Next Article in Journal
Phenotype Transformation of PitNETs
Previous Article in Journal
FDG-PET in Chimeric Antigen Receptor T-Cell (CAR T-Cell) Therapy Toxicity: A Systematic Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hypoxia-Inducible Factor-Dependent and Independent Mechanisms Underlying Chemoresistance of Hypoxic Cancer Cells

by
Peter Wai Tik Lee
1,
Lina Rochelle Koseki
1,
Takao Haitani
1,2,3,
Hiroshi Harada
1,2,* and
Minoru Kobayashi
1,2,*
1
Laboratory of Cancer Cell Biology, Graduate School of Biostudies, Kyoto University, Kyoto 606-8501, Japan
2
Department of Genome Repair Dynamics, Radiation Biology Center, Graduate School of Biostudies, Kyoto University, Kyoto 606-8501, Japan
3
Department of Urology, Graduate School of Medicine, Kyoto University, Kyoto 606-8507, Japan
*
Authors to whom correspondence should be addressed.
Cancers 2024, 16(9), 1729; https://doi.org/10.3390/cancers16091729
Submission received: 18 April 2024 / Revised: 25 April 2024 / Accepted: 26 April 2024 / Published: 29 April 2024

Abstract

:

Simple Summary

In solid tumors, oxygen concentration varies between different regions. Generally, while oxygen supply is sufficient in the vicinity of blood vessels, the concentration of oxygen gradually drops as the distance from the blood vessel increases. In the regions with low oxygen content (hypoxic regions), cancer cells acquire various malignant properties, like invasiveness, altered metabolism, and therapy resistance, leading to recurrence and poor clinical outcomes of patients. Hypoxia-inducible factor (HIF) is a major regulator of hypoxia responses. Herein, we summarized how tumor hypoxia activates different mechanisms, in both HIF-dependent and HIF-independent manners, and contributes to the acquisition of chemotherapy resistance. We also discussed the involvement of epigenetic regulation in hypoxia-induced chemoresistance, with a specific example of ATAD2 protein degradation inducing drug resistance under hypoxia. Finally, we briefly reviewed some current clinical trials that target HIF and tumor hypoxia for cancer treatment or therapy sensitization.

Abstract

In hypoxic regions of malignant solid tumors, cancer cells acquire resistance to conventional therapies, such as chemotherapy and radiotherapy, causing poor prognosis in patients with cancer. It is widely recognized that some of the key genes behind this are hypoxia-inducible transcription factors, e.g., hypoxia-inducible factor 1 (HIF-1). Since HIF-1 activity is suppressed by two representative 2-oxoglutarate-dependent dioxygenases (2-OGDDs), PHDs (prolyl-4-hydroxylases), and FIH-1 (factor inhibiting hypoxia-inducible factor 1), the inactivation of 2-OGDD has been associated with cancer therapy resistance by the activation of HIF-1. Recent studies have also revealed the importance of hypoxia-responsive mechanisms independent of HIF-1 and its isoforms (collectively, HIFs). In this article, we collate the accumulated knowledge of HIF-1-dependent and independent mechanisms responsible for resistance of hypoxic cancer cells to anticancer drugs and briefly discuss the interplay between hypoxia responses, like EMT and UPR, and chemoresistance. In addition, we introduce a novel HIF-independent mechanism, which is epigenetically mediated by an acetylated histone reader protein, ATAD2, which we recently clarified.

