Skip to main content

Unraveling the triad of hypoxia, cancer cell stemness, and drug resistance

Abstract

In the domain of addressing cancer resistance, challenges such as limited effectiveness and treatment resistance remain persistent. Hypoxia is a key feature of solid tumors and is strongly associated with poor prognosis in cancer patients. Another significant portion of the development of acquired drug resistance is attributed to tumor stemness. Cancer stem cells (CSCs), a small tumor cell subset with self-renewal and proliferative abilities, are crucial for tumor initiation, metastasis, and intra-tumoral heterogeneity. Studies have shown a significant association between hypoxia and CSCs in the context of tumor resistance. Recent studies reveal a strong link between hypoxia and tumor stemness, which together promote tumor survival and progression during treatment. This review elucidates the interplay between hypoxia and CSCs, as well as their correlation with resistance to therapeutic drugs. Targeting pivotal genes associated with hypoxia and stemness holds promise for the development of novel therapeutics to combat tumor resistance.

Introduction

Each year, over 20 million people are diagnosed with cancer, resulting in around 10 million deaths, or roughly one-sixth of all global fatalities [1]. The limited effectiveness of current therapies against drug resistance is one of the main reasons for the high mortality rates associated with cancer [2]. Drug resistance can be attributed to the tumor microenvironment (TME) and the considerable intra-tumoral heterogeneity [3, 4].

Studies have shown that various TME components contributing to drug resistance development [5, 6]. Hypoxia, a common TME characteristic of solid tumors, arises because the rapid and unregulated growth of tumors restricts the supply of oxygen [7]. In response to hypoxic conditions, cancer cells adapt and transform into a more aggressive phenotype that is resistant to pharmaceutical interventions [8].

Cancer stem cells (CSCs), characterized by self-renewal, differentiation potential, and resistance to chemotherapy and radiotherapy, contribute significantly to tumor progression [9,10,11,12]. Furthermore, studies demonstrate that hypoxia further impacts CSC pluripotency, driving tumor progression, metastasis, and resistance to therapies [7]. However, current research has not definitively proven a causal relationship between hypoxia, pluripotency, and drug resistance, and there is a notable absence of clinical strategies addressing both hypoxia and drug resistance concurrently [13, 14].

This review elucidates the relationship between hypoxia and CSCs, underscoring their contribution to the development of resistance to cancer therapies. Furthermore, it explores the underlying mechanisms and identifies prospective therapeutic targets to guide future research initiatives.

The hypoxic TME enhances the stemness of cancer stem cells

Tumors have the ability to attract surrounding endogenous stromal cells [15], which play a role in facilitating extracellular matrix modification, angiogenesis, cellular movement, invasion, and resistance to drugs. Moreover, stromal cells generate a TME that can evade immune surveillance through the production and release of diverse chemokines, cytokines, and growth factors [16]. There are at least six distinct cellular origins for stromal cells associated with tumors, including fibroblasts, pericytes, mesenchymal stem cells (MSCs) from bone marrow, adipocytes, and endothelial and tumor cells post-EMT. Hypoxia, a key feature of the TME, arises from rapid cancer cell growth and heightened metabolic demands, requiring significant energy to sustain proliferation [17]. Tumor expansion can stimulate angiogenesis, but disorganized blood vessel formation creates spatial gaps exceeding the oxygen diffusion capacity, leading to localized hypoxia [18]. During hypoxia, significant changes occur in non-cancerous components, including increased activation and growth of stromal cells such as stellate cells and cancer-associated fibroblasts (CAFs), as well as an accumulation of stromal elements such as fibrin [19]. These modifications contribute to the morphological restructuring of cancer, including the compression of blood vessels, which may lead to impaired circulation and insufficient oxygen supply. This sequence of events can ultimately lead to thrombosis and tissue hypoxia, which plays a critical role in regulating cancer cell populations, particularly CSCs [7].

CSCs were first identified in 1997 and are characterized by two fundamental properties: self-renewal and pluripotency [20]. These characteristics are essential for understanding intra-tumoral heterogeneity within the CSC hypothesis, which is a key concept in the hierarchical model of tumor development [21]. CSCs are typically characterized by their robust proliferative capacity [22]. Cancer cell proliferation is strongly influenced by the activation of key signaling pathways, such as AKT, mTOR, and MAPK/ERK, which upregulate proteins that control the cell cycle [23]. Conversely, interventions aimed at inhibiting stemness tend to diminish the proliferative potential of these cells [24]. CSCs are also responsible for resistance to chemotherapy, radiotherapy, and immunotherapy [14]. The survival of CSCs enables the replenishment of tumor cell populations, ultimately contributing to cancer recurrence [9].

Hypoxia-inducible factors (HIFs) are responsible for regulating gene expression in response to reduced oxygen levels in both normal tissues and cancer cells. Their activation can facilitate the advancement of tumors through the modulation of cellular metabolism and the promotion of blood vessel formation [25]. HIFs are composed of oxygen-sensitive α-subunits and stable β-subunits, with the primary regulation of their activity being controlled by the stability of the α-subunits. Under normoxic conditions, HIF-α interacts with the von Hippel-Lindau (VHL) protein, triggering the activation of the ubiquitin ligase system and subsequent proteasomal degradation of HIF-α. Conversely, under hypoxic conditions, the inactivation of prolyl hydroxylase (PHD) inhibits the PHD-dependent binding of HIF-α to VHL, stabilizing HIF-α. This stabilized form then binds to HIF-β subunits to form heterodimers that regulate gene expression by binding to hypoxia response elements (HREs) on DNA [26]. The HIFs family comprises HIF-1, HIF-2, and HIF-3, which regulate various biological processes through their target genes, such as angiogenesis, erythropoiesis, metastasis through EMT, autophagy, and the modulation of metabolic pathways. These alterations are associated with the maintenance of tumor stemness.

Hypoxia promotes stemness by regulating glucose metabolism

Under low oxygen conditions, cancer cells, including CSCs, shift from oxygen-dependent mitochondrial oxidative phosphorylation (OXPHOS) to oxygen-independent glycolysis to maintain ATP production [27, 28]. Hypoxia primarily regulates glucose metabolism by influencing the expression of glucose transporters and glycolytic enzymes and suppressing the tricarboxylic acid (TCA) cycle pathway. The increase in glycolytic activity has been associated with the characteristics of tumor stemness.

In hypoxic conditions, the glucose transporter (GLUT) family members GLUT1 [29], GLUT3 [30] and GLUT5 [31] have increased expression levels, facilitating increased intracellular glucose transportation to fuel the energy demands of cancerous cells. HIF-1 directly interacts with HREs in the promoter regions of genes encoding glycolytic enzymes, increasing the expression of key enzymes such as M2 pyruvate kinase (PKM2) [32], aldolase A [33], enolase (ENO) [33] and lactate dehydrogenase A [33]. NF-κB, FOXO4, and numerous non-coding RNAs are significant contributors to the modulation of glucose metabolism under hypoxic conditions [29, 34,35,36,37,38,39]. Hypoxia impedes OXPHOS by decreasing the influx of metabolites into the TCA cycle via the activation of pyruvate dehydrogenase kinase 1 (PDK1) in a HIF-1-dependent manner [40]. Additional pathways through which HIF-1 suppresses oxidative metabolism have been reported, such as the inhibition of fatty acid oxidation [41], the induction of selective autophagy in mitochondria [42], and the suppression of activity in electron transport chain (ETC) complex I [43].

Alterations in glucose metabolism under hypoxic conditions are associated with elevated levels of CSC markers and improved capacity for self-renewal [40, 44,45,46]. The precise mechanism by which alterations in glucose metabolism impact the maintenance of stemness remains uncertain. One hypothesis posits that CSCs suppress the generation of ROS within cells by enhancing the flow of glucose metabolism, thereby preserving REDOX balance and promoting cell viability. Reducing glucose uptake and glycolysis in primary glioblastoma stem cells (GSCs) elevates ROS levels, impairing proliferation and stemness, while ROS elimination reverses these effects [47]. In hepatocellular carcinoma, the accumulation of ROS is increased through the targeting of glutaminase 1, leading to a decrease in the stem cell-like properties of CSCs [48]. Various branches of glycolysis, such as the pentose phosphate pathway, serine pathway, and folic acid pathway, are interconnected with the intracellular antioxidant system [49, 50]. These pathways are selectively upregulated to varying extents in response to hypoxic conditions. In breast CSCs, HIF-1 promotes the expression of phosphoglycerate dehydrogenase (PHGDH) and the serine synthesis pathway, along with five enzymes in the folate cycle, supporting CSC proliferation and enrichment [51]. Glucose hexaphosphate, produced from glycogen breakdown, is processed through the pentose phosphate pathway, regulating ROS levels via the NADPH-glutathione system [44]. The reliance of CSCs on ROS exhibits variability across distinct cancer types [52]. In the case of liver CSCs, there is a reduction in mitochondrial ROS production facilitated by NANOG activity, which in turn promotes stemness [53]. Conversely, in aggressive triple-negative breast CSCs, elevated levels of ROS are sustained through the upregulation of mitochondrial biogenesis. This process is regulated by MYC and MCL1, which facilitate mammosphere formation by stabilizing HIF1α [54]. This challenges the theory that uses ROS to explain the link between hypoxic metabolism and stemness, though no superior alternative explanation exists at present. Also, there has been no comprehensive research conducted to substantiate this theory.

In summary, hypoxia influences tumor cell metabolism by upregulating glucose transporters, enhancing glycolytic enzymes, inhibiting the TCA cycle and fatty acid oxidation, suppressing ETC complex I activity, and inducing selective mitochondrial autophagy. However, the connection between hypoxia-induced metabolic alterations and tumor stemness has been explored primarily through correlation studies rather than direct mechanistic investigations. The reduction in intracellular ROS production may be the underlying mechanism. Inhibiting glycolysis is linked to increased ROS levels, and the upregulation of ROS scavenging systems has been observed in CSCs. Further research is needed to comprehensively elucidate the interplay among hypoxia, glucose metabolism, and stemness.

Hypoxia promotes stemness by regulating EMT

EMT is a phenomenon characterized by the reversible transformation of epithelial cells into mesenchymal cells. This process involves alterations in the expression of cell adhesion molecules and cytoskeletal components [55]. The suppression of epithelial markers, such as E-cadherin, and the increased expression of mesenchymal markers, such as vimentin and N-cadherin, represent common alterations in genetic expression associated with EMT [56, 57]. Hypoxia has been shown to facilitate the progression of EMT in cancer cells, leading to increased acquisition of mesenchymal and stem cell-like properties [58]. The primary regulatory mechanism governing EMT involves the suppression of the E-cadherin promoter by transcription factors [59]. The E-cadherin promoter in humans is characterized by the presence of three regulatory elements known as “e-box”. Transcription factors associated with epithelial-mesenchymal transition (EMT-TFs), including SNAIL, SLUG, ZEB1, ZEB2, and TWIST, are capable of binding to the e-box elements, thereby suppressing the transcription of E-cadherin [60,61,62,63].

HIF-1 triggers EMT by increasing the expression of EMT-associated transcription factors or reducing the expression of inhibitory factors. This process involves the activation of EMT-related signaling pathways and the modulation of EMT-associated inflammatory cytokines. HIF-1 induces EMT by directly controlling the expression of E-cadherin, SNAIL, ZEB1, TWIST, and TCF3 [64, 65]. In breast cancer cells, hypoxia can lead to the degradation of the biological clock gene Period2, triggering the activation of EMT-related genes such as TWIST1 and SNAIL [66]. HIF-1 may also regulate EMT through long non-coding RNAs (lncRNAs), microRNA [67,68,69], and calcium signaling pathways [70]. Multiple signaling pathways are involved in the EMT process, and hypoxia can activate NF-κB [71], Notch signaling [72], β-catenin signaling [73] and Hh signaling [74] to promote EMT. Tumor-associated macrophages (TAMs), the primary pro-inflammatory cells within tumors, exhibit increased secretion of IL-1β in response to moderate hypoxia. This leads to the up-regulation of HIF-1α synthesis and EMT in hepatocellular carcinoma cells via cyclooxygenase-2 [75].

The process of EMT has the potential to confer self-renewal capabilities to cancer cells, leading to the transformation of non-CSCs into CSCs [76]. In breast cancer, elevated expression of Twist, Snail, or FOXC2 is associated with increased CD44+/CD24- markers linked to breast CSCs and improved breast globule formation [77, 78]. Disseminated cancer cells demonstrate characteristics of a mesenchymal phenotype and possess properties akin to those of stem cells [79]. The connection between EMT and CSCs at the molecular level is unclear. One possible explanation is that EMT-induced changes in cancer cell secretory proteins establish autocrine signaling loops involving key pathways like TGF-β and Wnt-β-catenin, which are crucial for maintaining stem cell characteristics [80]. It has been proposed that transcription factors involved in regulating EMT may also influence the maintenance of stem cell properties. For example, ZEB1 has been shown to promote the splicing of the CD44 subtype (CD44s) in pancreatic cancer by suppressing the epithelial splicing regulator ESRP1, leading to elevated ZEB1 levels. This newly identified interaction between CD44s and ZEB1 impacts various aspects of cancer cell behavior, such as enhanced tumor sphere formation and increased metastatic potential [81]. Also in pancreatic cancer cells, the absence of ZEB1 impedes stemness and colonizing capabilities [82]. ZEB1 suppresses the activity of the miR-200 family, which strongly promotes epithelial differentiation. Additionally, its potential targets include stemness-associated factors such as SRY-box 2 (SOX2) and Klf4 [82]. In human colorectal cancer tissue, Snail plays a role in controlling the expression of IL-8 and stimulating the activity of CSCs [83]. These findings provide evidence of an intricate relationship between the stemness and metastatic properties of tumors. CSCs play a significant role in the EMT process, and the activation of EMT in tumor cells enhances their stem-like characteristics, invasiveness, and ability to metastasize.

In summary, hypoxia influences the regulation of EMT-TFs through various mechanisms including transcriptional control and the modulation of microRNA and lncRNA processing. While pathways like Notch, NF-κB, and Hh signaling pathways have not been definitively linked to EMT-TFs in this context, EMT-TFs are crucial in mediating the interplay among hypoxia, EMT, and stemness. These factors not only serve as significant transcriptional regulators of EMT but also function as modulators of stem cell-like properties.

Hypoxia promotes stemness by inducing tumor angiogenesis

When a solid tumor grows beyond a certain volume where diffusion can no longer adequately supply oxygen and nutrients to the cells, a deprived and hypoxic TME will develop [84]. Hypoxia has the potential to stimulate the reformation of blood vessels within the tumor in response to environmental conditions, thereby ensuring the continued provision of essential nutrients and oxygen.

HIFs are widely recognized as significant regulators of angiogenesis, a process closely linked to the preservation of tumor stemness [85]. HIF-1 induces the expression of a series of pro-angiogenic factors inside tumor cells. These factors include vascular endothelial growth factor (VEGF), placenta growth factor (PGF), angiopoietin-2 (ANGPT-2), chemokine C-X-C motif ligand-12 (CXCL-12), and stem cell factor (SCF) [86,87,88,89,90]. These factors are crucial for promoting tumor angiogenesis by interacting with receptors on the surfaces of endothelial cells, pericytes, and vascular smooth muscle cells to stimulate angiogenesis. In addition to regulating the secretion of factors that activate endothelial cells, HIF-1 also controls the intrinsic expression of numerous genes in hypoxic endothelial cells. A significant portion of these genes encode cell surface receptors that enable endothelial cells to react to angiogenic cytokines induced by hypoxia [89]. The expression of miR-23a is notably increased in exosomes derived from lung cancer cells exposed to hypoxic conditions. This upregulation leads to the direct suppression of proline hydroxylase 1 and 2 by miR-23a, resulting in the accumulation of HIF1α in endothelial cells [90]. Furthermore, research has demonstrated a strong correlation between EMT and angiogenesis. In particular, the process of epithelial-endothelial transformation (EET) in CSCs is notably enhanced in hypoxic environments [91]. Studies conducted in three-dimensional cell cultures have revealed that melanoma cells exhibiting angiogenic mimicry express the CD133 marker associated with CSCs. Notably, the suppression of this marker leads to a substantial decrease in the capacity of cells to form an angiogenic mimicry network [92].

Angiogenesis generates vascular niches to support and maintain the stemness of CSCs. The proximity of CSCs to the vascularized area underscores the importance of vascular niches in supporting CSC function [93,94,95]. A reduction in vascular endothelial cells has been shown to increase the self-renewal capacity of CD133+ brain CSCs in vivo by suppressing the ERBB2 or VEGF signaling pathways. This depletion also leads to a decrease in the number of CSCs and effectively hinders the growth of xenograft tumors [93]. The unique metabolic conditions in the perivascular niche, marked by concentrated mTOR activity in glioblastoma, play a key role in shaping the gene expression of GSCs and promoting GSC-specific traits [96]. Furthermore, endothelial cells increase the expression of genes associated with CSCs, including Olig2, Bmi1, and Sox2 [95]. The Notch signaling pathway in GSCs can be activated by nitric oxide from endothelial cells via the NO/cGMP/PKG pathway, enhancing glioma stem cell traits and accelerating tumor progression in murine models [97].

In brief, hypoxia stimulates the formation of new blood vessels in the tumor area through its effects on both tumor cells and endothelial cells. When exposed to low oxygen levels, cancer cells trigger the production of angiogenic factors that interact with specific receptors on cell surfaces, thereby facilitating angiogenesis. Additionally, endothelial cells respond directly to hypoxia signals by upregulating the expression of diverse angiogenic genes. These hypoxia-induced signals contribute to the establishment of vascular niches within the tumor microenvironment, which support the maintenance of CSC stemness properties.

Hypoxia promotes autophagy and thus maintains CSC activity

Autophagy serves as a stress response mechanism that cancer cells utilize to endure nutrient scarcity or hypoxia during advanced stages of tumor development. The autophagy process is governed by over 40 autophagy-related genes (ATGs), whose proteins initiate the process upon receiving signals associated with autophagosome formation. Methylation of ULK, an oxygen-sensitive process, promotes the assembly of the initiation complex ULK-1-ATG13-FIP200, which is essential for autophagy induction in hypoxic environments [98]. Autophagy-associated proteins, including ATG2A, ATG5, ATG7, ATG9A, ATG14, and Beclin1, exhibit increased expression levels in response to hypoxic conditions, thereby facilitating the induction of autophagy in cancer cells [99,100,101,102,103]. Furthermore, Beclin1 has been identified as a crucial intermediary protein involved in the control of apoptosis and autophagy pathways within cancer cells. Beclin1 plays a pivotal role in determining the cellular response by interacting with proteins associated with autophagy or apoptosis. Under hypoxic conditions, tumor apoptosis is inhibited through the HIF-1α/BCL2/adenovirus E1B 19 kDa protein-3/Beclin 1 signaling pathway, while tumor autophagy is promoted [104].

Mitochondrial autophagy plays a crucial role in maintaining the self-renewal capacity of human stem cells. This mechanism involves the turnover of respiratory mitochondria to support the maintenance of hematopoietic stem cells in a glycolytic state with reduced levels of oxidative metabolism [105]. Suppression of mitochondrial autophagy in esophageal squamous cell carcinoma cells leads to a decrease in the presence of CD44, a recognized marker of CSCs [106]. One study reported that in contrast to cancer cells cultured in a two-dimensional environment, mammary spheres enriched with CSCs had notably elevated expression levels of Beclin1 and increased autophagy flux [107]. Investigating the impact of autophagy on stem cell characteristics is a significant research area, yet there is limited empirical support available. Studies in gliomas using transcriptomics and proteomics have revealed that chaperone-mediated autophagy (CMA) and its primary receptor LAMP2A, located on the lysosomal membrane, modulate CSC behavior through mechanisms such as extracellular matrix interactions, mitochondrial metabolism, and immune system pathways [108]. In addition, autophagy plays a role in regulating the secretion of IL-6 [109], and the IL-6-Jak2-STAT3 signaling pathway is important in the process of transforming non-CSCs to CSCs [110]. CD44 serves as an indicator of various CSCs and plays a crucial role in controlling the stem cell properties of CSCs [111]. Autophagy may contribute to the upregulation of CD44 and ADAM17, the latter being responsible for the cleavage of exomeric domain of CD44. This process involves extracellular matrix remodeling, cancer invasion, and metastasis [112]. When mitochondrial autophagy is increased, p53 localizes to mitochondria and is subsequently eliminated in a mitochondrial autophagy-dependent manner. This process may represent a mechanism through which autophagy influences CSC stemness [113]. In a murine model, autophagy was shown to play a role in CSC regulation through the EGFR/STAT3 and TGF-β/Smad signaling pathways in two distinct populations of breast cancer stem-like cells [114]. Autophagy is associated with angiogenesis, as evidenced by the autophagy-induced activation of KDR/VEGFR-2, facilitating the development of vasogenic mimicry in GSCs [115].

