cancer agent through two main mechanisms. a) Doxorubicin interferes with DNA, and in a process known as intercalation, dampens the biosynthesis of numerous macromolecules as well as disrupting the progress of topoisomerase II. In other words, doxorubicin stabilizes the DNA following the DNA breaks created by topoisomerase II, therefore stops the release of the DNA double helix, which further prohibits the DNA replication [232]. b) Doxorubicin stimulates free radicals which further triggers damages in DNA, proteins, and membrane [231].

Mechanisms of doxorubicin in cell death through signaling cascades

Doxorubicin is oxidized into an unstable metabolite, namely semiquinone, a process in which the reactive oxygen species (ROS) are generated. ROS damages organelle membranes, DNA, and proteins, to say the least. It also causes oxidative stress, increases the alkylation and lipid peroxidation [231], and induces apoptotic pathways [233]. For instance, The apoptotic pathway is triggered when cells fail to repair the damaged DNA, which leads to cell cycle arrest in the G1 and G2 phases [231].

Moreover, doxorubicin induces autophagy in response to DNA damage. The activation of the nuclear enzyme, namely poly (ADP-ribose) polymerase-1 (PARP-1), is known to be crucial in order to determine whether cells undergo autophagy or not. DNA stresses can stimulate PARP-1 hyperactivation, which further results in the shortage in NAD+ and ATP resources. Consequently, cells would experience energy-resource depletion, which leads them toward apoptosis, providing that it is irreversible [234]. These findings reinforce that the normal-dose administration of doxorubicin, but not in high-dose, directs cells toward autophagy and necrosis due to the cellular energy collapse subsequent to PARP-1 hyperactivation [231].

Besides, doxorubicin stimulates the activation of AMP-activated protein kinase (AMPK). This activation is accomplished by activating ROS-dependent liver kinase B1 (LKB1), which supplies AMPK with necessary upstream signals [6]. Activated AMPK stimulates the activation and phosphorylation of serine 15 in P53, which further triggers apoptosis in B16 melanoma cells [235]. Therefore, AMPK induces cell death pathways through activating JNK kinase in liver and insulin-secretor β-Langerhans cells [236, 237].

Moreover, studies on MCF-7 have shown that doxorubicin diminishes the concentration of Bcl-2, an anti-apoptotic agent, whereas it enhances the amount of Bax, which acts against Bcl-2 [238]. It was first conjectured that doxorubicin decreases the mRNA level of Bcl-2 in a P53-independent manner. However, according to MCF-7 studies [238], it has been suggested that doxorubicin impacts the expression level of Bcl-2 through P53, owing to the fact that the ratio of Bcl-2/Bax must be altered in order for caspases to be activated [6].

Besides, some argued that doxorubicin acts partly through recruiting some components in Fas/Fas-ligand apoptotic pathway; nonetheless, others provided contradictory outcomes [239, 240].

Mechanisms of doxorubicin resistance through signaling cascades and how to be tackled

Autophagy can either stimulate or dampen chemoresistance, depending on the tumor's nature, the drug properties, the treatment duration, and how drugs influence metabolic stress [6]. The high mobility group box 1 protein (HMGB1) is a crucial regulator of both selective and non-selective autophagy. It also plays a critical role in DNA replication and DNA repair mechanisms [241]. Different studies indicated that tumor cells become sensitive to doxorubicin by the time both autophagy and HMGB1 are blocked, meaning that HMGB1 contributes to doxorubicin resistance during tumor progression. In other words, subsequent to doxorubicin administration, the mRNA and protein levels of HMGB1 are enhanced. Besides, HMGB1 competes with Bcl-2 in binding to BECN1; therefore, establishing the BECN1-PtdIns3KC3 complex induces autophagosome formation and triggers autophagy. Moreover, the upstream signal, namely ULK1-mATG12-FIP200, is required for the interplay between HMGB1 and BECN1 [6].

Ceramides also have provoked interest in contributing to doxorubicin resistance. Ceramide is known as a cellular lipid messenger and plays a distinct role in regulating doxorubicin-induced cell death. Results indicate that doxorubicin induces the cellular level of ceramides, resulting in the upregulation of glucosylceramide synthase (GCS). Increased glycosylation of ceramides and cellular stress caused by doxorubicin administration results in cellular drug resistance. Possible mechanisms in which doxorubicin increases the level of ceramides remains elusive to some part; nonetheless, some believe it is either accomplished through activating enzymes responsible for ceramide synthesis or by stimulating sphingomyelinase [242].

Furthermore, it has been reported that doxorubicin resistance in melanoma cells is considerably modulated through ABCB8, which seems to play a role in the preservation of the mitochondrial genome; nevertheless, the exact mechanisms are yet to be discovered. It is worth mentioning that, although ATP-binding cassette (ABC) proteins are involved in numerous chemoresistant tumors, ABCB8 is explicitly involved in doxorubicin resistance, but not in other chemotherapy agents. It has been reported that ABCB8 knockdown using specific shRNA could diminish doxorubicin resistance [243, 244]. Besides, other ABC transporters have been shown to partially correlate with doxorubicin resistance including, ABCB1 (MDR1, Pgp) [245] ABCC1 (MRP1) [246], as well as ABCC2, ABCC3, ABCG2, and RAPBP1 [247], [248], [249].

Besides, replication of the TOP2A gene is the other defined mechanism in doxorubicin resistance [250]. The amplification of this gene has been reported to influence doxorubicin response [251, 252]. Interestingly, TOP2A has a neighboring interplay with HER-2, a marker for breast cancer progression and chemotherapy response. It is now clear that the amplification of HER-2 also impacts doxorubicin response [252]. Fig. 5.

Fig. 5.

Fig. 5

Doxorubicin forms adducts with Topoisomerase and DNA in a process, namely intercalation. When entered in to the target cell, DOX is oxidized into semiquinone which ROS are generated in its process. ROS further, causes lipid peroxidation, organelles’ membrane damage, DNA damage, and protein damage. Following the DNA damage PARP-1 is activated which leads cell toward either autophagy (providing that ATP and NAD+ resources are low), or G1/S arrest and apoptosis. Protein damages also lead to G2/M arrest which gives rise to apoptosis. Moreover, DOX can phosphorylate and activate p53 through JNK,AMPK, and p38-MAPK. However, increase in the protein or mRNA level of HMGB1 attenuates apoptosis and favors cell survival and resistance.

Etoposide

Etoposide mechanism of action

By and large, etoposide is an anti-topoisomerase II and a semi-synthetic podophyllotoxin derivative agent that interrupts DNA replication and other mechanisms, in which topoisomerase II is required including, chromatin remodeling, DNA transcription, and DNA repair. Two essential transesterification reactions enable topoisomerase to create a transient break within the double helix strain. The tyrosine motif at the active site of the enzyme establishes a covalent bond with one phosphate residue within the DNA backbone, resulting in a temporary interruption in DNA strain integrity during the first step. The second reaction takes place to re-ligate the double helix by releasing the DNA break. The cleaved double-helix DNA resulting from the first reaction is temporary; however, it can be stabilized using toxic agents [253].

Etoposide is one of the recommended agents, which poisons the topoisomerase-II cleavage complex (TopoIIcc) and prevents the second transesterification reaction, namely re-ligation. To be precise, etoposide has a low affinity toward naked DNA, whereas it shows a high affinity toward the Topo-II-DNA complex by stabilizing the short-term-topoisomerase-cleaved DNA. Topoisomerase II has two isotypes, namely Topo-IIα and Topo-IIβ. Topo-IIα serves cell cycle events such as DNA replication, DNA repair, and chromosome segregation; nonetheless, Topo-IIβ is mainly involved in transcription and developmental progression [254], [255], [256], [257]. Etoposide interferes with specific amino acid residues within the structure of both isotypes of topoisomerase to enter the Topo-II-DNA complex as a Topo-inhibitor [258, 259].

Moreover, it has been recently reported that etoposide has a high affinity toward histones and chromatin, which brings about other possible mechanisms of action [260].

Additionally, it has been reported that etoposide has other impacts on replication machinery as well. In other words, the distribution of replication proteins undergoes a progressive impact by etoposide during the S phase. This spreads the replication sites along the DNA strain, creating huge nuclear foci containing single-strand DNA binding protein RPA [261].

There is also another mechanism of action for etoposide, in which it can negatively impact the transcription of some specific genes. Takami, et al. have revealed that transcription factor E2F-4 binds to etoposide, leading to the downregulation and inhibition in the transcription of some genes mediated by the heterodimeric E2F-4/DP complexes in the nucleus [262].

Mechanisms of etoposide in cell death through signaling cascades

A 5′-phosphodiesterase has been identified that can excise the Topo-II-DNA bonds and repair topoisomerase-mediated DNA damages [263]. To illustrate this, TRAF and TNF receptor-associated protein (TTRAP), which has recently been renamed to TDP2 (tyrosine phosphodiesterase 2), is an Mg2+/Mn2+-dependent member of the phosphodiesterase group, enabling the religation of 5′-phosphate residues in the cleaved double helix strain. It has been demonstrated that TDP2 depletion leads to an increased cellular sensitivity to etoposide-induced double-strand breaks [264, 265], therefore, suggesting the possible contribution of TTRAP/TDP2 in etoposide chemotherapeutic response [263].

Besides, Adachi, et al. have demonstrated that the low-fidelity non-homologous end-joining (NHEJ) is the primary pathway in which cells repair the etoposide-induced DNA damage through [266]. In other words, cells repair DNA damages through various pathways based on the different cell cycle phases that the damage has occurred in [267]. For instance, G1 damages are repaired through NHEJ, whereas those in the S and G2 phases are repaired by homologous recombination (HR). It has been shown that dampening the NHEJ pathway via knocking out the DNA ligase IV and Ku70 subunit (LIG4−/−, Ku70−/−) in DT40 cells makes them severely sensitive to etoposide. However, RAD54−/− (a DNA repair and recombination protein) knocked-out, as HR-defective cells, are much less sensitive toward etoposide [253, 267]. Malik et al, have lent support to this idea by representing the importance of Ku70 and Ku80 existence for NHEJ and cell survival after etoposide treatment [268]. Furthermore, Chen and colleagues have also acknowledged that dysfunctional Ku70 decreases the DNA repair capacity in cells exposed to etoposide earlier. Moreover, they have mentioned the impact of Ku70 acetylation in histone deacetylase (HDAC) inhibitors. To be precise, Ku70 mutations are shown to mimic the acetylation of some lysine residues, rendering Ku70 unable to bind DNA, which ultimately makes prostate cancer cells more sensitive to etoposide [269].

In addition, etoposide stimulates the activation of ataxia-telangiectasia mutated (ATM) and its downstream kinase, namely checkpoint kinase 2 (Chk2). The activation of ATM kinase consequently causes the construction of the Mre11/Rad50/NSB1 (MRN) complex and ionizing radiation-induced foci (IRIF), which altogether signify severe DNA damages [270]. ATM mutations lead to etoposide hypersensitivity. Besides, loss of G2/M checkpoint in AT cells allows mitosis to occur, although DNA breaks and chromosomal abnormalities have happened following etoposide administration [271, 272].

Furthermore, etoposide administration increases the interaction between transcription factor E2F1 and bridging integrator-1 (BIN1) promoter. It has been shown that dampening BIN1 using an antisense-RNA diminishes the cell-death rate mediated through the interaction of E2F1 and etoposide [273].

Moreover, the alternative splicing mechanism is also affected by etoposide. Etoposide treatment stimulates the dephosphorylation of serine/arginine-rich splicing factor-1 (SRSF1), a splicing element, which monitors and regulates whether the anti- or pro-apoptotic genes such as Ron oncogene and tumor suppressor BIN1 should be alternatively spliced. Moreover, alteration in the phosphorylation of SRSF1 subsequent to DNA damages alters the alternative splicing pattern in caspase 9. This signifies the active influence of etoposide on alternative splicing mechanisms and its correlation with the modulation of apoptosis in target cells [274], [275], [276], [277].

Various studies have supported that high concentrations of etoposide stimulate cytochrome C-caspase-9-mediated pathways. Activating Apaf1, it constructs an apoptosome complex capable of cleaving pro-caspase 9 into the active caspase 9, which ultimately leads to cellular apoptosis through the cleavage and activation of caspase 3 and 7 [278].

Moreover, new pieces of evidence signify the correlation of Fas/FasL in etoposide-related apoptosis [279]. To illustrate this, etoposide administration induces the interaction between FasL and its receptor, namely FasR, resulting in the construction of a death-inducing signaling complex (DISK), where FADD activates pro-caspase 8 into functional caspase 8, which further activates caspase 7.

Etoposide also triggers DNA damage responses (DDRs) in tumor cells and resting cells like T lymphocytes. ATM is phosphorylated in this regard, which results in the phosphorylation of H2AX and p53. Ultimately, this leads to the activation of a pro-apoptotic protein, namely p53-upregulated modulator of apoptosis (PUMA). It has been reported that inhibition of ATM functions in dormant cells, using Ku55933, dampens DDR and apoptosis by diminishing the expression level of PUMA and decreasing the activation of caspases, whereas performing the same procedure in tumor cells increases the cytotoxic impact of etoposide [280].

It should be noted that p53 plays a distinct role in etoposide-induced cell apoptosis. To be precise, Nemo-like serine/threonine kinase (NLK) is essential for the activation of p53 following the DDR caused by etoposide administration [281]. Although the exact mechanism is still elusive, NLK is upregulated subsequent to DDR. NLK stabilizes p53 through the inhibition of protein degradation mediated by Double minute 2 (MDM2) and ubiquitin-proteasome system [281, 282]. It has also been suggested that NLK enhances the cellular sensitivity to etoposide through dampening some transcriptional factors such as NF-KB [283, 284].

Furthermore, studies have shown that etoposide treatment induces some stress pathways, including JNK and MAPK. These pathways are supposed to have double-edged functions in anti- and pro-apoptotic pathways. WWOX is an oxidoreductase and a tumor suppressor that acts through stabilizing p53, resulting in cell death. It has been reported that WWOX is increased following the etoposide-induced DDR [285].

Mechanisms of etoposide resistance through signaling cascades and how to be tackled

Likewise other chemotherapeutic regimens, chemoresistance is one of the complications during etoposide treatment. MDM2 gene, which encodes ubiquitin ligase involved in protein degradation, contributes to etoposide resistance. In other words, a single-nucleotide polymorphism, known as SNP 309 T/G, that is located within the MDM2 promoter causes its upregulation, which results in a diminished response to numerous DNA-damaging drugs such as etoposide. It has been reported that homozygote-SNP309 cell lines are etoposide resistant. MDM2 in these cell lines represented an increased affinity toward binding and degrading Topo-II. Furthermore, by the time MDM2 was knocked-down by RNAi, Topo-II was stabilized, and etoposide resistance was decreased [286].

Furthermore, Zhang, et al. have reported that a proteasome-mediated degradation procedure is involved in etoposide resistance. Using knock-out-Topo-IIβ mice, and knock-down-Topo-IIβ, they have represented that etoposide-mediated DNA damages were attenuated by MG132 dampening proteasome [287].

Shreds of evidence have indicated the correlation of MAGE family proteins with regulating cell survival. It has been reported that MageA2 recruits HDACs to the P53 transcription sites, which stimulate histone hypoacetylation, therefore, suppressing its activity in melanoma cells. Hence, melanoma cells expressing MAGE-A genes are resistant to etoposide-induced apoptosis [253, 288].

Shreds of evidence suggest that etoposide is also involved in autophagy. Katayama et al. have represented the feasibility of autophagy-dependent ATP production in glioma cells. Surprisingly, not only it does not kill them, but also it favors their survival; therefore, it might contribute to etoposide resistance [289]. It has been indicated that this crisis is irreversible even when cells experience glucose starvation. Nonetheless, it can be inhibited by the preincubation with autophagy inhibitor 3-methyladenine (3-MA), mitochondrial inhibitor oligomycin, as well as through the siRNA-mediated downregulation of beclin 1, which reduce the autophagy-induced ATP level and lead cells to non-apoptotic cell death [290].

Besides, Alpsoy et al. [291] have demonstrated that MRP1 (ABCC1 transporter) mainly contributes to etoposide efflux and chemoresistance. Besides, they have reported the mismatch repair mechanisms to be involved in MCF7-etoposide resistant cells. In other words, they argued that MLH1 and MSH2, which are involved in DNA damage repair, have significantly downregulated in etoposide-resistant cells. Moreover, they have investigated the expression rate of topoisomerase IIβ-binding protein 1 (TOPBP1) and E3 ubiquitin-protein ligase EDD. TOPBP1 functions in DNA replication, proliferation, and DNA damage response signaling pathways. It causes cell cycle arrest and ignites the apoptosis pathway [292, 293]. EDD is a tumor suppressor protein and one of its interplay partners contributing to checkpoint responses and DNA damage signaling [294]. Alpsoy and colleagues found a decrease in the expression of both TOPBP1 and EDD. Altogether, it addresses resistance in etoposide-treated cells [291]. Fig. 6.

Fig. 6.

Fig. 6

Etoposide, an anti-topoisomerase II agent, poisons the TopoIIcc and prevents the religation of DNA strands. Following the persistent DNA damage caused by etoposide, ATM/Chk2 are recruited that by the assist of Mre11/Rad50/NSB1 cause the S phase arrest which leads to apoptosis. Also through alternative signaling, namely NLK, ATM is phosphorylated which ultimately phosphorylates and activates p53 leading to p53. Nonetheless, SNP mutations like SNP 309 T/G forms a different subtype of MDM2 which inactivates p53 and accounts for etoposide resistance in some cases. Cyto C/Apaf-1/Cas-9 as well as Fas/FasL/Cas-8,10 are among other possible mechanisms described to play role in etoposide-cell death mechanisms.

The role of mTOR/Akt/PI3K signaling pathways in human cancer

The mammalian target of rapamycin (mTOR) and phosphatidylinositol-3-kinase/Akt signaling cascades act as a crucial double-edged sword within intracellular signaling pathways under both physiological and pathological conditions, including cell survival, cell proliferation, apoptosis, and invasion. These pathways are neatly interconnected with numerous cell signaling pathways such as HIF and NF-KB, to say the least [295].

PI3K/Akt signaling cascade is considered the essential modulator favoring cell survival under stressful circumstances [296]. To illustrate this, PI3K is formed of a lipid kinase family that can phosphorylate inositol rings 3′-OH in inositol phospholipids [297]. Class-I PI3K is formed of catalytic subunits (CAT), namely p110, and adaptor-regulatory domain, known as p85. In this regard, the signaling cascade is commenced by the stimulation and phosphorylation of tyrosine kinases in growth factor receptors. Thereafter, PI3K is recruited to the cell membrane, where it binds to tyrosine residues in receptors via SH2 domains within the adaptor subunit. This brings about the stimulation of CAT residue, which further leads to the generation of phosphatidylinositol-3,4,5-triphosphate (PI3,4,5-P3), namely the second messenger. Then, PI3,4,5-P3 recruits other signaling proteins including, protein serine/threonine kinase-3′-phosphoinositide-dependent kinase 1 (PDK1) and Akt/protein kinase B (PKB), which then regulate cell survival and cell cycle progression [295, 297, 298].

Regarding cell survival, Akt/PKB inactivates pro-caspase9 and Bad, besides it dampens related transcription factors that trigger the expression of apoptotic elements such as Fas/FasL [299, 300]. Moreover, Akt/PKB has been introduced to perform as a resistant-inducing agent through modulating TNF-related apoptosis-induced ligand (TRAIL)/APO-2L [301]. It also activates the survival signaling element in NF-KB signaling, namely IKB kinase (IKK) [302].

Regarding the cell cycle progression, numerous protein synthesis procedures, glycogen metabolism pathways, and cell cycle regulators have close interplay with Akt downstream signaling pathways, including mTOR, glycogen synthase kinase-3 (GSK-3), insulin receptor substrate-1 (IRS-1), p21, p27, and Raf-1 [298, 303].

On the other hand, Akt kinases themselves are members of the AGC family that correlate with AMP/GMP kinases and protein kinase C (PKC). To be precise, Akt kinases are constructed of three main domains, including a PH domain (N-terminal), CAT domain (central domain), and a regulatory hydrophobic residue (C-terminal). CAT domain is highly conserved among all Akt family and is highly related to PKC, PKA, and SGK [304].

MTOR is exceptionally considered as the most vital element within the realm of cell signaling, because it serves different aspects of cell fate. In other words, mTOR acts through its two main complexes. The mTOR complex-1 (mTORC1) is constructed of mTOR, Raptor, mLST8, and PRAS40. It is highly sensitive to rapamycin and generally stimulates S6K and dampens 4E-BP1, which leads to the translation of proteins and cell growth [305]. The mTORC2, which is less sensitive to rapamycin, on the other hand, is formed of mTOR, Rictor, Sin 1, and mLST8. The mTORC2 is appreciated for its stimulatory impact on Akt, which promotes cell survival and proliferation. The canonical pathway in mTOR signaling is believed to act through PI3K/Akt and Ras/MEK/ERK [306].

Different upstream signaling pathways regulate mTOR in normal cells [305]. These regulators are subcategorized into positive and negative modulators. Regarding positive regulators, for instance, insulin growth factor-1 (IGF-1) and its receptor, human epidermal growth factor receptors (HER family) and their ligands, and vascular endothelial growth factor receptors (VEGFRs) and their ligands, channel signals through PI3K/Akt. Nonetheless, the negative regulators include an energy-sensing element, namely LKB1, phosphatase and tensin homolog (PTEN), tuberous sclerosis complex-1 (TSC-1), and -2 (TSC-2). In other words, PTEN performs its negative impacts through PI3K/Akt, and TCS-2 also releases its inhibitory effects on mTOR by the time it is phosphorylated by Akt [295, 307].

PTEN acts as the regulator of PI3K/Akt/mTOR signaling cascades [295]. PTEN has impacts on both lipids and proteins. It is a tumor-suppressor agent that assists in inhibiting cell growth and leads target cells toward apoptosis [308]. However, PTEN is mutated in numerous cancers; for instance, PTEN is mutated in Cowden's syndrome, leaving the patient at high risk of being diagnosed with multiple cancers simultaneously [309]. PTEN indeed acts as the negative regulator of PI3K/Akt signaling; nevertheless, by losing the PTEN activity, cells experience non-stop and uncontrolled activation of PI3K/Akt signaling, which leads to excessive and unnecessary cell survival and proliferation [295].

As the mTORC1 signaling is concerned, it is activated by external and internal stimuli, including nutrients, hormones, and growth factors, as well as intracellular accumulation of DNA damages, amino acids, glucose, ATP, and oxygen. For instance, subsequent to DNA damages, mTORC1 is dampened through p53, which induces the activation of TSC-2 [310]. Besides, AMPK is stimulated following a period of energy exhaustion, leading to the formation of the TSC1/2 complex that ultimately inhibits mTORC1 through the phosphorylation of Raptor [311].