1. Brief Introduction

The oxygen microenvironment inside of a solid tumor tissue is highly heterogeneous; the partial oxygen pressure (pO2) varies among different regions. In the regions approximately 70~100 μm away from tumor blood vessels, a hypoxic environment exists in which cancer cells are chronically not supplied with sufficient oxygen due to the imbalance between the oxygen consumption by cancer cells and the oxygen supply from blood vessels, which is caused by a low vascular density and an abnormal tumor vessel structure [1]. In addition, the transient occlusion of tortuous and immature tumor blood vessels give rise to an acutely hypoxic environment even in the proximal regions of blood vessels. Previous studies have shown that cancer cells in hypoxic regions demonstrate higher resistance to chemotherapy and radiotherapy; moreover, a higher intratumoral hypoxic fraction has been shown to be correlated with poor prognosis in multiple types of patients with cancer [2,3]. To improve the outcome of cancer treatment, it is necessary to elucidate the molecular mechanisms by which cancer cells acquire resistance to conventional therapies by exploiting the hypoxia response mechanism and to apply these findings to the establishment of novel therapeutic strategies.
In this article, we will first overview the basis of hypoxia-induced therapy resistance due to the physical distance from tumor blood vessels and the resultant decrease in pO2, and then we will summarize the knowledge on HIF-dependent molecular mechanisms which has been elucidated from the viewpoints of hypoxia and HIF biology. We will also introduce some HIF-independent mechanisms by which hypoxic cancer cells acquire resistance to chemotherapy. Finally, we will briefly discuss epigenetic gene regulation in relation to hypoxia and present an update on our discovery of an epigenetic mechanism mediated by the HIF-independent function of the ATPase family AAA domain containing 2 (ATAD2) that recognizes acetylated histone H4 [4].

2. Hypoxic Microenvironment in Malignant Solid Tumors and Its Association with Therapy Resistance

Cancer cells accessible to sufficient nutrients and oxygen (normoxic tumor cells) actively proliferate around tumor blood vessels in solid tumor tissues, resulting in high oxygen consumption [1,5,6]. In addition, tumor blood vessels are fragile and leakier than normal blood vessels, which causes higher interstitial pressure within tumor tissues and reduced oxygen diffusion. For these reasons, oxygen consumption greatly exceeds oxygen supply within tumor tissues, and this imbalance generates chronically hypoxic regions in the regions approximately 70–100 μm away from tumor blood vessels [1]. Moreover, because tumor blood vessels are tortuous and immature, they often fall into an occluded state, and this induces an acutely hypoxic environment even in the proximal regions of vessels [1].
The high interstitial pressure also results in an insufficient diffusion of anticancer drugs within a tumor tissue [1,7]. Additionally, hypoxia impedes drug delivery as it promotes intratumoral fibrosis. Particularly in pancreatic cancers, there is extensive lysyl oxidase (LOX)-mediated crosslinking of collagen fibers in the tumor stroma, and hypoxia also stimulates stromal cells’ collagen production [8,9,10]. As a result, the highly dense extracellular matrix acts as a physical barrier that impairs drug diffusion from blood vessels.
Moreover, in such hypoxic microenvironments, many chemotherapeutic drugs which mechanistically target actively growing cells, including alkylating agents, platinum-based drugs, and taxanes, do not exhibit full efficacy as hypoxic tumor cells do not actively proliferate and are in a state of cell cycle delay [7,11].
Tumor hypoxia remains a roadblock for not only chemotherapy, but also radiotherapy, especially those involving X-rays and other types of radiation with low linear energy transfer (LET) [6,12,13,14]. The half-life and the amount of cytotoxic reactive oxygen species (ROS) produced by ionizing radiation are increased in the presence of oxygen; conversely, under hypoxic conditions, the lack of oxygen reduces ROS generation/activity and thus hampers the induction of DNA damage by radiation [6,12,15]. It is also known that the ends of DNA double-strand breaks caused by radiation are efficiently oxidized and are less likely to be repaired under normoxia, whereas hypoxia represses such oxidation and results in more transient DNA damage [6,12,15,16]. Thus, it has been confirmed that cancer cells acquire resistance to radiation therapy in hypoxic regions.
In addition to these well-established reasons behind therapy resistance that we have come to understand over the years, the importance of the biological mechanism mediated by hypoxia-responsive genes has also become recognized, as described below.

