International Journal of Molecular Medicine is an international journal devoted to molecular mechanisms of human disease.
International Journal of Oncology is an international journal devoted to oncology research and cancer treatment.
Covers molecular medicine topics such as pharmacology, pathology, genetics, neuroscience, infectious diseases, molecular cardiology, and molecular surgery.
Oncology Reports is an international journal devoted to fundamental and applied research in Oncology.
Experimental and Therapeutic Medicine is an international journal devoted to laboratory and clinical medicine.
Oncology Letters is an international journal devoted to Experimental and Clinical Oncology.
Explores a wide range of biological and medical fields, including pharmacology, genetics, microbiology, neuroscience, and molecular cardiology.
International journal addressing all aspects of oncology research, from tumorigenesis and oncogenes to chemotherapy and metastasis.
Multidisciplinary open-access journal spanning biochemistry, genetics, neuroscience, environmental health, and synthetic biology.
Open-access journal combining biochemistry, pharmacology, immunology, and genetics to advance health through functional nutrition.
Publishes open-access research on using epigenetics to advance understanding and treatment of human disease.
An International Open Access Journal Devoted to General Medicine.
The incidence of tumors is rising due to population growth, aging and advancements in tumor diagnostics (1), making malignant tumors a significant threat to human health and life (2,3). Cancer treatment options include surgery, radiotherapy, chemotherapy, targeted therapy, hormone therapy and immunotherapy (1), with chemotherapy being the most widely used across various cancer types. However, chemoresistance remains a complex clinical challenge, leading to the majority of cancer-related deaths (4,5). Understanding the mechanisms of chemoresistance is therefore essential for enhancing therapeutic efficacy, and the present article reviewed the latest developments in tumor chemoresistance. Notably, since numerous studies have already detailed the drug resistance mechanisms of molecular targeted therapies, the present review did not delve into those aspects (6–8).
Anatomical barriers that limit or obstruct the entry of foreign substances into target tissues or cells play a significant role in tumor chemoresistance. For example, in mammals, the blood-brain barrier, composed of tightly connected endothelial cells, basement membranes and astrocytes, restricts the passage of most chemotherapeutic drugs, thereby reducing their efficacy against central nervous system cancers. Mechanosensitive Sox2+ tumor cells, for instance, develop chemoresistance by constructing a blood-tumor barrier around capillaries (9). Similarly, the blood-testis barrier, formed by testicular supporting cells, protects germ cells from toxic substances. Sertoli cells, which are resistant to apoptosis despite strong inflammatory stimuli (for example, LPS and IL-18), play a pivotal role in maintaining the homeostasis of the testicular environment, and this anti-apoptotic and anti-inflammatory capability further enhances their resistance to chemotherapy (10).
Chemotherapy resistance in cancer can be categorized into inherent and secondary resistance based on its origin. Inherent resistance refers to the presence of drug-resistant cells within the tumor tissue before treatment begins (11). This concept suggests that drug resistance is a ‘fait accompli’, with relapse occurring once preexisting resistant cells repopulate (12). By contrast, secondary drug resistance arises when tumor cells develop resistance to chemotherapeutic drugs due to intracellular changes induced by the drugs themselves (Fig. 1).
Chemotherapy resistance can also be classified into primary drug resistance and multidrug resistance, depending on the response of tumor cells to chemotherapy agents (13). Primary drug resistance occurs when tumor cells develop resistance to a therapeutic drug after treatment, commonly due to decreased drug uptake or increased drug efflux. Multidrug resistance, on the other hand, involves tumor cells becoming resistant not only to the therapeutic agent but also to other chemotherapeutic drugs to which they have not been previously exposed. This phenomenon is driven by mechanisms such as enhanced drug efflux, increased DNA repair, neutralization of chemotherapeutic agents, closure of nuclear pores and tumor cell dormancy.
Chemotherapeutic agents encompass a wide range of categories, including alkylating agents, antimetabolites, antibiotics and antitumor phytopharmaceuticals, among others. Although these agents employ diverse mechanisms of action, the majority must penetrate the cell nucleus to exert their therapeutic effects (Fig. 2).
Alkylating agents, such as cyclophosphamide and cisplatin, are capable of targeting both proliferative and non-proliferative cells. These agents possess active alkylating groups that generate electrophilic groups with positive carbon ions. These ions form covalent bonds with various nucleophilic groups within cells, leading to cross-linking between DNA molecules or between DNA and proteins, ultimately causing disruptions in DNA replication and transcription, culminating in cell death (14,15). Furthermore, alkylating agents have been shown to induce protein oxidation and acetylation, which can impair the lipid homeostasis of the nuclear membrane and destabilize the cellular genome (15). Cisplatin, a cis-alkylating agent, specifically induces cross-links within DNA double strands, resulting in the cessation of DNA replication and transcription, or even causing DNA strand breaks and apoptosis (16). As the first metal-based chemotherapeutic drug, cisplatin has been extensively used, though its side effects, including ototoxicity and nephrotoxicity, have posed significant challenges in its clinical application (17,18).
Antimetabolites, structurally similar to natural metabolites but lacking their corresponding biological functions, competitively inhibit and disrupt nucleic acid and protein synthesis, ultimately leading to cell death (19). The most notable example is 5-fluorouracil, which inhibits thymidine synthetase, thereby impairing DNA synthesis within the body (20).
Antitumor antibiotics, derived from microorganisms, exert their effects by directly damaging DNA or intercalating within it, thereby disrupting transcription (21). A key example is doxorubicin, which is extensively employed in the treatment of hematological malignancies, breast cancer and lung cancer (22).
Plant-derived antitumor agents, such as paclitaxel and vincristine, inhibit tumor cell proliferation by targeting microtubule proteins. Paclitaxel primarily prevents the depolymerization of microtubules during mitosis, whereas vincristine inhibits microtubule polymerization (23). Despite the clinical success of these plant-derived anticancer drugs, tumor resistance to them remains prevalent, particularly in some gastrointestinal cancers that exhibit resistance from the onset of chemotherapy.
Plant-derived chemotherapeutic agents, such as paclitaxel and vincristine, possess favorable lipid solubility, enabling them to permeate cells via passive diffusion through the lipid bilayer of the cell membrane (24,25). By contrast, cisplatin, a metal-based anticancer drug, increases its water solubility by replacing chloride ions with Pt(NH3)2(H2O)2. However, this substitution process occurs slowly, necessitating that cisplatin primarily enters cells through active transport mechanisms (26).
Research has identified that defective expression of copper ion transporter protein 1 (CTR1) correlates with reduced platinum accumulation in tissues, leading to decreased therapeutic efficacy of platinum-based drugs in various tumors (27). Under basal conditions, CTR1 is predominantly localized in the perinuclear region of cultured cells, but in the presence of copper complexes, CTR1 is highly expressed on the cell membrane, facilitating cisplatin's active transport. The uptake of cisplatin mediated by CTR1 depends on its metal-bound extracellular region, which is connected to the N-glycan chain at Asn15 and the O-glycan chain at Thr27 (28). Although cisplatin does not alter the subcellular localization of CTR1, co-localization studies using fluorescent cisplatin derivatives revealed their presence in CTR1-associated vesicular structures, indicating that cisplatin is endocytosed concurrently with CTR1 and transported into the cell (29).
Organic cation transporters (OCTs) are essential for the cellular uptake of endogenous cationic substrates, hydrophilic exogenous compounds, and platinum-based anticancer drugs. OCTs are highly expressed in the renal basement membranes of humans and mice, playing a significant role in mediating cisplatin uptake (30). Elevated levels of OCT2 in pre-chemotherapy biopsy samples have been associated with improved prognosis when treated with cisplatin-inclusive regimens (31). Conversely, in OCT2 knockout (−/-) mouse models, both nephrotoxicity and ototoxicity were significantly reduced following cisplatin treatment (32). Although overexpressing transport proteins exogenously is not yet feasible in clinical practice, current strategies focus on developing novel carriers, such as nanomaterial conjugates, to enhance the intracellular delivery of chemotherapeutic drugs. This approach aims to increase the intracellular concentration of these drugs, thereby enhancing their efficacy in tumor cell eradication.
Members of the ABC protein transporter family, primarily located in cell membranes, are responsible for expelling cytotoxic substances, including chemotherapeutic drugs, out of the cell by utilizing ATP hydrolysis for energy (33). This process impedes the accumulation of therapeutic drug concentrations in target cells or organs, thereby contributing to chemoresistance (34). Moreover, ABC transporter proteins have been implicated in promoting tumor cell migration and invasion (35). The ABC transporter family is categorized into seven subgroups (ABCA-ABCG) comprising 48 proteins, based on sequence homology (36). All subfamilies, except the ABCF family, have been linked to tumor chemoresistance, with key members including ABCB1, ABCC1 and ABCG2.
The ABCB1 protein, also known as P-glycoprotein (P-gp), serves a physiological role in safeguarding sensitive tissues and the fetus from endogenous and exogenous toxins. It primarily facilitates the transport of lipid-soluble drugs, such as paclitaxel, vincristine and adriamycin (34,36–38). Although P-gp binding substrates like everolimus have shown potential in reversing chemoresistance in vitro and in animal models by interacting with extracellular binding epitopes, clinical trials have not yielded significant results, leaving the underlying mechanisms unclear. Developing inhibitors that induce conformational changes in ATP transport proteins remains a promising avenue for future research (34,39–41).