However, by the time mTORC1 is activated, as it was mentioned earlier, it transfers signals to S6K and 4E-BP1 [312]. Following the involvement of S6K and 4E-BP1, eukaryotic initiation factor (eIF)-4E and eukaryotic initiation factor-3 (eIF-3) are recruited, which promotes ribosomal biogenesis [313]. To illustrate this, the low activity of mTORC1 phosphorylates and stimulates 4E-BP1, which ignites protein translation. Besides, S6K phosphorylates elF-4B and S6 ribosomal protein (S6RP), which allow the translation process to continue and commence the elongation phase [314], [315], [316]. Thereafter, elF-4E constructs the elF-4F complex that stimulates protein translation, a stage that is vital, providing that the G1/S transition is concerned [312]. Therefore, it is conjectured that the activation of mTORC1 favors the translation of oncogenic proteins, which give rise to cell invasion, metastasis, and proliferation [317]. Moreover, mTORC1 activation also modulates other elements, including HIF-α, Protein phosphatase 2A (PP2A), and STAT3, which assists the biogeneration of essential lipids, proteins, and nucleotides in malignant cells, facilitating their survival, progression, and invasion [312].

On the other hand, the mechanisms in which mTORC2 performs its regulatory impacts remain partially elusive, probably owing to the fact that distinguishing between mTORC1 and mTORC2 is quite elaborate [318]. The mTORC2 is modulated by mTORC1 and PI3K/Akt signaling. In this regard, PI3K stimulates mTORC2 to bind to ribosomes under both physiological and pathological circumstances [319]. Akt is highly correlated with mTORC2, and studies have lent support to its increased expression level in various malignancies, where it collects signals from PI3K/mTORC2 and PI3K/PDK1, which induces cell survival and proliferation. Moreover, Akt impacts on mTORC1 in other complicated approaches [320].

It is well appreciated that mTORC1 negatively modulates mTORC2. To be precise, S6K1 stimulates the degradation of insulin receptor substrate-1 (IRS-1), which results in the inhibition of mTORC2 and PI3K/Akt signaling. Besides, Grb-10 is also recruited by mTORC1, which blocks mTORC2 [321], [322], [323].

Regarding the downstream pathways involved in mTORC2-related signaling cascades, glucocorticoid kinase (SGK) and protein kinase C (PKC) are considered as two key phosphorylation substrates, which lead cell survival machinery under hypoxia and malnutrition or even during PI3K blockage. It has been demonstrated that mTORC2 modulates cytoskeletal reorganizations and cell movements, necessary for cell invasion and tumorigenesis, by exploiting various PKC members [312].

To summarize, therefore, since mTOR regulates the production of numerous vital proteins and is involved in various signaling, including cell survival, cell proliferation, glucose metabolism, cell cycle, as well as protein and lipid synthesis, it is closely related to different tumors pathology. Moreover, mTOR plays a role in chemoresistance. In other words, some elements including, cyclin D1 [324], and HIF [325], are among essential proteins that interact with the aforementioned signaling pathways and give rise to the survival and expansion of the tumors by allowing the progression of the cell cycle and provoking the expression of angiogenic factors like VEGF, respectively. More importantly, mTOR and its upstream and downstream signaling pathways have been reported to be directly and indirectly the reason for the onset of different malignancies [312]. Fig. 7.

Fig. 7.

Fig. 7

MTOR is considered as the vital signaling cascade and as the crossroads within the realm of intracellular signaling owing to the fact that it serves different cellular functions under both physiological and pathological circumstances. It mainly acts through two complexes, namely mTORC1 and mTORC2. These complexes have a close interaction with upstream signaling cascades including, PI3K/Akt, Ras/ERK, and AMPK. The phosphorylation of Akt by the upstream signaling dampens the inhibitory function of TCS 2 which allows mTORC1 to perform its activities. The mTORC1 further regulates $E-BP1, S6K1, SREBP, STAT3, HIF-α, and PP2A in order to bring about lipid, protein, and nucleotide synthesis, as well as cell cycle progression and angiogenesis. The mTORC2, on the other hand, is modulated by mTORC1 through Grb-10 (in a feedback loop) and IRS-1 which ultimately favors cytoskeletal reorganization, cell movements, cell migration, and cell proliferation.

Recent outcomes in the application of combination chemotherapy

As discussed earlier, single-chemotherapy applications have been limited by chemoresistance and tumor relapse in the majority of cases. Every chemotherapeutic agent targets various intracellular-signaling cascades and different phases in the cell cycle. Therefore, the application of combination chemotherapy seems to boost chemotherapy responses; nonetheless, antagonistic impacts, variations in pharmacokinetics, and different drug distribution patterns have limited this application [326, 327].

For instance, gemcitabine was administered in combination with cisplatin on numerous solid tumors. In-vitro studies in this regard demonstrated synergetic impacts between gemcitabine and cisplatin, which seems to be due to an increase in platinum-DNA adducts. The synergism was also witnessed in gemcitabine-etoposide combined therapy in ovarian and lung cancer [173]. Moreover, the impact of gemcitabine-nab-paclitaxel, nanoliposomal irinotecan-5-fluorouracil-leucovorin, and gemcitabine-capecitabine was analyzed on elderly people diagnosed with pancreatic cancer. The results showed beneficiary outcomes for these combination therapies [328].

Furthermore, in a retrospective study, 113 advanced gastric cancer patients were subjected to combination chemotherapy of paclitaxel, 5-fluorouracil, and leucovorin (TFL) as their first-line treatment protocol, which represented 43.4% overall response with moderate toxicity. Therefore, Wan-Cai et al. have claimed TFL as an active and safe approach for these patients [329]. However, Choi, et al. have mentioned that the results of combination chemotherapy in patients with recurrent or primary metastatic gastric are somewhat conflicting. They further discussed, although combination chemotherapy is recommended for these patients, single chemotherapy administration can be considered as the logical approach for certain cases, especially elderly patients [330].

In another retrospective study, patients with advanced gastric cancer were subjected to a combination of paclitaxel and oxaliplatin as their first-line treatment every 14 days. The disease control was about 80%, making this combination therapy a new, effective, and safe regimen for these patients [331]. Besides, since peritoneal metastasis in advanced or recurrent gastric cancer is the common cause of death among these patients, better chemotherapy approaches are essential; however, the current treatments maintained defective. Ohnuma et al. have suggested the prescription of docetaxel, cisplatin, and S-1 as feasible and effective combination chemotherapy for these patients [332].

Pancreatic cancer has been treated with gemcitabine for many years. Nonetheless, chemoresistance against this regimen has significantly limited its application. Li et al. [333] have co-administered gemcitabine with valproic acid (VPA) and reported their synergistic impact in a dose-dependent manner. It has been reported that this combination therapy, with high-dose VPA, could successfully increase the sensitivity of pancreatic cancer cells to gemcitabine. However, they further indicated that low-dose VPA combined with gemcitabine promoted the migration and invasion potency in pancreatic cancer cells through increasing the level of ROS, as well as the activation of Akt, STAT3, and Bmi1. Whereas, administration of high-dose VPA-gemcitabine enhanced excessive ROS retention, which stimulated the activation of p38 resulting in the inactivation of STAT3 and Bmi1.

Besides, advanced non-small cell lung cancer patients were subjected to bevacizumab combined with gemcitabine-cisplatin (GC) combination chemotherapy in a study. Dividing the cases into two groups, one received GC as the control group and the other one received GC-bevacizumab as the observation group. The results were somewhat promising. Duan et al. reported that the efficiency rate and disease control were approximately 41% and 71%, respectively, in the control group, and 71% and 90%, respectively, in the observation group. Moreover, the level of tumor markers, namely CEA and CYFRA21-1, as well as the concentration of serum vascular endothelial growth factor (VEGF) were significantly lower [334].

Combination chemotherapy has represented promising outcomes in multidrug-resistant cancers as well. To illustrate this, Polymeric nanogels were used to encapsulate cisplatin and doxorubicin. They could successfully deliver more concentration of drug into MCF7/ADR cells and killed more cancer cells, which altogether lends support to its synergetic impacts [335]. Furthermore, multidrug delivery was also performed using magnetic nano-carriers. Rahimi et al. have developed dendritic chitosan grafted mPEG coated magnetic nanoparticles for delivering doxorubicin and methotrexate into MCF7 cells. They demonstrated that several peripheral bloodstream proteins could bind to these nano-carriers and improve drug delivery. Besides, they confirmed that the anticancer activity of combined drugs was significantly more in comparison to free drugs [336].

In another study, the combination of paclitaxel and doxorubicin were co-encapsulated using recombinant high-density lipoprotein nanoparticles (rHDL). The results showed an enhanced intracellular accumulation of drugs, and improved cytotoxicity properties [326]. Also, Zhu et al. have utilized bilayered folate (FA) receptor-targeted polymersomes to encapsulate paclitaxel and doxorubicin. The results demonstrated more sustainability in drug release and dampened cell growth more sufficiently compared to free drug cocktail [327]. This technique has also been experimented on breast cancers that metastasize into the brain. This subtype cannot be cured with current chemotherapies due to poor drug delivery to the brain. Therefore, Bao et al. [337] have encapsulated oleanolic acid (OA), an excellent anti-tumor agent with penetration potency to the brain, and paclitaxel (PTX) in nanoparticles. They reported the synergistic efficacy of PTX-OA-NPs combination chemotherapy to effectively dampen breast and metastasized brain cancer progression.

Polymers seem to be interesting for encapsulation drug delivery approaches as well. An armed amphiphilic star copolymer was utilized to encapsulate doxorubicin and avasimibe. The cytotoxicity assays showed considering synergistic anticancer activity. Moreover, co-administration of avasimibe could reduce the required dose of doxorubicin, which reduced its side effects, suggesting a promising combination therapy against K562 and HeLa cells [338].

There are crucial risk factors in approaching any combination chemotherapy protocol that needs to be addressed in advance. For instance, although the combined gemcitabine-cisplatin (GC) chemotherapy has become standard for patients with urothelial cancer (UC), the subsequent hematological toxicity is still one major limitation in this regard. The results of a retrospective study indicated grade four neutropenia in roughly 48% of patients and grade 3 thrombocytopenia in 21% of all cases. Although some believed that age is the risk factor, they further discussed neutrophil count, platelet count, and K level as prominent risk factors among UC patients receiving GC combined therapy-induced hematological toxicity [339]. Moreover, in a study by Toffalorio et al. [340] they discussed cN-II expression level correlating with gemcitabine-platinum combination chemotherapy fate in patients suffering from non-small-cell lung cancer.

Besides, as discussed earlier, multiple-drug resistance (MDR) is a challenging event in chemotherapies. Therefore, it is important to tackle this crisis by increasing the intracellular accumulation of chemotherapeutic drugs. Regarding this issue, celastrol (CST) and doxorubicin (DOX) were co-encapsulated into carrier-free nanoparticles (CST/DOX NPs) in order to overcome DOX-resistance. Hopefully, this formulation has increased the water solubility and decreased the required DX dosage. Therefore, it could significantly enhance the drug concentration within the target cells, activate heat-shock factor-1 (HSF-1), dampen NF-KB to suppress P-gp expression, which altogether induced apoptosis and autophagy through ROS/JNK signaling cascade in DOX-resistant cells [341].

Paclitaxel has been prescribed for numerous cancers, yet chemoresistance and side effects such as neuro-, hepato-, and cardio-toxicity have limited its application. Ashrafizadeh et al. [342] have combined curcumin as an anti-tumor and anti-inflammation agent with paclitaxel and reported this combination chemotherapy application successful in enhancing the anti-tumor potency and decreasing the primary side effects of single-paclitaxel administration.

The results of combination chemotherapy applications are not just limited to patients with a promising prognosis. Interestingly, a 79-year-old man diagnosed with hilar cholangiocarcinoma which had metastasized into his lymph nodes, was subjected to neoadjuvant gemcitabine/cisplatin/S-1 combination in a case-report study. Although they could not save the liver due to massive impairments, biopsy and cytology results showed no local cancer cells after resection [343]. However, in another case study, a 61-year-old man diagnosed with stage IV papillary and anaplastic thyroid cancer was subjected to the combination of dabrafenib and trametinib, which resulted in life-threatening arrhythmia. The patient's condition was severe enough was undergone plasmapheresis to remove the chemotherapeutic drugs [344].

The list of combination chemotherapy applications is not limited to the regimens mentioned above. These results reinforce the fact that this approach can be a potential approach against numerous cancers. Although it might leave disappointing outcomes in some cases, there are beneficiary responses as well. However, there is still room for understanding the exact mechanisms of action underlying the function of combined drugs in order to develop combinations with fewer side effects and favorable responses.

The contribution of precision medicine in chemotherapy

Cancers might represent pathologically identical, yet even the same cancer subtypes might respond to a particular chemotherapeutic drug differently [345]. Although current chemotherapy selections are mostly based on the cellular and genetic mechanisms underlying chemoresistance, various tumor-specific and patient-specific criteria contribute to chemotherapy response. Therefore, in addition to an optimized drug, specific measurements are required to select chemotherapeutic drugs and their administration schedule based on the specialized pharmacokinetics and pharmacodynamics for each patient to achieve the best possible response [8].

Although new chemotherapeutic drugs have been developed and combination chemotherapies are becoming commonplace, the responses are not entirely satisfactory in some cases. This is due to the heterogeneity of causes underlying the onset of each cancer case in different patients. However, current signs of progress regarding genomics and next-generation sequencing (NGS) can help to identify genetic variations in different patients to narrow the development of drugs into precise and single-patient designed approaches for their specific biological, molecular, and cellular variations rather than stochastic chemotherapy applications [346, 347]. Furthermore, the diversity of natural compounds has contributed to developing new chemotherapeutic agents for about a half-century. Using the combined chemistry and highthrouput technology will help predict the behavior of each molecule and design personalized chemotherapy approaches and overcome chemoresistance based on the specifications in each case [348].

Mathematical models and omics technologies play key roles in this regard. For instance, shreds of evidence have been provided by McKenna, et al. that mathematical models can predict, specify, and improve chemotherapies in the realm of precision medicine when applying to breast cancer [8].

Besides, oral squamous cell carcinoma is a complex malignancy representing tumor heterogeneity and plasticity in different cases. Omics technologies are being used to address these variations. Omics are highthrouput technologies capable of screening different target molecules qualitatively and quantitatively. In other words, genomics, transcriptomics, proteomics, and metabolomics are used to find personalized biomarkers in tumor biopsies, circulating tumor cells, or body fluids like saliva [349]. Dissimilar to conventional chemotherapies that merely target one signaling cascade, using omics, one can acquire a thorough and unbiased perspective through specific genomics and proteomics, and perhaps target cells and molecules more specifically [350].

Intriguingly, Cammarota et al. have discussed the application of gut microbiota, big-data mining, and machine learning in the context of precision medicine as well. The gut microbiome is identical to fingerprints for individuals and has complicated cancer therapies by stimulating tumor progression, tolerability, and modulating immune-system responses to cancers. Using omics, we can collect beneficiary data, and by data mining, one can design more precise approaches. Moreover, machine learning is quickly developing, and it is becoming one crucial element in precision medicine. This technology goes beyond merely genomics and proteomics, yet it collects all data, and by designing different algorithms, it can assist in precision medicine [351]. For instance, Lee et al. have represented a promising approach in identifying molecular characteristics to develop more precise targeted therapies for acute myeloid leukemia (AML). To illustrate this, collecting genome-wide gene expression data, in-vitro investigation of the drug sensitivity of 160 chemotherapeutic drugs, as well as multi-omic computational algorithms, they were able to find SMARCA4 as a marker and driver of sensitivity to topoisomerase II inhibitors, mitoxantrone, and etoposide in AML [345].

Finding predictive and specific biomarkers have impacted the chemotherapy approaches for ovarian cancer patients as well. There are specific molecular characteristics even within a particular histological subtype of ovarian cancer. Therefore, treating ovarian cancer has altered from one-size-fits-all to a precise methodology, including surgery, chemotherapy, and targeted therapy. Moreover, the progressions in NGS hope to find various distinct biomarkers and variations in genomics, which altogether assist in developing new approaches for each ovarian cancer patient [352].

Immunotherapy is another approach that has been accepted to benefit cancer cases; nonetheless, one major disadvantage is that they give rise to unnecessary inflammatory responses and autoimmune diseases by the upregulation of the immune system [353]. Having said that, their applications and clinical trials are expanding due to their increased survival rate in advanced cancers [354]. Tumor mutational burden (TMB) is a robust approach for predicting specific biomarkers in the context of personalized medicine. Non-small-cell lung cancer is an example regarding the application of precision medicine, that numerous FDA-approved targeted therapies, such as immune-checkpoint inhibitors, have been developed based on identified biomarkers, like EGFR, ALK, ROS1, and BRAF. TMB can be assessed using NGS and whole-exome sequencing (WES), although NGS is more affordable globally. Shreds of evidence have lent support to the application of TMB in the selection of possible personalized immune-checkpoint-inhibitor treatments [355].

Moreover, the treatment of head and neck squamous cell carcinoma (HNSCC) is a sophisticated procedure due to loads of mutations without any specifications. Although surgery and radiation are the established therapy against HNSCC, administration of checkpoint inhibitors, targeted immunotherapy, including anti-PD-L1 and anti-CTLA-4, as well as precision medicine, can shed light on developing novel and more effective approaches against HNSCC [356].

Promising chemotherapeutic drugs, immunotherapies, and related clinical trials

Currently, numerous chemotherapeutic and immunotherapeutic candidates are under clinical trials. Numerous reports have indicated their responses, and some provided a comprehensive comparison between the application of them either alone or in combination [353]. In this regard, we have summarized some of the recent clinical trials regarding the comparison between immunotherapies and chemotherapies (Table 2.A) [353], and combination therapies (Table 2.B), as well as their target cancer and study phase. Table 2.

Table 2.

Clinical trials for chemotherapies, immunotherapies, and combination chemotherapies.

A. Immunotherapies Vs. Chemotherapies in recent clinical trials on solid tumors
Clinical trial name Target cancer Phase of study Intervention and dose Control treatment Ref
No-name Melanoma or non-small-cell lung cancer with untreated brain metastases II Pembrolizumab 10 mg/kg Investigator's choice Goldberg, et al, 2016 [376]
No-name Melanoma with active brain metastases II Pembrolizumab 10 mg/kg Investigator's choice Kluger, et al, 2019 [381]
GETUG-AFU 26 NIVOREN Stage IV-metastatic brain form clear cell renal cell carcinoma II Nivolumab 3 mg/kg VEGFR-directed therapy Flippot, et al,2019 [374]
JAVELIN Solid Tumor Stage IIIC or IV unresectable melanome Ib Avelumab 10 mg/kg Investigator's choice Keilholz, et al, 2019 [380]
Javelin Gastric 300 Gastric or gastroesophageal junction adenocarcinoma III Avelumab 10 mg/kg Investigator's choice Bang, et al, 2018 [362]
Javelin Lung 200 Non-small-cell lung cancer III Avelumab 10 mg/kg Docetaxel Barlesi, et al, 2018 [363]
KEYNOTE-045 Urothelial carcinoma III Pembrolizumab 200 mg Investigator's choice Bellmunt, et al, 2017 [364]
CheckMate 057 Non-small-cell lung cancer III Nivolumab 3 mg/kg Docetaxel Borghaei, et al, 2015 [366]
CheckMate 015 Small-cell lung cancer III Nivolumab 3 mg/kg Docetaxel Brahmer, et al, 2015 [367]
CheckMate 026 Non-small-cell lung cancer or small-cell lung cancer III Nivolumab 3 mg/kg Platinum-based Carbone, et al, 2017 [368]
KEYNOTE-040 Head and neck squamous cell carcinoma III Pembrolizumab 200 mg Investigator's choice Cohen, et al, 2019 [369]
POPLAR Non-small-cell lung cancer II Atezolizumab 1200 mg Docetaxel Fehrenbacher, et al, 2016 [371]
CheckMate 141 Head and neck squamous cell carcinoma III Nivolumab 3 mg/kg Investigator's choice Ferris, et al, 2016 [372]
CheckMate 227 Non-small-cell lung cancer III Nivolumab 3 mg/kg Platinum-doublet therapy Hellmann, et al, 2018 [378]
KEYNOTE-010 Non-small-cell lung cancer II/III Pembrolizumab 2 mg/kg Docetaxel Herbst, et al, 2016 [379]
KEYNOTE-042 Non-small-cell lung cancer III Pembrolizumab 200 mg Investigator's choice Mok, et al, 2019 [386]
IMvigor 211 Urothelial carcinoma III Atezolizumab 1200 mg Investigator's choice Powles, et al, 2018 [389]
IFCT-1603 Small-cell lung cancer II Atezolizumab 1200 mg Investigator's choice Pujol, et al, 2019 [390]
KEYNOTE-024 Non-small-cell lung cancer III Pembrolizumab 200 mg Investigator's choice Reck, et al, 2016 [391]
No-name Melanoma III Tremelimumab 15 mg/kg Investigator's choice Ribas, et al, 2013 [392]
KEYNOTE-002 Melanoma II Pembrolizumab 2 mg/kg Investigator's choice Ribas, et al, 2015 [393]
OAK Non-small-cell lung cancer III Atezolizumab 1200 mg Investigator's choice Rittmeyer, et al, 2017 [394]
CheckMate 066 Melanoma III Nivolumab 3 mg/kg Dacarbazine Robert, et al, 2015 [395]
KEYNOTE-061 Gastric or gastroesophageal junction adenocarcinoma III Pembrolizumab 200 mg Paclitaxel Shitara, et al, 2018 [396]
CheckMate 037 Melanoma III Nivolumab 3 mg/kg Investigator's choice Weber, et al, 2015 [400]
CheckMate 078 Non-small-cell lung cancer III Nivolumab 3 mg/kg Docetaxel Wu, et al, 2019 [401]

B. Combination therapies in recent clinical trials
Clinical trial name Target cancer Phase of study Intervention and dose Control treatment Ref

JASPAC 04 Resectable pancreatic ductal adenocarcinoma II IV Gemcitabine 1000 mg/m2
+ Oral S-1 50 mg
Chemoradiotherapy Toyama, et al, 2020 [398]
SOBIC Metastatic colorectal cancer II FL: S-1+ Oxaliplatin+ Bevacizumab (SOX+Bmab)
SL: S-1+ Irinotecan+ Cetuximab (IRIS+Cmab)
None Nakamoto, et al, 2020 [388]
SOPP Metastatic or recurrent gastric cancer III S-1 80 mg/m2 + Oxaliplatin 130 mg/m2 (SOX) S-1 80 mg/m2 + Cisplatin 60 mg/m2 (SP) Lee, et al,
2021 [383]
No-name Platinum-resistant recurrent epithelial ovarian, primary peritoneal, or fallopian tube cancer II IV Bevacizumab 15 mg/kg + IV Gemcitabine 1000 mg/m2 Platinum-based Nagao, et al, 2020 [387]
NRG- GI004/SWOG-S1610 Deficient DNA mismatch repair (dMMR) colorectal cancer III mFOLFOX6/ Bevacizumab, Atezolizumab monotherapy, or mFOLFOX6/Bevacizumab + Atezolizumab Investigator's choice Lee, et al.,
2019 [382]
GOG-0213 Platinum-sensitive, recurrent ovarian cancer III Platinum-based combination chemotherapy (with or without Bevacizumab) Investigator's choice Coleman, et al., 2018 [370]
DESMOPAZ Progressive desmoid tumor II Oral Pazopanib 800 mg per day
or
IV combined Methotrexate-Vinblastine (5 mg/m2, 30 mg/m2)
Investigator's choice Toulmonde, et al., 2019 [397]
PICCA Advanced biliary tract cancer II cisplatin 25 mg/m2 and gemcitabine 1000 mg/m2 with or without Panitumumab 9 mg/kg Investigator's choice Vogel, et al, 2018 [399]
No-name Advanced bone and soft tissue sarcomas II Gemcitabine 900 mg/m2 + Docetaxel 70 mg/m2 None Hara, et al., 2019 [377]
POUT Upper tract urothelial carcinoma III IV cisplatin or carboplatin 70 mg/m2 + IV gemcitabine 1000 mg/m2 Investigator's choice Birtle, et al., 2020 [365]
CheckMate 016 Advanced or metastatic renal cell carcinoma I Nivolumab + Sunitinib (50 mg/day)
or
Nivolumab + Pazopanib (800 mg/day)
None Amin, et al., 2018 [360]
TBCRC 022 HER-2-positive breast cancer with brain metastases II Oral Neratinib (240 mg/day) + Capecitabine (750 mg/m2) Lapatinib-naÏve or Lapatinib-treated Freedman, et al., 2019 [375]
LANDSCAPE HER-2-positive breast cancer with brain metastases (not treated with WBRT) II Oral Lapatinib (1250 mg) + Capecitabine (2000 mg/m2) None Bachelot, et al., 2013 [361]
No-name HER-2-positive relapsed or metastatic breast cancer treated with taxanes, anthracyclines, trastuzumab II Capecitabine (1000 mg/m2) + Pyrotinib (400 mg) or Lapatinib (1250 mg) None Ma, et al.,
2019 [385]
BrighTNess Stage-II or III-triple-negative breast cancer III Segment 1 regimen: IV Paclitaxel (80 mg/m2) + IV Carboplatin (6 mg/ml per min) + oral Veliparib (go mg)
Segment 2 regimen: Doxorubicin + Cyclophosphamide
paclitaxel plus carboplatin plus veliparib placebo, or paclitaxel plus carboplatin placebo plus veliparib placebo Loibl, et al., 2018 [384]
No-name ER-positive, HER-2-negative advanced breast cancer III Palbociclib (125 mg) + Letrozole (2.5 mg) Placebo + Letrozole Finn, et al.,
2016 [373]

IV: intravenous, FL: first-line, SL: second line, HER-2: human epidermal growth factor-2, WBRT: whole brain radiotherapy, ER: estrogen receptor

Conclusion

Although numerous chemotherapeutic drugs have been introduced and increased the life expectancy of cancer patients, chemoresistance and tumor relapse have limited their application. Mutations seem to be one of the game-changer criteria in this regard. Stochastic mutations allow tumor cells to become resistant and continue growing, expanding, and invading, despite the chemotherapy in some cases [357]. Therefore, malignant cells have found their ways to proliferate and adapt themselves in most cases [358].