3. HIF-Mediated Mechanisms behind Chemotherapy Resistance of Cancer Cells under Hypoxia

3.1. The Molecular Mechanisms behind the Regulation of HIFs’ Activity

HIF, the master regulator of hypoxic responses, is known to promote hypoxia-mediated chemoresistance. HIF functions as a heterodimeric transcription factor, which consists of one α subunit (HIF-1α, HIF-2α, or HIF-3α; hereafter, HIF-αs) and one β subunit (ARNT/HIF-1β), to regulate the transcription of thousands of downstream genes in response to hypoxia stimuli [17]. While the β subunit is stably and constitutively expressed regardless of the oxygen conditions, the expression level of HIF-αs is regulated post-translationally in an oxygen-dependent manner. Under normoxic conditions, prolyl-4-hydroxylase domain (PHD) proteins, which belong to the O2/Fe2+/2-oxoglutarate (2-OG)-dependent dioxygenases (2-OGDD) superfamily, are catalytically active and hydroxylate HIF-αs at two proline residues located in the oxygen-dependent degradation (ODD) domain (HIF-1α: P402 and P564; HIF-2α: P405 and P531) [18,19,20,21,22]. The hydroxylated HIF-αs are then recognized and ubiquitinated by the E3 ubiquitin ligase containing von Hippel–Lindau protein (pVHL) and subsequently degraded by the 26S proteasome machinery [23,24,25,26]. Additionally, in normoxic environments, another 2-OGDD protein, factor inhibiting HIF-1 (FIH-1), can suppress the transactivation activity of HIF; FIH-1 hydroxylates the asparagine-803 residue of HIF-1α (N847 of HIF-2α) and thereby obstructs HIFs’ interaction with other transcription co-factors, including p300 and CREB-binding protein (CBP) [27,28,29]. Together, these maintain the low levels of HIF-α proteins as well as HIF activity under normoxia. Contrarily, under hypoxic conditions, the oxygen-dependent PHDs and FIH become inactive, leading to the accumulation of HIF-αs, which interact with HIF-1β and p300/CBP to initiate the transcription of downstream genes. This allows HIFs to promote malignant properties, like angiogenesis [30,31,32,33,34], tumor cell migration/invasion [35,36,37], and metabolic reprogramming to the anaerobic glycolytic pathways [38], under hypoxia.