The ABCC1 protein, also known as multidrug resistance-associated protein 1 (MRP1), was the second drug resistance transporter identified after P-gp. MRP1 transports a wide array of pathophysiological substrates, including folic acid, bilirubin, anthracyclines, glutathione (GSH) and glucuronide conjugates, though it is less effective at transporting paclitaxel-like drugs compared with other ABC transporters (36,41–43). Tumor cells overexpressing MRP1 exhibit lower intracellular GSH levels and higher GSH efflux, leading to its designation as a GSH transporter protein. After GSH binds to chemotherapeutic drugs to form a complex, MRP1-mediated co-extrusion reduces the intracellular concentration of these drugs, contributing to tumor cell drug resistance (41,44).
ABCG2, also known as breast cancer resistance protein, was first identified in breast cancer-resistant tumors and functions as a hemi-transport protein, requiring homo- or heterodimerization to act as an effective transporter (45). Cryo-electron microscopy has revealed that ABCG2 can bind and transport chemotherapeutic agents such as topotecan and mitoxantrone, as well as P-gp inhibitors such as tariquidar (46).
Currently, the primary strategy for modulating the ABC transporter protein family involves targeting overexpressed proteins, but clinical benefits have been limited. Future research may focus on developing drugs that induce conformational changes in ABC transporter proteins, potentially offering a new direction for overcoming chemoresistance.
GSH, a tripeptide consisting of glutamate, cysteine and glycine, is a critical component of the cellular antioxidant system, essential for tumor cells in scavenging reactive oxygen species (ROS) and neutralizing chemotherapeutic agents (47). Predominantly existing in its reduced form within the cell, GSH reacts with oxidizing agents such as ROS, converting to GSH disulfide (GSSG). Excess GSSG can be expelled from the cell or reduced back to GSH via GSH reductase. The intracellular GSH concentration typically ranges from 2–10 mM, ~1,000-fold higher than in the extracellular environment (2–10 µM) (47).
GSH also plays a critical role in ROS scavenging as part of the cellular antioxidant system. Elevated GSH levels in tumor cells safeguard them from ROS-induced DNA damage and apoptosis by neutralizing ROS during tumor growth. Research indicates that reducing GSH levels in tumor cells can elevate ROS levels, thereby enhancing the cytotoxic effects of therapeutic agents on these cells (48,49) (Fig. 3).
Moreover, GSH is involved in the neutralization of cytotoxic substances under physiological conditions, with numerous chemotherapeutic drugs, such as cisplatin and adriamycin, acting as binding substrates for GSH (50). Catalyzed by GSH-S-transferase (GST), GSH directly binds to electrophilic xenobiotics in the cytoplasm, preventing these chemotherapeutic drugs from entering the nucleus and exerting their inhibitory effects. The resulting GSH-drug conjugates are then exported from the cell via MRP1, thereby reducing the intracellular concentration of these drugs and contributing to the development of chemoresistance in tumor cells (51). Experimental evidence suggests that decreasing intracellular GSH levels can reverse cisplatin resistance and enhance its therapeutic efficacy (52).
Intracellular GSH levels are modulated by various factors. Transforming growth factor β (TGF-β) is a well-known inhibitor of normal epithelial cell proliferation, and its suppression can increase the susceptibility of epithelial tissues to cancer. Interestingly, while high TGF-β signaling expression in the skin prevents benign papillary tumors from progressing to malignancy, it paradoxically facilitates the malignant transformation of squamous cell carcinoma stem cells into squamous carcinoma, thereby promoting metastasis (53). Upon activation, TGF-β ligands bind to TbRII, which phosphorylates TbRI. The activated TbRI then transmits signals by phosphorylating intracellular effectors SMAD2 and SMAD3 (SMAD2/3). These phosphorylated SMAD proteins form a complex with SMAD4, which interacts with other transcriptional regulators to activate downstream target genes. Notably, TGF-β reporter gene-positive basal cells exhibit high expression of genes involved in GSH metabolism and significantly lower ROS levels (53). In RNA-seq analysis of tyrosine kinase receptor inhibitor (TKI)-resistant non-small cell lung cancer, aldo-keto reductase 1B1 (AKR1B1) was found to be highly expressed in multiple drug-resistant cell lines, promoting GSH biosynthesis through STAT3-mediated upregulation of SLC7A11-dependent cystine uptake, leading to TKI resistance (54).
Both GSH and GST levels are elevated in tumor cells compared with normal cells, closely correlating with drug resistance. GSH levels can be measured using gold nanoparticles detected by a lateral flow plasma biosensor or even visually, with quantification achieved through automated analysis software. This method has shown a strong positive correlation between GSH levels and temozolomide resistance in GBM cells (55).
Beyond the GSH antioxidant system, eukaryotes possess another critical antioxidant mechanism-the thioredoxin system-which manages intracellular oxidative stress and supports cell proliferation when GSH is depleted (56). Cytosolic flavin adenine dinucleotide oxidoreductase 1 (TXNRD1), a selenoprotein with extensive antioxidant and redox regulatory functions, is overexpressed in numerous cancers and plays a key role in their growth and survival. Inhibiting TXNRD1, especially in the context of GSH depletion, increases oxidative stress within tumor cells, leading to their death (57). Given the protective role of the GSH system in shielding tumor cells from chemotherapeutic agents and ROS, exploring the regulatory mechanisms of this system and developing strategies to deplete GSH in tumor cells using molecularly targeted drugs or nanomaterials holds significant therapeutic promise.
The nuclear pore complex is a basket-like structure embedded within the inner and outer nuclear membranes, featuring apertures of ~70-80 nm and a channel diameter of ~9 nm, facilitating the exchange of substances between the nucleus and the cytoplasm. Numerous chemotherapeutic agents, such as platinum and fluoropyrimidines, which inhibit DNA replication, must traverse the nuclear pore complex to reach their targets and exert their cytotoxic effects (58). Vault particles, barrel-shaped structures primarily composed of major vault protein (MVP), vADP-ribose polymerase, and telomerase-associated protein 1, are considered to interact with the nuclear pore complex. MVP, which constitutes 70% of the vault's mass, is considered to mediate the translocation of macromolecules, potentially rerouting chemotherapeutic drugs away from their subcellular targets and contributing to multidrug resistance in tumor cells (59,60). However, this hypothesis requires further validation.
Lung resistance-related protein (LRP), also known as major vault protein, is a key component of vault particles and nuclear pore complexes in humans (61). Studies have demonstrated that LRP is highly expressed in numerous tumor cells that lack P-gp expression, and its elevated levels are associated with reduced chemotherapy sensitivity (62). In a lung adenocarcinoma cell model, gefitinib-resistant cells exhibited significant upregulation of YB-1, which promotes downstream AKT signaling and activates epithelial-mesenchymal transition (EMT), thereby increasing resistance to gefitinib through direct activation of MVP (63). CD73, a glycosylphosphatidylinositol-anchored plasma membrane protein, physiologically hydrolyzes extracellular adenosine monophosphate into adenosine and inorganic phosphate. Evidence suggests that CD73 interacts with MVP, activating the SRC-AKT pathway and affecting gemcitabine chemosensitivity in pancreatic ductal adenocarcinoma (64).
While LRP is known to enhance tumor cell resistance to chemotherapeutic agents, MVP deficiency has been associated with a reduced likelihood of tumor progression. In HBV-encoded protein X (HBx)-transgenic (TG) mice crossed with LRP-deficient mice, significant reductions in tumor number, size, liver-to-body weight ratio, alanine aminotransferase levels and alpha-fetoprotein levels were observed compared with HBx-TG mice carrying the HBx, indicating that LRP loss greatly diminishes tumor progression and extends survival time (65).
Vault RNAs (vtRNAs), non-coding RNAs comprising ~5% of vault particles, have been found to show high expression in cell lines with elevated EBV protein levels, particularly those expressing high levels of LMP1. Overexpression of vtRNAs has been associated with increased EBV expression and the inhibition of apoptosis through the overexpression of NOL3 and BCL-xl (59). Although most chemotherapeutic agents, including platinum compounds, antibiotics and phytochemicals, require nuclear entry to function effectively, the mechanisms governing their translocation across the nuclear membrane remain underexplored. Further investigation into the nuclear transport mechanisms of chemotherapeutic drugs and other macromolecules is essential for advancing the understanding of tumor chemoresistance.
The DNA damage response (DDR) pathway plays a pivotal role in recognizing, signaling and repairing DNA damage caused by endogenous or exogenous factors, including chemotherapeutic agents. It also regulates cell cycle progression through DNA repair mechanisms, mitigating damage and preventing apoptosis in the case of unrepaired lesions. As such, the DDR pathway is a critical target in cancer therapy (66–68). Numerous chemotherapeutic drugs, such as platinum compounds, alkylating agents and antibiotics, induce DNA double-strand breaks by directly binding to DNA. These breaks are among the most lethal forms of DNA damage; failure to repair them triggers cell death. Enhanced DNA repair mechanisms are a key factor in the development of multidrug resistance in tumor cells (58,69–71).
Two primary pathways are responsible for repairing DNA double-strand breaks (DSBs): non-homologous end-joining (NHEJ) and homologous recombination (HR). HR typically requires sister chromatids as templates, limiting its activity to the S and G2 phases of the cell cycle. The efficiency of DNA repair largely depends on HR-promoting factors including BRCA1 and NHEJ-promoting proteins such as TP53BP1 (72). BRCA1 and BRCA2 are essential for repairing DSBs, playing a key role in inhibiting tumor cell proliferation and contributing to drug resistance (73–75). Cells with defective BRCA1 or BRCA2 genes exhibit reduced homologous recombination repair capacity, making them more sensitive to DNA-damaging agents such as cisplatin and poly (ADP-ribose) polymerase (PARP) inhibitors (75).