Despite all the prosperities in chemotherapeutic drugs, it is assumed that tumor cells are one step ahead. Tumors either through acquired mutations or exploiting the vital resources essential for our normal cells, immune cells, in particular, aggressively proliferate and survive even under stressful circumstances, such as hypoxia, malnutrition, and low pH, to say the least [359].

Since single chemotherapies have represented less satisfactory outcomes, combination chemotherapies have absorbed the attention of scientists to improve the chemotherapy responses. As discussed earlier, although they have had disappointing results in some cases, promising responses in the rest of cancers have shed light on developing novel regimens for various cancers.

Moreover, the role of precision medicine should be appreciated. Genomic and proteomic screenings have become easier due to the recent progressions and availabilities in highthrouput technologies such as omics, NGS, and WES. Therefore, using big data extracted from the aforementioned technologies will probably facilitate developing personalized chemotherapy regimens based on the genetic variations in the human population.

Finally, our aim must be to evaluate and study the exact mechanisms, including intracellular signaling and molecular interactions in both resistance and cell-death. We should also utilize highthrouput technologies to investigate the distinct variations that influence the response in different cancer cases. This way, we might develop potential drugs, enhance life-expectancy, and improve the quality of life in patients suffering from cancers.

CRediT authorship contribution statement

Mojtaba Mollaei: Project administration, Supervision, Writing – original draft, Writing – review & editing, Investigation, Validation, Conceptualization. Zuhair Mohammad Hassan: Writing – review & editing, Investigation, Validation, Conceptualization. Fatemeh Khorshidi: Writing – review & editing, Investigation, Validation. Ladan Langroudi: Writing – review & editing.

Declaration of Competing Interest

The authors declare no conflict of interest.