3.2. HIF-Mediated Mechanisms behind Chemotherapy Resistance

Extensive research has demonstrated the involvement of HIFs in cancer therapy resistance. Below, the major mechanisms through which HIF contributes to chemoresistance will be introduced (Figure 1).
First, HIFs have been reported to promote the chemoresistance of cancer cells under hypoxia by enhancing the expression of ATP Binding Cassette (ABC) transporters, which directly efflux chemotherapeutic agents out of cancer cells [39]. The expression of several ABC transporters is induced under hypoxic conditions in a HIF-1/2-dependent manner, including ABCB1 (Multi-Drug Resistance Protein 1 (MDR1) or P-glycoprotein 1), ABCC1 (Multi-Drug Resistance-Associated Protein 1 (MRP1)), and ABCG2 (Breast Cancer Resistance Protein (BCRP)) [39,40,41,42,43]. In particular, high HIF-1α and MDR1 expression levels were correlated with a higher resistance to 5-FU-based chemotherapy in patients, and sensitivities to various anticancer drugs were restored upon the suppression of HIF-1α/2α [39,40,41,44].
Second, HIF-1 can upregulate the expression of the inhibitors of apoptosis protein (IAP) family in cancer cells to enhance cell viability and drug resistance [45]. It has been reported that HIF-1 induces the expression of a member of the IAP family protein, Survivin, and conversely suppresses the expression of Caspase-9, which is involved in the induction of cell death via the intrinsic apoptosis pathway, thereby enhancing the anti-apoptotic potential of hypoxic cancer cells and leading to resistance to anticancer drugs [45]. In addition, the inhibition of Survivin and HIF-1α in vitro has been shown to sensitize cancer cells to cisplatin treatment by enhancing the activity of the execution caspase, caspase-3 [46]. Recent reports have also elucidated the regulatory role of HIF-1α in microRNA expression to circumvent chemotherapy-induced apoptosis, either by directly targeting factors involved in the caspase-mediated apoptosis pathway or indirectly targeting upstream signaling. For example, under hypoxia, HIF upregulates the expression level of both miR-675-5p, which targets the 3′-UTR region of the caspase-3 mRNA, and miR-27a, which targets the mRNA of APAF1 (apoptotic protease activating factor 1), the major component of apoptosome, and it enhances 5-FU resistance in colorectal cancer cells and paclitaxel resistance in ovarian cancer in vitro. In addition, HIF has been reported to promote docetaxel resistance in triple-negative breast cancer cells by suppressing miR-494 expression to enhance Survivin expression [47]. Furthermore, HIF is known to cause oxaliplatin resistance in colorectal cancer by inducing miR-338-5p expression under hypoxia to promote IL-6 signaling, which, in turn, suppresses apoptosis via the STAT3/BCL2 pathway [48,49,50].
Third, the expressions of SLC7A11/xCT, a cysteine transporter, and Glutamate-Cysteine Ligase Regulatory Subunit (GCLM), an enzyme responsible for glutathione (GSH) synthesis from cysteine, are both induced by chemotherapeutic agents in a HIF-1-dependent manner. This maintains a high level of the reduced form of glutathione to mitigate oxidative stress in cancer cells, rendering higher tolerance to therapeutic drugs [51].