DNA topoisomerase-2 (TOP2) is an enzyme essential for DNA replication, transcription and recombination, with a known role in promoting cancer across various tumors (76). In hepatocellular carcinoma cells, continuous exposure to regorafenib leads to upregulation of TOP2A expression, while silencing TOP2A increases the sensitivity of these cells to regorafenib (77).
PARP, a poly ADP-ribose polymerase, is involved in chromatin modification, DNA replication, transcription and DNA repair, particularly through base excision repair of single-strand DNA (ssDNA) damage. In the absence of functional PARP, ssDNA breaks remain unrepaired, leading to DSBs during subsequent cell cycles as DNA replication forks encounter ssDNA regions. Cells with intact DSB repair pathways can repair such breaks and survive. PARP inhibitors (PARPi) specifically target BRCA1/2-deficient tumors, inducing apoptosis (73). BRCA1/2 mutant cells are particularly vulnerable to PARPi due to their inability to repair DSBs effectively (73). Treatment with the PARP inhibitor olaparib has been shown to restore sensitivity to conventional chemotherapy in patients with prostate cancer resistant to standard treatments and carrying defective DNA repair genes, such as BRCA1/2 (78).
The phosphatase and tensin homolog (PTEN) gene is critical in regulating DNA damage repair, chromosome stability and cell cycle progression through phosphatase-independent mechanisms. Phosphorylation at tyrosine 240 (pY240-PTEN) is frequently observed in patients with tumors undergoing radiotherapy. This phosphorylated form of PTEN is highly expressed, associates with chromatin via interaction with Ki-67, facilitates RAD51 recruitment, and enhances DNA repair processes, contributing to chemoresistance and poor prognosis (70).
Schlafen family member 11 (SLFN11) is another key regulator of DNA damage repair and is linked to the cellular response to DNA-damaging agents in vitro. Replication protein A (RPA), a heterotrimeric complex, is the primary eukaryotic single-stranded DNA-binding protein, crucial for various DNA metabolic pathways, including DNA replication, recombination, damage checkpoints and repair. SLFN11 is recruited to sites of DNA damage in an RPA-dependent manner, destabilizing the RPA-ssDNA complex. Elevated levels of SLFN11 result in defects in checkpoint maintenance and homologous recombination repair, thereby sensitizing cells to DNA-damaging agents (79). Enhancer of zeste homolog 2 (EZH2) induces the H3K27me3 histone modification, leading to the silencing of SLFN11 in vivo. The use of EZH2 inhibitors in combination with standard cytotoxic therapies has been identified to prevent secondary resistance and improve the efficacy of chemotherapy in both chemo-sensitive and chemoresistant small-cell lung cancer models (80).
Mutations in DNA methyltransferase 3A (DNMT3A), particularly at arginine 882 (DNMT3Amut), are commonly associated with poor response to erythromycin chemotherapy. DNMT3Amut cells exhibit impaired nucleosome expulsion and chromatin remodeling in response to anthracyclines, which in turn hinders the recruitment of the histone chaperone SPT-16. This defect impairs the cell's ability to detect and repair DNA damage, exacerbating the DNMT3Amut phenotype. Additionally, DNMT3Amut cells display reduced CHK1 phosphorylation, along with diminished downstream p53 phosphorylation/stabilization and apoptotic signaling, further contributing to chemoresistance (81).
DNA mismatch repair (MMR) is a conserved process that identifies and corrects spontaneously misincorporated bases during DNA replication, ensuring genomic integrity (82). Impairment in MMR leads to microsatellite instability (MSI), a hallmark found in >20 different tumor types (82). Werner helicase (WRN), part of the RecQ family of DNA helicases, is critical in maintaining genomic stability, DNA repair, replication, transcription and telomere maintenance. WRN has been identified as a synthetic lethal target in deficient (d)MMR/MSI-H cancers, being essential for the survival of dMMR/MSI-H cells both in vitro and in vivo. Knockdown of WRN in these cells induces double-stranded DNA breaks and significant genomic instability, leading to apoptosis. WRN inhibition has proven effective in dMMR colorectal cancer models that develop secondary resistance to broad-spectrum chemotherapeutic agents such as irinotecan, oxaliplatin, or 5-FU, as well as combinations involving epidermal growth factor receptor (EGFR) monoclonal antibodies and BRAF or NTRK inhibitors (82).
While enhanced DNA repair can contribute to chemotherapy resistance, molecular alterations in MMR-related genes can also drive tumorigenesis and progression (83). Studies have revealed that cells exposed to targeted therapies (for example, EGFR inhibitors or BRAF inhibitors) temporarily downregulate the expression of DDR-related genes, including those involved in MMR and HR. This transient suppression of DDR capacity is reversible, with gene expression returning to normal levels after the cessation of targeted therapy (84). The presence of DNA damage and repair mechanisms within tumor cells-and whether these mechanisms are actively engaged to prevent apoptosis-presents a strategic opportunity for combination therapies. Exploiting the transient DDR deficiencies that tumors experience during treatment could enhance the effectiveness of anticancer therapies.
The tumor microenvironment (TME) is characterized by several distinct features that differentiate it from normal tissues, including acidic pH, hypoxia, elevated ROS, upregulated antioxidant systems and overexpression of specific enzymes (47). Hypoxia, a hallmark of tumors, arises from the imbalance between oxygen consumption and vascular oxygen supply in tumor tissues and is a critical factor in promoting tumorigenesis and progression. This condition drives metabolic reprogramming in tumor cells, allowing them to adapt to the hypoxic TME (85). While platinum-based chemotherapeutic drugs target rapidly proliferating cancer cells, the quiescent cell population, often associated with hypoxia, remains largely unaffected by such treatments. Hypoxia-inducible factor 1 (HIF-1), composed of alpha and beta subunits, is a key regulator of cellular hypoxic responses. HIF-1α specifically modulates the cellular response to hypoxia, influencing processes such as apoptosis, proliferation, vasodilation, energy metabolism and angiogenesis. Salidroside has been identified to promote the degradation of HIF-1α, and its administration in combination with platinum drugs can reverse platinum resistance and inhibit metastasis induced by the hypoxic TME (86).
Hypoxia also contributes to tumor growth by affecting exosome secretion. Exosomes, nanoscale extracellular vesicles (30–150 nm in diameter), facilitate the transfer of proteins, RNA and other molecules between cells within the TME, thereby influencing the behavior of surrounding cells. In the context of EGFR-mutant lung cancer, a common target in clinical therapy, drug resistance remains a significant challenge. It has been demonstrated that cells with wild-type EGFR can be internalized by EGFR-mutant cancer cells via clathrin-dependent endocytosis, leading to the acquisition of a wild-type EGFR phenotype. This phenotypic change activates downstream PI3K/AKT and MAPK signaling pathways, thereby triggering drug resistance (87). Additionally, the EGFR-targeted drug oxitinib has been shown to promote exosome release by upregulating Rab GTPase, further contributing to drug resistance (87).
In acute myeloid leukemia (AML), the apoptosis repressor with caspase recruitment domain (ARC) protein serves as a potent independent marker of poor prognosis. ARC activates NF-κB, leading to increased IL1β expression in AML cells, which in turn elevates the expression of CCL2, CCL4 and CXCL12 in mesenchymal stromal cells (MSCs) (88). When AML cells are co-cultured with MSCs, IL1β expression is further elevated, driving AML cell migration towards CCL2, CCL4 and CXCL12. Inhibition of IL1β has been revealed to reduce AML cell migration (88). Moreover, co-cultures of AML and MSCs have been found to increase Cox-2 expression in MSCs through PGE2-mediated signaling in an ARC/IL1β-dependent manner, thereby modulating ARC expression and enhancing the chemoresistance of AML cells (89).
Microorganisms within the TME are significant contributors to cancer progression and treatment resistance. A study on ovarian epithelial cancer revealed that antibiotic use during chemotherapy was linked to poorer overall survival, primarily due to bacterial imbalance. Stool analysis indicated that co-treatment with antibiotics and cisplatin disrupted 49 non-resistant intestinal microbes in mice, leading to accelerated tumor growth and cisplatin resistance compared with mice treated with chemotherapy alone. This resistance was characterized by reduced apoptosis, increased DNA damage repair and enhanced angiogenesis. However, transplanting cecal microbes from control mice into the co-treated mice restored cisplatin sensitivity (90). Similarly, L-lactic acid produced by Lactobacillus iners in tumors reprograms cervical tumor metabolism, enhancing resistance to gemcitabine combined with 5-FU chemotherapy (91). Fusobacterium nucleatum and its metabolite succinic acid induce resistance to anti-PD-1 monoclonal antibody immune checkpoint blockade therapy in colorectal cancer by inhibiting key immune pathways and reducing CD8+ T cell migration into the TME (92). In addition, telomelysin OBP-301, a telomerase-specific, replication-competent oncolytic adenovirus with an hTERT promoter upstream of the E1 gene, has been demonstrated to enhance Akt phosphorylation in hepatocellular carcinoma cells. However, the histone deacetylase inhibitor AR42 reduces telomerase-induced Akt phosphorylation and enhances telomerase-induced apoptosis. Combined treatment with telomelysin and AR42 demonstrated synergistic anti-hepatocellular carcinoma effects (93).
Tumor cells do not exist in isolation; their characteristics are intricately linked to the TME. The interactions between tumor cells and other cells, including bacteria within the microenvironment, as well as the unique pH and hypoxic conditions, profoundly influence tumor cell characteristics and drug sensitivity. Inflammation within the TME promotes drug resistance by activating pro-inflammatory cytokines and signaling pathways. This chronic inflammatory response not only drives tumor growth and metastasis but also significantly reduces cancer cell sensitivity to chemotherapy through the regulation of resistance protein expression (94). Investigating how the microenvironment contributes to chemotherapy resistance may provide new therapeutic strategies that focus on targeting the TME, rather than the tumor cells alone.