References

  • 1.Raimondi S., Maisonneuve P., Lowenfels A.B. Epidemiology of pancreatic cancer: an overview. Nat. Rev. Gastroenterol. Hepatol. 2009;6:699. doi: 10.1038/nrgastro.2009.177. [DOI] [PubMed] [Google Scholar]
  • 2.Midha S., Chawla S., Garg P.K. Modifiable and non-modifiable risk factors for pancreatic cancer: a review. Cancer Lett. 2016;381:269–277. doi: 10.1016/j.canlet.2016.07.022. [DOI] [PubMed] [Google Scholar]
  • 3.Anand P., Kunnumakara A.B., Sundaram C., Harikumar K.B., Tharakan S.T., Lai O.S., Sung B., Aggarwal B.B. Cancer is a preventable disease that requires major lifestyle changes. Pharm. Res. 2008;25:2097–2116. doi: 10.1007/s11095-008-9661-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Kushi L.H., Doyle C., McCullough M., Rock C.L., Demark-Wahnefried W., Bandera E.V., Gapstur S., Patel A.V., Andrews K., Gansler T. American cancer society guidelines on nutrition and physical activity for cancer prevention: reducing the risk of cancer with healthy food choices and physical activity. CA: Cancer J. Clin. 2012;62:30–67. doi: 10.3322/caac.20140. [DOI] [PubMed] [Google Scholar]
  • 5.Mentella M.C., Scaldaferri F., Ricci C., Gasbarrini A., Miggiano G.A.D. Cancer and Mediterranean diet: a review. Nutrients. 2019;11:2059. doi: 10.3390/nu11092059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Tacar O., Sriamornsak P., Dass C.R. Doxorubicin: an update on anticancer molecular action, toxicity and novel drug delivery systems. J. Pharm. Pharmacol. 2013;65:157–170. doi: 10.1111/j.2042-7158.2012.01567.x. [DOI] [PubMed] [Google Scholar]
  • 7.A.s.o.c. cancer.org.
  • 8.McKenna M.T., Weis J.A., Brock A., Quaranta V., Yankeelov T.E. Precision medicine with imprecise therapy: computational modeling for chemotherapy in breast cancer. Transl. Oncol. 2018;11:732–742. doi: 10.1016/j.tranon.2018.03.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Huang C.-Y., Ju D.-T., Chang C.-F., Reddy P.M., Velmurugan B.K. A review on the effects of current chemotherapy drugs and natural agents in treating non–small cell lung cancer. Biomedicine. 2017;7 doi: 10.1051/bmdcn/2017070423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Gillet J.-P., Gottesman M.M. Springer; 2010. Mechanisms of Multidrug Resistance in Cancer, Multi-drug Resistance in Cancer; pp. 47–76. [DOI] [PubMed] [Google Scholar]
  • 11.Galanski M., Jakupec M.A., Keppler B.K. Update of the preclinical situation of anticancer platinum complexes: novel design strategies and innovative analytical approaches. Curr. Med. Chem. 2005;12:2075–2094. doi: 10.2174/0929867054637626. [DOI] [PubMed] [Google Scholar]
  • 12.Nagai N., Okuda R., Kinoshita M., Ogata H. Decomposition kinetics of cisplatin in human biological fluids. J. Pharm. Pharmacol. 1996;48:918–924. doi: 10.1111/j.2042-7158.1996.tb06002.x. [DOI] [PubMed] [Google Scholar]
  • 13.Ishida S., Lee J., Thiele D.J., Herskowitz I. Uptake of the anticancer drug cisplatin mediated by the copper transporter Ctr1 in yeast and mammals. Proc. Natl. Acad. Sci. 2002;99:14298–14302. doi: 10.1073/pnas.162491399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Fuertes M.A., Alonso C., Pérez J.M. Biochemical modulation of cisplatin mechanisms of action: enhancement of antitumor activity and circumvention of drug resistance. Chem. Rev. 2003;103:645–662. doi: 10.1021/cr020010d. [DOI] [PubMed] [Google Scholar]
  • 15.Payet D., Gaucheron F., Sip M., Leng M. Instability of the monofunctional adducts in cis-[Pt (NH 3) 2 (N7-N-methyl-2-diazapyrenium) CI] 2+;-modified DNA: rates of cross-linking reactions in cis-platinummodified DNA. Nucl. Acids Res. 1993;21:5846–5851. doi: 10.1093/nar/21.25.5846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.S.M. Cohen, S.J. Lippard, Cisplatin: from DNA damage to cancer chemotherapy, (2001). [DOI] [PubMed]
  • 17.Yen H.C., Tang Y.C., Chen F.Y., Chen S.W., Majima H.J. Vol. 1042. Annals of the New York Academy of Sciences; 2005. pp. 516–522. (Enhancement of Cisplatin-Induced Apoptosis and Caspase 3 Activation by Depletion of Mitochondrial DNA in a Human Osteosarcoma Cell Line). [DOI] [PubMed] [Google Scholar]
  • 18.Imamura T., Izumi H., Nagatani G., Ise T., Nomoto M., Iwamoto Y., Kohno K. Interaction with p53 enhances binding of cisplatin-modified DNA by high mobility group 1 protein. J.f Biol. Chem. 2001;276:7534–7540. doi: 10.1074/jbc.M008143200. [DOI] [PubMed] [Google Scholar]
  • 19.Zamble D.B., Mikata Y., Eng C.H., Sandman K.E., Lippard S.J. Testis-specific HMG-domain protein alters the responses of cells to cisplatin. J. Inorg. Biochem. 2002;91:451–462. doi: 10.1016/s0162-0134(02)00472-5. [DOI] [PubMed] [Google Scholar]
  • 20.Siddik Z.H. Cisplatin: mode of cytotoxic action and molecular basis of resistance. Oncogene. 2003;22:7265–7279. doi: 10.1038/sj.onc.1206933. [DOI] [PubMed] [Google Scholar]
  • 21.Petrović M., Todorović D. Biochemical and molecular mechanisms of action of cisplatin in cancer cells. Facta Univ., Ser.: Med. Biol. 2016;18 [Google Scholar]
  • 22.Vaisman A., Varchenko M., Umar A., Kunkel T.A., Risinger J.I., Barrett J.C., Hamilton T.C., Chaney S.G. The role of hMLH1, hMSH3, and hMSH6 defects in cisplatin and oxaliplatin resistance: correlation with replicative bypass of platinum-DNA adducts. Cancer Res. 1998;58:3579–3585. [PubMed] [Google Scholar]
  • 23.Gumulec J., Balvan J., Sztalmachova M., Raudenska M., Dvorakova V., Knopfova L., Polanska H., Hudcova K., Ruttkay-Nedecky B., Babula P. Cisplatin-resistant prostate cancer model: Differences in antioxidant system, apoptosis and cell cycle. Int. J. Oncol. 2014;44:923–933. doi: 10.3892/ijo.2013.2223. [DOI] [PubMed] [Google Scholar]
  • 24.Saad S.Y., Najjar T.A., Alashari M. Role of non-selective adenosine receptor blockade and phosphodiesterase inhibition in cisplatin-induced nephrogonadal toxicity in rats. Clin. Exp. Pharmacol. Physiol. 2004;31:862–867. doi: 10.1111/j.1440-1681.2004.04127.x. [DOI] [PubMed] [Google Scholar]
  • 25.Nuñez G., Benedict M.A., Hu Y., Inohara N. Caspases: the proteases of the apoptotic pathway. Oncogene. 1998;17:3237–3245. doi: 10.1038/sj.onc.1202581. [DOI] [PubMed] [Google Scholar]
  • 26.Salvesen G.S., Abrams J.M. Caspase activation–stepping on the gas or releasing the brakes? Lessons from humans and flies. Oncogene. 2004;23:2774–2784. doi: 10.1038/sj.onc.1207522. [DOI] [PubMed] [Google Scholar]
  • 27.Elmore S. Apoptosis: a review of programmed cell death. Toxicol. Pathol. 2007;35:495–516. doi: 10.1080/01926230701320337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Kohno K., Wang K.-Y., Takahashi M., Kurita T., Yoshida Y., Hirakawa M., Harada Y., Kuma A., Izumi H., Matsumoto S. Mitochondrial transcription factor A and mitochondrial genome as molecular targets for cisplatin-based cancer chemotherapy. Int. J. Mol. Sci. 2015;16:19836–19850. doi: 10.3390/ijms160819836. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Chang L., Karin M. Mammalian MAP kinase signalling cascades. Nature. 2001;410:37–40. doi: 10.1038/35065000. [DOI] [PubMed] [Google Scholar]
  • 30.Johnson G.L., Lapadat R. Mitogen-activated protein kinase pathways mediated by ERK, JNK, and p38 protein kinases. Science. 2002;298:1911–1912. doi: 10.1126/science.1072682. [DOI] [PubMed] [Google Scholar]
  • 31.Bogoyevitch M.A., Court N.W. Counting on mitogen-activated protein kinases—ERKs 3, 4, 5, 6, 7 and 8. Cell. Signal. 2004;16:1345–1354. doi: 10.1016/j.cellsig.2004.05.004. [DOI] [PubMed] [Google Scholar]
  • 32.Yoon S., Seger R. The extracellular signal-regulated kinase: multiple substrates regulate diverse cellular functions. Growth Factors. 2006;24:21–44. doi: 10.1080/02699050500284218. [DOI] [PubMed] [Google Scholar]
  • 33.Dempke W., Voigt W., Grothey A., Hill B.T., Schmoll H.-J. Cisplatin resistance and oncogenes-a review. Anti-Cancer Drugs. 2000;11:225–236. doi: 10.1097/00001813-200004000-00001. [DOI] [PubMed] [Google Scholar]
  • 34.DeHaan R.D., Yazlovitskaya E.M., Persons D.L. Regulation of p53 target gene expression by cisplatin-induced extracellular signal-regulated kinase. Cancer Chemother. Pharmacol. 2001;48:383–388. doi: 10.1007/s002800100318. [DOI] [PubMed] [Google Scholar]
  • 35.Basu A., Tu H. Activation of ERK during DNA damage-induced apoptosis involves protein kinase Cδ. Biochem. Biophys. Res. Commun. 2005;334:1068–1073. doi: 10.1016/j.bbrc.2005.06.199. [DOI] [PubMed] [Google Scholar]
  • 36.Jones E.V., Dickman M.J., Whitmarsh A.J. Regulation of p73-mediated apoptosis by c-Jun N-terminal kinase. Biochem. J. 2007;405:617–623. doi: 10.1042/BJ20061778. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Lafarga V., Cuadrado A., Nebreda A.R. p18Hamlet mediates different p53-dependent responses to DNA damage inducing agents. Cell Cycle. 2007;6:2319–2322. doi: 10.4161/cc.6.19.4741. [DOI] [PubMed] [Google Scholar]
  • 38.Winograd-Katz S., Levitzki A. Cisplatin induces PKB/Akt activation and p38 MAPK phosphorylation of the EGF receptor. Oncogene. 2006;25:7381–7390. doi: 10.1038/sj.onc.1209737. [DOI] [PubMed] [Google Scholar]
  • 39.Jamieson E.R., Lippard S.J. Structure, recognition, and processing of cisplatin− DNA adducts. Chem. Rev. 1999;99:2467–2498. doi: 10.1021/cr980421n. [DOI] [PubMed] [Google Scholar]
  • 40.Florea A.-M., Büsselberg D. Cisplatin as an anti-tumor drug: cellular mechanisms of activity, drug resistance and induced side effects. Cancers. 2011;3:1351–1371. doi: 10.3390/cancers3011351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Kartalou M., Essigmann J.M. Mechanisms of resistance to cisplatin. Mutat. Res./Fundam. Mol. Mech. Mutagen. 2001;478:23–43. doi: 10.1016/s0027-5107(01)00141-5. [DOI] [PubMed] [Google Scholar]
  • 42.Rabik C.A., Dolan M.E. Molecular mechanisms of resistance and toxicity associated with platinating agents. Cancer Treat. Rev. 2007;33:9–23. doi: 10.1016/j.ctrv.2006.09.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Perez R. Cellular and molecular determinants of cisplatin resistance. Eur. J. Cancer. 1998;34:1535–1542. doi: 10.1016/s0959-8049(98)00227-5. [DOI] [PubMed] [Google Scholar]
  • 44.Niedner H., Christen R., Lin X., Kondo A., Howell S. Identification of genes that mediate sensitivity to cisplatin. Mol. Pharmacol. 2001;60:1153–1160. [PubMed] [Google Scholar]
  • 45.Mansouri A., Ridgway L.D., Korapati A.L., Zhang Q., Tian L., Wang Y., Siddik Z.H., Mills G.B., Claret F.X. Sustained activation of JNK/p38 MAPK pathways in response to cisplatin leads to Fas ligand induction and cell death in ovarian carcinoma cells. J. Biol. Chem. 2003;278:19245–19256. doi: 10.1074/jbc.M208134200. [DOI] [PubMed] [Google Scholar]
  • 46.Sève P., Dumontet C. Chemoresistance in non-small cell lung cancer. Curr. Med. Chem.-Anti-Cancer Agents. 2005;5:73–88. doi: 10.2174/1568011053352604. [DOI] [PubMed] [Google Scholar]
  • 47.Kelland L. The resurgence of platinum-based cancer chemotherapy. Nat. Rev. Cancer. 2007;7:573–584. doi: 10.1038/nrc2167. [DOI] [PubMed] [Google Scholar]
  • 48.Holzer A.K., Katano K., Klomp L.W., Howell S.B. Cisplatin rapidly down-regulates its own influx transporter hCTR1 in cultured human ovarian carcinoma cells. Clin. Cancer Res. 2004;10:6744–6749. doi: 10.1158/1078-0432.CCR-04-0748. [DOI] [PubMed] [Google Scholar]
  • 49.Holzer A.K., Howell S.B. The internalization and degradation of human copper transporter 1 following cisplatin exposure. Cancer Res. 2006;66:10944–10952. doi: 10.1158/0008-5472.CAN-06-1710. [DOI] [PubMed] [Google Scholar]
  • 50.Holzer A.K., Manorek G.H., Howell S.B. Contribution of the major copper influx transporter CTR1 to the cellular accumulation of cisplatin, carboplatin, and oxaliplatin. Mol. Pharmacol. 2006;70:1390–1394. doi: 10.1124/mol.106.022624. [DOI] [PubMed] [Google Scholar]
  • 51.Nakayama K., Miyazaki K., Kanzaki A., Fukumoto M., Takebayashi Y. Expression and cisplatin sensitivity of copper-transporting P-type adenosine triphosphatase (ATP7B) in human solid carcinoma cell lines. Oncol. Rep. 2001;8:1285–1287. doi: 10.3892/or.8.6.1285. [DOI] [PubMed] [Google Scholar]
  • 52.De Luca A., Parker L.J., Ang W.H., Rodolfo C., Gabbarini V., Hancock N.C., Palone F., Mazzetti A.P., Menin L., Morton C.J. A structure-based mechanism of cisplatin resistance mediated by glutathione transferase P1-1. Proc. Natl. Acad. Sci. 2019;116:13943–13951. doi: 10.1073/pnas.1903297116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Jansen B.A., Brouwer J., Reedijk J. Glutathione induces cellular resistance against cationic dinuclear platinum anticancer drugs. J. Inorg. Biochem. 2002;89:197–202. doi: 10.1016/s0162-0134(02)00381-1. [DOI] [PubMed] [Google Scholar]
  • 54.Welsh C., Day R., McGurk C., Masters J.R., Wood R.D., Köberle B. Reduced levels of XPA, ERCC1 and XPF DNA repair proteins in testis tumor cell lines. Int. J. Cancer. 2004;110:352–361. doi: 10.1002/ijc.20134. [DOI] [PubMed] [Google Scholar]
  • 55.Siegsmund M.J., Marx C., Seemann O., Schummer B., Steidler A., Toktomambetova L., Köhrmann K.-U., Rassweiler J., Alken P. Cisplatin-resistant bladder carcinoma cells: enhanced expression of metallothioneins. Urol. Res. 1999;27:157–163. doi: 10.1007/s002400050103. [DOI] [PubMed] [Google Scholar]
  • 56.Meijer C., Timmer A., De Vries E.G., Groten J.P., Knol A., Zwart N., Dam W.A., Sleijfer D.T., Mulder N.H. Role of metallothionein in cisplatin sensitivity of germ-cell tumours. Int. J. Cancer. 2000;85:777–781. doi: 10.1002/(sici)1097-0215(20000315)85:6<777::aid-ijc6>3.0.co;2-d. [DOI] [PubMed] [Google Scholar]
  • 57.Moggs J.G., Szymkowski D.E., Yamada M., Karran P., Wood R.D. Differential human nucleotide excision repair of paired and mispaired cisplatin-DNA adducts. Nucl. Acids Res. 1997;25:480–490. doi: 10.1093/nar/25.3.480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Reardon J.T., Vaisman A., Chaney S.G., Sancar A. Efficient nucleotide excision repair of cisplatin, oxaliplatin, and bis-aceto-ammine-dichloro-cyclohexylamine-platinum (IV)(JM216) platinumIntrastrand DNA diadducts. Cancer Res. 1999;59:3968–3971. [PubMed] [Google Scholar]
  • 59.Selvakumaran M., Pisarcik D.A., Bao R., Yeung A.T., Hamilton T.C. Enhanced cisplatin cytotoxicity by disturbing the nucleotide excision repair pathway in ovarian cancer cell lines. Cancer Res. 2003;63:1311–1316. [PubMed] [Google Scholar]
  • 60.Metzger R., Bollschweiler E., Hölscher A.H., Warnecke-Eberz U. ERCC1: impact in multimodality treatment of upper gastrointestinal cancer. Fut. Oncol. 2010;6:1735–1749. doi: 10.2217/fon.10.140. [DOI] [PubMed] [Google Scholar]
  • 61.Woźniak K., Błasiak J. Recognition and repair of DNA-cisplatin adducts. Acta Biochim. Pol. 2002;49:583–596. [PubMed] [Google Scholar]
  • 62.Wu X., Obata T., Khan Q., Highshaw R., White R.D.V., Sweeney C. The phosphatidylinositol-3 kinase pathway regulates bladder cancer cell invasion. BJU Int. 2004;93:143–150. doi: 10.1111/j.1464-410x.2004.04574.x. [DOI] [PubMed] [Google Scholar]
  • 63.Garcia J.A., Danielpour D. Mammalian target of rapamycin inhibition as a therapeutic strategy in the management of urologic malignancies. Mol. Cancer Ther. 2008;7:1347–1354. doi: 10.1158/1535-7163.MCT-07-2408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Pinto-Leite R., Botelho P., Ribeiro E., Oliveira P.A., Santos L. Effect of sirolimus on urinary bladder cancer T24 cell line. J. Exp. Clin. Cancer Res. 2009;28:3. doi: 10.1186/1756-9966-28-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Romeo Y., Moreau J., Zindy P.-J., Saba-El-Leil M., Lavoie G., Dandachi F., Baptissart M., Borden K.L., Meloche S., Roux P.P. RSK regulates activated BRAF signalling to mTORC1 and promotes melanoma growth. Oncogene. 2013;32:2917–2926. doi: 10.1038/onc.2012.312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Houles T., Roux P.P. Seminars in Cancer Biology. Elsevier; 2018. Defining the role of the RSK isoforms in cancer; pp. 53–61. [DOI] [PubMed] [Google Scholar]
  • 67.Romeo Y., Zhang X., Roux P.P. Regulation and function of the RSK family of protein kinases. Biochem. J. 2012;441:553–569. doi: 10.1042/BJ20110289. [DOI] [PubMed] [Google Scholar]
  • 68.Kim H.S., Kim S.-J., Bae J., Wang Y., Park S.Y., Min Y.S., Je H.D., Sohn U.D. The p90rsk-mediated signaling of ethanol-induced cell proliferation in HepG2 cell line. Korean J. Physiol. Pharmacol. 2016;20:595–603. doi: 10.4196/kjpp.2016.20.6.595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Neuss N., Mallett G., Brannon D., Mabe J., Horton H., Huckstep L. Vinca alkaloids XXXIII [1].Microbiological conversions of vincaleukoblastine (VLB, vinblastine), an antitumor alkaloid from Vinca rosea. Linn, Helvetica Chim. Acta. 1974;57:1886–1890. doi: 10.1002/hlca.19740570643. [DOI] [PubMed] [Google Scholar]
  • 70.Rowinsky E.K., Onetto N., Canetta R., Arbuck S. Taxol: the first of the taxanes, an important new class of antitumor agents. Sem. Oncol. 1992:646–662. [PubMed] [Google Scholar]
  • 71.ter Haar E., Kowalski R.J., Hamel E., Lin C.M., Longley R.E., Gunasekera S.P., Rosenkranz H.S., Day B.W. Discodermolide, a cytotoxic marine agent that stabilizes microtubules more potently than taxol. Biochemistry. 1996;35:243–250. doi: 10.1021/bi9515127. [DOI] [PubMed] [Google Scholar]
  • 72.Eckardt J.R. Antitumor activity of docetaxel. Am. J. Health-Syst. Pharm. 1997;54:S2–S6. doi: 10.1093/ajhp/54.suppl_2.S2. [DOI] [PubMed] [Google Scholar]
  • 73.Gligorov J., Lotz J.P. Preclinical pharmacology of the taxanes: implications of the differences. Oncologist. 2004;9:3–8. doi: 10.1634/theoncologist.9-suppl_2-3. [DOI] [PubMed] [Google Scholar]
  • 74.Hennequin C., Giocanti N., Favaudon V. S-phase specificity of cell killing by docetaxel (Taxotere) in synchronised HeLa cells. Br. J. Cancer. 1995;71:1194–1198. doi: 10.1038/bjc.1995.232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Haldar S., Basu A., Croce C.M. Bcl2 is the guardian of microtubule integrity. Cancer Res. 1997;57:229–233. [PubMed] [Google Scholar]
  • 76.Schiff P.B., Horwitz S.B. Taxol stabilizes microtubules in mouse fibroblast cells. Proc. Natl. Acad. Sci. 1980;77:1561–1565. doi: 10.1073/pnas.77.3.1561. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Schiff P.B., Fant J., Horwitz S.B. Promotion of microtubule assembly in vitro by taxol. Nature. 1979;277:665–667. doi: 10.1038/277665a0. [DOI] [PubMed] [Google Scholar]
  • 78.Kampan N.C., Madondo M.T., McNally O.M., Quinn M., Plebanski M. Paclitaxel and its evolving role in the management of ovarian cancer. BioMed Res. Int. 2015;2015 doi: 10.1155/2015/413076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Rao S., Orr G.A., Chaudhary A.G., Kingston D.G., Horwitz S.B. Characterization of the Taxol Binding Site on the Microtubule 2-(m-azidobenzoyl) taxol photolabels a peptide (amino acids 217-231) of β-tubulin. J. Biol. Chem. 1995;270:20235–20238. doi: 10.1074/jbc.270.35.20235. [DOI] [PubMed] [Google Scholar]
  • 80.Zhang D., Yang R., Wang S., Dong Z. Paclitaxel: new uses for an old drug. Drug Des., Dev. Therapy. 2014;8:279. doi: 10.2147/DDDT.S56801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Kreis T., Vale R. Oxford University Press; 1993. Guidebook to the Cytoskeletal and Motor Proteins. [Google Scholar]
  • 82.Sevko A., Kremer V., Falk C., Umansky L., Shurin M.R., Shurin G.V., Umansky V. Application of paclitaxel in low non-cytotoxic doses supports vaccination with melanoma antigens in normal mice. J. Immunotoxicol. 2012;9:275–281. doi: 10.3109/1547691X.2012.655343. [DOI] [PubMed] [Google Scholar]
  • 83.Szakács G., Paterson J.K., Ludwig J.A., Booth-Genthe C., Gottesman M.M. Targeting multidrug resistance in cancer. Nat. Rev. Drug Discov. 2006;5:219–234. doi: 10.1038/nrd1984. [DOI] [PubMed] [Google Scholar]
  • 84.Wang T., Chan Y., Chen C., Kung W., Lee Y., Wang S., Chang T., Wang H. Paclitaxel (Taxol) upregulates expression of functional interleukin-6 in human ovarian cancer cells through multiple signaling pathways. Oncogene. 2006;25:4857–4866. doi: 10.1038/sj.onc.1209498. [DOI] [PubMed] [Google Scholar]
  • 85.Yakirevich E., Sabo E., Naroditsky I., Sova Y., Lavie O., Resnick M.B. Multidrug resistance-related phenotype and apoptosis-related protein expression in ovarian serous carcinomas. Gynecol. Oncol. 2006;100:152–159. doi: 10.1016/j.ygyno.2005.08.050. [DOI] [PubMed] [Google Scholar]
  • 86.Pfannenstiel L.W., Lam S.S., Emens L.A., Jaffee E.M., Armstrong T.D. Paclitaxel enhances early dendritic cell maturation and function through TLR4 signaling in mice. Cell. Immunol. 2010;263:79–87. doi: 10.1016/j.cellimm.2010.03.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Belotti D., Vergani V., Drudis T., Borsotti P., Pitelli M.R., Viale G., Giavazzi R., Taraboletti G. The microtubule-affecting drug paclitaxel has antiangiogenic activity. Clin. Cancer Res. 1996;2:1843–1849. [PubMed] [Google Scholar]
  • 88.Klauber N., Parangi S., Flynn E., Hamel E., D'Amato R.J. Inhibition of angiogenesis and breast cancer in mice by the microtubule inhibitors 2-methoxyestradiol and taxol. Cancer Res. 1997;57:81–86. [PubMed] [Google Scholar]
  • 89.Lau D.H., Xue L., Young L.J., Burke P.A., Cheung A.T. Paclitaxel (Taxol): an inhibitor of angiogenesis in a highly vascularized transgenic breast cancer. Cancer Biotherapy Radiopharmaceut. 1999;14:31–36. doi: 10.1089/cbr.1999.14.31. [DOI] [PubMed] [Google Scholar]
  • 90.Jiang X. Harnessing the immune system for the treatment of breast cancer. J. Zhejiang Univ. Sci. B. 2014;15:1–15. doi: 10.1631/jzus.B1300264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Alexandre J., Hu Y., Lu W., Pelicano H., Huang P. Novel action of paclitaxel against cancer cells: bystander effect mediated by reactive oxygen species. Cancer Res. 2007;67:3512–3517. doi: 10.1158/0008-5472.CAN-06-3914. [DOI] [PubMed] [Google Scholar]
  • 92.Hadzic T., Aykin-Burns N., Zhu Y., Coleman M.C., Leick K., Jacobson G.M., Spitz D.R. Paclitaxel combined with inhibitors of glucose and hydroperoxide metabolism enhances breast cancer cell killing via H2O2-mediated oxidative stress. Free Radical Biol. Med. 2010;48:1024–1033. doi: 10.1016/j.freeradbiomed.2010.01.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Javeed A., Ashraf M., Riaz A., Ghafoor A., Afzal S., Mukhtar M.M. Paclitaxel and immune system. Eur. J. Pharm. Sci. 2009;38:283–290. doi: 10.1016/j.ejps.2009.08.009. [DOI] [PubMed] [Google Scholar]
  • 94.Bracci L., Schiavoni G., Sistigu A., Belardelli F. Immune-based mechanisms of cytotoxic chemotherapy: implications for the design of novel and rationale-based combined treatments against cancer. Cell Death Differ. 2014;21:15–25. doi: 10.1038/cdd.2013.67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Byrd-Leifer C.A., Block E.F., Takeda K., Akira S., Ding A. The role of MyD88 and TLR4 in the LPS-mimetic activity of Taxol. Eur. J. Immunol. 2001;31:2448–2457. doi: 10.1002/1521-4141(200108)31:8<2448::aid-immu2448>3.0.co;2-n. [DOI] [PubMed] [Google Scholar]