Fourth, the effects of conventional anticancer drugs are attenuated because of the suppression of cancer cell proliferation in the hypoxic microenvironment; congruently, HIF-1 has been shown to negatively regulate cyclin D1 and c-Myc expression and increase p21CIP1 protein levels, thereby halting the progression of the cell cycle and negating the effect of chemotherapeutics [52,53]. Intriguingly, it has also been found that chemotherapy-induced senescence, which can be characterized by irreversible cell cycle arrest, is subdued under hypoxia in a HIF-1-dependent manner. This also allows hypoxic cancer cells to survive as well as to retain the potential to proliferate after anticancer drug treatments, indicating that the cell cycle may actually be stringently controlled by HIF under hypoxia via intricate mechanisms, rather than by simple go–stop signals [54].
Fifth, hypoxia can promote autophagy-mediated chemoresistance by the induction of autophagy-triggering proteins, like eIF5A2 and VMP1, both of which are HIF-1 downstream genes [55,56,57]. In addition, HIF has been reported to regulate the key factors involved in various steps of autophagosome formation to promote drug resistance. Beclin-1 (also named ATG6), which is important for the formation of the pre-autophagosomal structure and isolation membrane (autophagy initiation), is induced under hypoxia by the well-known HIF target genes, BNIP3 and FOXO3a, in gemcitabine-resistant bladder cancer and sorafenib-resistant hepatocellular carcinoma, respectively [58,59]. HIF-1 has also been found to upregulate ATG5 expression, which promotes phagophore elongation, and thus induce autophagy-mediated chemoresistance in glioblastoma and lung cancer both in vitro and in vivo [60,61]. Other regulatory mechanisms of hypoxia-induced autophagy, like the PTBP3-ATG12 axis, have also been indicated in chemotherapy-resistant pancreatic cancer in relation to HIF [62]. Of note, the role of autophagy in cancer progression and therapy resistance still remains elusive, as both tumor-suppressing and tumor-promoting mechanisms have been proposed and are highly dependent on the exact conditions of cells and the environment [63,64]. Nonetheless, in the context of hypoxia, a vast majority of reports to date have underpinned the importance of autophagy-mediated chemoresistance.
In addition, HIF has been shown to ameliorate DNA damage induced by chemotherapeutic agents. Breast cancer cells and prostate cancer cells exposed to hypoxia in vitro had reduced levels of topoisomerase IIα, a DNA binding protein essential for DNA-damaging agents to introduce double strand breaks, and such decrease was reversed by the knockdown of HIF-1α. Whilst the detailed mechanism by which HIF regulated topoisomerase IIα expression remains unclear, this study demonstrated that HIF can reduce chemotherapy-induced double strand break and thus causes resistance to etoposide treatment in hypoxic cancer cells [65].
Furthermore, the presence of cancer stem cells (CSCs) in tumors is known to confer chemotherapy resistance, and HIF has been found to foster a CSC enrichment. Intriguingly, HIF-2α increased c-Myc, Oct-4, and Nanog expressions and engendered a stem cell phenotype in breast cancer cells, contributing to Paclitaxel resistance in vitro and in vivo. Conversely, HIF-2α inhibition suppressed stemness and restored sensitivity to Paclitaxel treatment [66,67]. Others, using in vitro and in vivo models, have also demonstrated the involvement of HIF-1 in recruiting tumor-associated macrophages (TAMs) that produce protumor cytokines, e.g., growth/differentiation factor 15 (GDF15), and evoke chemoresistance in a paracrine manner in gastric and colorectal cancer cells. TAMs under low oxygen conditions have also been shown to induce temozolomide resistance in glioblastoma cells by VEGF signaling [68,69,70]. To summarize, numerous research studies have accentuated the role of HIFs in conferring chemotherapy resistance to hypoxic cancer cells via various pathways and mechanisms (Figure 1).