In summary, autophagy is a significant link between hypoxia and stem cell characteristics. Numerous ATGs are elevated in hypoxic environments, and intermediary molecules that connect autophagy and apoptosis exist. Increased autophagy levels in cancer cells are associated with increased apoptotic tendencies. Research has demonstrated that autophagy enhances the stemness of CSCs, although the underlying mechanism remains to be explored in mouse and cellular models.

Hypoxia promotes key transcription factors in stemness-related pathways

The effects of hypoxia on cell stemness are interconnected, and many genes such as NANOG [116, 117], EGF-like domain 7 (EGFL7) [118], Krüppel-like factor 5 (KLF5) [119], SOX2 [120], and OCT4 [121], are up-regulated under hypoxia conditions and affect the stemness of CSCs through multiple mechanisms (Fig. 1).

Fig. 1
figure 1

Hypoxia promotes key transcription factors in the stemness-related pathways. Under hypoxia conditions, the expression of KLF5, EGFL7, OCT4, SOX2, and NANOG are up-regulated, and these genes act on multiple target genes, which have effects on EMT, immune escape, metabolic changes, apoptosis, proliferation, and other aspects. Abbreviation: EGFL7, EGF-like domain 7; KLF5, Krüppel-like factor 5; OCT4, octamer-binding transcription factor 4; SOX2, SRY-box 2

NANOG, KLF5, and SOX2 proteins promote cell proliferation, with NANOG exerting its effects through cyclin D1, whereas KLF5 is influenced by both cyclin D1 and cyclin B1 [122,123,124]. Additionally, SOX2 is responsible for the upregulation of cyclin D3 expression [125]. In terms of metabolism, the hypoxia-induced NANOG/SOX9 pathway enhances lactic acid production [126], and KLF5 promotes the expression of phospholipase PLA2G16 and thus promotes glycolysis in pancreatic cancer [127]. NANOG, EGFL7, KLF5, OCT4 and SOX2 have also been demonstrated to promote the EMT process in cancer cells, thereby affecting tumor metastasis and invasion [128,129,130,131,132,133,134]. As a key angiogenic factor, EGFL7 promotes angiogenesis by activating the MAPK/ERK, PI3K/Akt, and JAK/STAT3 pathways, while also inhibiting Notch signaling and its target genes to further facilitate angiogenesis [118]. SOX2 and NANOG function as significant regulators of autophagy [135]. NANOG plays a crucial role in facilitating the immune evasion of cancer cells by activating the TGF-β1 and Tcl1a/Akt signaling pathways, thereby allowing tumor cells to evade immune surveillance and avoid destruction by the immune system [136, 137]. Similarly, OCT4 can facilitate evasion of the immune system through the Tcl1a/Akt signaling pathway [138]. EGFL7 suppresses intercellular adhesion molecule-1 and vascular cell adhesion molecule-1, which are crucial endothelial adhesion molecules facilitating the attachment and infiltration of immune cells [139]. In lung cancer cells, SOX2 has been shown to increase the expression of survivin, a protein that suppresses the apoptosis of cancer cells by acting as an inhibitor of programmed cell death [140]. Additionally, OCT4, NANOG, and SOX2 interact to form a self-sustaining network that maintains CSC characteristics by promoting self-renewal genes and suppressing differentiation genes [141].

Cancer stem cells have the potential to drive resistance to various cancer therapies.

The development of resistance to both conventional chemotherapeutic agents and targeted therapies is a significant contributor to mortality associated with cancer [142]. Cells that endure therapeutic interventions possess the capacity for proliferation, which may result in disease recurrence or metastatic dissemination. Both outcomes significantly diminish the overall survival rates of patients [143]. CSCs are frequently identified as a major contributing factor to the development of multi-drug resistance and the recurrence of cancer [14]. Detailed mechanisms for CSC-induced chemoresistance, radioresistance, and resistance to immunotherapy will be introduced in the following sections.

Mechanisms of stemness-related resistance to conventional cancer treatments

Chemotherapy drugs are genotoxic agents that cause DNA damage directly or indirectly. Similarly, radiotherapy, which employs ionizing radiation, affects the DNA structure by inducing breaks in the DNA strands, particularly double-strand breaks. Resistance to radiotherapy has been consistently noted in solid tumors and leukemia-derived CSCs [144]. Clinical research has demonstrated a notable increase in the presence of CSC markers in small-cell lung cancer patients following chemotherapy treatment [145]. The effectiveness of chemotherapy was found to be linked to multidrug resistance mediated by ATP-binding cassette (ABC) transporters. The resistance of CSCs to radiation primarily arises from mechanisms such as reactive oxygen species (ROS) scavenging, DNA repair, and cell cycle arrest (Fig. 2).

Fig. 2
figure 2

Mechanism of drug resistance and relapse of CSCs. Chemotherapy and radiotherapy primarily target specialized cells and decrease tumor size in the immediate period. Nevertheless, CSCs can acquire resistance to chemoradiotherapy and radiotherapy via various mechanisms, including ABC transporter activity, cellular quiescence, upregulated DNA repair mechanisms, increased autophagy, and decreased apoptosis rates. Abbreviation: ABC, ATP-binding cassette; ROS, Reactive oxygen species

Enriched quiescence and slow-cycling

Cell quiescence refers to the state in which stem cells enter a reversible G0 phase where they neither undergo cell death nor actively proliferate [146]. CSCs typically exist in dormant or slowly proliferating conditions, rendering them less responsive to therapies designed to attack rapidly dividing cells. As a key angiogenic factor, EGFL7 promotes angiogenesis by activating the MAPK/ERK, PI3K/Akt, and JAK/STAT3 pathways, while also inhibiting Notch signaling and its target genes to further facilitate angiogenesis [147]. In murine models, after chemotherapy, CSCs can be activated and their process of moving toward a proliferative state can be accelerated [148]. The precise mechanisms of the quiescence and drug resistance of CSCs remain to be elucidated; however, existing evidence suggests an association with Notch signaling [149], SET domain-containing protein 4 (SETD4) [150], and the c-Yes/YAP axis [151]. Advances in techniques for isolating and identifying CSCs are expected to enhance research into the dormant characteristics and drug resistance mechanisms of CSCs.

Increased efflux pump activity

CSCs exhibit elevated levels of efflux pumps, which expel chemotherapy drugs from the cell, reducing their efficacy by lowering intracellular drug concentrations [152]. ABC transporters, a diverse family of membrane proteins that can extrude a variety of chemotherapy drugs from the cell, are of particular interest [153]. Within this family, 12 glycoproteins have been identified as potential drug transporters, including P-glycoprotein encoded by ABCB1, multidrug resistance-associated protein 1 (MRP1) encoded by ABCC1, and ABC subfamily G member 2 (ABCG2) encoded by ABCG2 [154]. Their expression levels exhibit a significant correlation with various aspects of cancer cell behavior, including tumorigenicity, proliferation, drug resistance, and metastasis [155]. The expression levels of ABC transporters are elevated in CSCs in vivo [156, 157]. For example, ovarian CSCs have elevated levels of ABCB1 [158], whereas ABCB5 is prominently expressed in the initial cells of malignant melanoma [159]. The small interfering RNA (siRNA) treatment suppressed ABCG2 expression in CD133 + colorectal CSCs, leading to reduced self-renewal capacity of CSCs and increased apoptosis induced by chemotherapy [160]. A research investigation established a vesicle co-delivery system utilizing non-ionic surfactants, designed to concurrently siRNA targeting ABCG2 and the anti-apoptotic gene BCL2, in conjunction with doxorubicin, to CSCs [161]. The findings indicated a marked enhancement in the cytotoxic efficacy of doxorubicin against CSCs.

In addition to expelling chemotherapy drugs, ABC transporters act as efflux pumps for various signaling molecules, including hormones, offering CSCs an additional survival mechanism [162]. Importantly, while some ABC transporters can eliminate specific drugs, not all chemotherapy agents are substrates for these transporters. Furthermore, targeting a single ABC transporter may be insufficient to address multidrug resistance in tumors. For instance, tepotinib can reverse multidrug resistance caused by ABCB1, but not by ABCC1 or ABCG2 [163]. Hence, owing to limited knowledge of mutual compensation among ABC transporters and the regulatory pathways governing their expression, no specific inhibitor targeting all ABC transporters is currently available in clinical settings.

Enhanced DNA repair mechanisms

CSCs have effective DNA repair mechanisms [164], enabling them to recover from DNA damage caused by treatments. As a result, drugs that target cell DNA that induce tumor cell death, including cisplatin, oxaliplatin (DNA crosslinking agent), methotrexate (DNA synthesis inhibitor), doxorubicin and daunorubicin (topoisomerase inhibitor), have limited effectiveness in killing CSCs. The enhanced DNA repair process in CSCs involves various mechanisms, such as base excision repair (BER), nucleotide excision repair (NER), post-replication repair (PRR), non-homologous end joining (NHEJ), and homologous recombination (HR). Numerous DNA repair genes are overexpressed in CSCs (Table 1). Most target proteins are overexpressed in the HR pathway, which is the primary mechanism through which CSCs develop drug resistance by repairing DNA damage. The central player in HR is the strand-exchange protein, Rad51, which is highly expressed in clinical samples and patient-derived GSCs [165, 166]. The HR system includes additional elements for genetic recombination, such as the MRE11-RAD50-NBS1 (MRN) complex, which triggers the activation of ataxia-telangiectasia mutated (ATM) kinase, early checkpoint response kinases, and BRCA-related proteins [167].

Table 1 DNA-repair proteins related to CSCs and their therapeutic resistance

Increased anti-apoptotic pathways

CSCs have the potential to enhance the activation of anti-apoptotic pathways, thereby reducing their vulnerability to programmed cell death induced by stress from therapeutic interventions [168]. CSCs can evade apoptosis signals through various mechanisms, including alterations in cell cycle regulation, an imbalance in pro-apoptotic and anti-apoptotic proteins, down-regulation of death receptors coupled with up-regulation of c-FLIT, heightened capacity for clearing oxygen free radicals, and elevated expression of inhibitors of apoptosis family proteins (IAPs). Additionally, CSCs extend the G2/M phase of the cell cycle by increasing the expression of the checkpoint proteins checkpoint kinase 1 (CHK1) and CHK2. This extended phase is linked to DNA repair and resistance to programmed cell death [169].

The B lymphoblastoma-2 (Bcl-2) protein family comprises two main categories: anti-apoptotic members, such as Bcl-2, Bcl-XL, and Mcl-1, and pro-apoptotic proteins such as multi-domain proteins (e.g., Bax and Bak), and BH3-only molecules (e.g., Bim, Noxa, and Puma) [170]. The direct regulation of apoptosis is mediated by high levels of anti-apoptotic proteins which predominate [171, 172]. Hepatic CSCs show increased expression of the anti-apoptotic protein Bcl-2 and activation of the PI3K and ERK signaling pathways, which contribute to resistance to radiotherapy [173]. Death receptors, which belong to the TNF receptor superfamily, can induce cell death upon activation [174]. These receptors are distinguished by an intracellular death domain that aids in the transmission of cell death signals from the cell membrane to intracellular signaling molecules [175]. The reduced expression of death receptors and elevated levels of c-FLIP in CSCs impede the activation of the death receptor pathway and block apoptosis progression [176, 177]. An increased capacity for eliminating free radicals may confer increased resistance to cancer therapy in CSCs. In a study of breast CSCs in mice, researchers found that these cells resisted genotoxic stress by reducing ROS production and efficiently clearing existing ROS after treatment [178]. CSCs can increase the production of the antioxidant NADPH by increasing the expression of ALDH, a recognized marker of CSCs. This process aids in the removal of ROS generated along with oxidative stress. Following the inhibition of ALDH with disulfiram, the apoptotic response triggered by cisplatin notably intensifies [179]. The gene expression products involved in ROS clearance, including superoxide dismutase, glutathione peroxidase, and catalase, are increased in breast CSCs [178]. IAP proteins encompass a group of natural caspase inhibitors that impede apoptosis signaling pathways at critical junctures. In addition to the documented role of various IAP proteins in inhibiting cell death, elevated expression levels of IAP protein expression have been observed in CSCs. For example, the Baculoviral inhibitor of apoptosis repeat containing 5 (BIRC5), an inhibitor of the apoptosis protein family, is highly expressed in lung CSCs and GSCs, suggesting its potential as a therapeutic target [180].

CSCs possess a strong ability to evade programmed cell death induced by detachment (known as anoikis), allowing them to survive under adverse metastatic conditions characterized by extracellular matrix (ECM) deprivation, which is promoted during chemotherapy or radiotherapy [181]. The amplified expression of breast CSC markers ALDH1 and CD44+/CD24 − can induce STAT3-mediated anoikis resistance [182]. CSCs acquire mesenchymal characteristics by promoting the EMT process, which is closely related to anoikis resistance [183]. A separate study demonstrated that CSCs exhibit elevated expression of β1 integrin, which confer protection against anoikis through the activation of the survivin signaling pathway [184].

Overall, CSCs have evolved mechanisms to evade apoptosis signals, thereby bolstering their resistance to therapeutic agents and radiation. Targeting the apoptosis pathway may prove to be a promising approach to overcoming drug resistance in CSCs.

Increased autophagy levels

Investigations into breast CSCs have revealed that the maintenance of autophagy homeostasis is a fundamental feature of their capacity to differentiate into diverse cell phenotypes under varying pathological states [185]. Comprehensive research suggests that increased autophagy not only promotes the life of tumors but also contributes to their resistance to drugs across various types of tumors [186]. Autophagy functions by removing damaged macromolecules or organelles caused by chemotherapy and radiotherapy, thereby shielding cancer cells from apoptosis. In estrogen receptor (ER)-positive breast cancer, inhibiting autophagy enhances the sensitivity of resistant tumors to tamoxifen-induced cell death [187]. Combined treatment of epidermal growth factor receptor-tyrosin kinase inhibitor and autophagy inhibitor inhibits the stemness and restored the toxicity of osimertinib [188]. Similarly, in prostate cancer, autophagy inhibition can overcome resistance to enzalutamide [189]. Salinomycin by inhibiting autophagy has also been reported to reduce the proportion of breast CSC population [190].

Nevertheless, the precise mechanism underlying resistance to autophagy is not fully understood. Research has indicated that the inhibition of autophagy leads to the accumulation of the autophagy-regulating transcription factor FoxO3a in cancer cells. This accumulation results in increased expression of the pro-apoptotic target gene Puma, ultimately triggering programmed cell death [191]. Autophagy also triggers an intensified DNA damage response through the HR repair pathway, which is a primary mechanism for mending double-strand breaks (DSBs) [192]. Additionally, in oral CSCs, autophagy has been demonstrated to stimulate the upregulation of the drug efflux protein ABCB1 [112]. Another study on gastric cancer revealed that autophagy plays a role in controlling the resistance of gastric CSCs to chemotherapy via the Notch signaling pathway [193]. Autophagy-mediated drug resistance represents a multifaceted phenomenon that involves a range of factors, such as gene repair mechanisms, the replenishment of cytoplasmic components, alterations in drug concentration and metabolism, as well as modifications in the activity and expression of essential proteins [186]. The increased sensitivity of GSCs to radiotherapy can also be achieved through the suppression of autophagy [194]. Further comprehensive experiments are needed to clarify the mechanisms underlying autophagy, its role in CSC resistance, and potential therapeutic targets related to autophagy and CSCs.

Mechanisms of stemness-related resistance to immunotherapy

The ongoing advancements in immune checkpoint inhibitors and cellular immunotherapy have led to the development of novel approaches for target tumors. Since the identification of CSCs, numerous studies have investigated their immunological properties extensively. These findings indicate that CSCs possess immune-privileged characteristics that contribute significantly to the three phases of immunoediting, namely elimination, equilibrium, and escape [195, 196]. The inherent low immunogenicity and quiescent nature of CSCs enable them to evade immune surveillance within their microenvironment, thus protecting them from elimination by the immune system. Studies of melanoma cell lines reveal that the immunogenic tumor-associated antigen MART-1 is exclusively present in differentiated melanoma cells and not in melanoma CSCs [197]. CSCs possess distinct immune privileges and exhibit enhanced immune evasion capabilities through various mechanisms. The antigen processing and presentation processes involving the interaction between cytotoxic T lymphocyte (CTL) receptors and major histocompatibility complex class I (MHC-I)/antigen complexes are impaired in CSCs. CSCs can increase the level of CD47 expression. CD47, in turn, can bind to the inhibitory immune receptor SIRPα found on myeloid cells, leading to a “do not eat me” signal. This mechanism serves to protect CSCs from being engulfed by macrophages and dendritic cells [198, 199]. Inhibition of CD47 has been demonstrated to enhance the process of macrophage-mediated phagocytosis of CSCs in pancreatic ductal adenocarcinoma, acute myeloid leukemia, and hepatocellular carcinoma, consequently aiding in their eradication [198,199,200]. CSCs derived from melanoma, glioblastoma multiforme, and breast cancer have been demonstrated to reduce the expression of MHC-I molecules, consequently evading detection by T cells [201,202,203]. A reduction in the number of components involved in the antigen processing machinery, including low-molecular-weight proteins, antigen processing-related transporters, and beta-macroglobulin, has been observed in GSCs [202]. These findings indicate that CSCs can interfere with antigen processing mechanisms.

In several types of cancer, including glioblastoma, lung cancer, and breast cancer, M1 macrophages, which are classically activated and pro-inflammatory, are recruited to the tumor site through chemotaxis in response to cytokines released by CSCs. Upon arrival at the tumor microenvironment, these M1 macrophages undergo a transformation into M2 macrophages, which secrete factors such as TGF-β, IL-10, IL-23, and arginase 1, thereby contributing to the establishment of an immune-suppressive microenvironment that facilitates tumor growth [204,205,206]. CSCs facilitate the evasion of the immune system by suppressing immune-related signaling pathways, including toll-like receptor 4 (TLR4) [207] and the signal transducer and activator of transcription 3 (STAT3) pathway [208]. Furthermore, breast CSCs can specifically avoid elimination by natural killer (NK) cells and antibody-dependent cell-mediated cytotoxicity (ADCC) [209]. CSCs facilitate the infiltration of regulatory T cells (Treg cells) in glioblastoma via the release of the costimulatory molecule PD-L1, soluble Galectin-3, and TGF-β [94]. An increasing body of research has shown that CSCs play an important role in immune avoidance by altering the phenotypes of Dendritic cells (DCs) and inhibiting their recruitment to TME [210]. CSCs are not optimal targets for immune surveillance because of their ability to impede immune responses through the diverse expression of immune checkpoint molecules. The immune checkpoint protein CD276 (B7-H3) is significantly upregulated in the CSCs of head and neck squamous cell carcinoma, aiding in the evasion of host immune responses [211]. The EMT/β-catenin signaling pathway can induce significant PD-L1 expression in CSCs, thereby enabling CSCs to evade detection by T cells [212]. CSCs can also upregulate the production of the immunosuppressive molecule CD200, potentially leading to a shift in lymphocyte behavior from a Th1-like immune response to a Th2-like immune response [213].