Tumor recurrence years after chemotherapy or surgery is often attributed to the presence of dormant cancer cells within the tumor mass. Dormancy refers to a reversible state in which cells cease division, exhibit low metabolic activity, and reduce mRNA synthesis, yet remain responsive to external stimuli (95–97). Dormancy can occur in early metastatic sites as well as in residual lesions following chemotherapy (98–100). Traditionally, cellular dormancy was viewed as a quiescent state associated with the G0 phase of the cell cycle (101). However, it was recently suggested that these non-dividing cells exist in a slow-cycling survival mode, akin to embryonic stasis, and are therefore also known as slow-cycling or drug-resistant persister cells (100).
Dormant tumor cells differ from tumor stem-like cells, though there is overlap between the two. Tumor stem-like cells are generally considered a subset of dormant tumor cells, possessing the ability to remain dormant, yet not all dormant cells possess stem cell properties such as self-renewal and differentiation.
Most chemotherapeutic agents target proliferating cells, allowing tumor cells to evade therapeutic stress by entering a dormant state, thereby contributing to drug resistance (102,103). Although some reactivated cells may remain sensitive to chemotherapy, the prognosis for patients with recurrent tumors is often poor (100,104). The induction of dormancy in response to drug treatment underscores the importance of accurately identifying dormant tumor cells to improve guidance of clinical drug use (105).
Next-generation sequencing and lentiviral barcoding experiments have revealed that xenograft tumors arising after chemotherapy-induced dormancy do not exhibit significant reductions in genetic or barcode complexity, indicating that dormancy is a non-genetic state (100). Tumor cell dormancy functions as a survival strategy, enabling cells to adapt to external stress rather than representing a distinct subpopulation; all tumor cells have the potential to enter a dormant state (100). Currently, there are no specific markers for identifying dormant tumor cells. Detection primarily relies on immunohistochemistry, immunofluorescence and western blot analysis of relevant indicators, typically associated with tumor stem-like cells, such as CK, Sox-2 and CD133, as well as dormancy activation markers including Ki-67, cyclin D1, C-myc, VEGF and proliferating cell nuclear antigen. TUNEL staining is also commonly used (106). Additionally, tracking the mitotic kinetics of dormant tumor cells with lipophilic fluorescent dyes serves as another method for determining cellular dormancy (107). One of the key features of dormant cells is reduced mRNA synthesis activity. Since mRNA is transcribed by RNA polymerase II (RNApII), low RNApII phosphorylation is highly specific to dormant cells, with cyclin-dependent kinase 9 (CDK9) being critical for RNApII-dependent gene transcription. The Optical Stem Cell Activity Reporter system distinguishes dormant cells from proliferating cells by detecting intracellular transcriptional status (95). In vitro immobilization techniques can also be employed to isolate and recover dormant cancer cells. This can be achieved using a microfluidic flow-focusing device that allows individual cells coated with agarose to be immobilized and survive on silica gel wells. Typically, actively proliferating cells do not survive the immobilization process, but dormant cells can be re-awakened by in situ digestion of the agarose gel and effectively recovered through magnetic separation of the silica gel (108) (Fig. 4).
In breast cancer, the tyrosine kinase receptor TIE2 induces cell dormancy and resistance to 5-FU and adriamycin by activating the expression of cyclin-dependent kinase inhibitors CDKN1A and CDKN1B (109). The treatment of diabetes with metformin or a high-fat diet can promote the survival of dormant ER+ breast cancer cells through the upregulation of the AMPK signaling pathway and the activation of fatty acid oxidation (105). FBX8, a member of the F-box protein family, plays a role in maintaining tumor cell dormancy under the pressure of chemotherapy with drugs such as oxaliplatin and 5-FU. Overexpression of FBX8 in dormant cells enhances the degradation of HIF-1α, CDK4 and C-Myc via the ubiquitin-proteasome pathway, thereby prolonging the dormancy period, whereas FBX8 knockdown shortens this period (106). In paclitaxel-treated non-small cell lung cancer, regulator of G protein signaling 2 (RGS2) mediates translational arrest and dormancy in tumor cells through the proteasomal degradation of activating transcription factor 4, while antagonizing RGS2 in the endoplasmic reticulum pathway induces apoptosis in dormant cells (110). CXCL10 has been identified as a factor that can reactivate the proliferation of dormant tumor cells within micro-metastases, making it a potential therapeutic target for addressing tumor dormancy (111). Elevated intracellular copper levels are generally associated with tumor progression; however, copper carrier drugs such as elesclomol can induce cell death in a copper-ion-dependent manner. Consequently, blocking copper uptake can inhibit tumor cell proliferation and induce dormancy, reducing overall cell mortality (112,113).
The MAPK/ERK signaling pathway, a classical proliferation-related pathway, also plays a role in the induction of cell dormancy. In EGFR-mutated non-small cell lung cancer, the combination of EGFR inhibitors and TKIs drives tumor cells into a senescence-like dormant state characterized by high YAP/TEAD activity, owing to ERK1/2 blockade. Inhibiting YAP and TEAD in conjunction with EGFR/MEK inhibition has been shown to enhance apoptosis, effectively depleting dormant cells (114).
Myc is a critical gene that regulates tumor cell growth and proliferation. Inhibition of Myc or bromodomain-containing protein 4 prompts tumor cells to enter dormancy by reducing the initiation of apoptosis. Conversely, inducing Myc expression increases chemotherapy sensitivity, suggesting that maintaining dormancy through Myc inhibition post-chemotherapy or disrupting dormancy by inhibiting CDK9 could be promising therapeutic strategies against dormant tumor cells (115).
Autophagy plays a pivotal role in cellular stress response by eliminating damaged organelles, misfolded proteins and abnormal protein aggregates, regulating mitochondrial mass, and preventing the accumulation of ROS, thereby aiding cell survival (102). Studies have revealed that when tumor cells enter a dormant state, the activity of the autophagy-related mTOR pathway decreases, while autophagy-associated phenotypes increase (102,116). Conversely, inhibiting autophagy with drugs such as chloroquine can reawaken dormant cells, prompting them to re-enter the proliferative state (117). The combination of autophagy inhibitors with chemotherapeutic agents has been found to induce the death of dormant cells (100,118). Thus, autophagy is integral not only to the initiation and maintenance of dormancy but also to the transition from dormancy to proliferation.
The TME significantly contributes to inducing tumor cell dormancy. In a mouse model of breast cancer metastasis, Ki-67 immunofluorescence revealed that tumor cells metastasizing to the lungs entered a dormant state. This metastatic dormancy was dependent on the presence of T cells, particularly CD39+PD-1+CD8+ T cells, which induced cell cycle arrest by secreting IFNγ and TNF-α, thereby promoting a dormant phenotype (119). Hypoxia, another key feature of the TME, also influences dormancy. Under hypoxic conditions, the expression of CSN8, a subunit of the COP9 signalosome, is elevated. This upregulation is accompanied by increased levels of dormancy markers (NR2F1, DEC2 and p27) and hypoxia markers (HIF-1α and GLUT1), along with decreased expression of the proliferation marker Ki-67, suggesting that CSN8 plays a role in regulating hypoxia-induced dormancy (120).
While most chemotherapeutic agents target proliferating cells and are ineffective against dormant cells, nimustine has shown efficacy in targeting dormant cells in BRCA1-deficient mice. Platinum-induced intra-strand crosslinks can be repaired by nucleotide excision during the G0-G1 phase; however, in BRCA1-deficient tumor cells, nimustine-induced inter-strand crosslink repair is impaired, leading to cell death (98).
The study of tumor cell dormancy and its role in chemotherapy resistance remains an emerging field. The definitions, markers and specific mechanisms of tumor cell dormancy remain incompletely understood. However, continued research is essential to elucidate how tumor cells evade chemotherapy and survive as micro-metastases, ultimately improving therapeutic strategies.
In recent decades, the molecular mechanisms underlying chemoresistance have become increasingly well-understood. These mechanisms include reduced drug uptake into cells, increased drug efflux, GSH-mediated drug neutralization, nuclear pore closure, enhanced DNA repair, promotion of tumor cell dormancy, and the complex interactions between tumor cells and the TME. Although the mechanisms underlying chemotherapy resistance in tumor cells are increasingly well understood, their complexity in clinical practice presents significant challenges. For example, the use of inhibitors targeting multidrug resistance transporter proteins such as MDR1 is complicated by their role in immune function. Literature indicates that CD8+ T cells secrete various cytokines to mediate immune responses and exert cytotoxic effects on viruses and tumor cells, processes in which MDR1 plays a critical role. MDR1 is essential for the development of naive CD8+ T cells, regulated primarily by Runt-Related transcription factors, which help inhibit oxidative stress, enhance cell survival, and protect the mitochondrial function of nascent CD8+ cytotoxic T lymphocytes. Therefore, inhibiting MDR1 in patients with cancers could impair immune function, potentially leading to treatment failure (121).