  • 96.Zhu Y., Liu N., Xiong S., Zheng Y., Chu Y. CD4+ Foxp3+ regulatory T-cell impairment by paclitaxel is independent of toll-like receptor 4. Scand. J. Immunol. 2011;73:301–308. doi: 10.1111/j.1365-3083.2011.02514.x. [DOI] [PubMed] [Google Scholar]
  • 97.Wang A.C., Ma Y.B., Wu F.X., Ma Z.F., Liu N.F., Gao R., Gao Y.S., Sheng X.G. TLR4 induces tumor growth and inhibits paclitaxel activity in MyD88-positive human ovarian carcinoma in vitro. Oncol. Lett. 2014;7:871–877. doi: 10.3892/ol.2013.1759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Kaneno R., Shurin G.V., Tourkova I.L., Shurin M.R. Chemomodulation of human dendritic cell function by antineoplastic agents in low noncytotoxic concentrations. J. Transl. Med. 2009;7:58. doi: 10.1186/1479-5876-7-58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Kubo M., Morisaki T., Matsumoto K., Tasaki A., Yamanaka N., Nakashima H., Kuroki H., Nakamura K., Nakamura M., Katano M. Paclitaxel probably enhances cytotoxicity of natural killer cells against breast carcinoma cells by increasing perforin production. Cancer Immunol., Immunother. 2005;54:468–476. doi: 10.1007/s00262-004-0617-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Sako T., Burioka N., Yasuda K., Tomita K., Miyata M., Kurai J., Chikumi H., Watanabe M., Suyama H., Fukuoka Y. Cellular immune profile in patients with non-small cell lung cancer after weekly paclitaxel therapy. Acta Oncol. 2004;43:15–19. doi: 10.1080/02841860310016226. [DOI] [PubMed] [Google Scholar]
  • 101.Markasz L., Stuber G., Vanherberghen B., Flaberg E., Olah E., Carbone E., Eksborg S., Klein E., Skribek H., Szekely L. Effect of frequently used chemotherapeutic drugs on the cytotoxic activity of human natural killer cells. Mol. Cancer Ther. 2007;6:644–654. doi: 10.1158/1535-7163.MCT-06-0358. [DOI] [PubMed] [Google Scholar]
  • 102.Michels T., Shurin G.V., Naiditch H., Sevko A., Umansky V., Shurin M.R. Paclitaxel promotes differentiation of myeloid-derived suppressor cells into dendritic cells in vitro in a TLR4-independent manner. J. Immunotoxicol. 2012;9:292–300. doi: 10.3109/1547691X.2011.642418. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Kemp R.A., Ronchese F. Tumor-specific Tc1, but not Tc2, cells deliver protective antitumor immunity. J. Immunol. 2001;167:6497–6502. doi: 10.4049/jimmunol.167.11.6497. [DOI] [PubMed] [Google Scholar]
  • 104.Zhang L., Dermawan K., Jin M., Liu R., Zheng H., Xu L., Zhang Y., Cai Y., Chu Y., Xiong S. Differential impairment of regulatory T cells rather than effector T cells by paclitaxel-based chemotherapy. Clin. Immunol. 2008;129:219–229. doi: 10.1016/j.clim.2008.07.013. [DOI] [PubMed] [Google Scholar]
  • 105.Vicari A.P., Luu R., Zhang N., Patel S., Makinen S.R., Hanson D.C., Weeratna R.D., Krieg A.M. Paclitaxel reduces regulatory T cell numbers and inhibitory function and enhances the anti-tumor effects of the TLR9 agonist PF-3512676 in the mouse. Cancer Immunol., Immunother. 2009;58:615–628. doi: 10.1007/s00262-008-0586-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Liu N., Zheng Y., Zhu Y., Xiong S., Chu Y. Selective impairment of CD4+ CD25+ Foxp3+ regulatory T cells by paclitaxel is explained by Bcl-2/Bax mediated apoptosis. Int. Immunopharmacol. 2011;11:212–219. doi: 10.1016/j.intimp.2010.11.021. [DOI] [PubMed] [Google Scholar]
  • 107.Ferlini C., Cicchillitti L., Raspaglio G., Bartollino S., Cimitan S., Bertucci C., Mozzetti S., Gallo D., Persico M., Fattorusso C. Paclitaxel directly binds to Bcl-2 and functionally mimics activity of Nur77. Cancer Res. 2009;69:6906–6914. doi: 10.1158/0008-5472.CAN-09-0540. [DOI] [PubMed] [Google Scholar]
  • 108.Agarwal R., Kaye S.B. Ovarian cancer: strategies for overcoming resistance to chemotherapy. Nat. Rev. Cancer. 2003;3:502–516. doi: 10.1038/nrc1123. [DOI] [PubMed] [Google Scholar]
  • 109.Sherman-Baust C.A., Becker K.G., Wood W.H., III, Zhang Y., Morin P.J. Gene expression and pathway analysis of ovarian cancer cells selected for resistance to cisplatin, paclitaxel, or doxorubicin. J. Ovarian Res. 2011;4:21. doi: 10.1186/1757-2215-4-21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Acker T., Plate K.H. Role of hypoxia in tumor angiogenesis—molecular and cellular angiogenic crosstalk. Cell Tissue Res. 2003;314:145–155. doi: 10.1007/s00441-003-0763-8. [DOI] [PubMed] [Google Scholar]
  • 111.Höckel M., Vaupel P. Seminars in Oncology. Elsevier; 2001. Biological consequences of tumor hypoxia; pp. 36–41. [PubMed] [Google Scholar]
  • 112.Huang L., Zhang Q.-H., Ao Q.-L., Xing H., Lu Y.-P., Ma D. Effect of hypoxia on the chemotherapeutic sensitivity of human ovarian cancer cells to paclitaxel and its mechanism. Zhonghua zhong liu za zhi Chin. J. Oncol. 2007;29:96–100. [PubMed] [Google Scholar]
  • 113.Rohwer N., Cramer T. Hypoxia-mediated drug resistance: novel insights on the functional interaction of HIFs and cell death pathways. Drug Resist. Updates. 2011;14:191–201. doi: 10.1016/j.drup.2011.03.001. [DOI] [PubMed] [Google Scholar]
  • 114.Aggarwal B.B., Vijayalekshmi R., Sung B. Targeting inflammatory pathways for prevention and therapy of cancer: short-term friend, long-term foe. Clin. Cancer Res. 2009;15:425–430. doi: 10.1158/1078-0432.CCR-08-0149. [DOI] [PubMed] [Google Scholar]
  • 115.Duan Z., Foster R., Bell D.A., Mahoney J., Wolak K., Vaidya A., Hampel C., Lee H., Seiden M.V. Signal transducers and activators of transcription 3 pathway activation in drug-resistant ovarian cancer. Clin. Cancer Res. 2006;12:5055–5063. doi: 10.1158/1078-0432.CCR-06-0861. [DOI] [PubMed] [Google Scholar]
  • 116.Orr G.A., Verdier-Pinard P., McDaid H., Horwitz S.B. Mechanisms of Taxol resistance related to microtubules. Oncogene. 2003;22:7280–7295. doi: 10.1038/sj.onc.1206934. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Gottesman M.M., Ludwig J., Xia D., Szakacs G. Defeating drug resistance in cancer. Discov. Med. 2009;6:18–23. [PubMed] [Google Scholar]
  • 118.Giannakakou P., Sackett D.L., Kang Y.-K., Zhan Z., Buters J.T., Fojo T., Poruchynsky M.S. Paclitaxel-resistant human ovarian cancer cells have mutant β-tubulins that exhibit impaired paclitaxel-driven polymerization. J. Biol. Chem. 1997;272:17118–17125. doi: 10.1074/jbc.272.27.17118. [DOI] [PubMed] [Google Scholar]
  • 119.Kavallaris M., Kuo D., Burkhart C.A., Regl D.L., Norris M.D., Haber M., Horwitz S.B. Taxol-resistant epithelial ovarian tumors are associated with altered expression of specific beta-tubulin isotypes. J. Clin. Investig. 1997;100:1282–1293. doi: 10.1172/JCI119642. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Mozzetti S., Ferlini C., Concolino P., Filippetti F., Raspaglio G., Prislei S., Gallo D., Martinelli E., Ranelletti F.O., Ferrandina G. Class III β-tubulin overexpression is a prominent mechanism of paclitaxel resistance in ovarian cancer patients. Clin. Cancer Res. 2005;11:298–305. [PubMed] [Google Scholar]
  • 121.Traxler P. Tyrosine kinases as targets in cancer therapy–successes and failures. Expert Opin. Ther. Targ. 2003;7:215–234. doi: 10.1517/14728222.7.2.215. [DOI] [PubMed] [Google Scholar]
  • 122.Fabian M.A., Biggs W.H., Treiber D.K., Atteridge C.E., Azimioara M.D., Benedetti M.G., Carter T.A., Ciceri P., Edeen P.T., Floyd M. A small molecule–kinase interaction map for clinical kinase inhibitors. Nat. Biotechnol. 2005;23:329–336. doi: 10.1038/nbt1068. [DOI] [PubMed] [Google Scholar]
  • 123.Meric J.-B., Faivre S., Monnerat C., Vago N.A., Chevalier T.Le, Armand J.-P., Raymond E., Le ZD. 1839:«Iressa». Bull. Cancer. 2000;87:873–876. [PubMed] [Google Scholar]
  • 124.Anderson N.G., Ahmad T., Chan K., Dobson R., Bundred N.J. ZD1839 (Iressa), a novel epidermal growth factor receptor (EGFR) tyrosine kinase inhibitor, potently inhibits the growth of EGFR-positive cancer cell lines with or without erbB2 overexpression. Int. J. Cancer. 2001;94:774–782. doi: 10.1002/ijc.1557. [DOI] [PubMed] [Google Scholar]
  • 125.Lee S.J., Lee H.S., Choi J.S., Na J.O., Seo K.H., Oh M.H., Jou S.S. Remarkable effect of gefitinib retreatment in a lung cancer patient with lepidic predominat adenocarcinoma who had experienced favorable results from initial treatment with gefitinib: a case report. J. Clin. Med. Res. 2012;4:216. doi: 10.4021/jocmr816e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Schiff B.A., McMurphy A.B., Jasser S.A., Younes M.N., Doan D., Yigitbasi O.G., Kim S., Zhou G., Mandal M., Bekele B.N. Epidermal growth factor receptor (EGFR) is overexpressed in anaplastic thyroid cancer, and the EGFR inhibitor gefitinib inhibits the growth of anaplastic thyroid cancer. Clin. Cancer Res. 2004;10:8594–8602. doi: 10.1158/1078-0432.CCR-04-0690. [DOI] [PubMed] [Google Scholar]
  • 127.Han S.-Y., Zhao M.-B., Zhuang G.-B., Li P.-P. Marsdenia tenacissima extract restored gefitinib sensitivity in resistant non-small cell lung cancer cells. Lung Cancer. 2012;75:30–37. doi: 10.1016/j.lungcan.2011.06.001. [DOI] [PubMed] [Google Scholar]
  • 128.Li F., Zhu T., Cao B., Wang J., Liang L. Apatinib enhances antitumour activity of EGFR-TKIs in non-small cell lung cancer with EGFR-TKI resistance. Eur. J. Cancer. 2017;84:184–192. doi: 10.1016/j.ejca.2017.07.037. [DOI] [PubMed] [Google Scholar]
  • 129.Wang Y., Schmid-Bindert G., Zhou C. Erlotinib in the treatment of advanced non-small cell lung cancer: an update for clinicians. Ther. Adv. Med. Oncol. 2012;4:19–29. doi: 10.1177/1758834011427927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Abera M.B., Kazanietz M.G. Protein kinase Cα mediates erlotinib resistance in lung cancer cells. Mol. Pharmacol. 2015;87:832–841. doi: 10.1124/mol.115.097725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Hidalgo M., Bloedow D. Seminars in Oncology. Elsevier; 2003. Pharmacokinetics and pharmacodynamics: maximizing the clinical potential of Erlotinib (Tarceva) pp. 25–33. [PubMed] [Google Scholar]
  • 132.Schettino C., Bareschino M.A., Ricci V., Ciardiello F. Erlotinib: an EGF receptor tyrosine kinase inhibitor in non-small-cell lung cancer treatment. Exp. Rev. Respir. Med. 2008;2:167–178. doi: 10.1586/17476348.2.2.167. [DOI] [PubMed] [Google Scholar]
  • 133.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. doi: 10.1016/1040-8428(94)00144-i. [DOI] [PubMed] [Google Scholar]
  • 134.Ciardiello F., Bianco R., Damiano V., Fontanini G., Caputo R., Pomatico G., De Placido S., Bianco A.R., Mendelsohn J., Tortora G. Antiangiogenic and antitumor activity of anti-epidermal growth factor receptor C225 monoclonal antibody in combination with vascular endothelial growth factor antisense oligonucleotide in human GEO colon cancer cells. Clin. Cancer Res. 2000;6:3739–3747. [PubMed] [Google Scholar]
  • 135.Yarden Y., Sliwkowski M.X. Untangling the ErbB signalling network. Nat. Rev. Mol. Cell Biol. 2001;2:127–137. doi: 10.1038/35052073. [DOI] [PubMed] [Google Scholar]
  • 136.Wang Y.-N., Hung M.-C. Nuclear functions and subcellular trafficking mechanisms of the epidermal growth factor receptor family. Cell Biosci. 2012;2:1–10. doi: 10.1186/2045-3701-2-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Lynch T.J., Bell D.W., Sordella R., Gurubhagavatula S., Okimoto R.A., Brannigan B.W., Harris P.L., Haserlat S.M., Supko J.G., Haluska F.G. Activating mutations in the epidermal growth factor receptor underlying responsiveness of non–small-cell lung cancer to gefitinib. N Engl. J. Med. 2004;350:2129–2139. doi: 10.1056/NEJMoa040938. [DOI] [PubMed] [Google Scholar]
  • 138.Paez J.G., Jänne P.A., Lee J.C., Tracy S., Greulich H., Gabriel S., Herman P., Kaye F.J., Lindeman N., Boggon T.J. EGFR mutations in lung cancer: correlation with clinical response to gefitinib therapy. Science. 2004;304:1497–1500. doi: 10.1126/science.1099314. [DOI] [PubMed] [Google Scholar]
  • 139.Shigematsu H., Lin L., Takahashi T., Nomura M., Suzuki M., Wistuba I.I., Fong K.M., Lee H., Toyooka S., Shimizu N. Clinical and biological features associated with epidermal growth factor receptor gene mutations in lung cancers. J. Natl. Cancer Inst. 2005;97:339–346. doi: 10.1093/jnci/dji055. [DOI] [PubMed] [Google Scholar]
  • 140.Giaccone G. Epidermal growth factor receptor inhibitors in the treatment of non–small-cell lung cancer. J. Clin. Oncol. 2005;23:3235–3242. doi: 10.1200/JCO.2005.08.409. [DOI] [PubMed] [Google Scholar]
  • 141.Johnson B.E., Jänne P.A. Epidermal growth factor receptor mutations in patients with non–small cell lung cancer. Cancer Res. 2005;65:7525–7529. doi: 10.1158/0008-5472.CAN-05-1257. [DOI] [PubMed] [Google Scholar]
  • 142.Moulder S.L., Yakes F.M., Muthuswamy S.K., Bianco R., Simpson J.F., Arteaga C.L. Epidermal growth factor receptor (HER1) tyrosine kinase inhibitor ZD1839 (Iressa) inhibits HER2/neu (erbB2)-overexpressing breast cancer cells in vitro and in vivo. Cancer Res. 2001;61:8887–8895. [PubMed] [Google Scholar]
  • 143.Segovia-Mendoza M., Díaz L., González-González M.E., Martínez-Reza I., García-Quiroz J., Prado-Garcia H., Ibarra-Sánchez M.J., Esparza-López J., Larrea F., García-Becerra R. Calcitriol and its analogues enhance the antiproliferative activity of gefitinib in breast cancer cells. J. Steroid Biochem. Mol. Biol. 2015;148:122–131. doi: 10.1016/j.jsbmb.2014.12.006. [DOI] [PubMed] [Google Scholar]
  • 144.Anido J., Matar P., Albanell J., Guzmán M., Rojo F., Arribas J., Averbuch S., Baselga J. ZD1839, a specific epidermal growth factor receptor (EGFR) tyrosine kinase inhibitor, induces the formation of inactive EGFR/HER2 and EGFR/HER3 heterodimers and prevents heregulin signaling in HER2-overexpressing breast cancer cells. Clin. Cancer Res. 2003;9:1274–1283. [PubMed] [Google Scholar]
  • 145.Krol J., Francis R.E., Albergaria A., Sunters A., Polychronis A., Coombes R.C., Lam E.W.-F. The transcription factor FOXO3a is a crucial cellular target of gefitinib (Iressa) in breast cancer cells. Mol. Cancer Ther. 2007;6:3169–3179. doi: 10.1158/1535-7163.MCT-07-0507. [DOI] [PubMed] [Google Scholar]
  • 146.Dragowska W.H., Weppler S.A., Wang J.C., Wong L.Y., Kapanen A.I., Rawji J.S., Warburton C., Qadir M.A., Donohue E., Roberge M. Induction of autophagy is an early response to gefitinib and a potential therapeutic target in breast cancer. PloS One. 2013;8:e76503. doi: 10.1371/journal.pone.0076503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Pal A., Bains M., Agredo A., Kasinski A. Identification of microRNAs that promote erlotinib resistance in non-small cell lung cancer. Biochem. Pharmacol. 2020 doi: 10.1016/j.bcp.2020.114154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Shen H., Zhu F., Liu J., Xu T., Pei D., Wang R., Qian Y., Li Q., Wang L., Shi Z. Alteration in Mir-21/PTEN expression modulates gefitinib resistance in non-small cell lung cancer. PLoS One. 2014;9 doi: 10.1371/journal.pone.0103305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Zhou J.-Y., Chen X., Zhao J., Bao Z., Chen X., Zhang P., Liu Z.-F., Zhou J.-Y. MicroRNA-34a overcomes HGF-mediated gefitinib resistance in EGFR mutant lung cancer cells partly by targeting MET. Cancer Lett. 2014;351:265–271. doi: 10.1016/j.canlet.2014.06.010. [DOI] [PubMed] [Google Scholar]
  • 150.Zhang W., Lin J., Wang P., Sun J. miR-17-5p down-regulation contributes to erlotinib resistance in non-small cell lung cancer cells. J. Drug Target. 2017;25:125–131. doi: 10.1080/1061186X.2016.1207647. [DOI] [PubMed] [Google Scholar]
  • 151.Chen J., Cui J.d., Guo X.t., Cao X., Li Q. Increased expression of miR-641 contributes to erlotinib resistance in non-small-cell lung cancer cells by targeting NF 1. Cancer Med. 2018;7:1394–1403. doi: 10.1002/cam4.1326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Thomson S., Buck E., Petti F., Griffin G., Brown E., Ramnarine N., Iwata K.K., Gibson N., Haley J.D. Epithelial to mesenchymal transition is a determinant of sensitivity of non–small-cell lung carcinoma cell lines and xenografts to epidermal growth factor receptor inhibition. Cancer Res. 2005;65:9455–9462. doi: 10.1158/0008-5472.CAN-05-1058. [DOI] [PubMed] [Google Scholar]
  • 153.Suda K., Tomizawa K., Fujii M., Murakami H., Osada H., Maehara Y., Yatabe Y., Sekido Y., Mitsudomi T. Epithelial to mesenchymal transition in an epidermal growth factor receptor-mutant lung cancer cell line with acquired resistance to erlotinib. J. Thoracic Oncol. 2011;6:1152–1161. doi: 10.1097/JTO.0b013e318216ee52. [DOI] [PubMed] [Google Scholar]
  • 154.Padua D., Massagué J. Roles of TGFβ in metastasis. Cell Res. 2009;19:89–102. doi: 10.1038/cr.2008.316. [DOI] [PubMed] [Google Scholar]
  • 155.Toonkel R.L., Borczuk A.C., Powell C.A. TGF-² signaling pathway in lung adenocarcinoma invasion. J. Thoracic Oncol. 2010;5:153–157. doi: 10.1097/JTO.0b013e3181c8cc0c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Muraoka-Cook R.S., Dumont N., Arteaga C.L. Dual role of transforming growth factor β in mammary tumorigenesis and metastatic progression. Clin. Cancer Res. 2005;11:937s–943s. [PubMed] [Google Scholar]
  • 157.Serizawa M., Takahashi T., Yamamoto N., Koh Y. Combined treatment with erlotinib and a transforming growth factor-β type I receptor inhibitor effectively suppresses the enhanced motility of erlotinib-resistant non–small-cell lung cancer cells. J. Thoracic Oncol. 2013;8:259–269. doi: 10.1097/JTO.0b013e318279e942. [DOI] [PubMed] [Google Scholar]
  • 158.Kanda R., Kawahara A., Watari K., Murakami Y., Sonoda K., Maeda M., Fujita H., Kage M., Uramoto H., Costa C. Erlotinib resistance in lung cancer cells mediated by integrin β1/Src/Akt-driven bypass signaling. Cancer Res. 2013;73:6243–6253. doi: 10.1158/0008-5472.CAN-12-4502. [DOI] [PubMed] [Google Scholar]
  • 159.Wang J., Zhou P., Wang X., Yu Y., Zhu G., Zheng L., Xu Z., Li F., You Q., Yang Q. Rab25 promotes erlotinib resistance by activating the β1 integrin/AKT/β-catenin pathway in NSCLC. Cell Prolif. 2019;52:e12592. doi: 10.1111/cpr.12592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Segovia-Mendoza M., González-González M.E., Barrera D., Díaz L., García-Becerra R. Efficacy and mechanism of action of the tyrosine kinase inhibitors gefitinib, lapatinib and neratinib in the treatment of HER2-positive breast cancer: preclinical and clinical evidence. Am. J. Cancer Res. 2015;5:2531. [PMC free article] [PubMed] [Google Scholar]
  • 161.Normanno N., Luca A.D., Maiello M.R., Campiglio M., Napolitano M., Mancino M., Carotenuto A., Viglietto G., Menard S. The MEK/MAPK pathway is involved in the resistance of breast cancer cells to the EGFR tyrosine kinase inhibitor gefitinib. J. Cell. Physiol. 2006;207:420–427. doi: 10.1002/jcp.20588. [DOI] [PubMed] [Google Scholar]
  • 162.Normanno N., Campiglio M., Maiello M.R., De Luca A., Mancino M., Gallo M., D'Alessio A., Menard S. Breast cancer cells with acquired resistance to the EGFR tyrosine kinase inhibitor gefitinib show persistent activation of MAPK signaling. Breast Cancer Res. Treat. 2008;112:25–33. doi: 10.1007/s10549-007-9830-2. [DOI] [PubMed] [Google Scholar]
  • 163.Ferrer-Soler L., Vazquez-Martin A., Brunet J., Menendez J.A., De Llorens R., Colomer R. An update of the mechanisms of resistance to EGFR-tyrosine kinase inhibitors in breast cancer: Gefitinib (Iressa™)-induced changes in the expression and nucleo-cytoplasmic trafficking of HER-ligands. Int. J. Mol. Med. 2007;20:3–10. [PubMed] [Google Scholar]
  • 164.Moasser M.M., Basso A., Averbuch S.D., Rosen N. The tyrosine kinase inhibitor ZD1839 (“Iressa”) inhibits HER2-driven signaling and suppresses the growth of HER2-overexpressing tumor cells. Cancer Res. 2001;61:7184–7188. [PubMed] [Google Scholar]
  • 165.Bianco R., Shin I., Ritter C.A., Yakes F.M., Basso A., Rosen N., Tsurutani J., Dennis P.A., Mills G.B., Arteaga C.L. Loss of PTEN/MMAC1/TEP in EGF receptor-expressing tumor cells counteracts the antitumor action of EGFR tyrosine kinase inhibitors. Oncogene. 2003;22:2812–2822. doi: 10.1038/sj.onc.1206388. [DOI] [PubMed] [Google Scholar]
  • 166.She Q.-B., Solit D., Basso A., Moasser M.M. Resistance to gefitinib in PTEN-null HER-overexpressing tumor cells can be overcome through restoration of PTEN function or pharmacologic modulation of constitutive phosphatidylinositol 3′-kinase/Akt pathway signaling. Clin. Cancer Res. 2003;9:4340–4346. [PubMed] [Google Scholar]
  • 167.Mueller K.L., Hunter L.A., Ethier S.P., Boerner J.L. Met and c-Src cooperate to compensate for loss of epidermal growth factor receptor kinase activity in breast cancer cells. Cancer Res. 2008;68:3314–3322. doi: 10.1158/0008-5472.CAN-08-0132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Kuzumaki N., Suzuki A., Narita M., Hosoya T., Nagasawa A., Imai S., Yamamizu K., Morita H., Suzuki T., Okada Y. Multiple analyses of G-protein coupled receptor (GPCR) expression in the development of gefitinib-resistance in transforming non-small-cell lung cancer. PLoS One. 2012;7:e44368. doi: 10.1371/journal.pone.0044368. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Pao W., Wang T.Y., Riely G.J., Miller V.A., Pan Q., Ladanyi M., Zakowski M.F., Heelan R.T., Kris M.G., Varmus H.E. KRAS mutations and primary resistance of lung adenocarcinomas to gefitinib or erlotinib. PLoS Med. 2005;2:e17. doi: 10.1371/journal.pmed.0020017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Hart S., Fischer O.M., Prenzel N., Zwick-Wallasch E., Schneider M., Hennighausen L., Ullrich A. GPCR-induced migration of breast carcinoma cells depends on both EGFR signal transactivation and EGFR-independent pathways. Biol. Chem. 2005;386:845–855. doi: 10.1515/BC.2005.099. [DOI] [PubMed] [Google Scholar]
  • 171.Chengbo H., Huawei Z., Jietao M., Yang Z., Jianzhu Z. Comparison of EGFR and KRAS status between primary non-small cell lung cancer and corresponding metastases: a systematic review and meta-analysis. Zhongguo Fei Ai Za Zhi. 2010;13 doi: 10.3779/j.issn.1009-3419.2010.09.09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Han C.-B., Li F., Ma J.-T., Zou H.-W. Concordant KRAS mutations in primary and metastatic colorectal cancer tissue specimens: a meta-analysis and systematic review, Cancer Investig. 2012;30:741–747. doi: 10.3109/07357907.2012.732159. [DOI] [PubMed] [Google Scholar]
  • 173.L. Toschi, G. Finocchiaro, S. Bartolini, V. Gioia, F. Cappuzzo, Role of gemcitabine in cancer therapy, (2005). [DOI] [PubMed]
  • 174.Huang P., Chubb S., Hertel L.W., Grindey G.B., Plunkett W. Action of 2′, 2′-difluorodeoxycytidine on DNA synthesis. Cancer Res. 1991;51:6110–6117. [PubMed] [Google Scholar]