As previously noted, CSCs possess the capacity to transition between quiescent and actively dividing states, a characteristic that aids in their evasion of immune surveillance. Quiescent CSCs derived from diverse types of cancer can suppress the antigen-presenting machinery, such as MHC class I, Tap proteins, and UL16-binding protein (ULBP) ligands. This modulation helps avoid the detection of CSCs by cytotoxic T lymphocytes and NK cells [214, 215]. In a model of tumor dormancy, leukemia cells exhibited prolonged survival by upregulating the expression of B7-H1 and B7.1, thereby evading CTL-mediated destruction [216]. Moreover, inactive cancer cells evade T-cell-mediated cell death by disrupting the suppressor of cytokine signaling 1 (SOCS1) cascade and increasing the expression of the oncogenic factor IL-3 [217].

The immune resistance of CSCs is enhanced by their limited immunogenicity, enabling them to evade detection and destruction by the immune system through various mechanisms. In conjunction with resistance to chemoradiotherapy, the drug resistance exhibited by CSCs is influenced by multiple factors and pathways. The similarity of the defense mechanisms of CSCs and normal stem cells highlights the potential of CSC-targeted treatments. The ability of CSCs to exhibit plasticity and regenerate highlights the importance of employing CSC-specific therapies alongside traditional treatments to inhibit their regeneration, replenishment, and redistribution.

Key pathways of cancer stem cells involved in resistance to conventional treatment

The potential mechanisms of CSC-related drug resistance include activation of growth signaling pathways such as the Wnt/β-catenin, Hedgehog, Notch, PI3/AKT, and Hippo pathways [204, 218].These pathways contribute to CSC drug resistance by inducing quiescence, enhancing DNA damage repair and ROS clearance, and promoting multidrug resistance (MDR).

Wnt pathway

The WNT/β-Catenin signaling pathway, known for its participation in a range of physiological functions and pathological conditions, is conserved across evolutionary lineages [219]. Hypoxia is a significant factor in the maintenance of stemness through the Wnt signaling pathway. A study demonstrated that Wnt/β-catenin signaling facilitates the hypoxia-induced self-renewal capabilities of colorectal CSCs by reactivating the expression of Id2 [220]. Under conditions of hypoxia, there is an observed overexpression of BCL9, which serves as a significant co-activator in the Wnt/β-catenin signaling pathway [221]. HIF-1-regulated miR-1275 directly interacts with several antagonists of the Wnt/β-catenin signaling pathways, such as DKK3, SFRP1, GSK3β, and RUNX3. This interaction leads to the activation of signaling pathways and the preservation of stem cell characteristics [222]. In liver cancer, the increased expression of EPH receptor B2 in CSCs plays a crucial role in regulating cancer stemness and resistance to drugs, facilitated by the initiation of the SRC/AKT/GSK3β/β-catenin signaling pathway cascade [223]. In cases of ovarian cancer, the c-Kit receptor associated with stem cell factors, triggers the activation of the Wnt/β-catenin and ABCG2 pathways, leading to chemotherapy resistance [224]. The Wnt signaling pathway is also linked to the upregulation of ABC transporters, facilitating drug efflux and contributing to drug resistance [225].

Notch pathway

The Notch signaling pathway plays a crucial role in regulating the survival and self-renewal of CSCs. Similarly, hypoxia plays a significant role in the maintenance of stemness via the Notch signaling pathway. MiR-1275, which is regulated by HIF-1, directly engages with various antagonists of the Notch signaling pathway, including NUMB. This interaction facilitates the activation of downstream signaling pathways, thereby maintaining the characteristics associated with stem cell-like properties [222]. The Notch pathway inhibitor L685-458 was found to counteract the overexpression of HIF-2α, which is associated with the development of paclitaxel resistance and the transformation towards a stem-like phenotype [226]. Also, hypoxia stimulates the expression of the Notch ligand Jagged2 in cancer cells, which subsequently triggers EMT and enhances stemness characteristics, thereby contributing to tumor resistance [72]. In colorectal cancer, Notch1 is highly expressed in chemotherapy-resistant cells that are rich in the CSC markers CD133 and CD44 [227]. In gliomas, genes associated with the Notch and Hedgehog pathways are significantly upregulated in CD133+ CSCs and are resistant to temozolomide [228]. The Notch signaling pathway, particularly the Notch3 pathway, plays a crucial role in regulating the maintenance of CSCs and their resistance to cisplatin in ovarian cancer [229]. In the context of lung adenocarcinoma, cisplatin treatment results in the enrichment of CD133+ cells, thereby promoting MDR through the activation of the Notch signaling pathway [230].

Hedgehog pathway

The Hedgehog pathway is associated with the regulation of CSCs [231]. Suppression of the Hedgehog pathway results in a decrease in the expression of stemness-associated genes, including NANOG, SOX2, and octamer-binding transcription factor 4 (OCT4), in glioma cells [232]. The activation of the Hedgehog signaling pathway contributes to the development of resistance to sorafenib in organoids derived from individuals with hepatocellular carcinoma [233]. The expression of GLI-1, a constituent of the Hedgehog signaling pathway, is increased in cancer cells that have developed resistance to 5-FU. Conversely, the decreased expression of GLI-1 can mitigate resistance to 5-FU [234]. In the gas-fluid organoids derived from colon cancer, the efficacy of 5-FU, irinotecan, and oxaliplatin is enhanced by hedgehog signaling inhibitors through the suppression of GLI-1 expression, thereby reducing drug resistance [234].

PI3K/AKT pathway

The PI3K/AKT signaling pathway is significantly involved in the development of resistance to chemotherapy. Studies have demonstrated a strong association between metastasis-associated colon cancer 1 (MACC1) and stem cell-like properties as well as resistance to 5-FU in CSCs. Furthermore, MACC1 expression is modulated by the PI3K/AKT pathway [235]. One of the functions of the PI3K/AKT pathway is to regulate ABCG2 activity by targeting it to the plasma membrane [236], thereby facilitating the efflux of drugs. In breast cancer, PD-L1 plays a role in controlling the activation of CSCs by influencing the expression of genes associated with stemness, such as OCT-4, NANOG, and BMI1, through the PI3K/AKT signaling pathway [237]. This interaction has been linked to the development of tumor immunoresistance.

Hippo-YAP/TAZ pathway

The activation of the YAP/TAZ signaling pathway causes cancer cells to revert to a less differentiated state. These cells exhibit CSC-like properties, such as self-renewal ability and resistance to chemotherapy. The Hippo coactivator YAP1 plays a role in promoting the overexpression of EGFR in esophageal cancer, leading to enhanced resistance to chemotherapy [238]. Colon cancer cells were found to evade cell death induced by 5-FU chemotherapy by transitioning into a state related to stemness and dormancy, which is linked to the c-Yes/YAP axis [151]. In hepatocellular carcinoma, the YAP/TAZ signaling pathway plays a significant role in suppressing ferroptosis, thereby serving as a crucial factor in resistance to sorafenib treatment [239]. Similarly, Yap1 has been identified as a mediator of trametinib resistance in head and neck squamous cell carcinoma [240].

Biomarkers of Cancer stem cells

CSCs are responsible for resistance to chemotherapy, radiotherapy, and immunotherapy [14]. The survival of CSCs enables the replenishment of tumor cell populations and contributes to cancer recurrence [9]. One of the most effective methods for detecting CSCs within tumors is the use of CSC-specific biomarkers. Based on their cellular distribution, CSC markers can be classified into intracellular markers and cell-surface markers. Intracellular markers include transcription factors that function in the nucleus and markers found in the cytoplasm. Current studies have reported various molecules as CSC markers in solid tumors as outlined in Table 2.

Table 2 Cancer stem cell markers for solid tumors

A diverse array of cell-surface proteins serve as potential markers for CSCs in solid tumors. Among these, C-X-C Chemokine Receptor Type 4 (CXCR4), also referred to as CD184, is a CXC chemokine receptor that is encoded by the CXCR4 gene [241]. CXCR4 plays a significant role in cancer progression by activating various signaling pathways, such as the PI3K/AKT, PLC, hedgehog, ERK1/2, and JAK/STAT pathways [242]. LGR5, also known as G-Protein Coupled Receptor 49 (GPR49) or G-Protein Coupled Receptor 67 (GPR67), is encoded by the LGR5 gene. LGR5 has been identified as a part of the WNT signaling complex that potentiates WNT/β-Catenin signaling [243]. The significant involvement of the WNT signaling pathway in the maintenance of CSC properties has led to the recognition of LGR5 as a cell-surface marker in various types of solid tumors. CD24, alternatively referred to as Heat Stable Antigen (HSA), is encoded by the CD24 gene and serves as a cell-cell adhesion molecule [244]. CD24 is involved in various signaling pathways that have the potential to increase the stemness of tumor cells [245]. Similarly, CD44, also known as Homing Cell Adhesion Molecule (HCAM) and Phagocytic Glycoprotein-1 (Pgp-1), induces cell-cell adhesion and interactions [246]. Additionally, it participates in the activation of the PI3K/AKT and Src/MAPK signaling pathways and functions as a co-receptor for c-Met [246]. Both CD24 and CD44 can individually or in combination mark CSCs in several solid tumor types. Additionally, the presence of the CD44+/CD24 − combination serves as a marker for CSCs in breast cancer, prostate cancer, head and neck squamous cell carcinoma, and ovarian cancer. Additional cell surface markers for CSCs are presented in Table 1, which is not elaborated upon in this discussion.

OCT4, SOX2, and Nanog are fundamental transcription factors responsible for governing the embryonic stem cell phenotype [247]. Through cooperative interactions, they facilitate the upregulation of their respective promoters, thereby initiating the expression of genes crucial for preserving embryonic stem cell identity while concurrently suppressing the activation of lineage-specific transcription factors. The presence of these transcription factors in cancer cells confers stem-like characteristics, thereby establishing them as conventional markers for CSCs [248,249,250].

Hypoxia induces drug resistance in tumors

Studies have indicated an increase in drug resistance in hypoxic tumors, with the activation of HIFs associated with tumor resistance and reduced survival rates [251]. Although the role of HIFs in tumor drug resistance has been recognized, the specific molecular pathways leading to reduced drug efficacy are still under investigation. This review summarizes the current knowledge of HIF activation and the regulation of tumor drug resistance and focuses on mechanisms, including enhancing drug efflux, preventing cell death, promoting cell survival, and repairing DNA damage (Fig. 3; Table 3).

Fig. 3
figure 3

Hypoxia induces chemoradiotherapy resistance through a pleiotropic mechanism. Reduced oxygen levels in the TME inhibit the enzymatic function of PHDs, which normally hydroxylate the oxygen-sensitive HIF-α leading to its heterodimerization with the HIF-β subunit. This process induces the expression of GLUTs and glycolytic enzymes, promoting anaerobic glycolysis, intracellular acidification, and acidification of the TME, which is regulated by the coordinated expression of CAIX. The decrease in oxidative phosphorylation metabolism and increase in mitochondrial autophagy during hypoxia reduce the levels of ROS. Under anoxic conditions, drug efflux is facilitated through ATP-binding cassette transporters such as ABCB1, ABCC1, and ABCG2. Upregulation of anti-apoptotic proteins Bcl-2 and IAP-3 leads to decreased apoptosis, while genes associated with EMT enhance the invasiveness of apoptotic cells. Additionally, increased expression of VEGF promotes new angiogenesis, which is a key characteristic of chemotherapy resistance in hypoxic tumors. Abbreviation: TME, tumor microenvironment; PHDs, proline hydroxylase dioxygenase; GLUTs, glucose transporters; CAIX, carbacylase IX; ROS, reactive oxygen species; Bcl-2, B-cell lymphona 2; IAP-3, inhibitor of apoptosis protein 3; EMT, epithelial mesenchymal transformation; VEGF, vascular endothelial growth factor

Table 3 Overview of HIF-1-mediated mechanisms in drug resistance

Hypoxia can induce tumor resistance directly

Hypoxia is a key driver in the development of drug resistance in tumors. Like resistance facilitated by CSCs, hypoxia can contribute to drug resistance by directly influencing various cellular processes. These include the upregulation of drug efflux proteins, the modulation of apoptotic pathways, the initiation of autophagy, the control of mitochondrial function, and the detoxification of ROS.

Hypoxia induces drug resistance in tumors by activating specific pathways. Like that of CSCs, hypoxia can influence the expression of various members of the ABC transporter family, thereby facilitating the efflux of drugs. The transporter genes directly regulated by HIF-1α are ABCB1, ABCC1, and ABCG2, and their promoters contain HREs sensitive to HIF-1α transcriptional regulation. ABCB1 and ABCC1 are linked to unfavorable outcomes in human non-small cell lung cancer in a HIF-1-dependent fashion with resistance to cisplatin and docetaxel [252]. In the case of pancreatic cancer, hypoxic conditions can stimulate HIF-1α binding to a specific gene sequence within the ABCG2 promoter, resulting in increased HIF-1α expression levels [253].

Apoptotic malfunctions contribute to the development of resistance to hypoxia-induced tumors, which is crucial for the efficacy of anti-cancer treatments that rely on the activation of apoptosis through various members of the caspase protease family [254]. The initiation of the intrinsic caspase cascade is intricately linked to components of the Bcl-2 protein family. Under hypoxia, the balance of Bcl-2 family proteins in tumors is disrupted, and the proapoptotic proteins BIM [255] and Bax [256] are inhibited, whereas the antiapoptotic proteins Bcl-2 [257], Bcl-xl [257], and Mcl-1 [258] are upregulated, thus increasing the apoptotic threshold and reducing the efficacy of chemotherapy and radiotherapy. The inhibition of apoptosis induced by hypoxia also involves the P53 and Ras signaling pathways [259].

When cancer cells are subjected to hypoxia, they activate processes that suppress DNA damage and enhance repair mechanisms, leading to resistance to anticancer treatments. Among the various types of DNA damage, DSBs are the most severe and serve as the molecular foundation for the efficacy of radiotherapy in eradicating cancer cells [260]. DSBs are identified through the action of ATM kinase, which regulates the cascade of signal transduction processes that follow [261]. ATM and ATR enzymes are responsible for phosphorylating numerous substrates that play crucial roles in the DNA damage response (DDR) and subsequent DNA repair processes [260, 262]. DSB repair predominantly takes place during the S and G2 phases of the cell cycle through either the NHEJ or HR pathway. The relationship between the DDR under hypoxic conditions and signaling pathways involving HIF-1, ATR, and ATM is well established [263], and cells lacking HIF-1α exhibit reduced expression of numerous DSB receptor proteins [264].

The intracellular removal of ROS plays a crucial role in preventing DNA damage in cancer cells. HIF-2α stimulates the transcription of superoxide dismutase 2 (SOD2) in low-oxygen environments, leading to a decrease in mitochondrial ROS levels [265]. Additionally, in response to the HIF-1α signal, the production of heme is elevated, resulting in increased activation and buildup of catalase, which in turn effectively reduces the levels of ROS [266].

Hypoxia can induce tumor resistance indirectly through stemness

Hypoxia has a regulatory effect on the stemness of CSCs. In the process of maintaining the stemness of tumors through cellular metabolic reprogramming, EMT, and autophagy induction, hypoxia imparts drug resistance to tumors.

Numerous compounds within the glycolysis pathway have been demonstrated to engage in an intricate network involving hypoxia, stem cell characteristics, and resistance to pharmaceutical agents. Among glucose transporters, GLUT family members, including GLUT1 [267], GLUT4 [268], and GLUT5 [269] have been shown to induce antitumor resistance. In terms of glycolytic enzymes, Hexokinase (HK) [270], 6-fructose-2-phosphate kinase (PFKFB) [271], aldolase A [272], phosphoglycerol kinase 1 (PGK1) [273], enolase [274], pyruvate kinase (PK) [275], and lactate dehydrogenase (LDH) [276] are involved in inducing antitumor resistance. Furthermore, enzymes that indirectly influence glycolysis metabolism are linked to drug resistance. The pyruvate dehydrogenase (PDH) complex, which mediates the glycolytic to tricarboxylic acid cycle pathway, is inhibited by overexpressed PDK1 under hypoxic conditions [40], and PDH inhibition [277] and PDK overexpression [278] have also been found to induce antitumor resistance. Importantly, resistance can arise from the direct influence of glucose transporters and glycolytic enzymes or from the promotion of glycolytic metabolism. The process of drug resistance caused by glycolysis and its intermediary elements encompasses a wide range of mechanisms associated with hypoxia, including antiapoptotic effects [271, 273, 279], the induction of autophagy [276, 280], and the promotion of drug effectiveness [281]. The increased glycolytic activity observed in cancer cells is believed to be linked to the initiation of EMT [282].

Hypoxia regulates drug resistance through EMT, a process linked to the tumor stemness phenotype. The increased genetic expression typically in stromal cells is closely linked to therapeutic resistance. This transcriptional upregulation is partially driven by the activation of the EMT program in cancer cells [283]. EMT-TFs play roles in the development of acquired resistance. Numerous promoters of ABC transporter genes contain EMT-TF binding sites. Proteins such as SNAIL, Twist, and ZEB1 are known to trigger EMT and are linked to the regulation of P-glycoprotein and other ABC transporters [284,285,286]. EMT-TFs also induce drug resistance independently of ABC transporters by increasing cell resistance to drug-induced apoptosis. Furthermore, following conventional treatment, CSCs may undergo EMT, leading to distant metastasis and the subsequent reacquisition of self-renewal capabilities after a certain duration, a phenomenon referred to as recurrence [287]. This occurrence poses a significant challenge to the efficacy of therapeutic interventions.

The SNAIL protein suppresses the transcription of cyclin D2, slowing cell cycle progression. This action results in enhanced resistance to both internal and external apoptotic pathways, ultimately contributing to cellular resilience against DNA damage responses [288]. EMT-TFs play crucial roles in the cellular response to DNA damage repair [289]. For example, Zeb1 plays a role in safeguarding cells against genotoxic stress induced by chemotherapy by activating CHK1 and facilitating DNA repair through recombination mechanisms [290]. The activation of P53 through phosphorylation has been demonstrated to phosphorylate TWIST, thereby controlling the expression of various genes associated with cell cycle arrest (such as p21) and apoptosis (including BAX and Puma) [291]. The increased expression of SNAIL, in conjunction with EMT, triggers metabolic reprogramming in cancer cells, which is characterized by increased glucose absorption and lactate generation, as well as decreased mitochondrial respiration [292, 293].

Autophagy has been demonstrated to increase the survival of tumor cells and contribute to treatment resistance. In preclinical studies of breast cancer, autophagy has been found to support the survival of quiescent disseminated cells and is essential for metastasis following the dormant phase [294]. Regulatory processes control hypoxia and autophagy, as well as the mechanisms by which autophagy is upregulated, and drug resistance is induced in CSCs. However, the current literature lacks a comprehensive analysis that fully integrates the pathways of hypoxia, autophagy, and drug resistance.

The process of developing drug resistance under hypoxic conditions can be succinctly described as involving the upregulation of drug efflux proteins, the modulation of apoptosis signaling pathways, the initiation of autophagy, and alterations in DNA damage and repair pathways. Furthermore, tumor stemness is intricately linked to drug resistance, with hypoxia playing a crucial role in maintaining stemness and contributing to tumor drug resistance. Alterations in hypoxia-induced stemness, such as modifications in glucose metabolism, EMT, and autophagy, have been shown to result in drug resistance. These processes operate through various mechanisms that can directly affect pathways associated with drug resistance or be modulated by the stemness of tumors. The relationship between hypoxia and stemness involves a complex interaction network involving multiple pathways implicated in tumor drug resistance.

Hypoxia leads to the development of immune resistance in tumors

Hypoxia has been shown to impact the efficacy of immunotherapy through various mechanisms. Specifically, a decrease in oxygen levels can lead to a diminished proportion of antitumor immune cells relative to immune-tolerant or immunosuppressive cells. Furthermore, hypoxia upregulates the expression and function of immune checkpoints (ICPs) and their ligands (ICPLs) in both immune and cancer cells. The concurrent presence of immunosuppressive cells, anergic effector cells, and immune evasion cancer cells significantly limits the effectiveness of immune checkpoint inhibitors (ICPIs).