The development of nanomaterials designed to deliver chemotherapeutic agents directly to specific tumor sites or intracellular targets is a rapidly expanding area of research aimed at reversing chemoresistance. For example, a nanomaterial containing an iron oxide core has been engineered to deliver the cytotoxic agent adriamycin and the TLR3 agonist polyinosinic: Polycytidylic acid (Poly IC) to both breast cancer and dendritic cells. The Endoglin-binding peptide on the nanomaterial targets triple-negative breast cancer cells, inducing apoptosis through multiple mechanisms, thereby inhibiting tumor growth and metastasis. This approach has shown significant success in prolonging the survival of mouse models with aggressive and resistant triple-negative breast cancer metastases (122). Current strategies to combat GSH-mediated cellular resistance focus on GSH depletion or targeting GST. For instance, ethacraplatin, a platinum prodrug, inhibits GST by releasing ethacrynic acid, and encapsulating this compound in nanomicelles has been identified to enhance the intracellular accumulation of cisplatin (123).
Some polysulfides have been developed to target high GSH levels in tumor cells, serving as a prodrug backbone that also reverses multidrug resistance (5,124). Additionally, nanocarriers such as dendritic mesoporous silica nanoparticles can integrate components such as ultrasmall Fe3O4 nanoparticles, Mn2+ ions and the glutaminase inhibitor Telaglenastat (CB-839) into their large mesopores to form nanodrugs. These nanodrugs exhibit peroxidase-mimetic activity under acidic conditions, catalyzing the decomposition of hydrogen peroxide (H2O2) into hydroxyl radicals (−OH) and depleting existing GSH while blocking endogenous GSH synthesis. This process enhances ROS-mediated tumor catalytic therapy (125). Given the ongoing challenges with traditional inhibitors in combating chemotherapy resistance, the development of new nanomaterials represents a promising strategy for overcoming multidrug resistance in tumors, and it is expected to be a key focus of future research.
Not applicable.
The present study was partially supported by the Natural Science Foundation of Hunan Province (2023JJ50135, 2023JJ30521), the Projects of the Health Commission of Hunan Province (202201043124), the Aid Program for Science and Technology Innovative Research Team in Higher Educational Institutions of Hunan Province [No. (2023)233] and the Research Foundation of Education Bureau of Hunan (grant no. 21B0444).
Not applicable.
XW and WHZ conceived and designed the analysis, conducted the research, and drafted the manuscript. LHX created the figures. LYZ and PL contributed to the revision and polishing of the manuscript. ZL, WJG, QQ, DLC and XZ were involved in the conception and design of the analysis. XZ made significant contributions to the discussion of the content and reviewed, edited and finalized the manuscript. All authors read and approved the final version of the manuscript. Data authentication is not applicable.
Not applicable.
Not applicable.
The authors declare that they have no competing interests.
|
Miller KD, Nogueira L, Devasia T, Mariotto AB, Yabroff KR, Jemal A, Kramer J and Siegel RL: Cancer treatment and survivorship statistics, 2022. CA Cancer J Clin. 72:409–436. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Siegel RL, Miller KD, Fuchs HE and Jemal A: Cancer statistics, 2022. CA Cancer J Clin. 72:7–33. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Zheng R, Zhang S, Zeng H, Wang S, Sun K, Chen R, Li L, Wei W and He J: Cancer incidence and mortality in China, 2016. J Natl Cancer Cent. 2:1–9. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Bukowski K, Kciuk M and Kontek R: Mechanisms of multidrug resistance in cancer chemotherapy. Int J Mol Sci. 21:32332020. View Article : Google Scholar : PubMed/NCBI | |
|
Xiao X, Wang K, Zong Q, Tu Y, Dong Y and Yuan Y: Polyprodrug with glutathione depletion and cascade drug activation for multi-drug resistance reversal. Biomaterials. 270:1206492021. View Article : Google Scholar : PubMed/NCBI | |
|
Bedard PL, Hyman DM, Davids MS and Siu LL: Small molecules, big impact: 20 Years of targeted therapy in oncology. Lancet. 395:1078–1088. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Li X, Li M, Huang M, Lin Q, Fang Q, Liu J, Chen X, Liu L, Zhan X, Shan H, et al: The multi-molecular mechanisms of tumor-targeted drug resistance in precision medicine. Biomed Pharmacother. 150:1130642022. View Article : Google Scholar : PubMed/NCBI | |
|
Phan TG and Croucher PI: The dormant cancer cell life cycle. Nat Rev Cancer. 20:398–411. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Chen X, Momin A, Wanggou S, Wang X, Min HK, Dou W, Gong Z, Chan J, Dong W, Fan JJ, et al: Mechanosensitive brain tumor cells construct blood-tumor barrier to mask chemosensitivity. Neuron. 111:30–48. e142023. View Article : Google Scholar : PubMed/NCBI | |
|
Inoue T, Aoyama-Ishikawa M, Uemura M, Kohama K, Fujisaki N, Murakami H, Yamada T and Hirata J: The role of death receptor signaling pathways in mouse Sertoli cell avoidance of apoptosis during LPS- and IL-18-induced inflammatory conditions. J Reprod Immunol. 158:1039702023. View Article : Google Scholar : PubMed/NCBI | |
|
Russo M, Crisafulli G, Sogari A, Reilly NM, Arena S, Lamba S, Bartolini A, Amodio V, Magrì A, Novara L, et al: Adaptive mutability of colorectal cancers in response to targeted therapies. Science. 366:1473–1480. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Diaz LA Jr, Williams RT, Wu J, Kinde I, Hecht JR, Berlin J, Allen B, Bozic I, Reiter JG, Nowak MA, et al: The molecular evolution of acquired resistance to targeted EGFR blockade in colorectal cancers. Nature. 486:537–540. 2012. View Article : Google Scholar : PubMed/NCBI | |
|
Zheng N, Fang J, Xue G, Wang Z, Li X, Zhou M, Jin G, Rahman MM, McFadden G and Lu Y: Induction of tumor cell autosis by myxoma virus-infected CAR-T and TCR-T cells to overcome primary and acquired resistance. Cancer Cell. 40:973–985.e7. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Hirota K, Ooka M, Shimizu N, Yamada K, Tsuda M, Ibrahim MA, Yamada S, Sasanuma H, Masutani M and Takeda S: XRCC1 counteracts poly(ADP ribose)polymerase (PARP) poisons, olaparib and talazoparib, and a clinical alkylating agent, temozolomide, by promoting the removal of trapped PARP1 from broken DNA. Genes Cells. 27:331–344. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Ovejero S, Soulet C and Moriel-Carretero M: The alkylating agent Methyl methanesulfonate triggers lipid alterations at the inner nuclear membrane that are independent from its DNA-damaging ability. Int J Mol Sci. 22:74612021. View Article : Google Scholar : PubMed/NCBI | |
|
Ghosh S: Cisplatin: The first metal based anticancer drug. Bioorg Chem. 88:1029252019. View Article : Google Scholar : PubMed/NCBI | |
|
Fetoni AR, Paciello F and Troiani D: Cisplatin chemotherapy and cochlear damage: Otoprotective and chemosensitization properties of polyphenols. Antioxid Redox Signal. 36:1229–1245. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Curry JN and McCormick JA: Cisplatin-induced kidney injury: Delivering the goods. J Am Soc Nephrol. 33:255–256. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Balboni B, El Hassouni B, Honeywell RJ, Sarkisjan D, Giovannetti E, Poore J, Heaton C, Peterson C, Benaim E, Lee YB, et al: RX-3117 (fluorocyclopentenyl cytosine): A novel specific antimetabolite for selective cancer treatment. Expert Opin Investig Drugs. 28:311–322. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Guo J, Yu Z, Das M and Huang L: Nano codelivery of oxaliplatin and folinic acid achieves synergistic chemo-immunotherapy with 5-fluorouracil for colorectal cancer and liver metastasis. Acs Nano. 14:5075–5089. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Li Z, Li C, Wu Q, Tu Y, Wang C, Yu X, Li B, Wang Z and Sun S and Sun S: MEDAG enhances breast cancer progression and reduces epirubicin sensitivity through the AKT/AMPK/mTOR pathway. Cell Death Dis. 12:972021. View Article : Google Scholar : PubMed/NCBI | |
|
Li J, Yu K, Pang D, Wang C, Jiang J, Yang S, Liu Y, Fu P, Sheng Y, Zhang G, et al: Adjuvant capecitabine with docetaxel and cyclophosphamide plus epirubicin for triple-negative breast cancer (CBCSG010): An open-label, randomized, multicenter, phase III trial. J Clin Oncol. 38:1774–1784. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
You JH, Lee J and Roh JL: PGRMC1-dependent lipophagy promotes ferroptosis in paclitaxel-tolerant persister cancer cells. J Exp Clin Cancer Res. 40:3502021. View Article : Google Scholar : PubMed/NCBI | |
|
Chen Y, Liu R, Li C, Song Y, Liu G, Huang Q, Yu L, Zhu D, Lu C, Lu A, et al: Nab-paclitaxel promotes the cancer-immunity cycle as a potential immunomodulator. Am J Cancer Res. 11:3445–3460. 2021.PubMed/NCBI | |