  • 175.Sarup J.C., Johnson M.A., Verhoef V., Fridland A. Regulation of purine deoxynucleoside phosphorylation by deoxycytidine kinase from human leukemic blast cells. Biochem. Pharmacol. 1989;38:2601–2607. doi: 10.1016/0006-2952(89)90544-3. [DOI] [PubMed] [Google Scholar]
  • 176.Heinemann V., Xu Y.-Z., Chubb S., Sen A., Hertel L.W., Grindey G., Plunkett W. Inhibition of ribonucleotide reduction in CCRF-CEM cells by 2′, 2′-difluorodeoxycytidine. Mol. Pharmacol. 1990;38:567–572. [PubMed] [Google Scholar]
  • 177.Heinemann V., Xu Y.-Z., Chubb S., Sen A., Hertel L.W., Grindey G.B., Plunkett W. Cellular elimination of 2′, 2′-difluorodeoxycytidine 5′-triphosphate: a mechanism of self-potentiation. Cancer Res. 1992;52:533–539. [PubMed] [Google Scholar]
  • 178.de Sousa Cavalcante L., Monteiro G. Gemcitabine: metabolism and molecular mechanisms of action, sensitivity and chemoresistance in pancreatic cancer. Eur. J. Pharmacol. 2014;741:8–16. doi: 10.1016/j.ejphar.2014.07.041. [DOI] [PubMed] [Google Scholar]
  • 179.Ferreira C.G., Span S.W., Peters G.J., Kruyt F.A., Giaccone G. Chemotherapy triggers apoptosis in a caspase-8-dependent and mitochondria-controlled manner in the non-small cell lung cancer cell line NCI-H460. Cancer Res. 2000;60:7133–7141. [PubMed] [Google Scholar]
  • 180.Chandler N.M., Canete J.J., Callery M.P. Caspase-3 drives apoptosis in pancreatic cancer cells after treatment with gemcitabine. J. Gastrointest. Surg. 2004;8:1072–1078. doi: 10.1016/j.gassur.2004.09.054. [DOI] [PubMed] [Google Scholar]
  • 181.von der Maase H., Hansen S., Roberts J., Dogliotti L., Oliver T., Moore M., Bodrogi I., Albers P., Knuth A., Lippert C. Gemcitabine and cisplatin versus methotrexate, vinblastine, doxorubicin, and cisplatin in advanced or metastatic bladder cancer: results of a large, randomized, multinational, multicenter, phase III study. J. Clin. Oncol. 2000;18:3068–3077. doi: 10.1200/JCO.2000.18.17.3068. [DOI] [PubMed] [Google Scholar]
  • 182.Burris H.r., Moore M.J., Andersen J., Green M.R., Rothenberg M.L., Modiano M.R., Christine Cripps M., Portenoy R.K., Storniolo A.M., Tarassoff P. Improvements in survival and clinical benefit with gemcitabine as first-line therapy for patients with advanced pancreas cancer: a randomized trial. J. Clin. Oncol. 1997;15:2403–2413. doi: 10.1200/JCO.1997.15.6.2403. [DOI] [PubMed] [Google Scholar]
  • 183.Crino L., Scagliotti G.V., Ricci S., De Marinis F., Rinaldi M., Gridelli C., Ceribelli A., Bianco R., Marangolo M., Di Costanzo F. Gemcitabine and cisplatin versus mitomycin, ifosfamide, and cisplatin in advanced non–small-cell lung cancer: a randomized phase III study of the Italian lung cancer project. J. Clin. Oncol. 1999;17:3522–3530. doi: 10.1200/JCO.1999.17.11.3522. [DOI] [PubMed] [Google Scholar]
  • 184.Carmichael J., Possinger K., Phillip P., Beykirch M., Kerr H., Walling J., Harris A.L. Advanced breast cancer: a phase II trial with gemcitabine. J. Clin. Oncol. 1995;13:2731–2736. doi: 10.1200/JCO.1995.13.11.2731. [DOI] [PubMed] [Google Scholar]
  • 185.Kummer J.L., Rao P.K., Heidenreich K.A. Apoptosis induced by withdrawal of trophic factors is mediated by p38 mitogen-activated protein kinase. J. Biol. Chem. 1997;272:20490–20494. doi: 10.1074/jbc.272.33.20490. [DOI] [PubMed] [Google Scholar]
  • 186.Habiro A., Tanno S., Koizumi K., Izawa T., Nakano Y., Osanai M., Mizukami Y., Okumura T., Kohgo Y. Involvement of p38 mitogen-activated protein kinase in gemcitabine-induced apoptosis in human pancreatic cancer cells. Biochem. Biophys. Res. Commun. 2004;316:71–77. doi: 10.1016/j.bbrc.2004.02.017. [DOI] [PubMed] [Google Scholar]
  • 187.Nakashima M., Adachi S., Yasuda I., Yamauchi T., Kawaguchi J., Itani M., Yoshioka T., Matsushima-Nishiwaki R., Hirose Y., Kozawa O. Phosphorylation status of heat shock protein 27 plays a key role in gemcitabine-induced apoptosis of pancreatic cancer cells. Cancer Lett. 2011;313:218–225. doi: 10.1016/j.canlet.2011.09.008. [DOI] [PubMed] [Google Scholar]
  • 188.Shi Z., Azuma A., Sampath D., Li Y.-X., Huang P., Plunkett W. S-Phase arrest by nucleoside analogues and abrogation of survival without cell cycle progression by 7-hydroxystaurosporine. Cancer Res. 2001;61:1065–1072. [PubMed] [Google Scholar]
  • 189.Karnitz L.M., Flatten K.S., Wagner J.M., Loegering D., Hackbarth J.S., Arlander S.J., Vroman B.T., Thomas M.B., Baek Y.-U., Hopkins K.M. Gemcitabine-induced activation of checkpoint signaling pathways that affect tumor cell survival. Mol. Pharmacol. 2005;68:1636–1644. doi: 10.1124/mol.105.012716. [DOI] [PubMed] [Google Scholar]
  • 190.Zou L., Cortez D., Elledge S.J. Regulation of ATR substrate selection by Rad17-dependent loading of Rad9 complexes onto chromatin. Genes Dev. 2002;16:198–208. doi: 10.1101/gad.950302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Bermudez V.P., Lindsey-Boltz L.A., Cesare A.J., Maniwa Y., Griffith J.D., Hurwitz J., Sancar A. Loading of the human 9-1-1 checkpoint complex onto DNA by the checkpoint clamp loader hRad17-replication factor C complex in vitro. Proc. Natl. Acad. Sci. 2003;100:1633–1638. doi: 10.1073/pnas.0437927100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Neesse A., Michl P., Frese K.K., Feig C., Cook N., Jacobetz M.A., Lolkema M.P., Buchholz M., Olive K.P., Gress T.M. Stromal biology and therapy in pancreatic cancer. Gut. 2011;60:861–868. doi: 10.1136/gut.2010.226092. [DOI] [PubMed] [Google Scholar]
  • 193.Onishi H., Katano M. Hedgehog signaling pathway as a therapeutic target in various types of cancer. Cancer Sci. 2011;102:1756–1760. doi: 10.1111/j.1349-7006.2011.02010.x. [DOI] [PubMed] [Google Scholar]
  • 194.Thayer S.P., di Magliano M.P., Heiser P.W., Nielsen C.M., Roberts D.J., Lauwers G.Y., Qi Y.P., Gysin S., Fernández-del Castillo C., Yajnik V. Hedgehog is an early and late mediator of pancreatic cancer tumorigenesis. Nature. 2003;425:851–856. doi: 10.1038/nature02009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Bailey J.M., Swanson B.J., Hamada T., Eggers J.P., Singh P.K., Caffery T., Ouellette M.M., Hollingsworth M.A. Sonic hedgehog promotes desmoplasia in pancreatic cancer. Clin. Cancer Res. 2008;14:5995–6004. doi: 10.1158/1078-0432.CCR-08-0291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Jones S., Zhang X., Parsons D.W., Lin J.C.-H., Leary R.J., Angenendt P., Mankoo P., Carter H., Kamiyama H., Jimeno A. Core signaling pathways in human pancreatic cancers revealed by global genomic analyses. Science. 2008;321:1801–1806. doi: 10.1126/science.1164368. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Olive K.P., Jacobetz M.A., Davidson C.J., Gopinathan A., McIntyre D., Honess D., Madhu B., Goldgraben M.A., Caldwell M.E., Allard D. Inhibition of Hedgehog signaling enhances delivery of chemotherapy in a mouse model of pancreatic cancer. Science. 2009;324:1457–1461. doi: 10.1126/science.1171362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Köpper F., Bierwirth C., Schön M., Kunze M., Elvers I., Kranz D., Saini P., Menon M.B., Walter D., Sørensen C.S. Damage-induced DNA replication stalling relies on MAPK-activated protein kinase 2 activity. Proc. Natl. Acad. Sci. 2013;110:16856–16861. doi: 10.1073/pnas.1304355110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Giovannetti E., Del Tacca M., Mey V., Funel N., Nannizzi S., Ricci S., Orlandini C., Boggi U., Campani D., Del Chiaro M. Transcription analysis of human equilibrative nucleoside transporter-1 predicts survival in pancreas cancer patients treated with gemcitabine. Cancer Res. 2006;66:3928–3935. doi: 10.1158/0008-5472.CAN-05-4203. [DOI] [PubMed] [Google Scholar]
  • 200.Farrell J.J., Elsaleh H., Garcia M., Lai R., Ammar A., Regine W.F., Abrams R., Benson A.B., Macdonald J., Cass C.E. Human equilibrative nucleoside transporter 1 levels predict response to gemcitabine in patients with pancreatic cancer. Gastroenterology. 2009;136:187–195. doi: 10.1053/j.gastro.2008.09.067. [DOI] [PubMed] [Google Scholar]
  • 201.Van Haperen V.W.R., Veerman G., Eriksson S., Boven E., Stegmann A.P., Hermsen M., Vermorken J.B., Pinedo H.M., Peters G.J. Development and molecular characterization of a 2′, 2′-difluorodeoxycytidine-resistant variant of the human ovarian carcinoma cell line A2780. Cancer Res. 1994;54:4138–4143. [PubMed] [Google Scholar]
  • 202.Ohhashi S., Ohuchida K., Mizumoto K., Fujita H., Egami T., Yu J., Toma H., Sadatomi S., Nagai E., Tanaka M. Down-regulation of deoxycytidine kinase enhances acquired resistance to gemcitabine in pancreatic cancer. Anticancer Res. 2008;28:2205–2212. [PubMed] [Google Scholar]
  • 203.Bergman A.M., Eijk P.P., van Haperen V.W.R., Smid K., Veerman G., Hubeek I., van den IJssel P., Ylstra B., Peters G.J. In vivo induction of resistance to gemcitabine results in increased expression of ribonucleotide reductase subunit M1 as the major determinant. Cancer Res. 2005;65:9510–9516. doi: 10.1158/0008-5472.CAN-05-0989. [DOI] [PubMed] [Google Scholar]
  • 204.Nakahira S., Nakamori S., Tsujie M., Takahashi Y., Okami J., Yoshioka S., Yamasaki M., Marubashi S., Takemasa I., Miyamoto A. Involvement of ribonucleotide reductase M1 subunit overexpression in gemcitabine resistance of human pancreatic cancer. Int. J. Cancer. 2007;120:1355–1363. doi: 10.1002/ijc.22390. [DOI] [PubMed] [Google Scholar]
  • 205.Nakano Y., Tanno S., Koizumi K., Nishikawa T., Nakamura K., Minoguchi M., Izawa T., Mizukami Y., Okumura T., Kohgo Y. Gemcitabine chemoresistance and molecular markers associated with gemcitabine transport and metabolism in human pancreatic cancer cells. Br. J. Cancer. 2007;96:457–463. doi: 10.1038/sj.bjc.6603559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Fusco A., Fedele M. Roles of HMGA proteins in cancer. Nat. Rev. Cancer. 2007;7:899–910. doi: 10.1038/nrc2271. [DOI] [PubMed] [Google Scholar]
  • 207.Arlt A., Gehrz A., Müerköster S., Vorndamm J., Kruse M.-L., Fölsch U.R., Schäfer H. Role of NF-κ B and Akt/PI3K in the resistance of pancreatic carcinoma cell lines against gemcitabine-induced cell death. Oncogene. 2003;22:3243–3251. doi: 10.1038/sj.onc.1206390. [DOI] [PubMed] [Google Scholar]
  • 208.Raffoul J.J., Heydari A.R., Hillman G.G. DNA repair and cancer therapy: targeting APE1/Ref-1 using dietary agents. J. Oncol. 2012;2012 doi: 10.1155/2012/370481. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Skrypek N., Duchêne B., Hebbar M., Leteurtre E., Van Seuningen I., Jonckheere N. The MUC4 mucin mediates gemcitabine resistance of human pancreatic cancer cells via the concentrative nucleoside transporter family. Oncogene. 2013;32:1714–1723. doi: 10.1038/onc.2012.179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Semenza G.L. Hypoxia-inducible factors: mediators of cancer progression and targets for cancer therapy. Trends Pharmacol. Sci. 2012;33:207–214. doi: 10.1016/j.tips.2012.01.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Eltzschig H.K., Abdulla P., Hoffman E., Hamilton K.E., Daniels D., Schönfeld C., Löffler M., Reyes G., Duszenko M., Karhausen J. HIF-1–dependent repression of equilibrative nucleoside transporter (ENT) in hypoxia. J. Exp. Med. 2005;202:1493–1505. doi: 10.1084/jem.20050177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Morote–Garcia J.C., Rosenberger P., Nivillac N.M., Coe I.R., Eltzschig H.K. Hypoxia-inducible factor–dependent repression of equilibrative nucleoside transporter 2 attenuates mucosal inflammation during intestinal hypoxia. Gastroenterology. 2009;136:607–618. doi: 10.1053/j.gastro.2008.10.037. [DOI] [PubMed] [Google Scholar]
  • 213.Teicher B.A., Fricker S.P. CXCL12 (SDF-1)/CXCR4 pathway in cancer. Clin. Cancer Res. 2010;16:2927–2931. doi: 10.1158/1078-0432.CCR-09-2329. [DOI] [PubMed] [Google Scholar]
  • 214.Singh S., Srivastava S., Bhardwaj A., Owen L., Singh A. CXCL12–CXCR4 signalling axis confers gemcitabine resistance to pancreatic cancer cells: a novel target for therapy. Br. J. Cancer. 2010;103:1671–1679. doi: 10.1038/sj.bjc.6605968. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Arora S., Bhardwaj A., Singh S., Srivastava S.K., McClellan S., Nirodi C.S., Piazza G.A., Grizzle W.E., Owen L.B., Singh A.P. An undesired effect of chemotherapy gemcitabine promotes pancreatic cancer cell invasiveness through reactive oxygen species-dependent, nuclear factor κB-and hypoxia-inducible factor 1α-mediated up-regulation of CXCR4. J. Biol. Chem. 2013;288:21197–21207. doi: 10.1074/jbc.M113.484576. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Kim M.-K., Jeon Y.-K., Woo J.-K., Choi Y., Choi D.-H., Kim Y.-H., Kim C.-W. The C-terminal region of Bfl-1 sensitizes non-small cell lung cancer to gemcitabine-induced apoptosis by suppressing NF-κB activity and down-regulating Bfl-1. Mol. Cancer. 2011;10:98. doi: 10.1186/1476-4598-10-98. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Singh A.P., Arora S., Bhardwaj A., Srivastava S.K., Kadakia M.P., Wang B., Grizzle W.E., Owen L.B., Singh S. CXCL12/CXCR4 Protein Signaling Axis Induces Sonic Hedgehog expression in pancreatic cancer cells via extracellular regulated kinase-and Akt kinase-mediated activation of nuclear factor κB implications for bidirectional tumor-stromal interactions. J. Biol. Chem. 2012;287:39115–39124. doi: 10.1074/jbc.M112.409581. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Nagano H., Tomimaru Y., Eguchi H., Hama N., Wada H., Kawamoto K., Kobayashi S., Mori M., Doki Y. MicroRNA-29a induces resistance to gemcitabine through the Wnt/β-catenin signaling pathway in pancreatic cancer cells. Int. J. Oncol. 2013;43:1066–1072. doi: 10.3892/ijo.2013.2037. [DOI] [PubMed] [Google Scholar]
  • 219.Yang W., Yan H.-X., Chen L., Liu Q., He Y.-Q., Yu L.-X., Zhang S.-H., Huang D.-D., Tang L., Kong X.-N. Wnt/β-catenin signaling contributes to activation of normal and tumorigenic liver progenitor cells. Cancer Res. 2008;68:4287–4295. doi: 10.1158/0008-5472.CAN-07-6691. [DOI] [PubMed] [Google Scholar]
  • 220.Flahaut M., Meier R., Coulon A., Nardou K., Niggli F., Martinet D., Beckmann J., Joseph J., Mühlethaler-Mottet A., Gross N. The Wnt receptor FZD1 mediates chemoresistance in neuroblastoma through activation of the Wnt/β-catenin pathway. Oncogene. 2009;28:2245–2256. doi: 10.1038/onc.2009.80. [DOI] [PubMed] [Google Scholar]
  • 221.Zeng G., Germinaro M., Micsenyi A., Monga N.K., Bell A., Sood A., Malhotra V., Sood N., Midda V., Monga D.K. Aberrant Wnt/β-catenin signaling in pancreatic adenocarcinoma. Neoplasia. 2006;8:279–289. doi: 10.1593/neo.05607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.di Magliano M.P., Biankin A.V., Heiser P.W., Cano D.A., Gutierrez P.J., Deramaudt T., Segara D., Dawson A.C., Kench J.G., Henshall S.M. Common activation of canonical Wnt signaling in pancreatic adenocarcinoma. PloS One. 2007;2:e1155. doi: 10.1371/journal.pone.0001155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Klaus A., Birchmeier W. Wnt signalling and its impact on development and cancer. Nat. Rev. Cancer. 2008;8:387–398. doi: 10.1038/nrc2389. [DOI] [PubMed] [Google Scholar]
  • 224.Kapinas K., Kessler C., Ricks T., Gronowicz G., Delany A.M. miR-29 modulates Wnt signaling in human osteoblasts through a positive feedback loop. J. Biol. Chem. 2010;285:25221–25231. doi: 10.1074/jbc.M110.116137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Harris A.L. Hypoxia—a key regulatory factor in tumour growth. Nat. Rev. Cancer. 2002;2:38–47. doi: 10.1038/nrc704. [DOI] [PubMed] [Google Scholar]
  • 226.Chen E.Y., Mazure N.M., Cooper J.A., Giaccia A.J. Hypoxia activates a platelet-derived growth factor receptor/phosphatidylinositol 3-kinase/Akt pathway that results in glycogen synthase kinase-3 inactivation. Cancer Res. 2001;61:2429–2433. [PubMed] [Google Scholar]
  • 227.Jiang Z., Zhang Y., Chen X., Lam P.Y., Yang H., Xu Q., Yu A.C.H. Activation of Erk1/2 and Akt in astrocytes under ischemia. Biochem. Biophys. Res. Commun. 2002;294:726–733. doi: 10.1016/S0006-291X(02)00540-5. [DOI] [PubMed] [Google Scholar]
  • 228.Yokoi K., Fidler I.J. Hypoxia increases resistance of human pancreatic cancer cells to apoptosis induced by gemcitabine. Clin. Cancer Res. 2004;10:2299–2306. doi: 10.1158/1078-0432.ccr-03-0488. [DOI] [PubMed] [Google Scholar]
  • 229.Takano S., Togawa A., Yoshitomi H., Shida T., Kimura F., Shimizu H., Yoshidome H., Ohtsuka M., Kato A., Tomonaga T. Annexin II overexpression predicts rapid recurrence after surgery in pancreatic cancer patients undergoing gemcitabine-adjuvant chemotherapy. Ann. Surg. Oncol. 2008;15:3157–3168. doi: 10.1245/s10434-008-0061-5. [DOI] [PubMed] [Google Scholar]
  • 230.Kagawa S., Takano S., Yoshitomi H., Kimura F., Satoh M., Shimizu H., Yoshidome H., Ohtsuka M., Kato A., Furukawa K. Akt/mTOR signaling pathway is crucial for gemcitabine resistance induced by Annexin II in pancreatic cancer cells. J. Surg. Res. 2012;178:758–767. doi: 10.1016/j.jss.2012.05.065. [DOI] [PubMed] [Google Scholar]
  • 231.Rivankar S. An overview of doxorubicin formulations in cancer therapy. J. Cancer Res. Ther. 2014;10:853. doi: 10.4103/0973-1482.139267. [DOI] [PubMed] [Google Scholar]
  • 232.Fornari F.A., Randolph J.K., Yalowich J.C., Ritke M.K., Gewirtz D.A. Interference by doxorubicin with DNA unwinding in MCF-7 breast tumor cells. Mol. Pharmacol. 1994;45:649–656. [PubMed] [Google Scholar]
  • 233.Doroshow J.H. Role of hydrogen peroxide and hydroxyl radical formation in the killing of Ehrlich tumor cells by anticancer quinones. Proc. Natl. Acad. Sci. 1986;83:4514–4518. doi: 10.1073/pnas.83.12.4514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Minotti G., Menna P., Salvatorelli E., Cairo G., Gianni L. Anthracyclines: molecular advances and pharmacologic developments in antitumor activity and cardiotoxicity. Pharmacol. Rev. 2004;56:185–229. doi: 10.1124/pr.56.2.6. [DOI] [PubMed] [Google Scholar]
  • 235.Chen M.B., Shen W.X., Yang Y., Wu X.Y., Gu J.H., Lu P.H. Activation of AMP-activated protein kinase is involved in vincristine-induced cell apoptosis in B16 melanoma cell. J. Cell. Physiol. 2011;226:1915–1925. doi: 10.1002/jcp.22522. [DOI] [PubMed] [Google Scholar]
  • 236.Meisse D., Van de Casteele M., Beauloye C., Hainault I., Kefas B.A., Rider M.H., Foufelle F., Hue L. Sustained activation of AMP-activated protein kinase induces c-Jun N-terminal kinase activation and apoptosis in liver cells. FEBS Lett. 2002;526:38–42. doi: 10.1016/s0014-5793(02)03110-1. [DOI] [PubMed] [Google Scholar]
  • 237.Kefas B., Cai Y., Ling Z., Heimberg H., Hue L., Pipeleers D., Van de Casteele M. AMP-activated protein kinase can induce apoptosis of insulin-producing MIN6 cells through stimulation of c-Jun-N-terminal kinase. J. Mol. Endocrinol. 2003;30:151–162. doi: 10.1677/jme.0.0300151. [DOI] [PubMed] [Google Scholar]
  • 238.Leung L.K., Wang T.T. Differential effects of chemotherapeutic agents on the Bcl-2/Bax apoptosis pathway in human breast cancer cell line MCF-7. Breast Cancer Res. Treat. 1999;55:73–83. doi: 10.1023/a:1006190802590. [DOI] [PubMed] [Google Scholar]
  • 239.Adams J.M., Cory S. The Bcl-2 protein family: arbiters of cell survival. Science. 1998;281:1322–1326. doi: 10.1126/science.281.5381.1322. [DOI] [PubMed] [Google Scholar]
  • 240.Mcgahon A.J., Costa Pereira A.P., Daly L., Cotter T.G. Chemotherapeutic drug-induced apoptosis in human leukaemic cells is independent of the Fas (APO-1/CD95) receptor/ligand system. Br. J. Haematol. 1998;101:539–547. doi: 10.1046/j.1365-2141.1998.00745.x. [DOI] [PubMed] [Google Scholar]
  • 241.Huang J., Liu K., Yu Y., Xie M., Kang R., Vernon P.J., Cao L., Tang D., Ni J. Targeting HMGB1-mediated autophagy as a novel therapeutic strategy for osteosarcoma. Autophagy. 2012;8:275–277. doi: 10.4161/auto.8.2.18940. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Liu Y.-Y., Yu J.Y., Yin D., Patwardhan G.A., Gupta V., Hirabayashi Y., Holleran W.M., Giuliano A.E., Jazwinski S.M., Gouaze-Andersson V. A role for ceramide in driving cancer cell resistance to doxorubicin. FASEB J. 2008;22:2541–2551. doi: 10.1096/fj.07-092981. [DOI] [PubMed] [Google Scholar]
  • 243.Harper M.E., Antoniou A., Villalobos-Menuey E., Russo A., Trauger R., Vendemelio M., George A., Bartholomew R., Carlo D., Shaikh A. Characterization of a novel metabolic strategy used by drug-resistant tumor cells. FASEB J. 2002;16:1550–1557. doi: 10.1096/fj.02-0541com. [DOI] [PubMed] [Google Scholar]
  • 244.Elliott A.M., Al-Hajj M.A. ABCB8 mediates doxorubicin resistance in melanoma cells by protecting the mitochondrial genome. Mol. Cancer Res. 2009;7:79–87. doi: 10.1158/1541-7786.MCR-08-0235. [DOI] [PubMed] [Google Scholar]
  • 245.Germann U. P-glycoprotein—a mediator of multidrug resistance in tumour cells. Eur. J. Cancer. 1996;32:927–944. doi: 10.1016/0959-8049(96)00057-3. [DOI] [PubMed] [Google Scholar]
  • 246.Cole S., Bhardwaj G., Gerlach J., Mackie J., Grant C., Almquist K., Stewart A., Kurz E., Duncan A., Deeley R.G. Overexpression of a transporter gene in a multidrug-resistant human lung cancer cell line. Science. 1992;258:1650–1654. doi: 10.1126/science.1360704. [DOI] [PubMed] [Google Scholar]
  • 247.Young L.C., Campling B.G., Cole S.P., Deeley R.G., Gerlach J.H. Multidrug resistance proteins MRP3, MRP1, and MRP2 in lung cancer: correlation of protein levels with drug response and messenger RNA levels. Clin. Cancer Res. 2001;7:1798–1804. [PubMed] [Google Scholar]
  • 248.Singhal S.S., Singhal J., Sharma R., Singh S.V., Zimniak P., Awasthi Y.C., Awasthi S. Role of RLIP76 in lung cancer doxorubicin resistance: I. The ATPase activity of RLIP76 correlates with doxorubicin and 4-hydroxynonenal resistance in lung cancer cells. Int. J. Oncol. 2003;22:365–375. [PubMed] [Google Scholar]
  • 249.Lal S., Mahajan A., Ning Chen W., Chowbay B. Pharmacogenetics of target genes across doxorubicin disposition pathway: a review. Curr. Drug Metab. 2010;11:115–128. doi: 10.2174/138920010791110890. [DOI] [PubMed] [Google Scholar]
  • 250.Burgess D.J., Doles J., Zender L., Xue W., Ma B., McCombie W.R., Hannon G.J., Lowe S.W., Hemann M.T. Topoisomerase levels determine chemotherapy response in vitro and in vivo. Proc. Natl. Acad. Sci. 2008;105:9053–9058. doi: 10.1073/pnas.0803513105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Pritchard K.I., Messersmith H., Elavathil L., Trudeau M., O'Malley F., Dhesy-Thind B. HER-2 and topoisomerase II as predictors of response to chemotherapy. J. Clin. Oncol. 2008;26:736–744. doi: 10.1200/JCO.2007.15.4716. [DOI] [PubMed] [Google Scholar]
  • 252.Oakman C., Moretti E., Galardi F., Santarpia L., Di Leo A. The role of topoisomerase IIα and HER-2 in predicting sensitivity to anthracyclines in breast cancer patients. Cancer Treat. Rev. 2009;35:662–667. doi: 10.1016/j.ctrv.2009.08.006. [DOI] [PubMed] [Google Scholar]