The hypoxic conditions within the TME significantly modulate the function of immunosuppressive cell populations. Through mechanisms like increased lactate secretion and carbonic anhydrase-driven carbonic acid production, hypoxia lowers extracellular pH to between 5.8 and 6.5, creating an acidic environment. Vascular irregularities associated with hypoxic regions impede the influx of circulating T lymphocytes. Moreover, increased collagen levels in the hypoxic tumor stroma hinder the outward movement and penetration of CD8 + T lymphocytes [295]. In instances where tumor cells persist under acidic conditions, the presence of lactic acid diminishes the functionality of CD8 + T cells by reducing their viability, proliferation rate, cytolytic activity, and secretion of proinflammatory cytokines [296,297,298]. Furthermore, the development of dendritic cells crucial for facilitating the proliferation of CD8 + T lymphocytes is impeded within an oxygen-deprived environment abundant in lactic acid [299]. Conversely, elevating the pH within the TME has the potential to relieve the suppression of T cells and increase the expression of IFN-γ in CD8 + T cells, thereby substantially enhancing the effectiveness of both active and adoptive immunotherapy strategies [300]. In instances of tumor hypoxia, continuous T cell receptor (TCR) activation expedites the transformation of effector CD8 + T cells into a state of exhaustion, facilitating the upregulation of immunosuppressive effector molecules [301].

Furthermore, in addition to its impact on T lymphocytes, HIF-1α hinders the functionality of NK cells by inhibiting the upregulation of key receptors involved in NK cell activation, including NKp46, NKp30, NKp44, and NKG2D [302]. Myeloid-derived suppressor cells (MDSCs), a distinct subset of immunosuppressive cells abundant in the hypoxic TME, diminish the activation of CD8 + T lymphocytes through the secretion of the inhibitory cytokines IL-10 and IL-6. Concurrently, the transcription factor HIF-1α upregulates the expression of the immune checkpoint molecules PD-L1 and PD-L2 on MDSCs, thereby transforming these cells into a central hub of immunosuppression [303]. M2-polarized TAMs are prevalent in hypoxic tumor environments and are driven by HIF-1α transcriptional programs that promote the transition from M1 to M2 polarization [304].

Treg cells present metabolic advantages in low-glucose, high-lactate environments. This metabolic advantage facilitates the promotion of peripheral immune tolerance and aids cancer cells in evading immune destruction within the TME [305]. Furthermore, hypoxia induces the build-up of the suppressive metabolite adenosine in Treg cells by facilitating the HIF-1α-mediated expression of the extracellular ATP-degrading enzymes CD39 and CD73 on Treg cells. The limitation of extracellular ATP restricts the activation of T cell receptors, whereas elevated levels of free adenosine suppress cytotoxicity by interacting with A2AR and A2BR receptors [306,307,308]. HIF-1α also regulates the degradation of the transcription factor forkhead box P3 (FoxP3), which physiologically converts effector T cells into Treg cells, thereby reducing the anticancer activity of tumor-infiltrating lymphocytes (TILs) [309].

Hypoxia directly regulates the expression and activity of ICP and its ligands. The PD-L1 promoter contains HREs and is a specific target of HIF-1α, which may be upregulated in hypoxic environments [303]. Research has demonstrated an increase in PD-L1 expression during the process of EMT, with PD-L1 signaling playing a role in sustaining EMT [310]. These findings indicate a reciprocal regulation between EMT and PD-L1, with EMT serving as a bridge between stemness and hypoxia. Furthermore, multiple investigations have revealed diverse interactions between PD-L1 and HIF-1α-dependent pathways, including the PI3K/mTOR [311] and NF-κB pathways(reviewed in [312]). Hypoxia impacts the structure and binding of ICP by inducing posttranslational modifications or modifying the lipid environment in which ICP is located. Additionally, various pathways influence the localization of ICP between the plasma membrane and endosomes, as discussed in a previous review [313].

In brief, hypoxia recruits and activates immunosuppressive cell subsets; elevates PD-1 expression on cancer, stromal, and immune cells; and facilitates the development of an immunosuppressive microenvironment within tumors. Consequently, tumors exhibit resistance to immune checkpoint inhibitor therapies under hypoxic conditions [314, 315].

Perspective

Hypoxia triggers aggressive tumor traits, including metabolic reprogramming, apoptosis inhibition, autophagy initiation, enhanced migration, and angiogenesis. Additionally, hypoxic tumors exhibit increased resistance to chemotherapy and radiotherapy (Fig. 4). The impact of hypoxia on tumors extends beyond cancer cells to encompass other tumor-associated entities such as endothelial cells and immune infiltrating cells. The reactions of each constituent are intricately interconnected and collaborate to engender tumors that are more aggressive and resilient to standard treatments. Endothelial cells respond to hypoxia by releasing angiogenic factors, which contribute to the establishment of new vascular environments and sustain tumor stemness. Immune infiltrating cells exhibit reduced antitumor cytotoxicity, increased immune checkpoint protein expression, and increased levels of immunotolerant or immunosuppressive cells. Alterations in the immune microenvironment represent significant mechanisms underlying tumor immune evasion and immune resistance. Particularly in advanced or metastatic cancer, a hypoxic TME is frequently established and plays a pivotal role in cancer progression. Investigations on hypoxia have offered novel perspectives on its mechanisms and set the groundwork for targeted therapies with potentially improved outcomes. Nevertheless, owing to the intimate association between hypoxia and tumor stemness, achieving optimal outcomes solely with targeted therapies for hypoxia or stemness remains challenging [13].

Fig. 4
figure 4

The cross-relationship between hypoxia, stemness, drug resistance, current research deficiencies, and potential treatment therapy. The interplay between hypoxia and stemness is intricate, and both elements play significant roles in influencing drug resistance. Current research in this field is constrained by several factors, including the absence of optimal biomarkers, insufficient investigation into associated mechanisms, and the necessity to identify more appropriate therapeutic targets. There exist targeted therapies aimed at addressing hypoxia and CSCs; however, the majority of these therapies are still undergoing clinical trials. Notably, the most promising interventions tend to be combined therapy. Abbreviation: CSCs, cancer stem cells; EMT, EMT, epithelial-mesenchymal transition; HIFs, hypoxia-inducible factors

Hypoxia impacts tumor cell behavior, initiating or exacerbating stemness-associated traits. Numerous targets of HIFs have been identified, with alterations in glucose metabolism, angiogenesis, increased autophagy, and activation of the EMT program being extensively linked to tumor stemness. Conversely, increased tumor stemness increases cancer cell survival, as increased metabolism and self-renewal can lead to local hypoxia. The reciprocal stimulation of hypoxia and CSCs establishes a positive feedback loop, rendering the tumor more resilient to conventional treatments. With respect to drug resistance mechanisms, hypoxia and CSCs comparably contribute to tumor resistance. ABC transporters facilitate drug efflux, whereas the modulation of apoptosis pathways, autophagy induction, and ROS clearance are common drug resistance mechanisms. Within the hypoxia regulatory network, molecules such as glycolytic enzymes, GLUTs, Snail, Twist, Zeb1, and other transcription factors are linked to drug resistance. This occurs either by reinforcing glycolytic and EMT pathways to enhance tumor stemness or through direct involvement in resistance mechanisms.

Given the significant impact of hypoxia on cancer development, interest in increasing oxygen levels within the TME as a potential therapeutic approach is increasing. However, the direct intravenous delivery of oxygen to cancer lesions is hindered by systemic exposure to ROS and the limited solubility of oxygen in the bloodstream. One proposed strategy to increase the oxygen supply involves augmenting blood flow to the tumor via vasodilatory drugs; however, practical challenges exist in how to effectively modulate blood perfusion to the tumor tissue [316]. To address this obstacle, researchers are utilizing nanoscale and bioengineering methodologies to facilitate the targeted delivery of oxygen to specific tumor sites, thereby increasing the efficacy of chemotherapy, radiation therapy, and immunotherapy [317, 318]. However, nearby healthy cell destruction is an inevitable consequence. Another effective approach is to block the activation of HIFs and their associated genes. Compounds that inhibit HIF-1α, such as topotecan and bortezomib (PS-341), as well as HIF-2α inhibitors, such as PT239, PT2385, and PT2977, have been shown to suppress the activity of HIF1/2α and the expression of their downstream genes. Clinical trials are also investigating the inhibition of downstream targets of HIF to achieve anticancer effects (Supplementary Table 1). The aggressive biological characteristics of CSCs are supported by a complex interplay of various signaling pathways, suggesting the potential importance of focusing on traditional pathways in the treatment of CSCs. Preclinical studies have confirmed the viability of targeting signaling pathways within CSCs. ICG-001, an inhibitor of the WNT signaling pathway, effectively diminishes the stemness and metastatic characteristics of colorectal cancer cells by inhibiting the expression of the downstream target gene of the WNT pathway, Myeloid Ecotropic Viral Insertion Site 1 (MEIS1) [319]. In a similar vein, pharmacological agents that target the Notch, hedgehog, PI3K/AKT, and NF-κB signaling pathways have demonstrated potential utility in addressing tumor resistance [320,321,322,323]. Cell surface markers, including CD13, CD44, and CD133, as well as intracellular markers, such as Nanog, ALDH1, and SOX2, serve as effective indicators for the identification of rare populations of CSCs. These markers are significant targets for the eradication of the diverse malignant biological behaviors exhibited by CSCs [324, 325]. In addition to focusing on the markers, signaling pathways, and niches associated with CSCs, alternative strategies for the eradication of CSCs may involve the modulation of genes related to stemness, the alteration of aberrant metabolic processes, and the manipulation of non-coding RNAs, among other approaches [326, 327]. Further research is required at the genetic, epigenetic, proteomic, and metabolic levels to identify potential targets for CSCs, helping to develop new therapies that are more effective to the existing challenge of cancer drug resistance [328].

The identification of CSC-specific antigens or biomarkers remains a challenge. Cell surface markers, especially transport proteins and signaling receptors, have attracted considerable interest because of their potential to enhance diagnostic capabilities and enable the targeted delivery of therapeutic agents to CSCs [329]. However, the non-specificity and low abundance of these markers present substantial barriers to their practical application. Currently identified surface markers do not exhibit specificity for any particular CSC type, as they are also expressed on non-CSCs or healthy cells, albeit at reduced levels [330]. Despite extensive research on the relationship between hypoxia and tumor stemness in recent years, the expected broad use of medical drugs aimed at these factors has not occurred as anticipated. Many studies have focused only on establishing correlations between hypoxia and tumor stemness rather than delving into the underlying mechanisms involved. For example, the key regulators of hypoxia, HIFs, govern a multitude of target genes, prompting inquiries about the potential presence of additional regulators in addition to HIFs. Similarly, while numerous stemness-related genes are actively expressed in CSCs, the intricate interactions among them that ultimately contribute to the stemness phenotype remain unclear. Furthermore, the possibility of undiscovered stemness-related genes and understanding the full functions of known stemness-related genes necessitates further investigation. Moreover, tumor stemness and hypoxia mutually reinforce each other. Therapies focused solely on hypoxia often fail to achieve optimal outcomes because they neglect the regulatory influence of local tumor hypoxia. Conversely, treatments that focus on tumor stemness often overlook the importance of regulating hypoxia. Hypoxia in tumors is highly heterogeneous and lacks a sensitive detection method, posing challenges in identifying potential patient populations for hypoxia-targeting drugs. Recent investigations have revealed that CSCs exhibit plasticity, allowing for phenotypic alterations under particular conditions. Nevertheless, contemporary research concerning the microenvironment and CSCs predominantly depends on tumor implantation studies in murine models. These models are unable to accurately mimic the microenvironment of primary tumors or the interactions between human CSCs and their respective microenvironments, thereby imposing certain constraints on the findings [14]. Consequently, addressing these complexities is crucial for advancing the development and efficacy of therapies targeting hypoxia and tumor stemness.

The divergent findings presented in numerous studies underscore the necessity for further exploration of the interplay between hypoxia, CSCs and drug resistance. For example, the variability in ROS levels among different types of CSCs raises questions about whether diminished ROS levels can entirely explain the preservation of the stem cell phenotype, particularly in light of the altered glucose metabolism characteristic of these cells [53, 54]. Indeed, numerous conclusions remain to be substantiated across various types of CSCs, as their characteristics are not entirely uniform. Furthermore, the function of autophagy in the drug resistance exhibited by CSCs remains ambiguous. Some viewpoints propose that autophagy may serve a dual role in this context, prompting inquiries into the observed elevation of autophagy levels in hypoxic environments [331].

Emerging technologies such as organoids, three-dimensional (3D) printing technology, AI-driven drug discovery, single-cell transcriptomics, and CRISPR/Cas9 gene editing are significantly advancing the ability to target both hypoxia and CSCs, two key factors driving tumor progression and therapy resistance. The simulation of TME and the interactions between CSCs and the TME present significant challenges that impede current research efforts. Organoids and 3D printing technology hold considerable potential for accurately replicating the in vivo environment. These advanced methodologies can yield critical insights into the impact of hypoxia on CSCs and offer a platform for the development of personalized therapeutic strategies. AI-driven drug discovery accelerates the identification of compounds that can simultaneously target hypoxia-induced signaling pathways and CSCs, while single-cell transcriptomics enables the detailed profiling of gene expression at the single-cell level, revealing new molecular drivers of stemness under hypoxic stress. Furthermore, the advancement of single-cell transcriptomics and associated cell isolation methodologies has made it possible to explore more efficient approaches for the isolation and identification of CSCs, as well as to discover additional significant protein markers. The CRISPR/Cas9 gene editing technology is especially valuable for elucidating the molecular mechanisms that govern the relationship between hypoxia and cancer stem cells. This approach facilitates the identification of novel therapeutic targets and potential mechanisms of resistance. These advancements are paving the way for more effective, targeted therapies that address the complexities of hypoxia and CSCs, with the potential to overcome resistance and reduce tumor relapse.

Conclusions

This review highlights the complex interplay among hypoxia, stemness, and drug resistance. The induction of tumor stemness within a hypoxic microenvironment complicates the efficacy of tumor treatments, while the mechanism by which hypoxia triggers tumor stemness offers insights for targeted therapeutic approaches. As HIFs serve as primary regulators of oxygen homeostasis under hypoxic conditions, targeting HIFs and their downstream pathways has emerged as a promising strategy in cancer treatment. The stabilization of HIFs in hypoxic cancer cells prompts the expression of specific target genes encoding proteins involved in processes such as angiogenesis, metabolic alterations, EMT, metastasis, and autophagy, thereby influencing tumor stemness. These proteins play crucial roles as mediators of hypoxia and stemness, and inhibiting their activity through targeted therapies has the potential to overcome tumor resistance. It is essential to utilize emerging technologies to model the interactions between the hypoxic TME and CSCs. Furthermore, it is viable to investigate novel markers and genes associated with CSCs, as well as to explore regulatory targets related to hypoxia or identify new hypoxia-inducible factors.

Data availability

No datasets were generated or analysed during the current study.

Abbreviations

ABC:

ATP-binding cassette

ABCG2:

ATP-binding cassette subfamily G member 2

ADCC:

Antibody-dependent cell-mediated cytotoxicity

AGGF1:

G-patch and FHA domains 1

ANGPT-2:

Angiopoietin-2

ATGs:

Autophagy-related genes

ATM:

Ataxia-telangiectasia mutated

BER:

Base excision repair

BIRC5:

Baculoviral inhibitor of apoptosis repeat containing 5

CAFs:

Cancer-associated fibroblasts

CD44s:

CD44 subtype

CHK1:

Checkpoint kinase 1

CMA:

Chaperone-mediated autophagy

CSCs:

Cancer stem cells

CTL:

Cytotoxic T lymphocyte

CXCL-12:

Chemokine C-X-C motif ligand-12

CXCR4:

C-X-C Chemokine Receptor Type 4

DC:

Dendritic cell

DDR:

DNA damage response

DSB:

Double strand breaks

ECM:

Extracellular matrix

EGFL7:

EGF-like domain 7

EMT:

Epithelial-mesenchymal transition

EMT-TFs:

Transcription factors associated with epithelial-mesenchymal transition

ENO:

Enolase

ER:

Estrogen receptor

ETC:

Electron transport chain

5-FU:

5-fluorouracil

FoxP3:

Forkhead box P3

GLUT:

Glucose transporter

GPR49:

G-Protein Coupled Receptor 49

GPR67:

G-Protein Coupled Receptor 67

GSCs:

Glioblastoma stem cells

HCAM:

Homing Cell Adhesion Molecule

HIFs:

Hypoxia-inducible factors

HK:

Hexokinase

HNSCC:

Head and neck squamous cell carcinoma

HR:

Homologous recombination

HREs:

Hypoxia response elements

IAPs:

Inhibitors of apoptosis family proteins

ICP:

Immune checkpoints

ICPI:

Immune checkpoint inhibitors

ICPLs:

The ligands of immune checkpoints

JAK:

Janus Kinase

KLF5:

Krüppel-like factor 5

LDH:

Lactate dehydrogenase

lncRNAs:

Long non-coding RNAs

MACC1:

Metastasis-associated colon cancer 1

MDR:

Multidrug resistance

MDSCs:

Myeloid-derived suppressor cells

MHC-I:

Major histocompatibility complex class I

MRN:

MRE11-RAD50-NBS1

MRP1:

Multidrug resistance-associated protein 1

MSCs:

Mesenchymal stem cells

NER:

Nucleotide excision repair

NHEJ:

Non-homologous end joining

NK:

Natural kill

OCT4:

Octamer-binding transcription factor 4

OXPHOS:

Oxygen-dependent mitochondrial oxidative phosphorylation

PDH:

Pyruvate dehydrogenase

PDK1:

Pyruvate dehydrogenase kinase 1

PFKFB:

6-fructose-2-phosphate kinase

PGF:

Placenta growth factor

PGK1:

Phosphoglycerol kinase 1

Pgp-1:

Phagocytic Glycoprotein-1

PHGDH:

Phosphoglycerate dehydrogenase

PK:

Pyruvate kinase

PKM2:

M2 pyruvate kinase

PRR:

Post-replication repair

ROS:

Reactive oxygen species

SCF:

Stem cell factor

SOCS1:

Suppressor of cytokine signaling 1

SOD2:

Superoxide dismutase 2

SOX2:

SRY-box 2

STAT3:

Signal transducer and activator of transcription 3

TAMs:

Tumor-associated macrophages

TCA:

Tricarboxylic acid

TCR:

T cell receptor

TILs:

Tumor-infiltrating lymphocytes

TME:

Tumor microenvironment

Treg cells:

Regulatory T cells

ULBP:

UL16-binding protein

VEGF:

Vascular endothelial growth factor

VHL:

Von Hippel-Lindau

References

  1. Bray Fea. Global cancer statistics 2022: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin 2024 May-Jun;74(3):229–63.

  2. Mullard A. Stemming the tide of drug resistance in cancer. Nat Rev Drug Discov. 2020;19(4):221–3.

    Article  CAS  PubMed  Google Scholar 

  3. Dagogo-Jack IS. Tumour heterogeneity and resistance to cancer therapies. Nat Rev Clin Oncol. 2018;15(2):81–94.

    Article  CAS  PubMed  Google Scholar 

  4. Quail DF, Joyce JA. Microenvironmental regulation of tumor progression and metastasis. Nat Med. 2013;19(11):1423–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Zou W, Green DR. Beggars banquet: metabolism in the tumor immune microenvironment and cancer therapy. Cell Metab. 2023;35(7):1101–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Kundu Mea. Modulation of the tumor microenvironment and mechanism of immunotherapy-based drug resistance in breast cancer. Mol Cancer. 2024;23(1):92.

    Article  Google Scholar 

  7. Chen Z, Han F, Du Y, Shi H, Zhou W. Hypoxic microenvironment in cancer: molecular mechanisms and therapeutic interventions. Signal Transduct Target Ther. 2023;8(1):70.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Jing X, et al. Role of hypoxia in cancer therapy by regulating the tumor microenvironment. Mol Cancer. 2019;18(1):157.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Loh JJ, Ma S. Hallmarks of cancer stemness. Cell Stem Cell. 2024;31(5):617–39.