|
Elshamy AM, Salem OM, Safa MAE, Barhoma RAE, Eltabaa EF, Shalaby AM, Alabiad MA, Arakeeb HM and Mohamed HA: Possible protective effects of CO Q10 against vincristine-induced peripheral neuropathy: Targeting oxidative stress, inflammation, and sarmoptosis. J Biochem Mol Toxicol. 36:e229762022. View Article : Google Scholar : PubMed/NCBI | |
|
Rajković S, Živković MD and Djuran MI: Reactions of dinuclear Platinum(II) complexes with peptides. Curr Protein Pept Sci. 17:95–105. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Kim ES, Tang X, Peterson DR, Kilari D, Chow CW, Fujimoto J, Kalhor N, Swisher SG, Stewart DJ, Wistuba II and Siddik ZH: Copper transporter CTR1 expression and tissue platinum concentration in non-small cell lung cancer. Lung Cancer. 85:88–93. 2014. View Article : Google Scholar : PubMed/NCBI | |
|
Lv X, Song J, Xue K, Li Z, Li M, Zahid D, Cao H, Wang L, Song W, Ma T, et al: Core fucosylation of copper transporter 1 plays a crucial role in cisplatin-resistance of epithelial ovarian cancer by regulating drug uptake. Mol Carcinog. 58:794–807. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Kalayda GV, Wagner CH and Jaehde U: Relevance of copper transporter 1 for cisplatin resistance in human ovarian carcinoma cells. J Inorg Biochem. 116:1–10. 2012. View Article : Google Scholar : PubMed/NCBI | |
|
Gao H, Zhang S, Hu T, Qu X, Zhai J, Zhang Y, Tao L, Yin J and Song Y: Omeprazole protects against cisplatin-induced nephrotoxicity by alleviating oxidative stress, inflammation, and transporter-mediated cisplatin accumulation in rats and HK-2 cells. Chem Biol Interact. 297:130–140. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Naka A, Takeda R, Shintani M, Ogane N, Kameda Y, Aoyama T, Yoshikawa T and Kamoshida S: Organic cation transporter 2 for predicting cisplatin-based neoadjuvant chemotherapy response in gastric cancer. Am J Cancer Res. 5:2285–2293. 2015.PubMed/NCBI | |
|
Hucke A, Rinschen MM, Bauer OB, Sperling M, Karst U, Köppen C, Sommer K, Schröter R, Ceresa C, Chiorazzi A, et al: An integrative approach to cisplatin chronic toxicities in mice reveals importance of organic cation-transporter-dependent protein networks for renoprotection. Arch Toxicol. 93:2835–2848. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Verhalen B, Dastvan R, Thangapandian S, Peskova Y, Koteiche HA, Nakamoto RK, Tajkhorshid E and Mchaourab HS: Energy transduction and alternating access of the mammalian ABC transporter P-glycoprotein. Nature. 543:738–741. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Alam A, Kowal J, Broude E, Roninson I and Locher KP: Structural insight into substrate and inhibitor discrimination by human P-glycoprotein. Science. 363:753–756. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Pasello M, Giudice AM and Scotlandi K: The ABC subfamily A transporters: Multifaceted players with incipient potentialities in cancer. Semin Cancer Biol. 60:57–71. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Hanssen KM, Haber M and Fletcher JI: Targeting multidrug resistance-associated protein 1 (MRP1)-expressing cancers: Beyond pharmacological inhibition. Drug Resist Updat. 59:1007952021. View Article : Google Scholar : PubMed/NCBI | |
|
Kim Y and Chen J: Molecular structure of human P-glycoprotein in the ATP-bound, outward-facing conformation. Science. 359:915–919. 2018. View Article : Google Scholar : PubMed/NCBI | |
|
Miyamoto S, Azuma K, Ishii H, Bessho A, Hosokawa S, Fukamatsu N, Kunitoh H, Ishii M, Tanaka H, Aono H, et al: Low-dose erlotinib treatment in elderly or frail patients with EGFR mutation-positive non-small cell lung cancer: A multicenter phase 2 trial. JAMA Oncol. 6:e2012502020. View Article : Google Scholar : PubMed/NCBI | |
|
Robey RW, Pluchino KM, Hall MD, Fojo AT, Bates SE and Gottesman MM: Revisiting the role of ABC transporters in multidrug-resistant cancer. Nat Rev Cancer. 18:452–464. 2018. View Article : Google Scholar : PubMed/NCBI | |
|
Carvalho DM, Richardson PJ, Olaciregui N, Stankunaite R, Lavarino C, Molinari V, Corley EA, Smith DP, Ruddle R, Donovan A, et al: Repurposing vandetanib plus everolimus for the treatment of ACVR1-mutant diffuse intrinsic pontine glioma. Cancer Discov. 12:416–431. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Cole SPC: Targeting multidrug resistance protein 1 (MRP1, ABCC1): Past, present, and future. Annu Rev Pharmacol Toxicol. 54:95–117. 2014. View Article : Google Scholar : PubMed/NCBI | |
|
Johnson ZL and Chen J: ATP binding enables substrate release from multidrug resistance protein 1. Cell. 172:81–89.e10. 2018. View Article : Google Scholar : PubMed/NCBI | |
|
Vulsteke C, Lambrechts D, Dieudonné A, Hatse S, Brouwers B, van Brussel T, Neven P, Belmans A, Schöffski P, Paridaens R and Wildiers H: Genetic variability in the multidrug resistance associated protein-1 (ABCC1/MRP1) predicts hematological toxicity in breast cancer patients receiving (neo-)adjuvant chemotherapy with 5-fluorouracil, epirubicin and cyclophosphamide (FEC). Ann Oncol. 24:1513–1525. 2013. View Article : Google Scholar : PubMed/NCBI | |
|
Cole SPC: Multidrug resistance protein 1 (MRP1, ABCC1), a ‘multitasking’ ATP-binding cassette (ABC) transporter. J Biol Chem. 289:30880–30888. 2014. View Article : Google Scholar : PubMed/NCBI | |
|
Taylor NMI, Manolaridis I, Jackson SM, Kowal J, Stahlberg H and Locher KP: Structure of the human multidrug transporter ABCG2. Nature. 546:504–509. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Kowal J, Ni D, Jackson SM, Manolaridis I, Stahlberg H and Locher KP: Structural basis of drug recognition by the multidrug transporter ABCG2. J Mol Biol. 433:1669802021. View Article : Google Scholar : PubMed/NCBI | |
|
Niu B, Liao K, Zhou Y, Wen T, Quan G, Pan X and Wu C: Application of glutathione depletion in cancer therapy: Enhanced ROS-based therapy, ferroptosis, and chemotherapy. Biomaterials. 277:1211102021. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang J, Bai J, Zhou Q, Hu Y, Wang Q, Yang L, Chen H, An H, Zhou C, Wang Y, et al: Glutathione prevents high glucose-induced pancreatic fibrosis by suppressing pancreatic stellate cell activation via the ROS/TGFβ/SMAD pathway. Cell Death Dis. 13:4402022. View Article : Google Scholar : PubMed/NCBI | |
|
Chen M, Zhao S, Zhu J, Feng E, Lv F, Chen W, Lv S, Wu Y, Peng X and Song F: Open-source and reduced-expenditure nanosystem with ROS self-amplification and glutathione depletion for simultaneous augmented chemodynamic/photodynamic therapy. ACS Appl Mater Interfaces. 14:20682–20692. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Cheng X, Xu HD, Ran HH, Liang G and Wu FG: Glutathione-depleting nanomedicines for synergistic cancer therapy. ACS Nano. 15:8039–8068. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Gao P, Yang X, Xue YW, Zhang XF, Wang Y, Liu WJ and Wu XJ: Promoter methylation of glutathione S-transferase pi1 and multidrug resistance gene 1 in bronchioloalveolar carcinoma and its correlation with DNA methyltransferase 1 expression. Cancer. 115:3222–3232. 2009. View Article : Google Scholar : PubMed/NCBI | |
|
Xu Y, Han X, Li Y, Min H, Zhao X, Zhang Y, Qi Y, Shi J, Qi S, Bao Y and Nie G: Sulforaphane mediates glutathione depletion via polymeric nanoparticles to restore cisplatin chemosensitivity. ACS Nano. 13:13445–13455. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Oshimori N, Oristian D and Fuchs E: TGF-β promotes heterogeneity and drug resistance in squamous cell carcinoma. Cell. 160:963–976. 2015. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang KR, Zhang YF, Lei HM, Tang YB, Ma CS, Lv QM, Wang SY, Lu LM, Shen Y, Chen HZ and Zhu L: Targeting AKR1B1 inhibits glutathione de novo synthesis to overcome acquired resistance to EGFR-targeted therapy in lung cancer. Sci Transl Med. 13:eabg64282021. View Article : Google Scholar : PubMed/NCBI | |
|
Pang HH, Ke YC, Li NS, Chen YT, Huang CY, Wei KC and Yang HW: A new lateral flow plasmonic biosensor based on gold-viral biomineralized nanozyme for on-site intracellular glutathione detection to evaluate drug-resistance level. Biosens Bioelectron. 165:1123252020. View Article : Google Scholar : PubMed/NCBI | |
|
Harris IS, Endress JE, Coloff JL, Selfors LM, McBrayer SK, Rosenbluth JM, Takahashi N, Dhakal S, Koduri V, Oser MG, et al: Deubiquitinases maintain protein homeostasis and survival of cancer cells upon glutathione depletion. Cell Metab. 29:1166–1181.e6. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Stafford WC, Peng X, Olofsson MH, Zhang X, Luci DK, Lu L, Cheng Q, Trésaugues L, Dexheimer TS, Coussens NP, et al: Irreversible inhibition of cytosolic thioredoxin reductase 1 as a mechanistic basis for anticancer therapy. Sci Transl Med. 10:eaaf74442018. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang X, Wang M, Feng J, Qin B, Zhang C, Zhu C, Liu W, Wang Y, Liu W, Huang L, et al: Multifunctional nanoparticles co-loaded with Adriamycin and MDR-targeting siRNAs for treatment of chemotherapy-resistant esophageal cancer. J Nanobiotechnology. 20:1662022. View Article : Google Scholar : PubMed/NCBI | |