  • 253.Montecucco A., Zanetta F., Biamonti G. Molecular mechanisms of etoposide. EXCLI J. 2015;14:95. doi: 10.17179/excli2015-561. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 254.Yang X., Li W., Prescott E.D., Burden S.J., Wang J.C. DNA topoisomerase IIβ and neural development. Science. 2000;287:131–134. doi: 10.1126/science.287.5450.131. [DOI] [PubMed] [Google Scholar]
  • 255.Lyu Y.L., Lin C.-P., Azarova A.M., Cai L., Wang J.C., Liu L.F. Role of topoisomerase IIβ in the expression of developmentally regulated genes. Mol. Cell. Biol. 2006;26:7929–7941. doi: 10.1128/MCB.00617-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 256.Nitiss J.L. DNA topoisomerase II and its growing repertoire of biological functions. Nat. Rev. Cancer. 2009;9:327–337. doi: 10.1038/nrc2608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.Tiwari V.K., Burger L., Nikoletopoulou V., Deogracias R., Thakurela S., Wirbelauer C., Kaut J., Terranova R., Hoerner L., Mielke C. Target genes of Topoisomerase IIβ regulate neuronal survival and are defined by their chromatin state. Proc. Natl. Acad. Sci. 2012;109:E934–E943. doi: 10.1073/pnas.1119798109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Wilstermann A.M., Bender R.P., Godfrey M., Choi S., Anklin C., Berkowitz D.B., Osheroff N., Graves D.E. Topoisomerase II− drug interaction domains: Identification of substituents on etoposide that interact with the enzyme. Biochemistry. 2007;46:8217–8225. doi: 10.1021/bi700272u. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Wu C.-C., Li T.-K., Farh L., Lin L.-Y., Lin T.-S., Yu Y.-J., Yen T.-J., Chiang C.-W., Chan N.-L. Structural basis of type II topoisomerase inhibition by the anticancer drug etoposide. Science. 2011;333:459–462. doi: 10.1126/science.1204117. [DOI] [PubMed] [Google Scholar]
  • 260.Chamani E., Rabbani-Chadegani A., Zahraei Z. Spectroscopic detection of etoposide binding to chromatin components: the role of histone proteins. Spectrochim. Acta Part A: Mol. Biomol. Spectrosc. 2014;133:292–299. doi: 10.1016/j.saa.2014.05.068. [DOI] [PubMed] [Google Scholar]
  • 261.Rossi R., Lidonnici M.R., Soza S., Biamonti G., Montecucco A. The dispersal of replication proteins after Etoposide treatment requires the cooperation of Nbs1 with the ataxia telangiectasia Rad3-related/Chk1 pathway. Cancer Res. 2006;66:1675–1683. doi: 10.1158/0008-5472.CAN-05-2741. [DOI] [PubMed] [Google Scholar]
  • 262.Takami M., Takakusagi Y., Kuramochi K., Tsukuda S., Aoki S., Morohashi K., Ohta K., Kobayashi S., Sakaguchi K., Sugawara F. A screening of a library of T7 phage-displayed peptide identifies E2F-4 as an etoposide-binding protein. Molecules. 2011;16:4278–4294. doi: 10.3390/molecules16054278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Ledesma F.C., El Khamisy S.F., Zuma M.C., Osborn K., Caldecott K.W. A human 5′-tyrosyl DNA phosphodiesterase that repairs topoisomerase-mediated DNA damage. Nature. 2009;461:674–678. doi: 10.1038/nature08444. [DOI] [PubMed] [Google Scholar]
  • 264.Zeng Z., Cortés-Ledesma F., El Khamisy S.F., Caldecott K.W. TDP2/TTRAP is the major 5′-tyrosyl DNA phosphodiesterase activity in vertebrate cells and is critical for cellular resistance to topoisomerase II-induced DNA damage. J. Biol. Chem. 2011;286:403–409. doi: 10.1074/jbc.M110.181016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Gao R., Schellenberg M.J., Shar-yin N.H., Abdelmalak M., Marchand C., Nitiss K.C., Nitiss J.L., Williams R.S., Pommier Y. Proteolytic degradation of topoisomerase II (Top2) enables the processing of Top2• DNA and Top2• RNA covalent complexes by tyrosyl-DNA-phosphodiesterase 2 (TDP2) J. Biol. Chem. 2014;289:17960–17969. doi: 10.1074/jbc.M114.565374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 266.Adachi N., Suzuki H., Iiizumi S., Koyama H. Hypersensitivity of nonhomologous DNA end-joining mutants to VP-16 and ICRF-193 implications for the repair of topoisomerase II-mediated DNA damage. J. Biol. Chem. 2003;278:35897–35902. doi: 10.1074/jbc.M306500200. [DOI] [PubMed] [Google Scholar]
  • 267.Kakarougkas A., Jeggo P. DNA DSB repair pathway choice: an orchestrated handover mechanism. Br. J. Radiol. 2014;87 doi: 10.1259/bjr.20130685. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Malik M., Nitiss K.C., Enriquez-Rios V., Nitiss J.L. Roles of nonhomologous end-joining pathways in surviving topoisomerase II–mediated DNA damage. Mol. Cancer Ther. 2006;5:1405–1414. doi: 10.1158/1535-7163.MCT-05-0263. [DOI] [PubMed] [Google Scholar]
  • 269.Chen C.-S., Wang Y.-C., Yang H.-C., Huang P.-H., Kulp S.K., Yang C.-C., Lu Y.-S., Matsuyama S., Chen C.-Y., Chen C.-S. Histone deacetylase inhibitors sensitize prostate cancer cells to agents that produce DNA double-strand breaks by targeting Ku70 acetylation. Cancer Res. 2007;67:5318–5327. doi: 10.1158/0008-5472.CAN-06-3996. [DOI] [PubMed] [Google Scholar]
  • 270.Maser R.S., Monsen K.J., Nelms B.E., Petrini J. hMre11 and hRad50 nuclear foci are induced during the normal cellular response to DNA double-strand breaks. Mol. Cell. Biol. 1997;17:6087–6096. doi: 10.1128/mcb.17.10.6087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 271.Caporossi D., Porfirio B., Nicoletti B., Palitti F., Degrassi F., De Salvia R., Tanzarella C. Hypersensitivity of lymphoblastoid lines derived from ataxia telangiectasia patients to the induction of chromosomal aberrations by etoposide (VP-16) Mutat. Res./Fundam. Mol. Mech. Mutagen. 1993;290:265–272. doi: 10.1016/0027-5107(93)90167-e. [DOI] [PubMed] [Google Scholar]
  • 272.Nakada S., Katsuki Y., Imoto I., Yokoyama T., Nagasawa M., Inazawa J., Mizutani S. Early G2/M checkpoint failure as a molecular mechanism underlying etoposide-induced chromosomal aberrations. J. Clin. Investig. 2006;116:80–89. doi: 10.1172/JCI25716. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Karni R., de Stanchina E., Lowe S.W., Sinha R., Mu D., Krainer A.R. The gene encoding the splicing factor SF2/ASF is a proto-oncogene. Nat. Struct. Mol. Biol. 2007;14:185–193. doi: 10.1038/nsmb1209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 274.Moore M.J., Wang Q., Kennedy C.J., Silver P.A. An alternative splicing network links cell-cycle control to apoptosis. Cell. 2010;142:625–636. doi: 10.1016/j.cell.2010.07.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 275.Shultz J.C., Goehe R.W., Wijesinghe D.S., Murudkar C., Hawkins A.J., Shay J.W., Minna J.D., Chalfant C.E. Alternative splicing of caspase 9 is modulated by the phosphoinositide 3-kinase/Akt pathway via phosphorylation of SRp30a. Cancer Res. 2010;70:9185–9196. doi: 10.1158/0008-5472.CAN-10-1545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276.Leva V., Giuliano S., Bardoni A., Camerini S., Crescenzi M., Lisa A., Biamonti G., Montecucco A. Phosphorylation of SRSF1 is modulated by replicational stress. Nucl. Acids Res. 2012;40:1106–1117. doi: 10.1093/nar/gkr837. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Montecucco A., Biamonti G. Pre-mRNA processing factors meet the DNA damage response. Front. Genet. 2013;4:102. doi: 10.3389/fgene.2013.00102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Wang P., Song J.H., Song D.K., Zhang J., Hao C. Role of death receptor and mitochondrial pathways in conventional chemotherapy drug induction of apoptosis. Cell. Signal. 2006;18:1528–1535. doi: 10.1016/j.cellsig.2005.12.004. [DOI] [PubMed] [Google Scholar]
  • 279.Kaufmann S.H., Earnshaw W.C. Induction of apoptosis by cancer chemotherapy. Exp. Cell Res. 2000;256:42–49. doi: 10.1006/excr.2000.4838. [DOI] [PubMed] [Google Scholar]
  • 280.Korwek Z., Sewastianik T., Bielak-Zmijewska A., Mosieniak G., Alster O., Moreno-Villaneuva M., Burkle A., Sikora E. Inhibition of ATM blocks the etoposide-induced DNA damage response and apoptosis of resting human T cells. DNA Repair. 2012;11:864–873. doi: 10.1016/j.dnarep.2012.08.006. [DOI] [PubMed] [Google Scholar]
  • 281.Zhang H., Li S., Zhang Z., Hu X., Hou P., Gao L., Du R., Zhang X. Nemo-like kinase is critical for p53 stabilization and function in response to DNA damage. Cell Death Differ. 2014;21:1656–1663. doi: 10.1038/cdd.2014.78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 282.Kruse J.-P., Gu W. Modes of p53 regulation. Cell. 2009;137:609–622. doi: 10.1016/j.cell.2009.04.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.Yasuda J., Yokoo H., Yamada T., Kitabayashi I., Sekiya T., Ichikawa H. Nemo-like kinase suppresses a wide range of transcription factors, including nuclear factor-kB. Cancer Sci. 2004;95:52–57. doi: 10.1111/j.1349-7006.2004.tb03170.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Fan Y., Dutta J., Gupta N., Fan G., Gélinas C. Programmed Cell Death in Cancer Progression and Therapy. Springer; 2008. Regulation of programmed cell death by NF-κB and its role in tumorigenesis and therapy; pp. 223–250. [DOI] [PubMed] [Google Scholar]
  • 285.Jamshidiha M., Habibollahi P., Ostad S., Ghahremani M. Primary WWOX phosphorylation and JNK activation during etoposide induces cytotoxicity in HEK293 cells. Daru: J. Fac. Pharm. 2010;18:141. [PMC free article] [PubMed] [Google Scholar]
  • 286.Nayak M.S., Yang J.-M., Hait W.N. Effect of a Single nucleotide polymorphism in the murine double minute 2 promoter (SNP309) on the sensitivity to topoisomerase II–targeting drugs. Cancer Res. 2007;67:5831–5839. doi: 10.1158/0008-5472.CAN-06-4533. [DOI] [PubMed] [Google Scholar]
  • 287.Zhang A., Lyu Y.L., Lin C.-P., Zhou N., Azarova A.M., Wood L.M., Liu L.F. A protease pathway for the repair of topoisomerase II-DNA covalent complexes. J. Biol. Chem. 2006;281:35997–36003. doi: 10.1074/jbc.M604149200. [DOI] [PubMed] [Google Scholar]
  • 288.Monte M., Simonatto M., Peche L.Y., Bublik D.R., Gobessi S., Pierotti M.A., Rodolfo M., Schneider C. MAGE-A tumor antigens target p53 transactivation function through histone deacetylase recruitment and confer resistance to chemotherapeutic agents. Proc. Natl. Acad. Sci. 2006;103:11160–11165. doi: 10.1073/pnas.0510834103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 289.Katayama M., Kawaguchi T., Berger M., Pieper R. DNA damaging agent-induced autophagy produces a cytoprotective adenosine triphosphate surge in malignant glioma cells. Cell Death Differ. 2007;14:548–558. doi: 10.1038/sj.cdd.4402030. [DOI] [PubMed] [Google Scholar]
  • 290.Xie B.-S., Zhao H.-C., Yao S.-K., Zhuo D.-X., Jin B., Lv D.-C., Wu C.-L., Ma D.-L., Gao C., Shu X.-M. Autophagy inhibition enhances etoposide-induced cell death in human hepatoma G2 cells. Int. J. Mol. Med. 2011;27:599–606. doi: 10.3892/ijmm.2011.607. [DOI] [PubMed] [Google Scholar]
  • 291.Alpsoy A., Yasa S., Gündüz U. Etoposide resistance in MCF-7 breast cancer cell line is marked by multiple mechanisms. Biomed. Pharmacother. 2014;68:351–355. doi: 10.1016/j.biopha.2013.09.007. [DOI] [PubMed] [Google Scholar]
  • 292.Going J., Nixon C., Dornan E., Boner W., Donaldson M., Morgan I. Aberrant expression of TopBP1 in breast cancer. Histopathology. 2007;50:418–424. doi: 10.1111/j.1365-2559.2007.02622.x. [DOI] [PubMed] [Google Scholar]
  • 293.Forma E., Krzeslak A., Bernaciak M., Romanowicz-Makowska H., Brys M. Expression of TopBP1 in hereditary breast cancer. Mol. Biol. Rep. 2012;39:7795–7804. doi: 10.1007/s11033-012-1622-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294.O'Brien P.M., Davies M.J., Scurry J., Smith A.N., Barton C.A., Henderson M.J., Saunders D.N., Gloss B., Patterson K.I., Clancy J.L. The E3 ubiquitin ligase EDD is an adverse prognostic factor for serous epithelial ovarian cancer and modulates cisplatin resistance in vitro. Br. J. Cancer. 2008;98:1085–1093. doi: 10.1038/sj.bjc.6604281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 295.Porta C., Paglino C., Mosca A. Targeting PI3K/Akt/mTOR signaling in cancer. Front. Oncol. 2014;4:64. doi: 10.3389/fonc.2014.00064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 296.Datta S.R., Brunet A., Greenberg M.E. Cellular survival: a play in three Akts. Genes Dev. 1999;13:2905–2927. doi: 10.1101/gad.13.22.2905. [DOI] [PubMed] [Google Scholar]
  • 297.Fruman D.A., Meyers R.E., Cantley L.C. Phosphoinositide kinases. Annu. Rev. 1998 doi: 10.1146/annurev.biochem.67.1.481. 4139 El Camino Way, PO Box 10139, Palo Alto, CA 94303-0139, USA. [DOI] [PubMed] [Google Scholar]
  • 298.Vara J.Á.F., Casado E., de Castro J., Cejas P., Belda-Iniesta C., González-Barón M. PI3K/Akt signalling pathway and cancer. Cancer Treat. Rev. 2004;30:193–204. doi: 10.1016/j.ctrv.2003.07.007. [DOI] [PubMed] [Google Scholar]
  • 299.Pawson T., Nash P. Protein–protein interactions define specificity in signal transduction. Genes Dev. 2000;14:1027–1047. [PubMed] [Google Scholar]
  • 300.Testa J.R., Bellacosa A. AKT plays a central role in tumorigenesis. Proc. Natl. Acad. Sci. 2001;98:10983–10985. doi: 10.1073/pnas.211430998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 301.Chen X., Thakkar H., Tyan F., Gim S., Robinson H., Lee C., Pandey S.K., Nwokorie C., Onwudiwe N., Srivastava R.K. Constitutively active Akt is an important regulator of TRAIL sensitivity in prostate cancer. Oncogene. 2001;20:6073–6083. doi: 10.1038/sj.onc.1204736. [DOI] [PubMed] [Google Scholar]
  • 302.Hussain A.R., Ahmed S.O., Ahmed M., Khan O.S., Al AbdulMohsen S., Platanias L.C., Al-Kuraya K.S., Uddin S. Cross-talk between NFkB and the PI3-kinase/AKT pathway can be targeted in primary effusion lymphoma (PEL) cell lines for efficient apoptosis. PloS One. 2012;7:e39945. doi: 10.1371/journal.pone.0039945. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 303.Liang J., Slingerland J.M. Multiple roles of the PI3K/PKB (Akt) pathway in cell cycle progression. Cell Cycle. 2003;2:336–342. [PubMed] [Google Scholar]
  • 304.Kumar C.C., Madison V. AKT crystal structure and AKT-specific inhibitors. Oncogene. 2005;24:7493–7501. doi: 10.1038/sj.onc.1209087. [DOI] [PubMed] [Google Scholar]
  • 305.Hay N., Sonenberg N. Upstream and downstream of mTOR. Genes Dev. 2004;18:1926–1945. doi: 10.1101/gad.1212704. [DOI] [PubMed] [Google Scholar]
  • 306.Memmott R.M., Dennis P.A. Akt-dependent and-independent mechanisms of mTOR regulation in cancer. Cell. Signal. 2009;21:656–664. doi: 10.1016/j.cellsig.2009.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Shaw R.J., Bardeesy N., Manning B.D., Lopez L., Kosmatka M., DePinho R.A., Cantley L.C. The LKB1 tumor suppressor negatively regulates mTOR signaling. Cancer Cell. 2004;6:91–99. doi: 10.1016/j.ccr.2004.06.007. [DOI] [PubMed] [Google Scholar]
  • 308.Lu Y., Lin Y.-Z., LaPushin R., Cuevas B., Fang X., Yu S.X., Davies M.A., Khan H., Furui T., Mao M. The PTEN/MMAC1/TEP tumor suppressor gene decreases cell growth and induces apoptosis and anoikis in breast cancer cells. Oncogene. 1999;18:7034–7045. doi: 10.1038/sj.onc.1203183. [DOI] [PubMed] [Google Scholar]
  • 309.Simpson L., Parsons R. PTEN: life as a tumor suppressor. Exp. Cell Res. 2001;264:29–41. doi: 10.1006/excr.2000.5130. [DOI] [PubMed] [Google Scholar]
  • 310.Feng Z., Zhang H., Levine A.J., Jin S. The coordinate regulation of the p53 and mTOR pathways in cells. Proc. Natl. Acad. Sci. 2005;102:8204–8209. doi: 10.1073/pnas.0502857102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 311.Gwinn D.M., Shackelford D.B., Egan D.F., Mihaylova M.M., Mery A., Vasquez D.S., Turk B.E., Shaw R.J. AMPK phosphorylation of raptor mediates a metabolic checkpoint. Mo. Cell. 2008;30:214–226. doi: 10.1016/j.molcel.2008.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 312.Tian T., Li X., Zhang J. mTOR signaling in cancer and mTOR inhibitors in solid tumor targeting therapy. Int. J. Mol. Sci. 2019;20:755. doi: 10.3390/ijms20030755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 313.Ma X.M., Blenis J. Molecular mechanisms of mTOR-mediated translational control. Nat. Rev. Mol. Cell Biol. 2009;10:307–318. doi: 10.1038/nrm2672. [DOI] [PubMed] [Google Scholar]
  • 314.Gingras A.-C., Kennedy S.G., O'Leary M.A., Sonenberg N., Hay N. 4E-BP1, a repressor of mRNA translation, is phosphorylated and inactivated by the Akt (PKB) signaling pathway. Genes Dev. 1998;12:502–513. doi: 10.1101/gad.12.4.502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Browne G.J., Proud C.G. A novel mTOR-regulated phosphorylation site in elongation factor 2 kinase modulates the activity of the kinase and its binding to calmodulin. Mol. Cell. Biol. 2004;24:2986–2997. doi: 10.1128/MCB.24.7.2986-2997.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 316.Holz M.K., Ballif B.A., Gygi S.P., Blenis J. mTOR and S6K1 mediate assembly of the translation preinitiation complex through dynamic protein interchange and ordered phosphorylation events. Cell. 2005;123:569–580. doi: 10.1016/j.cell.2005.10.024. [DOI] [PubMed] [Google Scholar]
  • 317.Hsieh A.C., Liu Y., Edlind M.P., Ingolia N.T., Janes M.R., Sher A., Shi E.Y., Stumpf C.R., Christensen C., Bonham M.J. The translational landscape of mTOR signalling steers cancer initiation and metastasis. Nature. 2012;485:55–61. doi: 10.1038/nature10912. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 318.Sarbassov D.D., Ali S.M., Sengupta S., Sheen J.-H., Hsu P.P., Bagley A.F., Markhard A.L., Sabatini D.M. Prolonged rapamycin treatment inhibits mTORC2 assembly and Akt/PKB. Mol. Cell. 2006;22:159–168. doi: 10.1016/j.molcel.2006.03.029. [DOI] [PubMed] [Google Scholar]
  • 319.Willems L., Tamburini J., Chapuis N., Lacombe C., Mayeux P., Bouscary D. PI3K and mTOR signaling pathways in cancer: new data on targeted therapies. Curr. Oncol. Rep. 2012;14:129–138. doi: 10.1007/s11912-012-0227-y. [DOI] [PubMed] [Google Scholar]
  • 320.Inoki K., Li Y., Zhu T., Wu J., Guan K.-L. TSC2 is phosphorylated and inhibited by Akt and suppresses mTOR signalling. Nat. Cell Biol. 2002;4:648–657. doi: 10.1038/ncb839. [DOI] [PubMed] [Google Scholar]
  • 321.Um S.H., Frigerio F., Watanabe M., Picard F., Joaquin M., Sticker M., Fumagalli S., Allegrini P.R., Kozma S.C., Auwerx J. Absence of S6K1 protects against age-and diet-induced obesity while enhancing insulin sensitivity. Nature. 2004;431:200–205. doi: 10.1038/nature02866. [DOI] [PubMed] [Google Scholar]
  • 322.Hsu P.P., Kang S.A., Rameseder J., Zhang Y., Ottina K.A., Lim D., Peterson T.R., Choi Y., Gray N.S., Yaffe M.B. The mTOR-regulated phosphoproteome reveals a mechanism of mTORC1-mediated inhibition of growth factor signaling. Science. 2011;332:1317–1322. doi: 10.1126/science.1199498. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 323.Yu Y., Yoon S.-O., Poulogiannis G., Yang Q., Ma X.M., Villén J., Kubica N., Hoffman G.R., Cantley L.C., Gygi S.P. Phosphoproteomic analysis identifies Grb10 as an mTORC1 substrate that negatively regulates insulin signaling. Science. 2011;332:1322–1326. doi: 10.1126/science.1199484. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 324.Grewe M., Gansauge F., Schmid R.M., Adler G., Seufferlein T. Regulation of cell growth and cyclin D1 expression by the constitutively active FRAP-p70s6K pathway in human pancreatic cancer cells. Cancer Res. 1999;59:3581–3587. [PubMed] [Google Scholar]
  • 325.Abraham R. Springer; 2004. mTOR as a Positive Regulator of Tumor Cell Responses to hypoxia, TOR; pp. 299–319. [DOI] [PubMed] [Google Scholar]
  • 326.Rui M., Xin Y., Li R., Ge Y., Feng C., Xu X. Targeted biomimetic nanoparticles for synergistic combination chemotherapy of paclitaxel and doxorubicin. Mol. Pharm. 2017;14:107–123. doi: 10.1021/acs.molpharmaceut.6b00732. [DOI] [PubMed] [Google Scholar]
  • 327.Zhu D., Wu S., Hu C., Chen Z., Wang H., Fan F., Qin Y., Wang C., Sun H., Leng X. Folate-targeted polymersomes loaded with both paclitaxel and doxorubicin for the combination chemotherapy of hepatocellular carcinoma. Acta Biomater. 2017;58:399–412. doi: 10.1016/j.actbio.2017.06.017. [DOI] [PubMed] [Google Scholar]
  • 328.Garcia G., Odaimi M. Systemic combination chemotherapy in elderly pancreatic cancer: a review. J. Gastrointest. Cancer. 2017;48:121–128. doi: 10.1007/s12029-017-9930-0. [DOI] [PubMed] [Google Scholar]
  • 329.Que W.-C., Huang Y.-F., Lin X.-Y., Lan Y.-Q., Gao X.-Y., Wang X.-L., Wu R.-P., Du B., Huang X.-B., Qiu H.-q. Paclitaxel, 5-fluorouracil, and leucovorin combination chemotherapy as first-line treatment in patients with advanced gastric cancer. Anti-Cancer Drugs. 2019;30:302–307. doi: 10.1097/CAD.0000000000000735. [DOI] [PubMed] [Google Scholar]
  • 330.Choi J.-H., Choi Y.W., Kang S.Y., Jeong G.S., Lee H.W., Jeong S.H., Park J.S., Ahn M.S., Sheen S.S. Combination versus single-agent as palliative chemotherapy for gastric cancer. BMC Cancer. 2020;20:1–10. doi: 10.1186/s12885-020-6666-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 331.Lan Y.-q., Kong L.-j., Lin X.-y., Xu Q., Gao X.-y., Wu R.-p., Wang X.-l., Zhong D.-t. Combination chemotherapy with paclitaxel and oxaliplatin as first-line treatment in patients with advanced gastric cancer. Cancer Chemother. Pharmacol. 2018;81:1007–1015. doi: 10.1007/s00280-018-3576-x. [DOI] [PubMed] [Google Scholar]
  • 332.Ohnuma H., Sato Y., Hirakawa M., Kikuchi S., Miyanishi K., Sagawa T., Takahashi Y., Nobuoka T., Okamoto K., Miyamoto H. Docetaxel, cisplatin and S-1 (DCS) combination chemotherapy for gastric cancer patients with peritoneal metastasis: a retrospective study. Cancer Chemother. Pharmacol. 2018;81:539–548. doi: 10.1007/s00280-018-3523-x. [DOI] [PubMed] [Google Scholar]
  • 333.Li H., Zhang Z., Gao C., Wu S., Duan Q., Wu H., Wang C., Shen Q., Yin T. Combination chemotherapy of valproic acid (VPA) and gemcitabine regulates STAT3/Bmi1 pathway to differentially potentiate the motility of pancreatic cancer cells. Cell Biosci. 2019;9:1–15. doi: 10.1186/s13578-019-0312-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 334.Duan J., Yang Z., Liu D., Shi Y. Clinical efficacy of bevacizumab combined with gemcitabine and cisplatin combination chemotherapy in the treatment of advanced non-small cell lung cancer. J. BUON. 2018;23:1402–1406. [PubMed] [Google Scholar]
  • 335.Wu H., Jin H., Wang C., Zhang Z., Ruan H., Sun L., Yang C., Li Y., Qin W., Wang C. Synergistic cisplatin/doxorubicin combination chemotherapy for multidrug-resistant cancer via polymeric nanogels targeting delivery. ACS Appl. Mater. Interfaces. 2017;9:9426–9436. doi: 10.1021/acsami.6b16844. [DOI] [PubMed] [Google Scholar]
  • 336.Rahimi M., Safa K.D., Salehi R. Co-delivery of doxorubicin and methotrexate by dendritic chitosan-g-mPEG as a magnetic nanocarrier for multi-drug delivery in combination chemotherapy. Polym. Chemi. 2017;8:7333–7350. [Google Scholar]
  • 337.Bao Y., Zhang S., Chen Z., Chen A.T., Ma J., Deng G., Xu W., Zhou J., Yu Z.-Q., Yao G. Synergistic chemotherapy for breast cancer and breast cancer brain metastases via paclitaxel-loaded oleanolic acid nanoparticles. Mol. Pharm. 2020;17:1343–1351. doi: 10.1021/acs.molpharmaceut.0c00044. [DOI] [PubMed] [Google Scholar]