    Article  CAS  PubMed  Google Scholar 

  10. Lawson DA, et al. Single-cell analysis reveals a stem-cell program in human metastatic breast cancer cells. Nature. 2015;526(7571):131–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Saw PE, et al. Cancer stem cell mimicry for immune evasion and therapeutic resistance. Cell Stem Cell. 2024;31(8):1101–12.

    Article  CAS  PubMed  Google Scholar 

  12. Mai Y, et al. The strategies to cure cancer patients by eradicating cancer stem-like cells. Mol Cancer. 2023;22(1):171.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Yuan X, Ruan W, Bobrow B, Carmeliet P, Eltzschig H. K. Targeting hypoxia-inducible factors: therapeutic opportunities and challenges. Nat Rev Drug Discov. 2024;23(3):175–200.

    Article  CAS  PubMed  Google Scholar 

  14. Chu Xea. Cancer stem cells: advances in knowledge and implications for cancer therapy. Signal Transduct Target Ther. 2024;9(1):170.

    Article  Google Scholar 

  15. Bussard KM, et al. Tumor-associated stromal cells as key contributors to the tumor microenvironment. Breast Cancer Res. 2016;18(1):84.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Liu T, et al. Cancer-associated fibroblasts: an emerging target of anti-cancer immunotherapy. J Hematol Oncol. 2019;12(1):86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Zhao Y, et al. ROS signaling under metabolic stress: cross-talk between AMPK and AKT pathway. Mol Cancer. 2017;16(1):79.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Wigerup C, Påhlman S, Bexell D. Therapeutic targeting of hypoxia and hypoxia-inducible factors in cancer. Pharmacol Ther. 2016;164:152–69.

    Article  CAS  PubMed  Google Scholar 

  19. Gilkes DM, Semenza GL, Wirtz D. Hypoxia and the extracellular matrix: drivers of tumour metastasis. Nat Rev Cancer. 2014;14(6):430–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Bonnet D, Dick JE. Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell. Nat Med. 1997;3(7):730–7.

    Article  CAS  PubMed  Google Scholar 

  21. Plaks V, Kong N, Werb Z. The cancer stem cell niche: how essential is the niche in regulating stemness of tumor cells? Cell Stem Cell. 2015;16(3):225–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Batlle E, Clevers H. Cancer stem cells revisited. Nat Med. 2017;23(10):1124–34.

    Article  CAS  PubMed  Google Scholar 

  23. Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144(5):646–74.

    Article  CAS  PubMed  Google Scholar 

  24. Li QS, Zheng PS. ESRRB inhibits the TGFβ signaling pathway to drive cell proliferation in cervical Cancer. Cancer Res. 2023;83(18):3095–114.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Keith B, Simon MC. Hypoxia-inducible factors, stem cells, and cancer. Cell. 2007;129(3):465–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Schödel J, Ratcliffe PJ. Mechanisms of hypoxia signalling: new implications for nephrology. Nat Rev Nephrol. 2019;15(10):641–59.

    Article  PubMed  Google Scholar 

  27. Kierans SJ, Taylor CT. Regulation of Glycolysis by the hypoxia-inducible factor (HIF): implications for cellular physiology. J Physiol. 2021;599(1):23–37.

    Article  CAS  PubMed  Google Scholar 

  28. Jan S, et al. Cancer drug resistance and metabolic reprogramming. In: Khan SU, Malik F, editors. Drug resistance in cancer: mechanisms and strategies. Singapore: Springer Nature Singapore; 2024. pp. 183–214.

    Chapter  Google Scholar 

  29. Li L, et al. Hypoxia-induced GBE1 expression promotes tumor progression through metabolic reprogramming in lung adenocarcinoma. Signal Transduct Target Ther. 2020;5(1):54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Jacquier V, et al. RIP140 inhibits glycolysis-dependent proliferation of breast cancer cells by regulating GLUT3 expression through transcriptional crosstalk between hypoxia induced factor and p53. Cell Mol Life Sci. 2022;79(5):270.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Hamann I, et al. Expression and function of hexose transporters GLUT1, GLUT2, and GLUT5 in breast cancer-effects of hypoxia. Faseb J. 2018;32(9):5104–18.

    Article  CAS  PubMed  Google Scholar 

  32. Wang JZ, et al. The role of the HIF-1α/ALYREF/PKM2 axis in Glycolysis and tumorigenesis of bladder cancer. Cancer Commun (Lond). 2021;41(7):560–75.

    Article  PubMed  Google Scholar 

  33. Semenza GL, et al. Hypoxia response elements in the aldolase A, enolase 1, and lactate dehydrogenase A gene promoters contain essential binding sites for hypoxia-inducible factor 1. J Biol Chem. 1996;271(51):32529–37.

    Article  CAS  PubMed  Google Scholar 

  34. Zheng F, et al. The HIF-1α antisense long non-coding RNA drives a positive feedback loop of HIF-1α mediated transactivation and Glycolysis. Nat Commun. 2021;12(1):1341.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Li Q, et al. Circular RNA MAT2B promotes Glycolysis and malignancy of hepatocellular carcinoma through the miR-338-3p/PKM2 Axis under hypoxic stress. Hepatology. 2019;70(4):1298–316.

    Article  CAS  PubMed  Google Scholar 

  36. Guo W, et al. MiR-199a-5p is negatively associated with malignancies and regulates Glycolysis and lactate production by targeting hexokinase 2 in liver cancer. Hepatology. 2015;62(4):1132–44.

    Article  CAS  PubMed  Google Scholar 

  37. Zhang LF, et al. Suppression of miR-199a maturation by HuR is crucial for hypoxia-induced glycolytic switch in hepatocellular carcinoma. Embo J. 2015;34(21):2671–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Zhao Q, et al. Hypoxia-induced circRNF13 promotes the progression and Glycolysis of pancreatic cancer. Exp Mol Med. 2022;54(11):1940–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Lin J, et al. Hypoxia-induced Exosomal circPDK1 promotes pancreatic cancer Glycolysis via c-myc activation by modulating miR-628-3p/BPTF axis and degrading BIN1. J Hematol Oncol. 2022;15(1):128.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Peng F, et al. Glycolysis gatekeeper PDK1 reprograms breast cancer stem cells under hypoxia. Oncogene. 2018;37(8):1062–74.

    Article  CAS  PubMed  Google Scholar 

  41. Huang D, et al. HIF-1-mediated suppression of acyl-CoA dehydrogenases and fatty acid oxidation is critical for cancer progression. Cell Rep. 2014;8(6):1930–42.

    Article  CAS  PubMed  Google Scholar 

  42. Sulkshane P, et al. Ubiquitination and receptor-mediated mitophagy converge to eliminate oxidation-damaged mitochondria during hypoxia. Redox Biol. 2021;45:102047.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Chan SY, et al. MicroRNA-210 controls mitochondrial metabolism during hypoxia by repressing the iron-sulfur cluster assembly proteins ISCU1/2. Cell Metab. 2009;10(4):273–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Tang K, et al. Hypoxia promotes breast Cancer cell growth by activating a glycogen metabolic program. Cancer Res. 2021;81(19):4949–63.

    Article  CAS  PubMed  Google Scholar 

  45. Ling S, et al. USP22 promotes hypoxia-induced hepatocellular carcinoma stemness by a HIF1α/USP22 positive feedback loop upon TP53 inactivation. Gut. 2020;69(7):1322–34.

    Article  CAS  PubMed  Google Scholar 

  46. Yang D, et al. Oxidized ATM promotes breast cancer stem cell enrichment through energy metabolism reprogram-mediated acetyl-CoA accumulation. Cell Death Dis. 2020;11(7):508.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Ji CC, et al. A ketogenic diet attenuates proliferation and stemness of glioma stem–like cells by altering metabolism resulting in increased ROS production. Int J Oncol. 2020;56(2):606–17.

    CAS  PubMed  Google Scholar 

  48. Li B, et al. Targeting glutaminase 1 attenuates stemness properties in hepatocellular carcinoma by increasing reactive oxygen species and suppressing Wnt/beta-catenin pathway. EBioMedicine. 2019;39:239–54.

    Article  PubMed  Google Scholar 

  49. Cheung EC, Vousden KH. The role of ROS in tumour development and progression. Nat Rev Cancer. 2022;22(5):280–97.

    Article  CAS  PubMed  Google Scholar 

  50. Semenza GL. Hypoxia-inducible factors: coupling glucose metabolism and redox regulation with induction of the breast cancer stem cell phenotype. Embo J. 2017;36(3):252–9.

    Article  CAS  PubMed  Google Scholar 

  51. Samanta D, et al. Correction: PHGDH expression is required for mitochondrial redox homeostasis, breast Cancer stem cell maintenance, and lung metastasis. Cancer Res. 2022;82(6):1153.

    Article  CAS  PubMed  Google Scholar 

  52. Khan SU, et al. Redox balance and autophagy regulation in cancer progression and their therapeutic perspective. Med Oncol. 2022;40(1):12.

    Article  PubMed  Google Scholar 

  53. Chen CL, et al. NANOG metabolically reprograms Tumor-Initiating Stem-like cells through tumorigenic changes in oxidative phosphorylation and fatty acid metabolism. Cell Metab. 2016;23(1):206–19.

    Article  CAS  PubMed  Google Scholar 

  54. Lee, KMea. MYC and MCL1 cooperatively promote Chemotherapy-Resistant breast Cancer stem cells via regulation of mitochondrial oxidative phosphorylation. Cell Metab. 2017;26(4):633–e6477.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Huang Y, Hong W, Wei X. The molecular mechanisms and therapeutic strategies of EMT in tumor progression and metastasis. J Hematol Oncol. 2022;15(1):129.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Yang J, Weinberg RA. Epithelial-mesenchymal transition: at the crossroads of development and tumor metastasis. Dev Cell. 2008;14(6):818–29.

    Article  CAS  PubMed  Google Scholar 

  57. Thiery JP, Acloque H, Huang R, Y, Nieto MA. Epithelial-mesenchymal transitions in development and disease. Cell. 2009;139(5):871–90.

    Article  CAS  PubMed  Google Scholar 

  58. Hapke RY, Haake SM. Hypoxia-induced epithelial to mesenchymal transition in cancer. Cancer Lett. 2020;487:10–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Thiery JP. Epithelial-mesenchymal transitions in tumour progression. Nat Rev Cancer. 2002;2(6):442–54.

    Article  CAS  PubMed  Google Scholar 

  60. Cano A, et al. The transcription factor snail controls epithelial-mesenchymal transitions by repressing E-cadherin expression. Nat Cell Biol. 2000;2(2):76–83.

    Article  CAS  PubMed  Google Scholar 

  61. Comijn J, et al. The two-handed E box binding zinc finger protein SIP1 downregulates E-cadherin and induces invasion. Mol Cell. 2001;7(6):1267–78.

    Article  CAS  PubMed  Google Scholar 

  62. Hajra KM, Chen DY, Fearon ER. The SLUG zinc-finger protein represses E-cadherin in breast cancer. Cancer Res. 2002;62(6):1613–8.

    CAS  PubMed  Google Scholar 

  63. Yang J, et al. Twist, a master regulator of morphogenesis, plays an essential role in tumor metastasis. Cell. 2004;117(7):927–39.

    Article  CAS  PubMed  Google Scholar 

  64. Moreno-Bueno G, Portillo F, Cano A. Transcriptional regulation of cell Polarity in EMT and cancer. Oncogene. 2008;27(55):6958–69.

    Article  CAS  PubMed  Google Scholar 

  65. Zhang XD, Dong XQ, Xu JL, Chen SC, Sun Z. Hypoxia promotes epithelial-mesenchymal transition of hepatocellular carcinoma cells via inducing Twist1 expression. Eur Rev Med Pharmacol Sci. 2017;21(13):3061–8.

    PubMed  Google Scholar 

  66. Hwang-Verslues WW, et al. Loss of corepressor PER2 under hypoxia up-regulates OCT1-mediated EMT gene expression and enhances tumor malignancy. Proc Natl Acad Sci U S A. 2013;110(30):12331–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Wang X, et al. Upregulation of MiR-205 under hypoxia promotes epithelial-mesenchymal transition by targeting ASPP2. Cell Death Dis. 2016;7(12):e2517.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Peng PH, et al. Hypoxia-induced LncRNA RP11-390F4.3 promotes epithelial-mesenchymal transition (EMT) and metastasis through upregulating EMT regulators. Cancer Lett. 2020;483:35–45.

    Article  CAS  PubMed  Google Scholar 

  69. Li H, et al. Long noncoding RNA NORAD, a novel competing endogenous RNA, enhances the hypoxia-induced epithelial-mesenchymal transition to promote metastasis in pancreatic cancer. Mol Cancer. 2017;16(1):169.

    Article  PubMed  PubMed Central  Google Scholar 

  70. Davis FM, et al. Induction of epithelial-mesenchymal transition (EMT) in breast cancer cells is calcium signal dependent. Oncogene. 2014;33(18):2307–16.

    Article  CAS  PubMed  Google Scholar 

  71. Zhang D, et al. Hypoxia modulates stem cell properties and induces EMT through N-glycosylation of EpCAM in breast cancer cells. J Cell Physiol. 2020;235(4):3626–33.

    Article  CAS  PubMed  Google Scholar 

  72. Xing F, et al. Hypoxia-induced Jagged2 promotes breast cancer metastasis and self-renewal of cancer stem-like cells. Oncogene. 2011;30(39):4075–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Liu J, et al. Hypoxia induced ferritin light chain (FTL) promoted epithelia mesenchymal transition and chemoresistance of glioma. J Exp Clin Cancer Res. 2020;39(1):137.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Lei J, et al. Hedgehog signaling regulates hypoxia induced epithelial to mesenchymal transition and invasion in pancreatic cancer cells via a ligand-independent manner. Mol Cancer. 2013;12:66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Zhang J, et al. Hypoxia-inducible factor-1α/interleukin-1β signaling enhances hepatoma epithelial-mesenchymal transition through macrophages in a hypoxic-inflammatory microenvironment. Hepatology. 2018;67(5):1872–89.

    Article  CAS  PubMed  Google Scholar 

  76. Shibue T, Weinberg RA. EMT, CSCs, and drug resistance: the mechanistic link and clinical implications. Nat Rev Clin Oncol. 2017;14(10):611–29.

    Article  PubMed  PubMed Central  Google Scholar 

  77. Mani SA, et al. The epithelial-mesenchymal transition generates cells with properties of stem cells. Cell. 2008;133(4):704–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Hollier BG, et al. FOXC2 expression links epithelial-mesenchymal transition and stem cell properties in breast cancer. Cancer Res. 2013;73(6):1981–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Rhim AD, et al. EMT and dissemination precede pancreatic tumor formation. Cell. 2012;148(1–2):349–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Scheel C, et al. Paracrine and autocrine signals induce and maintain mesenchymal and stem cell States in the breast. Cell. 2011;145(6):926–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Preca BT, et al. A self-enforcing CD44s/ZEB1 feedback loop maintains EMT and stemness properties in cancer cells. Int J Cancer. 2015;137(11):2566–77.

    Article  CAS  PubMed  Google Scholar 

  82. Krebs AM, et al. The EMT-activator Zeb1 is a key factor for cell plasticity and promotes metastasis in pancreatic cancer. Nat Cell Biol. 2017;19(5):518–29.

    Article  CAS  PubMed  Google Scholar 

  83. Hwang WL, et al. SNAIL regulates interleukin-8 expression, stem cell-like activity, and tumorigenicity of human colorectal carcinoma cells. Gastroenterology. 2011;141(1):279–91. 291.e1-5.

    Article  CAS  PubMed  Google Scholar 

  84. Jiang X, et al. The role of microenvironment in tumor angiogenesis. J Exp Clin Cancer Res. 2020;39(1):204.

    Article  PubMed  PubMed Central  Google Scholar 

  85. Eyler CE, Rich JN. Survival of the fittest: cancer stem cells in therapeutic resistance and angiogenesis. J Clin Oncol. 2008;26(17):2839–45.

    Article  CAS  PubMed  Google Scholar 

  86. Lugano R, Ramachandran M, Dimberg A. Tumor angiogenesis: causes, consequences, challenges and opportunities. Cell Mol Life Sci. 2020;77(9):1745–70.

    Article  CAS  PubMed  Google Scholar 

  87. Liao D, Johnson RS. Hypoxia: a key regulator of angiogenesis in cancer. Cancer Metastasis Rev. 2007;26(2):281–90.

    Article  CAS  PubMed  Google Scholar 

  88. Schito L. Hypoxia-Dependent angiogenesis and lymphangiogenesis in Cancer. Adv Exp Med Biol. 2019;1136:71–85.

    Article  CAS  PubMed  Google Scholar 

  89. Manalo DJ, et al. Transcriptional regulation of vascular endothelial cell responses to hypoxia by HIF-1. Blood. 2005;105(2):659–69.

    Article  CAS  PubMed  Google Scholar 

  90. Hsu YL, et al. Hypoxic lung cancer-secreted Exosomal miR-23a increased angiogenesis and vascular permeability by targeting Prolyl hydroxylase and tight junction protein ZO-1. Oncogene. 2017;36(34):4929–42.

    Article  CAS  PubMed  Google Scholar 

  91. Wei X, et al. Mechanisms of vasculogenic mimicry in hypoxic tumor microenvironments. Mol Cancer. 2021;20(1):7.

    Article  PubMed  PubMed Central  Google Scholar 

  92. Lai CY, Schwartz BE, Hsu MY. CD133 + melanoma subpopulations contribute to perivascular niche morphogenesis and tumorigenicity through vasculogenic mimicry. Cancer Res. 2012;72(19):5111–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Calabrese C, et al. A perivascular niche for brain tumor stem cells. Cancer Cell. 2007;11(1):69–82.

    Article  CAS  PubMed  Google Scholar 

  94. Beck B, et al. A vascular niche and a VEGF-Nrp1 loop regulate the initiation and stemness of skin tumours. Nature. 2011;478(7369):399–403.

    Article  CAS  PubMed  Google Scholar 

  95. Yan GN, et al. Endothelial cells promote stem-like phenotype of glioma cells through activating the Hedgehog pathway. J Pathol. 2014;234(1):11–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Kumar S, et al. Identification of vascular cues contributing to cancer cell stemness and function. Angiogenesis. 2022;25(3):355–71.

    Article  CAS  PubMed  Google Scholar 

  97. Charles N, et al. Perivascular nitric oxide activates Notch signaling and promotes stem-like character in PDGF-induced glioma cells. Cell Stem Cell. 2010;6(2):141–52.

    Article  CAS  PubMed  Google Scholar 

  98. Li J, et al. Oxygen-sensitive methylation of ULK1 is required for hypoxia-induced autophagy. Nat Commun. 2022;13(1):1172.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Li Q, et al. HIF-1α-induced expression of m6A reader YTHDF1 drives hypoxia-induced autophagy and malignancy of hepatocellular carcinoma by promoting ATG2A and ATG14 translation. Signal Transduct Target Ther. 2021;6(1):76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Feng X, et al. Hypoxia-induced acetylation of PAK1 enhances autophagy and promotes brain tumorigenesis via phosphorylating ATG5. Autophagy. 2021;17(3):723–42.

    Article  CAS  PubMed  Google Scholar 

  101. He Z, et al. Autophagy-associated circrna circATG7 facilitates autophagy and promotes pancreatic cancer progression. Cell Death Dis. 2022;13(3):233.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Liang Cea. Hypoxia induces Sorafenib resistance mediated by autophagy via activating FOXO3a in hepatocellular carcinoma. Cell Death Dis. 2020;11(11):1017.

    Article  Google Scholar 

  103. Abdul Rahim SA, et al. Regulation of hypoxia-induced autophagy in glioblastoma involves ATG9A. Br J Cancer. 2017;117(6):813–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Yang X, et al. Hypoxia-induced autophagy promotes gemcitabine resistance in human bladder cancer cells through hypoxia-inducible factor 1α activation. Int J Oncol. 2018;53(1):215–24.