|
Amort M, Nachbauer B, Tuzlak S, Kieser A, Schepers A, Villunger A and Polacek N: Expression of the vault RNA protects cells from undergoing apoptosis. Nat Commun. 6:70302015. View Article : Google Scholar : PubMed/NCBI | |
|
Liu S, Hao Q, Peng N, Yue X, Wang Y, Chen Y, Wu J and Zhu Y: Major vault protein: A virus-induced host factor against viral replication through the induction of type-I interferon. Hepatology. 56:57–66. 2012. View Article : Google Scholar : PubMed/NCBI | |
|
Bai H, Wang C, Qi Y, Xu J, Li N, Chen L, Jiang B, Zhu X, Zhang H, Li X, et al: Major vault protein suppresses lung cancer cell proliferation by inhibiting STAT3 signaling pathway. BMC Cancer. 19:4542019. View Article : Google Scholar : PubMed/NCBI | |
|
Shen W, Qiu Y, Li J, Wu C, Liu Z, Zhang X, Hu X, Liao Y and Wang H: IL-25 promotes cisplatin resistance of lung cancer cells by activating NF-κB signaling pathway to increase of major vault protein. Cancer Med. 8:3491–3501. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Lou L, Wang J, Lv F, Wang G, Li Y, Xing L, Shen H and Zhang X: Y-box binding protein 1 (YB-1) promotes gefitinib resistance in lung adenocarcinoma cells by activating AKT signaling and epithelial-mesenchymal transition through targeting major vault protein (MVP). Cell Oncol (Dordr). 44:109–133. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Yu X, Liu W, Wang Z, Wang H, Liu J, Huang C, Zhao T, Wang X, Gao S, Ma Y, et al: CD73 induces gemcitabine resistance in pancreatic ductal adenocarcinoma: A promising target with non-canonical mechanisms. Cancer Lett. 519:289–303. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Yu H, Li M, He R, Fang P, Wang Q, Yi Y, Wang F, Zhou L, Zhang Y, Chen A, et al: Major vault protein promotes hepatocellular carcinoma through targeting interferon regulatory factor 2 and decreasing p53 activity. Hepatology. 72:518–534. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Hanahan D and Weinberg RA: Hallmarks of cancer: The next generation. Cell. 144:646–674. 2011. View Article : Google Scholar : PubMed/NCBI | |
|
Pilié PG, Tang C, Mills GB and Yap TA: State-of-the-art strategies for targeting the DNA damage response in cancer. Nat Rev Clin Oncol. 16:81–104. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Gourley C, Balmaña J, Ledermann JA, Serra V, Dent R, Loibl S, Pujade-Lauraine E and Boulton SJ: Moving from poly (ADP-Ribose) polymerase inhibition to targeting DNA repair and DNA damage response in cancer therapy. J Clin Oncol. 37:2257–2269. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Pettitt SJ, Frankum JR, Punta M, Lise S, Alexander J, Chen Y, Yap TA, Haider S, Tutt ANJ and Lord CJ: Clinical BRCA1/2 reversion analysis identifies hotspot mutations and predicted neoantigens associated with therapy resistance. Cancer Discov. 10:1475–1488. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Ma J, Benitez JA, Li J, Miki S, Ponte de Albuquerque C, Galatro T, Orellana L, Zanca C, Reed R, Boyer A, et al: Inhibition of nuclear PTEN tyrosine phosphorylation enhances glioma radiation sensitivity through attenuated DNA repair. Cancer Cell. 35:504–518.e7. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Goodall J, Mateo J, Yuan W, Mossop H, Porta N, Miranda S, Perez-Lopez R, Dolling D, Robinson DR, Sandhu S, et al: Circulating cell-free DNA to guide prostate cancer treatment with PARP inhibition. Cancer Discov. 7:1006–1017. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Dev H, Chiang TW, Lescale C, de Krijger I, Martin AG, Pilger D, Coates J, Sczaniecka-Clift M, Wei W, Ostermaier M, et al: Shieldin complex promotes DNA end-joining and counters homologous recombination in BRCA1-null cells. Nat Cell Biol. 20:954–965. 2018. View Article : Google Scholar : PubMed/NCBI | |
|
Schlacher K: PARPi focus the spotlight on replication fork protection in cancer. Nat Cell Biol. 19:1309–1310. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Lord CJ and Ashworth A: Mechanisms of resistance to therapies targeting BRCA-mutant cancers. Nat Med. 19:1381–1388. 2013. View Article : Google Scholar : PubMed/NCBI | |
|
Ray Chaudhuri A, Callen E, Ding X, Gogola E, Duarte AA, Lee JE, Wong N, Lafarga V, Calvo JA, Panzarino NJ, et al: Replication fork stability confers chemoresistance in BRCA-deficient cells. Nature. 535:382–387. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Awah CU, Chen L, Bansal M, Mahajan A, Winter J, Lad M, Warnke L, Gonzalez-Buendia E, Park C, Zhang D, et al: Ribosomal protein S11 influences glioma response to TOP2 poisons. Oncogene. 39:5068–5081. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Wang Z, Zhu Q, Li X, Ren X, Li J, Zhang Y, Zeng S, Xu L, Dong X and Zhai B: TOP2A inhibition reverses drug resistance of hepatocellular carcinoma to regorafenib. Am J Cancer Res. 12:4343–4360. 2022.PubMed/NCBI | |
|
Mateo J, Carreira S, Sandhu S, Miranda S, Mossop H, Perez-Lopez R, Nava Rodrigues D, Robinson D, Omlin A, Tunariu N, et al: DNA-repair defects and olaparib in metastatic prostate cancer. N Engl J Med. 373:1697–1708. 2015. View Article : Google Scholar : PubMed/NCBI | |
|
Mu Y, Lou J, Srivastava M, Zhao B, Feng XH, Liu T, Chen J and Huang J: SLFN11 inhibits checkpoint maintenance and homologous recombination repair. EMBO Rep. 17:94–109. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Gardner EE, Lok BH, Schneeberger VE, Desmeules P, Miles LA, Arnold PK, Ni A, Khodos I, de Stanchina E, Nguyen T, et al: Chemosensitive relapse in small cell lung cancer proceeds through an EZH2-SLFN11 axis. Cancer Cell. 31:286–299. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Guryanova OA, Shank K, Spitzer B, Luciani L, Koche RP, Garrett-Bakelman FE, Ganzel C, Durham BH, Mohanty A, Hoermann G, et al: DNMT3A mutations promote anthracycline resistance in acute myeloid leukemia via impaired nucleosome remodeling. Nat Med. 22:1488–1495. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Picco G, Cattaneo CM, van Vliet EJ, Crisafulli G, Rospo G, Consonni S, Vieira SF, Rodríguez IS, Cancelliere C, Banerjee R, et al: Werner helicase is a synthetic-lethal vulnerability in mismatch repair-deficient colorectal cancer refractory to targeted therapies, chemotherapy, and immunotherapy. Cancer Discov. 11:1923–1937. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Germano G, Lamba S, Rospo G, Barault L, Magrì A, Maione F, Russo M, Crisafulli G, Bartolini A, Lerda G, et al: Inactivation of DNA repair triggers neoantigen generation and impairs tumour growth. Nature. 552:116–120. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Gay CM, Parseghian CM and Byers LA: This is our cells under pressure: Decreased DNA damage repair in response to targeted therapies facilitates the emergence of drug-resistant clones. Cancer Cell. 37:5–7. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Guo W, Qiao T, Dong B, Li T, Liu Q and Xu X: The effect of hypoxia-induced exosomes on anti-tumor immunity and its implication for immunotherapy. Front Immunol. 13:9159852022. View Article : Google Scholar : PubMed/NCBI | |
|
Qin Y, Liu HJ, Li M, Zhai DH, Tang YH, Yang L, Qiao KL, Yang JH, Zhong WL, Zhang Q, et al: Salidroside improves the hypoxic tumor microenvironment and reverses the drug resistance of platinum drugs via HIF-1α signaling pathway. EBioMedicine. 38:25–36. 2018. View Article : Google Scholar : PubMed/NCBI | |
|
Wu S, Luo M, To KKW, Zhang J, Su C, Zhang H, An S, Wang F, Chen D and Fu L: Intercellular transfer of exosomal wild type EGFR triggers osimertinib resistance in non-small cell lung cancer. Mol Cancer. 20:172021. View Article : Google Scholar : PubMed/NCBI | |
|