  • 338.Bai T., Zhu B., Du Z., Shi J., Shao D., Kong J. Amphiphilic star copolymers-mediated co-delivery of doxorubicin and avasimibe for effective combination chemotherapy. J. Mater. Sci. 2020;55:9525–9537. [Google Scholar]
  • 339.Takahashi N., Sunaga T., Fujimiya T., Kurihara T., Nagatani A., Yamagishi M., Watanabe T., Sasaki H., Ogawa Y., Sasaki T. Risk associated with severe hematological toxicity in patients with urothelial cancer receiving combination chemotherapy of gemcitabine and cisplatin. Chemotherapy. 2020;65:29–34. doi: 10.1159/000508805. [DOI] [PubMed] [Google Scholar]
  • 340.Toffalorio F., Santarpia M., Radice D., Jaramillo C.A., Spitaleri G., Manzotti M., Catania C., Jordheim L.P., Pelosi G., Peters G.J. 5′-nucleotidase cN-II emerges as a new predictive biomarker of response to gemcitabine/platinum combination chemotherapy in non-small cell lung cancer. Oncotarget. 2018;9:16437. doi: 10.18632/oncotarget.24505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 341.Xiao Y., Liu J., Guo M., Zhou H., Jin J., Liu J., Liu Y., Zhang Z., Chen C. Synergistic combination chemotherapy using carrier-free celastrol and doxorubicin nanocrystals for overcoming drug resistance. Nanoscale. 2018;10:12639–12649. doi: 10.1039/c8nr02700e. [DOI] [PubMed] [Google Scholar]
  • 342.Ashrafizadeh M., Zarrabi A., Hashemi F., Moghadam E.R., Hashemi F., Entezari M., Hushmandi K., Mohammadinejad R., Najafi M. Curcumin in cancer therapy: A novel adjunct for combination chemotherapy with paclitaxel and alleviation of its adverse effects. Life Sci. 2020 doi: 10.1016/j.lfs.2020.117984. [DOI] [PubMed] [Google Scholar]
  • 343.Higashiguchi M., Yamada D., Akita H., Eguchi H., Iwagami Y., Asaoka T., Noda T., Gotoh K., Kobayashi S., Sakai D. Successful R0 resection of hilar cholangiocarcinoma by extrahepatic bile duct resection due to accompanying liver dysfunction after neoadjuvant gemcitabine/cisplatin/S-1 combination chemotherapy-a case report, gan to kagaku ryoho. Cancer Chemother. 2019;46:342–344. [PubMed] [Google Scholar]
  • 344.Genovese L., Chandrabhatla T., Bagnola A.J., Tran H., Naderi N. Combination chemotherapy with a shocking result. J. Am. Coll. Cardiol. 2020;75:3242. 3242. [Google Scholar]
  • 345.Lee S.-I., Celik S., Logsdon B.A., Lundberg S.M., Martins T.J., Oehler V.G., Estey E.H., Miller C.P., Chien S., Dai J. A machine learning approach to integrate big data for precision medicine in acute myeloid leukemia. Nat. Commun. 2018;9:1–13. doi: 10.1038/s41467-017-02465-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 346.Dugger S.A., Platt A., Goldstein D.B. Drug development in the era of precision medicine. Nat. Rev. Drug Discov. 2018;17:183. doi: 10.1038/nrd.2017.226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 347.Krzyszczyk P., Acevedo A., Davidoff E.J., Timmins L.M., Marrero-Berrios I., Patel M., White C., Lowe C., Sherba J.J., Hartmanshenn C. The growing role of precision and personalized medicine for cancer treatment. Technology. 2018;6:79–100. doi: 10.1142/S2339547818300020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 348.Mazumder A., Cerella C., Diederich M. Natural scaffolds in anticancer therapy and precision medicine. Biotechnol. Adv. 2018;36:1563–1585. doi: 10.1016/j.biotechadv.2018.04.009. [DOI] [PubMed] [Google Scholar]
  • 349.Li C.-C., Shen Z., Bavarian R., Yang F., Bhattacharya A. Oral cancer: genetics and the role of precision medicine. Dent. Clin. 2018;62:29–46. doi: 10.1016/j.cden.2017.08.002. [DOI] [PubMed] [Google Scholar]
  • 350.Kim D.-H., Kim Y.-S., Son N.-I., Kang C.-K., Kim A.-R. Recent omics technologies and their emerging applications for personalised medicine. IET Syst. Biol. 2017;11:87–98. doi: 10.1049/iet-syb.2016.0016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 351.Cammarota G., Ianiro G., Ahern A., Carbone C., Temko A., Claesson M.J., Gasbarrini A., Tortora G. Gut microbiome, big data and machine learning to promote precision medicine for cancer. Nat. Rev. Gastroenterol. Hepatol. 2020;17:635–648. doi: 10.1038/s41575-020-0327-3. [DOI] [PubMed] [Google Scholar]
  • 352.Lheureux S., Braunstein M., Oza A.M. Epithelial ovarian cancer: evolution of management in the era of precision medicine. CA: Cancer J. Clin. 2019;69:280–304. doi: 10.3322/caac.21559. [DOI] [PubMed] [Google Scholar]
  • 353.Magee D., Hird A., Klaassen Z., Sridhar S., Nam R., Wallis C., Kulkarni G. Adverse event profile for immunotherapy agents compared with chemotherapy in solid organ tumors: a systematic review and meta-analysis of randomized clinical trials. Ann. Oncol. 2020;31:50–60. doi: 10.1016/j.annonc.2019.10.008. [DOI] [PubMed] [Google Scholar]
  • 354.Tan A., Porcher R., Crequit P., Ravaud P., Dechartres A. Differences in treatment effect size between overall survival and progression-free survival in immunotherapy trials: a meta-epidemiologic study of trials with results posted at ClinicalTrials. gov. J. Clin. Oncol. 2017;35:1686–1694. doi: 10.1200/JCO.2016.71.2109. [DOI] [PubMed] [Google Scholar]
  • 355.Steuer C.E., Ramalingam S.S. Tumor mutation burden: leading immunotherapy to the era of precision medicine. J. Clin. Oncol. 2018;36:631–632. doi: 10.1200/JCO.2017.76.8770. [DOI] [PubMed] [Google Scholar]
  • 356.Kaidar-Person O., Gil Z., Billan S. Precision medicine in head and neck cancer. Drug Resist. Updat. 2018;40:13–16. doi: 10.1016/j.drup.2018.09.001. [DOI] [PubMed] [Google Scholar]
  • 357.Kumar N., Cramer G.M., Dahaj S.A.Z., Sundaram B., Celli J.P., Kulkarni R.V. Stochastic modeling of phenotypic switching and chemoresistance in cancer cell populations. Sci. Rep. 2019;9:1–10. doi: 10.1038/s41598-019-46926-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 358.Fodale V., Pierobon M., Liotta L., Petricoin E. Mechanism of cell adaptation: when and how do cancer cells develop chemoresistance? Cancer J. (Sudbury, Mass.) 2011;17:89. doi: 10.1097/PPO.0b013e318212dd3d. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 359.Belisario D.C., Kopecka J., Pasino M., Akman M., De Smaele E., Donadelli M., Riganti C. Hypoxia dictates metabolic rewiring of tumors: implications for chemoresistance. Cells. 2020;9:2598. doi: 10.3390/cells9122598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 360.Amin A., Plimack E.R., Ernstoff M.S., Lewis L.D., Bauer T.M., McDermott D.F., Carducci M., Kollmannsberger C., Rini B.I., Heng D.Y. Safety and efficacy of nivolumab in combination with sunitinib or pazopanib in advanced or metastatic renal cell carcinoma: the CheckMate 016 study. J. Immunother. Cancer. 2018;6:1–12. doi: 10.1186/s40425-018-0420-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 361.Bachelot T., Romieu G., Campone M., Diéras V., Cropet C., Dalenc F., Jimenez M., Le Rhun E., Pierga J.-Y., Gonçalves A. Lapatinib plus capecitabine in patients with previously untreated brain metastases from HER2-positive metastatic breast cancer (LANDSCAPE): a single-group phase 2 study. Lancet Oncol. 2013;14:64–71. doi: 10.1016/S1470-2045(12)70432-1. [DOI] [PubMed] [Google Scholar]
  • 362.Bang Y.-J., Ruiz E.Y., Van Cutsem E., Lee K.-W., Wyrwicz L., Schenker M., Alsina M., Ryu M.-H., Chung H.-C., Evesque L. Phase III, randomised trial of avelumab versus physician's choice of chemotherapy as third-line treatment of patients with advanced gastric or gastro-oesophageal junction cancer: primary analysis of JAVELIN Gastric 300. Ann. Oncol. 2018;29:2052–2060. doi: 10.1093/annonc/mdy264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 363.Barlesi F., Vansteenkiste J., Spigel D., Ishii H., Garassino M., de Marinis F., Özgüroğlu M., Szczesna A., Polychronis A., Uslu R. Avelumab versus docetaxel in patients with platinum-treated advanced non-small-cell lung cancer (JAVELIN Lung 200): an open-label, randomised, phase 3 study. Lancet Oncol. 2018;19:1468–1479. doi: 10.1016/S1470-2045(18)30673-9. [DOI] [PubMed] [Google Scholar]
  • 364.Bellmunt J., De Wit R., Vaughn D.J., Fradet Y., Lee J.-L., Fong L., Vogelzang N.J., Climent M.A., Petrylak D.P., Choueiri T.K. Pembrolizumab as second-line therapy for advanced urothelial carcinoma. N. Engl. J. Med. 2017;376:1015–1026. doi: 10.1056/NEJMoa1613683. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 365.Birtle A., Johnson M., Chester J., Jones R., Dolling D., Bryan R.T., Harris C., Winterbottom A., Blacker A., Catto J.W. Adjuvant chemotherapy in upper tract urothelial carcinoma (the POUT trial): a phase 3, open-label, randomised controlled trial. Lancet. 2020;395:1268–1277. doi: 10.1016/S0140-6736(20)30415-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 366.Borghaei H., Paz-Ares L., Horn L., Spigel D.R., Steins M., Ready N.E., Chow L.Q., Vokes E.E., Felip E., Holgado E. Nivolumab versus docetaxel in advanced nonsquamous non–small-cell lung cancer. N. Engl. J. Med. 2015;373:1627–1639. doi: 10.1056/NEJMoa1507643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 367.Brahmer J., Reckamp K.L., Baas P., Crinò L., Eberhardt W.E., Poddubskaya E., Antonia S., Pluzanski A., Vokes E.E., Holgado E. Nivolumab versus docetaxel in advanced squamous-cell non–small-cell lung cancer. N. Engl. J. Med. 2015;373:123–135. doi: 10.1056/NEJMoa1504627. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 368.Carbone D.P., Reck M., Paz-Ares L., Creelan B., Horn L., Steins M., Felip E., van den Heuvel M.M., Ciuleanu T.-E., Badin F. First-line nivolumab in stage IV or recurrent non–small-cell lung cancer. N. Engl. J. Med. 2017;376:2415–2426. doi: 10.1056/NEJMoa1613493. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 369.Cohen E.E., Soulières D., Tourneau C.Le, Dinis J., Licitra L., Ahn M.-J., Soria A., Machiels J.-P., Mach N., Mehra R. Pembrolizumab versus methotrexate, docetaxel, or cetuximab for recurrent or metastatic head-and-neck squamous cell carcinoma (KEYNOTE-040): a randomised, open-label, phase 3 study. Lancet. 2019;393:156–167. doi: 10.1016/S0140-6736(18)31999-8. [DOI] [PubMed] [Google Scholar]
  • 370.Coleman R.L., Enserro D., Spirtos N., Herzog T.J., Sabbatini P., Armstrong D.K., Kim B., Fujiwara K., Walker J.L., Flynn P.J. American Society of Clinical Oncology; 2018. A Phase III Randomized Controlled Trial of Secondary Surgical Cytoreduction (SSC) Followed By Platinum-Based Combination Chemotherapy (PBC), with or without Bevacizumab (B) in platinum-sensitive, Recurrent Ovarian Cancer (PSOC): A NRG Oncology/Gynecologic Oncology Group (GOG) Study. [Google Scholar]
  • 371.Fehrenbacher L., Spira A., Ballinger M., Kowanetz M., Vansteenkiste J., Mazieres J., Park K., Smith D., Artal-Cortes A., Lewanski C. Atezolizumab versus docetaxel for patients with previously treated non-small-cell lung cancer (POPLAR): a multicentre, open-label, phase 2 randomised controlled trial. Lancet. 2016;387:1837–1846. doi: 10.1016/S0140-6736(16)00587-0. [DOI] [PubMed] [Google Scholar]
  • 372.Ferris R.L., Blumenschein Jr G., Fayette J., Guigay J., Colevas A.D., Licitra L., Harrington K., Kasper S., Vokes E.E., Even C. Nivolumab for recurrent squamous-cell carcinoma of the head and neck. N. Engl. J. Med. 2016;375:1856–1867. doi: 10.1056/NEJMoa1602252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 373.Finn R.S., Martin M., Rugo H.S., Jones S., Im S.-A., Gelmon K., Harbeck N., Lipatov O.N., Walshe J.M., Moulder S. Palbociclib and letrozole in advanced breast cancer. N. Engl. J. Med. 2016;375:1925–1936. doi: 10.1056/NEJMoa1607303. [DOI] [PubMed] [Google Scholar]
  • 374.Flippot R., Dalban C., Laguerre B., Borchiellini D., Gravis G., Négrier S., Chevreau C., Joly F., Geoffrois L., Ladoire S. Safety and efficacy of nivolumab in brain metastases from renal cell carcinoma: results of the GETUG-AFU 26 NIVOREN multicenter phase II study. J. Clin. Oncol. 2019;37:2008–2016. doi: 10.1200/JCO.18.02218. [DOI] [PubMed] [Google Scholar]
  • 375.Freedman R.A., Gelman R.S., Anders C.K., Melisko M.E., Parsons H.A., Cropp A.M., Silvestri K., Cotter C.M., Componeschi K.P., Marte J.M. TBCRC 022: a phase II trial of neratinib and capecitabine for patients with human epidermal growth factor receptor 2–positive breast cancer and brain metastases. J. Clin. Oncol. 2019;37:1081. doi: 10.1200/JCO.18.01511. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 376.Goldberg S.B., Gettinger S.N., Mahajan A., Chiang A.C., Herbst R.S., Sznol M., Tsiouris A.J., Cohen J., Vortmeyer A., Jilaveanu L. Pembrolizumab for patients with melanoma or non-small-cell lung cancer and untreated brain metastases: early analysis of a non-randomised, open-label, phase 2 trial. Lancet Oncol. 2016;17:976–983. doi: 10.1016/S1470-2045(16)30053-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 377.Hara H., Kawamoto T., Fukase N., Kawakami Y., Takemori T., Fujiwara S., Kitayama K., Nishida K., Kuroda R., Akisue T. Gemcitabine and docetaxel combination chemotherapy for advanced bone and soft tissue sarcomas: protocol for an open-label, non-randomised, Phase 2 study. BMC Cancer. 2019;19:1–6. doi: 10.1186/s12885-019-5923-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 378.Hellmann M.D., Ciuleanu T.-E., Pluzanski A., Lee J.S., Otterson G.A., Audigier-Valette C., Minenza E., Linardou H., Burgers S., Salman P. Nivolumab plus ipilimumab in lung cancer with a high tumor mutational burden. N. Engl. J. Med. 2018;378:2093–2104. doi: 10.1056/NEJMoa1801946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 379.Herbst R.S., Baas P., Kim D.-W., Felip E., Pérez-Gracia J.L., Han J.-Y., Molina J., Kim J.-H., Arvis C.D., Ahn M.-J. Pembrolizumab versus docetaxel for previously treated, PD-L1-positive, advanced non-small-cell lung cancer (KEYNOTE-010): a randomised controlled trial. Lancet. 2016;387:1540–1550. doi: 10.1016/S0140-6736(15)01281-7. [DOI] [PubMed] [Google Scholar]
  • 380.Keilholz U., Mehnert J.M., Bauer S., Bourgeois H., Patel M.R., Gravenor D., Nemunaitis J.J., Taylor M.H., Wyrwicz L., Lee K.-W. Avelumab in patients with previously treated metastatic melanoma: phase 1b results from the JAVELIN Solid Tumor trial. J. Immunother. Cancer. 2019;7:1–13. doi: 10.1186/s40425-018-0459-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 381.Kluger H.M., Chiang V., Mahajan A., Zito C.R., Sznol M., Tran T., Weiss S.A., Cohen J.V., Yu J., Hegde U. Long-term survival of patients with melanoma with active brain metastases treated with pembrolizumab on a phase II trial. J. Clin. Oncol. 2019;37:52. doi: 10.1200/JCO.18.00204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 382.Lee J.J., Yothers G., Jacobs S.A., Sanoff H.K., Cohen D.J., Guthrie K.A., Henry N.L., Ganz P.A., Kopetz S., Lucas P.C. American Society of Clinical Oncology; 2019. Colorectal Cancer Metastatic dMMR Immuno-Therapy (COMMIT) Study (NRG-GI004/SWOG-S1610): A randomized Phase III Study of Mfolfox6/Bevacizumab Combination Chemotherapy With Or Without Atezolizumab Or Atezolizumab Monotherapy in the First-Line Treatment of Patients (pts) With Deficient DNA Mismatch Repair (dMMR) Metastatic Colorectal Cancer (mCRC) [Google Scholar]
  • 383.Lee K.-W., Chung I.-J., Ryu M.-H., Park Y.I., Nam B.-H., Oh H.-S., Lee K.H., Han H.S., Seo B.-G., Jo J.-C. Multicenter phase III trial of S-1 and cisplatin versus S-1 and oxaliplatin combination chemotherapy for first-line treatment of advanced gastric cancer (SOPP trial) Gastric Cancer. 2021;24:156–167. doi: 10.1007/s10120-020-01101-4. [DOI] [PubMed] [Google Scholar]
  • 384.Loibl S., O'Shaughnessy J., Untch M., Sikov W.M., Rugo H.S., McKee M.D., Huober J., Golshan M., von Minckwitz G., Maag D. Addition of the PARP inhibitor veliparib plus carboplatin or carboplatin alone to standard neoadjuvant chemotherapy in triple-negative breast cancer (BrighTNess): a randomised, phase 3 trial. Lancet Oncol. 2018;19:497–509. doi: 10.1016/S1470-2045(18)30111-6. [DOI] [PubMed] [Google Scholar]
  • 385.Ma F., Ouyang Q., Li W., Jiang Z., Tong Z., Liu Y., Li H., Yu S., Feng J., Wang S. Pyrotinib or lapatinib combined with capecitabine in HER2–positive metastatic breast cancer with prior taxanes, anthracyclines, and/or trastuzumab: a randomized, phase II study. J. Clin. Oncol. 2019;37:2610–2619. doi: 10.1200/JCO.19.00108. [DOI] [PubMed] [Google Scholar]
  • 386.Mok T.S., Wu Y.-L., Kudaba I., Kowalski D.M., Cho B.C., Turna H.Z., Castro Jr G., Srimuninnimit V., Laktionov K.K., Bondarenko I. Pembrolizumab versus chemotherapy for previously untreated, PD-L1-expressing, locally advanced or metastatic non-small-cell lung cancer (KEYNOTE-042): a randomised, open-label, controlled, phase 3 trial. Lancet. 2019;393:1819–1830. doi: 10.1016/S0140-6736(18)32409-7. [DOI] [PubMed] [Google Scholar]
  • 387.Nagao S., Kogiku A., Suzuki K., Shibutani T., Yamamoto K., Jimi T., Kitai M., Shiozaki T., Matsuoka K., Yamaguchi S. A phase II study of the combination chemotherapy of bevacizumab and gemcitabine in women with platinum-resistant recurrent epithelial ovarian, primary peritoneal, or fallopian tube cancer. J. Ovarian Res. 2020;13:14. doi: 10.1186/s13048-020-0617-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 388.Nakamoto Y., Noda M., Mikami R., Tokunaga Y., Okumoto T., Kawamura T., Fujiwara H., Doi S., Tomita N. Phase II study of S-1-based sequential combination chemotherapy including oxaliplatin plus bevacizumab and irinotecan with or without cetuximab for metastatic colorectal cancer: the SOBIC trial. Int. J. Clin. Oncol. 2020:1–6. doi: 10.1007/s10147-020-01657-2. [DOI] [PubMed] [Google Scholar]
  • 389.Powles T., Durán I., Van Der Heijden M.S., Loriot Y., Vogelzang N.J., De Giorgi U., Oudard S., Retz M.M., Castellano D., Bamias A. Atezolizumab versus chemotherapy in patients with platinum-treated locally advanced or metastatic urothelial carcinoma (IMvigor211): a multicentre, open-label, phase 3 randomised controlled trial. Lancet. 2018;391:748–757. doi: 10.1016/S0140-6736(17)33297-X. [DOI] [PubMed] [Google Scholar]
  • 390.Pujol J.-L., Greillier L., Audigier-Valette C., Moro-Sibilot D., Uwer L., Hureaux J., Guisier F., Carmier D., Madelaine J., Otto J. A randomized non-comparative phase II study of anti-programmed cell death-ligand 1 atezolizumab or chemotherapy as second-line therapy in patients with small cell lung cancer: results from the IFCT-1603 trial. J. Thoracic Oncol. 2019;14:903–913. doi: 10.1016/j.jtho.2019.01.008. [DOI] [PubMed] [Google Scholar]
  • 391.Reck M., Rodríguez-Abreu D., Robinson A.G., Hui R., Csőszi T., Fülöp A., Gottfried M., Peled N., Tafreshi A., Cuffe S. Pembrolizumab versus chemotherapy for PD-L1–positive non–small-cell lung cancer. N. Engl. J. Med. 2016;375:1823–1833. doi: 10.1056/NEJMoa1606774. [DOI] [PubMed] [Google Scholar]
  • 392.Ribas A., Kefford R., Marshall M.A., Punt C.J., Haanen J.B., Marmol M., Garbe C., Gogas H., Schachter J., Linette G. Phase III randomized clinical trial comparing tremelimumab with standard-of-care chemotherapy in patients with advanced melanoma. J. Clin. Oncol. 2013;31:616. doi: 10.1200/JCO.2012.44.6112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 393.Ribas A., Puzanov I., Dummer R., Schadendorf D., Hamid O., Robert C., Hodi F.S., Schachter J., Pavlick A.C., Lewis K.D. Pembrolizumab versus investigator-choice chemotherapy for ipilimumab-refractory melanoma (KEYNOTE-002): a randomised, controlled, phase 2 trial. Lancet Oncol. 2015;16:908–918. doi: 10.1016/S1470-2045(15)00083-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 394.Rittmeyer A., Barlesi F., Waterkamp D., Park K., Ciardiello F., von Pawel J., Gadgeel S.M., Hida T., Kowalski D.M., Dols M.C. Atezolizumab versus docetaxel in patients with previously treated non-small-cell lung cancer (OAK): a phase 3, open-label, multicentre randomised controlled trial. Lancet. 2017;389:255–265. doi: 10.1016/S0140-6736(16)32517-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 395.Robert C., Long G.V., Brady B., Dutriaux C., Maio M., Mortier L., Hassel J.C., Rutkowski P., McNeil C., Kalinka-Warzocha E. Nivolumab in previously untreated melanoma without BRAF mutation. N. Engl. J. Med. 2015;372:320–330. doi: 10.1056/NEJMoa1412082. [DOI] [PubMed] [Google Scholar]
  • 396.Shitara K., Özgüroğlu M., Bang Y.-J., Di Bartolomeo M., Mandalà M., Ryu M.-H., Fornaro L., Olesiński T., Caglevic C., Chung H.C. Pembrolizumab versus paclitaxel for previously treated, advanced gastric or gastro-oesophageal junction cancer (KEYNOTE-061): a randomised, open-label, controlled, phase 3 trial. Lancet. 2018;392:123–133. doi: 10.1016/S0140-6736(18)31257-1. [DOI] [PubMed] [Google Scholar]
  • 397.Toulmonde M., Pulido M., Ray-Coquard I., Andre T., Isambert N., Chevreau C., Penel N., Bompas E., Saada E., Bertucci F. Pazopanib or methotrexate–vinblastine combination chemotherapy in adult patients with progressive desmoid tumours (DESMOPAZ): a non-comparative, randomised, open-label, multicentre, phase 2 study. Lancet Oncol. 2019;20:1263–1272. doi: 10.1016/S1470-2045(19)30276-1. [DOI] [PubMed] [Google Scholar]
  • 398.Toyama H., Sugiura T., Fukutomi A., Asakura H., Takeda Y., Yamamoto K., Hirano S., Satoi S., Matsumoto I., Takahashi S. American Society of Clinical Oncology; 2020. Randomized Phase II Trial of Chemoradiotherapy With S-1 Versus Combination Chemotherapy with Gemcitabine and S-1 as Neoadjuvant Treatment for Resectable Pancreatic Cancer (JASPAC 04) [DOI] [PubMed] [Google Scholar]
  • 399.Vogel A., Kasper S., Bitzer M., Block A., Sinn M., Schulze-Bergkamen H., Moehler M., Pfarr N., Endris V., Goeppert B. PICCA study: panitumumab in combination with cisplatin/gemcitabine chemotherapy in KRAS wild-type patients with biliary cancer—a randomised biomarker-driven clinical phase II AIO study. Eur. J. Cancer. 2018;92:11–19. doi: 10.1016/j.ejca.2017.12.028. [DOI] [PubMed] [Google Scholar]
  • 400.Weber J.S., D'Angelo S.P., Minor D., Hodi F.S., Gutzmer R., Neyns B., Hoeller C., Khushalani N.I., Miller Jr W.H., Lao C.D. Nivolumab versus chemotherapy in patients with advanced melanoma who progressed after anti-CTLA-4 treatment (CheckMate 037): a randomised, controlled, open-label, phase 3 trial. Lancet Oncol. 2015;16:375–384. doi: 10.1016/S1470-2045(15)70076-8. [DOI] [PubMed] [Google Scholar]
  • 401.Wu Y.-L., Lu S., Cheng Y., Zhou C., Wang J., Mok T., Zhang L., Tu H.-Y., Wu L., Feng J. Nivolumab versus docetaxel in a predominantly Chinese patient population with previously treated advanced NSCLC: checkMate 078 randomized phase III clinical trial. J. Thoracic Oncol. 2019;14:867–875. doi: 10.1016/j.jtho.2019.01.006. [DOI] [PubMed] [Google Scholar]

Articles from Translational Oncology are provided here courtesy of Neoplasia Press

RESOURCES