    CAS  PubMed  Google Scholar 

  105. Ho TT, et al. Autophagy maintains the metabolism and function of young and old stem cells. Nature. 2017;543(7644):205–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Whelan KA, et al. Autophagy supports generation of cells with high CD44 expression via modulation of oxidative stress and Parkin-mediated mitochondrial clearance. Oncogene. 2017;36(34):4843–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Gong C, et al. Beclin 1 and autophagy are required for the tumorigenicity of breast cancer stem-like/progenitor cells. Oncogene. 2013;32(18):2261–72. 2272e.1-11.

    Article  CAS  PubMed  Google Scholar 

  108. Auzmendi-Iriarte J, et al. Chaperone-Mediated autophagy controls proteomic and transcriptomic pathways to maintain glioma stem cell activity. Cancer Res. 2022;82(7):1283–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Maycotte P, Jones K, L, Goodall ML, Thorburn J, Thorburn A. Autophagy supports breast Cancer stem cell maintenance by regulating IL6 secretion. Mol Cancer Res. 2015;13(4):651–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Iliopoulos D, Hirsch HA, Wang G, Struhl K. Inducible formation of breast cancer stem cells and their dynamic equilibrium with non-stem cancer cells via IL6 secretion. Proc Natl Acad Sci U S A. 2011;108(4):1397–402.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Hassn Mesrati M, Syafruddin SE, Mohtar MA, Syahir A. CD44: A multifunctional mediator of Cancer progression. Biomolecules. 2021;11(12).

  112. Naik PP et al. Autophagy regulates cisplatin-induced stemness and chemoresistance via the upregulation of CD44, ABCB1 and ADAM17 in oral squamous cell carcinoma. Cell Prolif. 2018;51(1).

  113. Liu K, et al. Mitophagy controls the activities of tumor suppressor p53 to regulate hepatic Cancer stem cells. Mol Cell. 2017;68(2):281–e2925.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Yeo SK, Wen J, Chen S, Guan JL. Autophagy differentially regulates distinct breast Cancer Stem-like cells in murine models via EGFR/Stat3 and Tgfβ/Smad signaling. Cancer Res. 2016;76(11):3397–410.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Wu HB, et al. Autophagy-induced KDR/VEGFR-2 activation promotes the formation of vasculogenic mimicry by glioma stem cells. Autophagy. 2017;13(9):1528–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Zhang C, et al. Hypoxia induces the breast cancer stem cell phenotype by HIF-dependent and ALKBH5-mediated m⁶A-demethylation of NANOG mRNA. Proc Natl Acad Sci U S A. 2016;113(14):E2047–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Zhang W, Sui Y, Ni J, Yang T. Insights into the Nanog gene: A propeller for stemness in primitive stem cells. Int J Biol Sci. 2016;12(11):1372–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Nichol D, Stuhlmann H. EGFL7: a unique angiogenic signaling factor in vascular development and disease. Blood. 2012;119(6):1345–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Luo Y, Chen C. The roles and regulation of the KLF5 transcription factor in cancers. Cancer Sci. 2021;112(6):2097–117.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Mirzaei S, et al. SOX2 function in cancers: association with growth, invasion, stemness and therapy response. Biomed Pharmacother. 2022;156:113860.

    Article  CAS  PubMed  Google Scholar 

  121. Ervin EH, French R, Chang CH, Pauklin S. Inside the stemness engine: mechanistic links between deregulated transcription factors and stemness in cancer. Semin Cancer Biol. 2022;87:48–83.

    Article  CAS  PubMed  Google Scholar 

  122. Han J, et al. RNA interference-mediated Silencing of NANOG reduces cell proliferation and induces G0/G1 cell cycle arrest in breast cancer cells. Cancer Lett. 2012;321(1):80–8.

    Article  CAS  PubMed  Google Scholar 

  123. Chen C, et al. KLF5 promotes cell proliferation and tumorigenesis through gene regulation and the TSU-Pr1 human bladder cancer cell line. Int J Cancer. 2006;118(6):1346–55.

    Article  CAS  PubMed  Google Scholar 

  124. Nandan MO, Chanchevalap S, Dalton WB, Yang VW. Krüppel-like factor 5 promotes mitosis by activating the Cyclin B1/Cdc2 complex during oncogenic Ras-mediated transformation. FEBS Lett. 2005;579(21):4757–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Herreros-Villanueva M, et al. SOX2 promotes dedifferentiation and imparts stem cell-like features to pancreatic cancer cells. Oncogenesis. 2013;2(8):e61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Siu MKY et al. Hexokinase 2 regulates ovarian Cancer cell migration, invasion and stemness via FAK/ERK1/2/MMP9/NANOG/SOX9 signaling cascades. Cancers (Basel). 2019;11(6).

  127. Xia W, Bai H, Deng Y, Yang Y. PLA2G16 is a mutant p53/KLF5 transcriptional target and promotes Glycolysis of pancreatic cancer. J Cell Mol Med. 2020;24(21):12642–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Meng HM, et al. Over-expression of Nanog predicts tumor progression and poor prognosis in colorectal cancer. Cancer Biol Ther. 2010;9(4):295–302.

    Article  CAS  PubMed  Google Scholar 

  129. Luo BH, et al. Epidermal growth factor-like domain-containing protein 7 (EGFL7) enhances EGF receptor-AKT signaling, epithelial-mesenchymal transition, and metastasis of gastric cancer cells. PLoS ONE. 2014;9(6):e99922.

    Article  PubMed  PubMed Central  Google Scholar 

  130. Liu R, et al. Krüpple-like factor 5 is essential for mammary gland development and tumorigenesis. J Pathol. 2018;246(4):497–507.

    Article  CAS  PubMed  Google Scholar 

  131. Yin X, et al. Coexpression of gene Oct4 and Nanog initiates stem cell characteristics in hepatocellular carcinoma and promotes epithelial-mesenchymal transition through activation of Stat3/Snail signaling. J Hematol Oncol. 2015;8:23.

    Article  PubMed  PubMed Central  Google Scholar 

  132. Li X, et al. SOX2 promotes tumor metastasis by stimulating epithelial-to-mesenchymal transition via regulation of WNT/β-catenin signal network. Cancer Lett. 2013;336(2):379–89.

    Article  CAS  PubMed  Google Scholar 

  133. Pan X, et al. Site-specific disruption of the Oct4/Sox2 protein interaction reveals coordinated mesendodermal differentiation and the Epithelial-Mesenchymal transition. J Biol Chem. 2016;291(35):18353–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Chiou SH, et al. Coexpression of Oct4 and Nanog enhances malignancy in lung adenocarcinoma by inducing cancer stem cell-like properties and epithelial-mesenchymal transdifferentiation. Cancer Res. 2010;70(24):10433–44.

    Article  CAS  PubMed  Google Scholar 

  135. Sharif T, et al. Autophagic homeostasis is required for the pluripotency of cancer stem cells. Autophagy. 2017;13(2):264–84.

    Article  CAS  PubMed  Google Scholar 

  136. Noh KH, et al. Nanog signaling in cancer promotes stem-like phenotype and immune evasion. J Clin Invest. 2012;122(11):4077–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Hasmim M, et al. Cutting edge: Hypoxia-induced Nanog favors the intratumoral infiltration of regulatory T cells and macrophages via direct regulation of TGF-β1. J Immunol. 2013;191(12):5802–6.

    Article  CAS  PubMed  Google Scholar 

  138. Fiorenza MT, Rava A. The TCL1 function revisited focusing on metabolic requirements of stemness. Cell Cycle. 2019;18(22):3055–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Delfortrie S, et al. Egfl7 promotes tumor escape from immunity by repressing endothelial cell activation. Cancer Res. 2011;71(23):7176–86.

    Article  CAS  PubMed  Google Scholar 

  140. Chen S, et al. SOX2 regulates apoptosis through MAP4K4-survivin signaling pathway in human lung cancer cells. Carcinogenesis. 2014;35(3):613–23.

    Article  CAS  PubMed  Google Scholar 

  141. Ng HH, Surani MA. The transcriptional and signalling networks of pluripotency. Nat Cell Biol. 2011;13(5):490–6.

    Article  CAS  PubMed  Google Scholar 

  142. Visvader JE, Lindeman GJ. Cancer stem cells: current status and evolving complexities. Cell Stem Cell. 2012;10(6):717–28.

    Article  CAS  PubMed  Google Scholar 

  143. Pantel K, Brakenhoff RH. Dissecting the metastatic cascade. Nat Rev Cancer. 2004;4(6):448–56.

    Article  CAS  PubMed  Google Scholar 

  144. Nassar D, Blanpain C. Cancer stem cells: basic concepts and therapeutic implications. Annu Rev Pathol. 2016;11:47–76.

    Article  CAS  PubMed  Google Scholar 

  145. Sarvi S, et al. CD133 + cancer stem-like cells in small cell lung cancer are highly tumorigenic and chemoresistant but sensitive to a novel neuropeptide antagonist. Cancer Res. 2014;74(5):1554–65.

    Article  CAS  PubMed  Google Scholar 

  146. de Morree A, Rando TA. Regulation of adult stem cell quiescence and its functions in the maintenance of tissue integrity. Nat Rev Mol Cell Biol. 2023;24(5):334–54.

    Article  PubMed  PubMed Central  Google Scholar 

  147. Li S, et al. Model of tumor dormancy/recurrence after short-term chemotherapy. PLoS ONE. 2014;9(5):e98021.

    Article  PubMed  PubMed Central  Google Scholar 

  148. Xie XP, et al. Quiescent human glioblastoma cancer stem cells drive tumor initiation, expansion, and recurrence following chemotherapy. Dev Cell. 2022;57(1):32–e468.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Zhou L, et al. NOTCH4 maintains quiescent mesenchymal-like breast cancer stem cells via transcriptionally activating SLUG and GAS1 in triple-negative breast cancer. Theranostics. 2020;10(5):2405–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Ye S, et al. SET Domain-Containing protein 4 epigenetically controls breast Cancer stem cell quiescence. Cancer Res. 2019;79(18):4729–43.

    Article  CAS  PubMed  Google Scholar 

  151. Touil Y, et al. Colon cancer cells escape 5FU chemotherapy-induced cell death by entering stemness and quiescence associated with the c-Yes/YAP axis. Clin Cancer Res. 2014;20(4):837–46.

    Article  CAS  PubMed  Google Scholar 

  152. Steinbichler TB, et al. Therapy resistance mediated by cancer stem cells. Semin Cancer Biol. 2018;53:156–67.

    Article  CAS  PubMed  Google Scholar 

  153. To KKW, Huang Z, Zhang H, Ashby CR, Jr, Fu L. Utilizing non-coding RNA-mediated regulation of ATP binding cassette (ABC) transporters to overcome multidrug resistance to cancer chemotherapy. Drug Resist Updat. 2024;73:101058.

    Article  CAS  PubMed  Google Scholar 

  154. Dean M, Annilo T. Evolution of the ATP-binding cassette (ABC) transporter superfamily in vertebrates. Annu Rev Genomics Hum Genet. 2005;6:123–42.

    Article  CAS  PubMed  Google Scholar 

  155. Nobili S, et al. Role of ATP-binding cassette transporters in cancer initiation and progression. Semin Cancer Biol. 2020;60:72–95.

    Article  CAS  PubMed  Google Scholar 

  156. Zhou S, et al. The ABC transporter Bcrp1/ABCG2 is expressed in a wide variety of stem cells and is a molecular determinant of the side-population phenotype. Nat Med. 2001;7(9):1028–34.

    Article  CAS  PubMed  Google Scholar 

  157. Ohashi R, et al. Expression of MRP1 and ABCG2 is associated with adverse clinical outcomes of papillary thyroid carcinoma with a solid component. Hum Pathol. 2017;67:11–7.

    Article  CAS  PubMed  Google Scholar 

  158. Eyre R, et al. Reversing Paclitaxel resistance in ovarian cancer cells via Inhibition of the ABCB1 expressing side population. Tumour Biol. 2014;35(10):9879–92.

    Article  CAS  PubMed  Google Scholar 

  159. Frank NY, et al. VEGFR-1 expressed by malignant melanoma-initiating cells is required for tumor growth. Cancer Res. 2011;71(4):1474–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Ma L, et al. ABCG2 is required for self-renewal and chemoresistance of CD133-positive human colorectal cancer cells. Tumour Biol. 2016;37(9):12889–96.

    Article  CAS  PubMed  Google Scholar 

  161. Sun M, et al. Enhanced efficacy of chemotherapy for breast cancer stem cells by simultaneous suppression of multidrug resistance and antiapoptotic cellular defense. Acta Biomater. 2015;28:171–82.

    Article  CAS  PubMed  Google Scholar 

  162. Wilson BJ, et al. ABCB5 maintains melanoma-initiating cells through a Proinflammatory cytokine signaling circuit. Cancer Res. 2014;74(15):4196–207.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Wu ZX, et al. Tepotinib reverses ABCB1-mediated multidrug resistance in cancer cells. Biochem Pharmacol. 2019;166:120–7.

    Article  CAS  PubMed  Google Scholar 

  164. Lord CJ, Ashworth A. The DNA damage response and cancer therapy. Nature. 2012;481(7381):287–94.

    Article  CAS  PubMed  Google Scholar 

  165. Krejci L, Altmannova V, Spirek M, Zhao X. Homologous recombination and its regulation. Nucleic Acids Res. 2012;40(13):5795–818.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. King HO, et al. RAD51 is a selective DNA repair target to radiosensitize glioma stem cells. Stem Cell Rep. 2017;8(1):125–39.

    Article  CAS  Google Scholar 

  167. Abad E, Graifer D, Lyakhovich A. DNA damage response and resistance of cancer stem cells. Cancer Lett. 2020;474:106–17.

    Article  CAS  PubMed  Google Scholar 

  168. Zhang L, et al. Apoptotic and autophagic pathways with relevant small-molecule compounds, in cancer stem cells. Cell Prolif. 2015;48(4):385–97.

    Article  PubMed  PubMed Central  Google Scholar 

  169. Wen Y, et al. EZH2 activates CHK1 signaling to promote ovarian cancer chemoresistance by maintaining the properties of cancer stem cells. Theranostics. 2021;11(4):1795–813.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Kelly PN, Strasser A. The role of Bcl-2 and its pro-survival relatives in tumourigenesis and cancer therapy. Cell Death Differ. 2011;18(9):1414–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Vera MB, et al. Noxa and Mcl-1 expression influence the sensitivity to BH3-mimetics that target Bcl-xL in patient-derived glioma stem cells. Sci Rep. 2022;12(1):17729.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Lu Y et al. Inhibition of Bcl-2 and Bcl-xL overcomes the resistance to the third-generation EGFR tyrosine kinase inhibitor osimertinib in non-small cell lung cancer. Mol Med Rep. 2021;23(1).

  173. Piao LS, et al. CD133 + liver cancer stem cells modulate radioresistance in human hepatocellular carcinoma. Cancer Lett. 2012;315(2):129–37.

    Article  CAS  PubMed  Google Scholar 

  174. Twomey JD, Kim SR, Zhao L, Bozza WP, Zhang B. Spatial dynamics of TRAIL death receptors in cancer cells. Drug Resist Updat. 2015;19:13–21.

    Article  PubMed  Google Scholar 

  175. Ashkenazi A. Targeting death and decoy receptors of the tumour-necrosis factor superfamily. Nat Rev Cancer. 2002;2(6):420–30.

    Article  CAS  PubMed  Google Scholar 

  176. Piggott L, et al. Suppression of apoptosis inhibitor c-FLIP selectively eliminates breast cancer stem cell activity in response to the anti-cancer agent, TRAIL. Breast Cancer Res. 2011;13(5):R88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Deng D, Shah K. TRAIL of hope meeting resistance in Cancer. Trends Cancer. 2020;6(12):989–1001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Diehn M, et al. Association of reactive oxygen species levels and radioresistance in cancer stem cells. Nature. 2009;458(7239):780–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Guo F, et al. Inhibitory effect on ovarian cancer ALDH + stem-like cells by Disulfiram and copper treatment through ALDH and ROS modulation. Biomed Pharmacother. 2019;118:109371.

    Article  CAS  PubMed  Google Scholar 

  180. Kahm YJ, Kim RK. BIRC5: A novel therapeutic target for lung cancer stem cells and glioma stem cells. Biochem Biophys Res Commun. 2023;682:141–7.

    Article  CAS  PubMed  Google Scholar 

  181. Khan SU, Fatima K, Malik F. Understanding the cell survival mechanism of anoikis-resistant cancer cells during different steps of metastasis. Clin Exp Metastasis. 2022;39(5):715–26.

    Article  PubMed  Google Scholar 

  182. An H, et al. Salinomycin promotes Anoikis and decreases the CD44+/CD24- Stem-Like population via Inhibition of STAT3 activation in MDA-MB-231 cells. PLoS ONE. 2015;10(11):e0141919.

    Article  PubMed  PubMed Central  Google Scholar 

  183. Wu Y, et al. Stabilization of snail by NF-kappaB is required for inflammation-induced cell migration and invasion. Cancer Cell. 2009;15(5):416–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Huaman J, Ogunwobi OO. Circulating tumor cell migration requires fibronectin acting through integrin B1 or SLUG. Cells. 2020;9(7).

  185. Han Y, et al. Role of autophagy in breast cancer and breast cancer stem cells (Review). Int J Oncol. 2018;52(4):1057–70.

    CAS  PubMed  Google Scholar 

  186. Khan SU, et al. Autophagy plays a dual role in drug resistance. In: Khan SU, Malik F, editors. Drug resistance in cancer: mechanisms and strategies. Singapore: Springer Nature Singapore; 2024. pp. 243–67.

    Chapter  Google Scholar 

  187. Qadir MA, et al. Macroautophagy Inhibition sensitizes tamoxifen-resistant breast cancer cells and enhances mitochondrial depolarization. Breast Cancer Res Treat. 2008;112(3):389–403.

    Article  CAS  PubMed  Google Scholar 

  188. Li L, et al. Protective autophagy decreases osimertinib cytotoxicity through regulation of stem cell-like properties in lung cancer. Cancer Lett. 2019;452:191–202.

    Article  CAS  PubMed  Google Scholar 

  189. Nguyen HG, et al. Targeting autophagy overcomes enzalutamide resistance in castration-resistant prostate cancer cells and improves therapeutic response in a xenograft model. Oncogene. 2014;33(36):4521–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Yue W, et al. Inhibition of the autophagic flux by salinomycin in breast cancer stem-like/progenitor cells interferes with their maintenance. Autophagy. 2013;9(5):714–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Fitzwalter BE, et al. Autophagy Inhibition mediates apoptosis sensitization in Cancer therapy by relieving FOXO3a turnover. Dev Cell. 2018;44(5):555–e5653.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Anand SK, Sharma A, Singh N, Kakkar P. Entrenching role of cell cycle checkpoints and autophagy for maintenance of genomic integrity. DNA Repair (Amst). 2020;86:102748.

    Article  CAS  PubMed  Google Scholar 

  193. Li LQ, et al. Autophagy regulates chemoresistance of gastric cancer stem cells via the Notch signaling pathway. Eur Rev Med Pharmacol Sci. 2018;22(11):3402–7.

    PubMed  Google Scholar 

  194. Lomonaco SL, et al. The induction of autophagy by gamma-radiation contributes to the radioresistance of glioma stem cells. Int J Cancer. 2009;125(3):717–22.

    Article  CAS  PubMed  Google Scholar 

  195. O’Donnell JS, Teng MWL, Smyth MJ. Cancer immunoediting and resistance to T cell-based immunotherapy. Nat Rev Clin Oncol. 2019;16(3):151–67.

    Article  PubMed  Google Scholar 

  196. Bruttel VS, Wischhusen J. Cancer stem cell immunology: key to Understanding tumorigenesis and tumor immune escape? Front Immunol. 2014;5:360.

    Article  PubMed  PubMed Central  Google Scholar 

  197. Schatton T, Frank MH. Antitumor immunity and cancer stem cells. Ann N Y Acad Sci. 2009;1176:154–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Lee TK, et al. Blockade of CD47-mediated cathepsin S/protease-activated receptor 2 signaling provides a therapeutic target for hepatocellular carcinoma. Hepatology. 2014;60(1):179–91.