Carter BZ, Mak PY, Chen Y, Mak DH, Mu H, Jacamo R, Ruvolo V, Arold ST, Ladbury JE, Burks JK, et al: Anti-apoptotic ARC protein confers chemoresistance by controlling leukemia-microenvironment interactions through a NFκB/IL1β signaling network. Oncotarget. 7:20054–20067. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Carter BZ, Mak PY, Wang X, Tao W, Ruvolo V, Mak D, Mu H, Burks JK and Andreeff M: An ARC-regulated IL1β/Cox-2/PGE2/β-catenin/ARC circuit controls leukemia-microenvironment interactions and confers drug resistance in AML. Cancer Res. 79:1165–1177. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Chambers LM, Esakov Rhoades EL, Bharti R, Braley C, Tewari S, Trestan L, Alali Z, Bayik D, Lathia JD, Sangwan N, et al: Disruption of the gut microbiota confers cisplatin resistance in epithelial ovarian cancer. Cancer Res. 82:4654–4669. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Johnston CD and Bullman S: Bacteria-derived L-lactate fuels cervical cancer chemoradiotherapy resistance. Trends Cancer. 10:97–99. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Jiang SS, Xie YL, Xiao XY, Kang ZR, Lin XL, Zhang L, Li CS, Qian Y, Xu PP, Leng XX, et al: Fusobacterium nucleatum-derived succinic acid induces tumor resistance to immunotherapy in colorectal cancer. Cell Host Microbe. 31:781–797.e9. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Lin ZZ, Hu MC, Hsu C, Wu YM, Lu YS, Ho JA, Yeh SH, Chen PJ and Cheng AL: Synergistic efficacy of telomerase-specific oncolytic adenoviral therapy and histone deacetylase inhibition in human hepatocellular carcinoma. Cancer Lett. 556:2160632023. View Article : Google Scholar : PubMed/NCBI | |
|
Vaidya FU, Sufiyan Chhipa A, Mishra V, Gupta VK, Rawat SG, Kumar A and Pathak C: Molecular and cellular paradigms of multidrug resistance in cancer. Cancer Rep (Hoboken). 5:e12912022. View Article : Google Scholar : PubMed/NCBI | |
|
Freter R, Falletta P, Omrani O, Rasa M, Herbert K, Annunziata F, Minetti A, Krepelova A, Adam L, Käppel S, et al: Establishment of a fluorescent reporter of RNA-polymerase II activity to identify dormant cells. Nat Commun. 12:33182021. View Article : Google Scholar : PubMed/NCBI | |
|
Summers MA, McDonald MM and Croucher PI: Cancer cell dormancy in metastasis. Cold Spring Harb Perspect Med. 10:a0375562020. View Article : Google Scholar : PubMed/NCBI | |
|
Yeh AC and Ramaswamy S: Mechanisms of cancer cell dormancy-another hallmark of cancer? Cancer Res. 75:5014–5022. 2015. View Article : Google Scholar : PubMed/NCBI | |
|
Pajic M, Blatter S, Guyader C, Gonggrijp M, Kersbergen A, Küçükosmanoğlu A, Sol W, Drost R, Jonkers J, Borst P and Rottenberg S: Selected alkylating agents can overcome drug tolerance of G0-like tumor cells and eradicate BRCA1-deficient mammary tumors in mice. Clin Cancer Res. 23:7020–7033. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Badia-Ramentol J, Linares J, Gómez-Llonin A and Calon A: Minimal residual disease, metastasis and immunity. Biomolecules. 11:1302021. View Article : Google Scholar : PubMed/NCBI | |
|
Rehman SK, Haynes J, Collignon E, Brown KR, Wang Y, Nixon AML, Bruce JP, Wintersinger JA, Singh Mer A, Lo EBL, et al: Colorectal cancer cells enter a diapause-like DTP state to survive chemotherapy. Cell. 184:226–242.e21. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Cackowski FC and Heath EI: Prostate cancer dormancy and recurrence. Cancer Lett. 524:103–108. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Akkoc Y, Peker N, Akcay A and Gozuacik D: Autophagy and cancer dormancy. Front Oncol. 11:6270232021. View Article : Google Scholar : PubMed/NCBI | |
|
Damen MPF, van Rheenen J and Scheele CLGJ: Targeting dormant tumor cells to prevent cancer recurrence. FEBS J. 288:6286–6303. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Lee J, Kim SH and Kang BJ: Prognostic factors of disease recurrence in breast cancer using quantitative and qualitative magnetic resonance imaging (MRI) parameters. Sci Rep. 10:75982020. View Article : Google Scholar : PubMed/NCBI | |
|
Hampsch RA, Wells JD, Traphagen NA, McCleery CF, Fields JL, Shee K, Dillon LM, Pooler DB, Lewis LD, Demidenko E, et al: AMPK activation by metformin promotes survival of dormant ER+ breast cancer cells. Clin Cancer Res. 26:3707–3719. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Zhu X, Wang F, Wu X, Li Z, Wang Z, Ren X, Zhou Y, Song F, Liang Y, Zeng Z, et al: FBX8 promotes metastatic dormancy of colorectal cancer in liver. Cell Death Dis. 11:6222020. View Article : Google Scholar : PubMed/NCBI | |
|
Quayle LA, Spicer A, Ottewell PD and Holen I: Transcriptomic profiling reveals novel candidate genes and signalling programs in breast cancer quiescence and dormancy. Cancers (Basel). 13:39222021. View Article : Google Scholar : PubMed/NCBI | |
|
Preciado J, Lam T, Azarin SM, Lou E and Aksan A: Induction of dormancy by confinement: An agarose-silica biomaterial for isolating and analyzing dormant cancer cells. J Biomed Mater Res B Appl Biomater. 109:2117–2130. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Drescher F, Juárez P, Arellano DL, Serafín-Higuera N, Olvera-Rodriguez F, Jiménez S, Licea-Navarro AF and Fournier PG: TIE2 induces breast cancer cell dormancy and inhibits the development of osteolytic bone metastases. Cancers (Basel). 12:8682020. View Article : Google Scholar : PubMed/NCBI | |
|
Cho J, Min HY, Lee HJ, Hyun SY, Sim JY, Noh M, Hwang SJ, Park SH, Boo HJ, Lee HJ, et al: RGS2-mediated translational control mediates cancer cell dormancy and tumor relapse. J Clin Invest. 131:e1719012021. View Article : Google Scholar | |
|
Clark AM, Heusey HL, Griffith LG, Lauffenburger DA and Wells A: IP-10 (CXCL10) can trigger emergence of dormant breast cancer cells in a metastatic liver microenvironment. Front Oncol. 11:6761352021. View Article : Google Scholar : PubMed/NCBI | |
|
Yu Z, Zhou R, Zhao Y, Pan Y, Liang H, Zhang JS, Tai S, Jin L and Teng CB: Blockage of SLC31A1-dependent copper absorption increases pancreatic cancer cell autophagy to resist cell death. Cell Prolif. 52:e125682019. View Article : Google Scholar : PubMed/NCBI | |
|
Tsvetkov P, Coy S, Petrova B, Dreishpoon M, Verma A, Abdusamad M, Rossen J, Joesch-Cohen L, Humeidi R, Spangler RD, et al: Copper induces cell death by targeting lipoylated TCA cycle proteins. Science. 375:1254–1261. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Kurppa KJ, Liu Y, To C, Zhang T, Fan M, Vajdi A, Knelson EH, Xie Y, Lim K, Cejas P, et al: Treatment-induced tumor dormancy through YAP-mediated transcriptional reprogramming of the apoptotic pathway. Cancer Cell. 37:104–122.e12. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Dhimolea E, de Matos Simoes R, Kansara D, Al'Khafaji A, Bouyssou J, Weng X, Sharma S, Raja J, Awate P, Shirasaki R, et al: An embryonic diapause-like adaptation with suppressed Myc activity enables tumor treatment persistence. Cancer Cell. 39:240–256.e11. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Hussein AM, Wang Y, Mathieu J, Margaretha L, Song C, Jones DC, Cavanaugh C, Miklas JW, Mahen E, Showalter MR, et al: Metabolic control over mTOR-dependent diapause-like state. Dev Cell. 52:236–250. e72020. View Article : Google Scholar : PubMed/NCBI | |
|
Lu Z, Luo RZ, Lu Y, Zhang X, Yu Q, Khare S, Kondo S, Kondo Y, Yu Y, Mills GB, et al: The tumor suppressor gene ARHI regulates autophagy and tumor dormancy in human ovarian cancer cells. J Clin Invest. 118:3917–3929. 2008.PubMed/NCBI | |
|
Anlaş AA and Nelson CM: Soft microenvironments induce chemoresistance by increasing autophagy downstream of integrin-linked kinase. Cancer Res. 80:4103–4113. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Tallón de Lara P, Castañón H, Vermeer M, Núñez N, Silina K, Sobottka B, Urdinez J, Cecconi V, Yagita H, Movahedian Attar F, et al: CD39+PD-1+CD8+ T cells mediate metastatic dormancy in breast cancer. Nat Commun. 12:7692021. View Article : Google Scholar : PubMed/NCBI | |
|
Ju S, Wang F, Wang Y and Ju S: CSN8 is a key regulator in hypoxia-induced epithelial-mesenchymal transition and dormancy of colorectal cancer cells. Mol Cancer. 19:1682020. View Article : Google Scholar : PubMed/NCBI | |
|
Chen ML, Sun A, Cao W, Eliason A, Mendez KM, Getzler AJ, Tsuda S, Diao H, Mukori C, Bruno NE, et al: Physiological expression and function of the MDR1 transporter in cytotoxic T lymphocytes. J Exp Med. 217:e201913882020. View Article : Google Scholar : PubMed/NCBI | |
|
Mu QG, Lin G, Jeon M, Wang H, Chang FC, Revia RA, Yu J and Zhang M: Iron oxide nanoparticle targeted chemo-immunotherapy for triple negative breast cancer. Mater Today (Kidlington). 50:149–169. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Li S, Li C, Jin S, Liu J, Xue X, Eltahan AS, Sun J, Tan J, Dong J and Liang XJ: Overcoming resistance to cisplatin by inhibition of glutathione S-transferases (GSTs) with ethacraplatin micelles in vitro and in vivo. Biomaterials. 144:119–129. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Ling X, Chen X, Riddell IA, Tao W, Wang J, Hollett G, Lippard SJ, Farokhzad OC, Shi J and Wu J: Glutathione-scavenging poly(disulfide amide) nanoparticles for the effective delivery of Pt(IV) prodrugs and reversal of cisplatin resistance. Nano Lett. 18:4618–4625. 2018. View Article : Google Scholar : PubMed/NCBI | |
|
Wu F, Du Y, Yang J, Shao B, Mi Z, Yao Y, Cui Y, He F, Zhang Y and Yang P: Peroxidase-like active nanomedicine with dual glutathione depletion property to restore oxaliplatin chemosensitivity and promote programmed cell death. ACS Nano. 16:3647–3663. 2022. View Article : Google Scholar : PubMed/NCBI |