    Article  CAS  PubMed  Google Scholar 

  199. Cioffi M, et al. Inhibition of CD47 effectively targets pancreatic Cancer stem cells via dual mechanisms. Clin Cancer Res. 2015;21(10):2325–37.

    Article  CAS  PubMed  Google Scholar 

  200. Majeti R, et al. CD47 is an adverse prognostic factor and therapeutic antibody target on human acute myeloid leukemia stem cells. Cell. 2009;138(2):286–99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Schatton T, et al. Modulation of T-cell activation by malignant melanoma initiating cells. Cancer Res. 2010;70(2):697–708.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Di Tomaso T, et al. Immunobiological characterization of cancer stem cells isolated from glioblastoma patients. Clin Cancer Res. 2010;16(3):800–13.

    Article  PubMed  PubMed Central  Google Scholar 

  203. Sultan M, et al. Epigenetic Silencing of TAP1 in Aldefluor(+) breast Cancer stem cells contributes to their enhanced immune evasion. Stem Cells. 2018;36(5):641–54.

    Article  CAS  PubMed  Google Scholar 

  204. Khan SU, et al. Unveiling the mechanisms and challenges of cancer drug resistance. Cell Commun Signal. 2024;22(1):109.

    Article  PubMed  PubMed Central  Google Scholar 

  205. Najafi M, Mortezaee K, Majidpoor J. Cancer stem cell (CSC) resistance drivers. Life Sci. 2019;234:116781.

    Article  CAS  PubMed  Google Scholar 

  206. Guo W, Qiao T, Li T. The role of stem cells in small-cell lung cancer: evidence from chemoresistance to immunotherapy. Semin Cancer Biol. 2022;87:160–9.

    Article  CAS  PubMed  Google Scholar 

  207. Alvarado AG, et al. Glioblastoma Cancer stem cells evade innate immune suppression of Self-Renewal through reduced TLR4 expression. Cell Stem Cell. 2017;20(4):450–e4614.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Marotta LL, et al. The JAK2/STAT3 signaling pathway is required for growth of CD44⁺CD24⁻ stem cell-like breast cancer cells in human tumors. J Clin Invest. 2011;121(7):2723–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Reim F, et al. Immunoselection of breast and ovarian cancer cells with trastuzumab and natural killer cells: selective escape of CD44high/CD24low/HER2low breast cancer stem cells. Cancer Res. 2009;69(20):8058–66.

    Article  CAS  PubMed  Google Scholar 

  210. Fatima K, et al. Role of Cancer stem cells in drug resistance. In: Khan SU, Malik F, editors. Drug resistance in cancer: mechanisms and strategies. Singapore: Springer Nature Singapore; 2024. pp. 77–120.

    Chapter  Google Scholar 

  211. Wang C, et al. CD276 expression enables squamous cell carcinoma stem cells to evade immune surveillance. Cell Stem Cell. 2021;28(9):1597–e16137.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  212. Hsu JM, et al. STT3-dependent PD-L1 accumulation on cancer stem cells promotes immune evasion. Nat Commun. 2018;9(1):1908.

    Article  PubMed  PubMed Central  Google Scholar 

  213. Kawasaki BT, Farrar WL. Cancer stem cells, CD200 and immunoevasion. Trends Immunol. 2008;29(10):464–8.

    Article  CAS  PubMed  Google Scholar 

  214. Agudo J, et al. Quiescent tissue stem cells evade immune surveillance. Immunity. 2018;48(2):271–e2855.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Malladi S, et al. Metastatic latency and immune evasion through autocrine Inhibition of WNT. Cell. 2016;165(1):45–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Saudemont A, Quesnel B. In a model of tumor dormancy, long-term persistent leukemic cells have increased B7-H1 and B7.1 expression and resist CTL-mediated Lysis. Blood. 2004;104(7):2124–33.

    Article  CAS  PubMed  Google Scholar 

  217. Saudemont A, et al. Dormant tumor cells develop cross-resistance to apoptosis induced by CTLs or Imatinib mesylate via methylation of suppressor of cytokine signaling 1. Cancer Res. 2007;67(9):4491–8.

    Article  CAS  PubMed  Google Scholar 

  218. Elgendy SM, Alyammahi SK, Alhamad DW, Abdin SM, Omar HA. Ferroptosis: an emerging approach for targeting cancer stem cells and drug resistance. Crit Rev Oncol Hematol. 2020;155:103095.

    Article  PubMed  Google Scholar 

  219. Rim EY, Clevers H, Nusse R. The Wnt pathway: from signaling mechanisms to synthetic modulators. Annu Rev Biochem. 2022;91:571–98.

    Article  CAS  PubMed  Google Scholar 

  220. Dong HJ, et al. The Wnt/β-catenin signaling/Id2 cascade mediates the effects of hypoxia on the hierarchy of colorectal-cancer stem cells. Sci Rep. 2016;6:22966.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Xu W, et al. Hypoxia activates Wnt/β-catenin signaling by regulating the expression of BCL9 in human hepatocellular carcinoma. Sci Rep. 2017;7:40446.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Jiang N, et al. HIF-1ɑ-regulated miR-1275 maintains stem cell-like phenotypes and promotes the progression of LUAD by simultaneously activating Wnt/β-catenin and Notch signaling. Theranostics. 2020;10(6):2553–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  223. Leung HW, et al. EPHB2 activates β-Catenin to enhance Cancer stem cell properties and drive Sorafenib resistance in hepatocellular carcinoma. Cancer Res. 2021;81(12):3229–40.

    Article  CAS  PubMed  Google Scholar 

  224. Chau WK, Ip CK, Mak AS, Lai HC, Wong A. S. c-Kit mediates chemoresistance and tumor-initiating capacity of ovarian cancer cells through activation of Wnt/β-catenin-ATP-binding cassette G2 signaling. Oncogene. 2013;32(22):2767–81.

    Article  CAS  PubMed  Google Scholar 

  225. Liu YS, Hsu HC, Tseng KC, Chen HC, Chen S. J. Lgr5 promotes cancer stemness and confers chemoresistance through ABCB1 in colorectal cancer. Biomed Pharmacother. 2013;67(8):791–9.

    Article  CAS  PubMed  Google Scholar 

  226. Yan Y, et al. HIF-2α promotes conversion to a stem cell phenotype and induces chemoresistance in breast cancer cells by activating Wnt and Notch pathways. J Exp Clin Cancer Res. 2018;37(1):256.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Huang R, et al. Colorectal cancer stem cell and chemoresistant colorectal cancer cell phenotypes and increased sensitivity to Notch pathway inhibitor. Mol Med Rep. 2015;12(2):2417–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  228. Ulasov IV, Nandi S, Dey M, Sonabend AM, Lesniak MS. Inhibition of Sonic Hedgehog and Notch pathways enhances sensitivity of CD133(+) glioma stem cells to Temozolomide therapy. Mol Med 2011 Jan-Feb;17(1–2):103–12.

  229. McAuliffe SM, et al. Targeting Notch, a key pathway for ovarian cancer stem cells, sensitizes tumors to platinum therapy. Proc Natl Acad Sci U S A. 2012;109(43):E2939–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  230. Liu YP, et al. Cisplatin selects for multidrug-resistant CD133 + cells in lung adenocarcinoma by activating Notch signaling. Cancer Res. 2013;73(1):406–16.

    Article  CAS  PubMed  Google Scholar 

  231. Cochrane CR, Szczepny A, Watkins DN, Cain JE. Hedgehog signaling in the maintenance of Cancer stem cells. Cancers (Basel). 2015;7(3):1554–85.

    Article  CAS  PubMed  Google Scholar 

  232. Clement V, Sanchez P, de Tribolet N, Radovanovic I, Ruiz i Altaba A. HEDGEHOG-GLI1 signaling regulates human glioma growth, cancer stem cell self-renewal, and tumorigenicity. Curr Biol. 2007;17(2):165–72.

    Article  CAS  PubMed  Google Scholar 

  233. Wang S, et al. Hedgehog signaling promotes Sorafenib resistance in hepatocellular carcinoma patient-derived organoids. J Exp Clin Cancer Res. 2020;39(1):22.

    Article  PubMed  PubMed Central  Google Scholar 

  234. Usui T et al. Hedgehog signals mediate Anti-Cancer drug resistance in Three-Dimensional primary colorectal Cancer organoid culture. Int J Mol Sci. 2018;19(4).

  235. Wang J, et al. MACC1 facilitates chemoresistance and cancer stem cell–like properties of colon cancer cells through the PI3K/AKT signaling pathway. Mol Med Rep. 2017;16(6):8747–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  236. Bleau AM, et al. PTEN/PI3K/Akt pathway regulates the side population phenotype and ABCG2 activity in glioma tumor stem-like cells. Cell Stem Cell. 2009;4(3):226–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  237. Almozyan S, et al. PD-L1 promotes OCT4 and Nanog expression in breast cancer stem cells by sustaining PI3K/AKT pathway activation. Int J Cancer. 2017;141(7):1402–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  238. Song S, et al. The Hippo coactivator YAP1 mediates EGFR overexpression and confers chemoresistance in esophageal Cancer. Clin Cancer Res. 2015;21(11):2580–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  239. Gao R, et al. YAP/TAZ and ATF4 drive resistance to Sorafenib in hepatocellular carcinoma by preventing ferroptosis. EMBO Mol Med. 2021;13(12):e14351.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  240. Mudianto T, et al. Yap1 mediates Trametinib resistance in head and neck squamous cell carcinomas. Clin Cancer Res. 2021;27(8):2326–39.

    Article  CAS  PubMed  Google Scholar 

  241. Alsayed R, et al. Epigenetic regulation of CXCR4 signaling in cancer pathogenesis and progression. Semin Cancer Biol. 2022;86(Pt 2):697–708.

    Article  CAS  PubMed  Google Scholar 

  242. Bianchi ME, Mezzapelle R. The chemokine receptor CXCR4 in cell proliferation and tissue regeneration. Front Immunol. 2020;11:2109.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  243. Carmon KS, Lin Q, Gong X, Thomas A, Liu Q. LGR5 interacts and cointernalizes with Wnt receptors to modulate Wnt/β-catenin signaling. Mol Cell Biol. 2012;32(11):2054–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  244. Zhao K, et al. From mechanism to therapy: the journey of CD24 in cancer. Front Immunol. 2024;15:1401528.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  245. Yang Y, Zhu G, Yang L, Yang Y. Targeting CD24 as a novel immunotherapy for solid cancers. Cell Commun Signal. 2023;21(1):312.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  246. Chen C, Zhao S, Karnad A, Freeman J. W. The biology and role of CD44 in cancer progression: therapeutic implications. J Hematol Oncol. 2018;11(1):64.

    Article  PubMed  PubMed Central  Google Scholar 

  247. Young RA. Control of the embryonic stem cell state. Cell. 2011;144(6):940–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  248. Novak D, et al. SOX2 in development and cancer biology. Semin Cancer Biol. 2020;67(Pt 1):74–82.

    Article  CAS  PubMed  Google Scholar 

  249. Villodre ES, Kipper FC, Pereira MB, Lenz G. Roles of OCT4 in tumorigenesis, cancer therapy resistance and prognosis. Cancer Treat Rev. 2016;51:1–9.

    Article  CAS  PubMed  Google Scholar 

  250. Najafzadeh B, et al. The oncogenic potential of NANOG: an important cancer induction mediator. J Cell Physiol. 2021;236(4):2443–58.

    Article  CAS  PubMed  Google Scholar 

  251. Beckers C, Pruschy M, Vetrugno I. Tumor hypoxia and radiotherapy: A major driver of resistance even for novel radiotherapy modalities. Semin Cancer Biol. 2024;98:19–30.

    Article  CAS  PubMed  Google Scholar 

  252. Salaroglio IC, et al. Mitochondrial ROS drive resistance to chemotherapy and immune-killing in hypoxic non-small cell lung cancer. J Exp Clin Cancer Res. 2022;41(1):243.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  253. He X, et al. Hypoxia regulates ABCG2 activity through the activivation of ERK1/2/HIF-1α and contributes to chemoresistance in pancreatic cancer cells. Cancer Biol Ther. 2016;17(2):188–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  254. Enari M, et al. A caspase-activated DNase that degrades DNA during apoptosis, and its inhibitor ICAD. Nature. 1998;391(6662):43–50.

    Article  CAS  PubMed  Google Scholar 

  255. Lu Y, et al. Hypoxia induces resistance to EGFR inhibitors in lung Cancer cells via upregulation of FGFR1 and the MAPK pathway. Cancer Res. 2020;80(21):4655–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  256. Chen JK, et al. Hypoxia-induced BAX overexpression and radiation killing of hypoxic glioblastoma cells. Radiat Res. 2005;163(6):644–53.

    Article  CAS  PubMed  Google Scholar 

  257. Ritter V, Krautter F, Klein D, Jendrossek V, Rudner J. Bcl-2/Bcl-xL inhibitor ABT-263 overcomes hypoxia-driven radioresistence and improves radiotherapy. Cell Death Dis. 2021;12(7):694.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  258. Bashari MH, et al. Mcl-1 confers protection of Her2-positive breast cancer cells to hypoxia: therapeutic implications. Breast Cancer Res. 2016;18(1):26.

    Article  PubMed  PubMed Central  Google Scholar 

  259. Zhang X, Qi Z, Yin H, Yang G. Interaction between p53 and Ras signaling controls cisplatin resistance via HDAC4- and HIF-1α-mediated regulation of apoptosis and autophagy. Theranostics. 2019;9(4):1096–114.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  260. Groelly FJ, Fawkes M, Dagg RA, Blackford A, N, Tarsounas M. Targeting DNA damage response pathways in cancer. Nat Rev Cancer. 2023;23(2):78–94.

    Article  CAS  PubMed  Google Scholar 

  261. Hayashi MT, Karlseder J. DNA damage associated with mitosis and cytokinesis failure. Oncogene. 2013;32(39):4593–601.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  262. Li J, et al. DNA damage response in breast cancer and its significant role in guiding novel precise therapies. Biomark Res. 2024;12(1):111.

    Article  PubMed  PubMed Central  Google Scholar 

  263. Hammond EM, Denko NC, Dorie MJ, Abraham RT, Giaccia A. J. Hypoxia links ATR and p53 through replication arrest. Mol Cell Biol. 2002;22(6):1834–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  264. Wirthner R, Wrann S, Balamurugan K, Wenger RH, Stiehl D. P. Impaired DNA double-strand break repair contributes to chemoresistance in HIF-1 alpha-deficient mouse embryonic fibroblasts. Carcinogenesis. 2008;29(12):2306–16.

    Article  CAS  PubMed  Google Scholar 

  265. Yan Y, et al. A novel HIF-2α targeted inhibitor suppresses hypoxia-induced breast cancer stemness via SOD2-mtROS-PDI/GPR78-UPR(ER) axis. Cell Death Differ. 2022;29(9):1769–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  266. Xiang L, et al. HIF-1-dependent Heme synthesis promotes gemcitabine resistance in human non-small cell lung cancers via enhanced ABCB6 expression. Cell Mol Life Sci. 2022;79(6):343.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  267. Kim J, et al. SALL4 promotes Glycolysis and chromatin remodeling via modulating HP1α-Glut1 pathway. Oncogene. 2017;36(46):6472–9.

    Article  CAS  PubMed  Google Scholar 

  268. Wei C, et al. Development of GLUT4-selective antagonists for multiple myeloma therapy. Eur J Med Chem. 2017;139:573–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  269. Shen Z, et al. GLUT5-KHK axis-mediated Fructose metabolism drives proliferation and chemotherapy resistance of colorectal cancer. Cancer Lett. 2022;534:215617.

    Article  CAS  PubMed  Google Scholar 

  270. Lee HJ. Non-proteolytic ubiquitination of hexokinase 2 by HectH9 controls tumor metabolism and cancer stem cell expansion. Nat Commun. 2019;10(1):2625.

    Article  PubMed  PubMed Central  Google Scholar 

  271. Gustafsson NMS, et al. Targeting PFKFB3 radiosensitizes cancer cells and suppresses homologous recombination. Nat Commun. 2018;9(1):3872.

    Article  PubMed  PubMed Central  Google Scholar 

  272. Li Q, et al. Rac1 activates non-oxidative Pentose phosphate pathway to induce chemoresistance of breast cancer. Nat Commun. 2020;11(1):1456.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  273. Zhou JW, Tang JJ, Sun W, Wang H. PGK1 facilities cisplatin chemoresistance by triggering HSP90/ERK pathway mediated DNA repair and methylation in endometrial endometrioid adenocarcinoma. Mol Med. 2019;25(1):11.

    Article  PubMed  PubMed Central  Google Scholar 

  274. Qian X, et al. Enolase 1 stimulates Glycolysis to promote chemoresistance in gastric cancer. Oncotarget. 2017;8(29):47691–708.

    Article  PubMed  PubMed Central  Google Scholar 

  275. Shang D, et al. Metformin increases sensitivity of osteosarcoma stem cells to cisplatin by inhibiting expression of PKM2. Int J Oncol. 2017;50(5):1848–56.

    Article  CAS  PubMed  Google Scholar 

  276. Das CK, Parekh A, Parida PK, Bhutia SK, Mandal M. Lactate dehydrogenase A regulates autophagy and Tamoxifen resistance in breast cancer. Biochim Biophys Acta Mol Cell Res. 2019;1866(6):1004–18.

    Article  CAS  PubMed  Google Scholar 

  277. Chakraborty PK, et al. MICU1 drives Glycolysis and chemoresistance in ovarian cancer. Nat Commun. 2017;8:14634.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  278. Shen YC, et al. Activating oxidative phosphorylation by a pyruvate dehydrogenase kinase inhibitor overcomes Sorafenib resistance of hepatocellular carcinoma. Br J Cancer. 2013;108(1):72–81.

    Article  CAS  PubMed  Google Scholar 

  279. Li FL, et al. Acetylation accumulates PFKFB3 in cytoplasm to promote Glycolysis and protects cells from cisplatin-induced apoptosis. Nat Commun. 2018;9(1):508.

    Article  PubMed  PubMed Central  Google Scholar 

  280. Liu X, et al. Elevated hexokinase II expression confers acquired resistance to 4-Hydroxytamoxifen in breast Cancer cells. Mol Cell Proteom. 2019;18(11):2273–84.

    Article  CAS  Google Scholar 

  281. Wartenberg M, et al. Glycolytic pyruvate regulates P-Glycoprotein expression in multicellular tumor spheroids via modulation of the intracellular redox state. J Cell Biochem. 2010;109(2):434–46.

    Article  CAS  PubMed  Google Scholar 

  282. Lee SY, et al. Oncogenic metabolism acts as a prerequisite step for induction of Cancer metastasis and Cancer stem cell phenotype. Oxid Med Cell Longev. 2018;2018:1027453.

    Article  PubMed  PubMed Central  Google Scholar 

  283. Farmer P, et al. A stroma-related gene signature predicts resistance to neoadjuvant chemotherapy in breast cancer. Nat Med. 2009;15(1):68–74.

    Article  CAS  PubMed  Google Scholar 

  284. Deng JJ, et al. Twist mediates an aggressive phenotype in human colorectal cancer cells. Int J Oncol. 2016;48(3):1117–24.

    Article  CAS  PubMed  Google Scholar 

  285. Saxena M, Stephens MA, Pathak H, Rangarajan A. Transcription factors that mediate epithelial-mesenchymal transition lead to multidrug resistance by upregulating ABC transporters. Cell Death Dis. 2011;2(7):e179.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  286. Tomono T, Yano K, Ogihara T. Snail-Induced Epithelial-to-Mesenchymal transition enhances P-gp-Mediated multidrug resistance in HCC827 cells. J Pharm Sci. 2017;106(9):2642–9.

    Article  CAS  PubMed  Google Scholar 

  287. Khan SU, et al. Cancer metastasis: molecular mechanisms and clinical perspectives. Pharmacol Ther. 2023;250:108522.

    Article  CAS  PubMed  Google Scholar 

  288. Vega S, et al. Snail blocks the cell cycle and confers resistance to cell death. Genes Dev. 2004;18(10):1131–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  289. Kciuk M, Gielecińska A, Kołat D, Kałuzińska Ż, Kontek R. Cancer-associated transcription factors in DNA damage response. Biochim Biophys Acta Rev Cancer. 2022;1877(4):188757.

    Article  CAS  PubMed  Google Scholar 

  290. Zhang P, et al. ATM-mediated stabilization of ZEB1 promotes DNA damage response and radioresistance through CHK1. Nat Cell Biol. 2014;16(9):864–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  291. Xing Y, et al. Autophagy Inhibition mediated by MCOLN1/TRPML1 suppresses cancer metastasis via regulating a ROS-driven TP53/p53 pathway. Autophagy. 2022;18(8):1932–54.

    Article  CAS  PubMed  Google Scholar 

  292. Lee SY, et al. Wnt/Snail signaling regulates cytochrome C oxidase and glucose metabolism. Cancer Res. 2012;72(14):3607–17.

    Article  CAS  PubMed  Google Scholar 

  293. Liu M et al. Snail-Overexpression induces Epithelial-mesenchymal transition and metabolic reprogramming in human pancreatic ductal adenocarcinoma and Non-tumorigenic ductal cells. J Clin Med. 2019;8(6).

  294. Vera-Ramirez L, Vodnala S, K, Nini R, Hunter KW, Green J. E. Autophagy promotes the survival of dormant breast cancer cells and metastatic tumour recurrence. Nat Commun. 2018;9(1):1944.

    Article  PubMed  PubMed Central  Google Scholar 

  295. Kuczek DE, et al. Collagen density regulates the activity of tumor-infiltrating T cells. J Immunother Cancer. 2019;7(1):68.

    Article  PubMed  PubMed Central  Google Scholar 

  296. Nakagawa Y, et al. Effects of extracellular pH and hypoxia on the function and development of antigen-specific cytotoxic T lymphocytes. Immunol Lett. 2015;167(2):72–86.

    Article  CAS  PubMed  Google Scholar 

  297. Fischer K, et al. Inhibitory effect of tumor cell-derived lactic acid on human T cells. Blood. 2007;109(9):3812–9.

    Article  CAS  PubMed  Google Scholar 

  298. Mendler AN, et al. Tumor lactic acidosis suppresses CTL function by Inhibition of p38 and JNK/c-Jun activation. Int J Cancer. 2012;131(3):633–40.

    Article  CAS  PubMed  Google Scholar 

  299. Sangsuwan R, et al. Lactate exposure promotes immunosuppressive phenotypes in innate immune cells. Cell Mol Bioeng. 2020;13(5):541–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  300. Calcinotto A, et al. Modulation of microenvironment acidity reverses anergy in human and murine tumor-infiltrating T lymphocytes. Cancer Res. 2012;72(11):2746–56.

    Article  CAS  PubMed  Google Scholar 

  301. Scharping NE, et al. Mitochondrial stress induced by continuous stimulation under hypoxia rapidly drives T cell exhaustion. Nat Immunol. 2021;22(2):205–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  302. Balsamo M, et al. Hypoxia downregulates the expression of activating receptors involved in NK-cell-mediated target cell killing without affecting ADCC. Eur J Immunol. 2013;43(10):2756–64.

    Article  CAS  PubMed  Google Scholar 

  303. Noman MZ, et al. PD-L1 is a novel direct target of HIF-1α, and its Blockade under hypoxia enhanced MDSC-mediated T cell activation. J Exp Med. 2014;211(5):781–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  304. Raggi F, et al. Regulation of human macrophage M1-M2 polarization balance by hypoxia and the triggering receptor expressed on myeloid Cells-1. Front Immunol. 2017;8:1097.

    Article  PubMed  PubMed Central  Google Scholar 

  305. Angelin A, et al. Foxp3 reprograms T cell metabolism to function in Low-Glucose, High-Lactate environments. Cell Metab. 2017;25(6):1282–e12937.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  306. Deaglio S, et al. Adenosine generation catalyzed by CD39 and CD73 expressed on regulatory T cells mediates immune suppression. J Exp Med. 2007;204(6):1257–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  307. Vignali PDA, et al. Hypoxia drives CD39-dependent suppressor function in exhausted T cells to limit antitumor immunity. Nat Immunol. 2023;24(2):267–79.

    Article  CAS  PubMed  Google Scholar 

  308. Synnestvedt K, et al. Ecto-5’-nucleotidase (CD73) regulation by hypoxia-inducible factor-1 mediates permeability changes in intestinal epithelia. J Clin Invest. 2002;110(7):993–1002.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  309. Dang EV, et al. Control of T(H)17/T(reg) balance by hypoxia-inducible factor 1. Cell. 2011;146(5):772–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  310. Jiang Y, Zhan H. Communication between EMT and PD-L1 signaling: new insights into tumor immune evasion. Cancer Lett. 2020;468:72–81.

    Article  PubMed  Google Scholar 

  311. Clark CA, et al. Tumor-Intrinsic PD-L1 signals regulate cell growth, pathogenesis, and autophagy in ovarian Cancer and melanoma. Cancer Res. 2016;76(23):6964–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  312. Antonangeli F, et al. Regulation of PD-L1 expression by NF-κB in Cancer. Front Immunol. 2020;11:584626.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  313. Kopecka J, et al. Hypoxia as a driver of resistance to immunotherapy. Drug Resist Updat. 2021;59:100787.

    Article  CAS  PubMed  Google Scholar 

  314. Zandberg DP et al. Tumor hypoxia is associated with resistance to PD-1 Blockade in squamous cell carcinoma of the head and neck. J Immunother Cancer. 2021;9(5).

  315. Scharping NE, Menk AV, Whetstone RD, Zeng X, Delgoffe GM. Efficacy of PD-1 Blockade is potentiated by Metformin-Induced reduction of tumor hypoxia. Cancer Immunol Res. 2017;5(1):9–16.

    Article  CAS  PubMed  Google Scholar 

  316. Jirtle RL. Chemical modification of tumour blood flow. Int J Hyperth. 1988 Jul-Aug;4(4):355–71.

  317. Wu J, et al. Oxygen microcapsules improve immune checkpoint Blockade by ameliorating hypoxia condition in pancreatic ductal adenocarcinoma. Bioact Mater. 2023;20:259–70.

    CAS  PubMed  Google Scholar 

  318. Li J, et al. Advanced nanomaterials targeting hypoxia to enhance radiotherapy. Int J Nanomed. 2018;13:5925–36.

    Article  CAS  Google Scholar 

  319. Choi JH et al. The Small-Molecule Wnt inhibitor ICG-001 efficiently inhibits colorectal Cancer stemness and metastasis by suppressing MEIS1 expression. Int J Mol Sci. 2021;22(24).

  320. Hayashi T, et al. Not all NOTCH is created equal: the oncogenic role of NOTCH2 in bladder Cancer and its implications for targeted therapy. Clin Cancer Res. 2016;22(12):2981–92.

    Article  CAS  PubMed  Google Scholar 

  321. Rodova M, Fu J, Watkins DN, Srivastava R, K, Shankar S. Sonic Hedgehog signaling Inhibition provides opportunities for targeted therapy by Sulforaphane in regulating pancreatic cancer stem cell self-renewal. PLoS ONE. 2012;7(9):e46083.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  322. Hongwiangchan N et al. Hydroquinone 5-O-Cinnamoyl ester of Renieramycin M suppresses lung Cancer stem cells by targeting Akt and destabilizes c-Myc. Pharmaceuticals (Basel). 2021;14(11).

  323. Chung SS, Vadgama JV. Curcumin and Epigallocatechin gallate inhibit the cancer stem cell phenotype via down-regulation of STAT3-NFκB signaling. Anticancer Res. 2015;35(1):39–46.

    CAS  PubMed  PubMed Central  Google Scholar 

  324. Kim HM, et al. Increased CD13 expression reduces reactive oxygen species, promoting survival of liver cancer stem cells via an epithelial-mesenchymal transition-like phenomenon. Ann Surg Oncol. 2012;19(Suppl 3):S539–48.

    Article  PubMed  Google Scholar 

  325. Han Y, Sun B, Cai H, Xuan Y. Simultaneously target of normal and stem cells-like gastric cancer cells via cisplatin and anti-CD133 CAR-T combination therapy. Cancer Immunol Immunother. 2021;70(10):2795–803.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  326. Hu C, et al. Lentivirus-mediated ShRNA targeting Nanog inhibits cell proliferation and attenuates cancer stem cell activities in breast cancer. J Drug Target. 2016;24(5):422–32.

    Article  CAS  PubMed  Google Scholar 

  327. Feng X, Li L, Wang L, Luo S, Bai X. Chromatin target of protein arginine methyltransferase regulates invasion, chemoresistance, and stemness in epithelial ovarian cancer. Biosci Rep. 2019;39(4).

  328. Jan S, et al. Introduction to drug resistance in Cancer. In: Khan SU, Malik F, editors. Drug resistance in cancer: mechanisms and strategies. Singapore: Springer Nature Singapore; 2024. pp. 1–30.

    Google Scholar 

  329. Gisina A, Kim Y, Yarygin K, Lupatov A. Can CD133 be regarded as a prognostic biomarker in oncology: pros and cons. Int J Mol Sci. 2023;24(24).

  330. Erisik D, et al. Differences and similarities between colorectal Cancer cells and colorectal Cancer stem cells: molecular insights and implications. ACS Omega. 2023;8(33):30145–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  331. Niu X, et al. Autophagy in cancer development, immune evasion, and drug resistance. Drug Resist Updat. 2025;78:101170.

    Article  CAS  PubMed  Google Scholar 

  332. Lou D, Zhu L, Ding H, Dai HY, Zou GM. Aberrant expression of redox protein Ape1 in colon cancer stem cells. Oncol Lett. 2014;7(4):1078–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  333. Karimi-Busheri F, Rasouli-Nia A, Mackey JR, Weinfeld M. Senescence evasion by MCF-7 human breast tumor-initiating cells. Breast Cancer Res. 2010;12(3):R31.

    Article  PubMed  PubMed Central  Google Scholar 

  334. Anuja K, et al. Radiation-induced DNA damage response and resistance in colorectal cancer stem-like cells. Int J Radiat Biol. 2019;95(6):667–79.

    Article  CAS  PubMed  Google Scholar 

  335. Zhao Y, et al. Stem cell-like side populations in esophageal cancer: a source of chemotherapy resistance and metastases. Stem Cells Dev. 2014;23(2):180–92.

    Article  CAS  PubMed  Google Scholar 

  336. Pudelko L, et al. Glioblastoma and glioblastoma stem cells are dependent on functional MTH1. Oncotarget. 2017;8(49):84671–84.

    Article  PubMed  PubMed Central  Google Scholar 

  337. Hasegawa D, et al. G-quadruplex ligand-induced DNA damage response coupled with telomere dysfunction and replication stress in glioma stem cells. Biochem Biophys Res Commun. 2016;471(1):75–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  338. Somasagara RR, et al. RAD6 promotes DNA repair and stem cell signaling in ovarian cancer and is a promising therapeutic target to prevent and treat acquired chemoresistance. Oncogene. 2017;36(48):6680–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  339. Ong DST, et al. PAF promotes stemness and radioresistance of glioma stem cells. Proc Natl Acad Sci U S A. 2017;114(43):E9086–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  340. Wang X, et al. PAF-Wnt signaling-induced cell plasticity is required for maintenance of breast cancer cell stemness. Nat Commun. 2016;7:10633.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  341. Yan X, et al. A positive feedback loop: RAD18-YAP-TGF-β between triple-negative breast cancer and macrophages regulates cancer stemness and progression. Cell Death Discov. 2022;8(1):196.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  342. Wang X, et al. Bmi-1 regulates stem cell-like properties of gastric cancer cells via modulating MiRNAs. J Hematol Oncol. 2016;9(1):90.

    Article  PubMed  PubMed Central  Google Scholar 

  343. Mei Y, Liu YB, Cao S, Tian ZW, Zhou HH. RIF1 promotes tumor growth and cancer stem cell-like traits in NSCLC by protein phosphatase 1-mediated activation of Wnt/β-catenin signaling. Cell Death Dis. 2018;9(10):942.

    Article  PubMed  PubMed Central  Google Scholar 

  344. Apostolou P, et al. Potential role for the Metnase transposase fusion gene in colon cancer through the regulation of key genes. PLoS ONE. 2014;9(10):e109741.

    Article  PubMed  PubMed Central  Google Scholar 

  345. Dussaussois-Montagne A, et al. SETMAR isoforms in glioblastoma: A matter of protein stability. Oncotarget. 2017;8(6):9835–48.

    Article  PubMed  Google Scholar 

  346. Manic G, et al. Control of replication stress and mitosis in colorectal cancer stem cells through the interplay of PARP1, MRE11 and RAD51. Cell Death Differ. 2021;28(7):2060–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  347. Venere M, et al. Therapeutic targeting of constitutive PARP activation compromises stem cell phenotype and survival of glioblastoma-initiating cells. Cell Death Differ. 2014;21(2):258–69.

    Article  CAS  PubMed  Google Scholar 

  348. Liu Y, et al. RAD51 mediates resistance of Cancer stem cells to PARP Inhibition in Triple-Negative breast Cancer. Clin Cancer Res. 2017;23(2):514–22.

    Article  CAS  PubMed  Google Scholar 

  349. Kim H, Lin Q, Yun Z. BRCA1 regulates the cancer stem cell fate of breast cancer cells in the context of hypoxia and histone deacetylase inhibitors. Sci Rep. 2019;9(1):9702.

    Article  PubMed  PubMed Central  Google Scholar 

  350. Gemenetzidis E, Gammon L, Biddle A, Emich H, Mackenzie I. C. Invasive oral cancer stem cells display resistance to ionising radiation. Oncotarget. 2015;6(41):43964–77.

    Article  PubMed  PubMed Central  Google Scholar 

  351. Cheng L, et al. L1CAM regulates DNA damage checkpoint response of glioblastoma stem cells through NBS1. Embo J. 2011;30(5):800–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  352. Abad E, Civit L, Potesil D, Zdrahal Z, Lyakhovich A. Enhanced DNA damage response through RAD50 in triple negative breast cancer resistant and cancer stem-like cells contributes to chemoresistance. Febs J. 2021;288(7):2184–202.

    Article  CAS  PubMed  Google Scholar 

  353. Zhou W, et al. Activation of the phosphorylation of ATM contributes to radioresistance of glioma stem cells. Oncol Rep. 2013;30(4):1793–801.

    Article  CAS  PubMed  Google Scholar 

  354. Wang WJ, et al. MYC regulation of CHK1 and CHK2 promotes radioresistance in a stem cell-like population of nasopharyngeal carcinoma cells. Cancer Res. 2013;73(3):1219–31.

    Article  CAS  PubMed  Google Scholar 

  355. Bartucci M, et al. Therapeutic targeting of Chk1 in NSCLC stem cells during chemotherapy. Cell Death Differ. 2012;19(5):768–78.

    Article  CAS  PubMed  Google Scholar 

  356. Kim WT, Ryu CJ. Cancer stem cell surface markers on normal stem cells. BMB Rep. 2017;50(6):285–98.

    Article  PubMed  PubMed Central  Google Scholar 

  357. Barzegar Behrooz A, Syahir A, Ahmad S. CD133: beyond a cancer stem cell biomarker. J Drug Target. 2019;27(3):257–69.

    Article  CAS  PubMed  Google Scholar 

  358. Moreno-Londoño AP, Robles-Flores M. Functional roles of CD133: more than stemness associated factor regulated by the microenvironment. Stem Cell Rev Rep. 2024;20(1):25–51.

    Article  PubMed  Google Scholar 

  359. Bhat-Nakshatri P, et al. Breast Cancer cell detection and characterization from breast Milk-Derived cells. Cancer Res. 2020;80(21):4828–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  360. Xie Y, et al. LGR5 promotes tumorigenicity and invasion of glioblastoma stem-like cells and is a potential therapeutic target for a subset of glioblastoma patients. J Pathol. 2019;247(2):228–40.

    Article  CAS  PubMed  Google Scholar 

  361. Cao HZ, Liu XF, Yang WT, Chen Q, Zheng P. S. LGR5 promotes cancer stem cell traits and chemoresistance in cervical cancer. Cell Death Dis. 2017;8(9):e3039.

    Article  PubMed  PubMed Central  Google Scholar 

  362. Rodriguez-Torres M, Allan AL. Aldehyde dehydrogenase as a marker and functional mediator of metastasis in solid tumors. Clin Exp Metastasis. 2016;33(1):97–113.

    Article  CAS  PubMed  Google Scholar 

  363. Wang MC, et al. BMI-1, a promising therapeutic target for human cancer. Oncol Lett. 2015;10(2):583–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  364. Kharas MG, Lengner CJ. Stem cells, cancer, and MUSASHI in blood and guts. Trends Cancer. 2017;3(5):347–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  365. Mohiuddin IS, Wei SJ, Kang MH. Role of OCT4 in cancer stem-like cells and chemotherapy resistance. Biochim Biophys Acta Mol Basis Dis. 2020;1866(4):165432.

    Article  CAS  PubMed  Google Scholar 

  366. He J, et al. Chemoresistance of colorectal cancer to 5-fluorouracil is associated with Silencing of the BNIP3 gene through aberrant methylation. J Cancer. 2017;8(7):1187–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  367. Hu Q, et al. MicroRNA–137 exerts protective effects on hypoxia–induced cell injury by inhibiting autophagy/mitophagy and maintaining mitochondrial function in breast cancer stem–like cells. Oncol Rep. 2020;44(4):1627–37.

    CAS  PubMed  PubMed Central  Google Scholar 

  368. Wang D, et al. Cisplatin-resistant NSCLC cells induced by hypoxia transmit resistance to sensitive cells through Exosomal PKM2. Theranostics. 2021;11(6):2860–75.

    Article  PubMed  PubMed Central  Google Scholar 

  369. Rouschop KM, et al. The unfolded protein response protects human tumor cells during hypoxia through regulation of the autophagy genes MAP1LC3B and ATG5. J Clin Invest. 2010;120(1):127–41.

    Article  CAS  PubMed  Google Scholar 

  370. An Y, et al. miR-23b-3p regulates the chemoresistance of gastric cancer cells by targeting ATG12 and HMGB2. Cell Death Dis. 2015;6(5):e1766.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This research was funded in part by the National Natural Science Foundation of China (82272938 to Jiaqi Liu), the Beijing Nova Program (20220484059 to Jiaqi Liu), and the CAMS Innovation Fund for Medical Sciences (2021-I2M-1-014 to Jiaqi Liu).

Funding

This research was funded in part by National Natural Science Foundation of China (82272938 to Jiaqi Liu), Beijing Nova Program (20220484059 to Jiaqi Liu), CAMS Innovation Fund for Medical Sciences (2021-I2M-1-014 to Jiaqi Liu), and CAMS Innovation Fund for Medical Sciences (2023-I2M-C&T-B-086 to Jiaqi Liu).

Author information

Authors and Affiliations

Authors

Contributions

Tongxuan Shang, Ziqi Jia, Jiayi Li, Heng Cao, and Jiaqi Liu wrote the first manuscript draft. Tongxuan Shang, Jiayi Li, and Ziqi Jia generated the tables and figures. Jiaqi Liu and Xiang Wang conceptualized and supervised this review. All the authors participated in revising the manuscript before submission and the formal revision, including literature searching, information extraction, and text modification. The corresponding authors attest that all listed authors meet authorship criteria and that no others meeting the criteria have been omitted.

Corresponding authors

Correspondence to Xiang Wang or Jiaqi Liu.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shang, T., Jia, Z., Li, J. et al. Unraveling the triad of hypoxia, cancer cell stemness, and drug resistance. J Hematol Oncol 18, 32 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13045-025-01684-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13045-025-01684-4

Keywords