3.3. HIF-Related Hypoxia-Responsive Non-Coding RNAs and Chemotherapy Resistance

In the past decade, with advancing RNA sequencing technology, tremendous effort has been devoted to interrogating the regulation of non-coding RNA in response to hypoxia, as well as the clinical relevance of aberrant non-coding RNA expression to therapy resistance [71,72,73,74,75,76]. Hypoxia-responsive non-coding RNAs themselves usually do not directly exhibit biochemical activity for chemoresistance; instead, they provoke changes through the regulation of downstream genes. MicroRNA (miRNA) functions by binding to the complementary sequences in target mRNA, which leads to mRNA degradation or block translation, both resulting in the decreased expression of the downstream target. On the other hand, the two other major types of non-coding RNA, long non-coding RNA (lncRNA) and circular RNA (circRNA), mainly work by sponging miRNAs from their target mRNA and thus interfere with the suppressive effect.
The role of miRNAs in drug resistance has been implicated in a variety of cancers [77,78,79,80,81,82,83,84,85,86]. By regulating the expression of different downstream targets, miRNAs impact a wide range of biological processes that contribute to therapy resistance [78,87,88,89,90,91]. Above, a few examples of HIF-dependent hypoxia-responsive miRNA involved in the regulation of apoptosis have already been introduced. Apart from those pathways, HIF-1α has been shown to induce miR-210-3p expression under hypoxia, which positively regulates TGF-β to enhance in vitro temozolomide resistance in glioma cells [92]. On the other hand, the expression of HIF-1α is also regulated by multiple miRNAs, some of which promote drug resistance [93]. For instance, miR-194-5p was downregulated upon hypoxia treatment, and this increased the expression of its downstream target, HIF-1α, contributing to the hypoxia-induced doxorubicin resistance in A549 non-small-cell lung cancer in vitro [94]. Conversely, miR-301a expression was increased under hypoxia, and it prevented the TAp63-mediated degradation of HIF-1α to induce gemcitabine resistance in MIA-PaCa-2 and BxPC-3 cells [95].
The involvement of lncRNA in therapy resistance has also been widely acknowledged, and there have been extensive studies in the literature oriented towards individual lncRNA or anticancer drugs [71,96,97,98,99,100,101,102,103]. Herein, instead of restating the details of each pathway, we summarized the lncRNAs which have been shown to induce chemoresistance under hypoxic conditions (Table 1). Importantly, whilst many of the existing studies are indicative of the non-coding RNAs being potentially involved in chemoresistance under hypoxia, caution should be taken when interpreting the results, as the functions/regulatory roles of these non-coding RNAs are not necessarily identical under normoxic and hypoxic conditions. Indeed, although there are plenty of studies that have shown the involvement of particular non-coding RNAs in drug resistance or the expression levels of various non-coding RNAs being altered under hypoxia, concrete evidence that comprehensively demonstrates the contribution of non-coding RNA to the chemoresistance of hypoxic cancer cells remains rather limited.
Compared to lncRNA, even less is currently known regarding the role of circRNA in cancer progression and therapy resistance, and there have been very limited reports about hypoxia-responsive circRNAs [104,105,106,107]. To date, only a few hypoxia-responsive circRNAs have been characterized to be relevant to chemoresistance. CircNRIP1 (hsa_circ_0004771) has been shown to be increased by hypoxia in gastric carcinoma cells in vitro, and such increase allowed the sponging of miR-138-5p, which targets the 3′UTR of HIF1A mRNA. This, in turn, upregulated HIF-1α expression and augmented HIF-mediated 5-FU resistance under hypoxia in a glycolysis-dependent manner [108]. Intriguingly, a recent report has demonstrated an alternative regulatory mechanism distinct from miRNA sponging that circRNA induces chemoresistance. In breast cancer cells, hypoxia increased circSTT3A (has_circ_0024760) in a HIF-1α-dependent manner; circSTT3A directly bound with the HSP70 protein to recruit and stabilize PGK1. This promoted the serine-S-adenosylmethionine (SAM)–H3K4 trimethylation axis and enhanced stemness, eventually contributing to doxorubicin resistance in a xenografted mouse model [109].
In addition to the non-coding RNAs expressed in hypoxic cancer cells, the involvement of non-coding RNAs from hypoxia-induced exosomes has also been gradually revealed. For example, in a study of pancreatic cancer, a circRNA microarray identified that circZNF91 is more abundant in exosomes from hypoxic cells. This circRNA sponges miR-23b-3p to enhance SIRT1-HIF-1α signaling in recipient cells and thereby induces gemcitabine resistance in cell culture and xenografted tumor models [110]. In a separate study, it has been shown that HIF-induced exosomal miR-21 from hypoxic CAFs activates the RAS-AKT-ERK pathway and promotes gemcitabine resistance in the recipient pancreatic cancer cells both in vitro and in vivo [111].
Together, these results demonstrate the role of hypoxia-responsive non-coding RNAs, acting either upstream or downstream of HIF, in hypoxia-induced chemoresistance.

3.4. HIF, EMT, and Chemotherapy Resistance

Epithelial-mesenchymal transition (EMT) initially describes the process of cell morphology changes wherein cells lose epithelial characteristics and acquire mesenchymal characteristics. However, with an increasing understanding of the molecular basis that governs EMT, it is recognized that a group of transcription factors (EMT-TFs), including SNAIL, SLUG, ZEB, and TWIST, plays a pivotal role in regulating EMT. The expression statuses of EMT-TFs and their downstream targets (EMT markers), rather than morphological changes, essentially became accepted for assessing the occurrence of EMT. Under hypoxia, EMT-TF expression and signaling are mainly upregulated by HIF-mediated transcriptional activation, as many of the EMT-TFs contain HRE in their promoter regions and are direct targets of HIF. In addition, the crosstalk with other pathways (e.g., NF-κB, NOTCH, and β-catenin) which further potentiate HIF-induced EMT has also been demonstrated [112,113,114,115].
In cancer, EMT contributes to disease progression not only by promoting invasion and metastasis, but also by affecting tumorigenesis, metabolic reprogramming, tumor dormancy, stemness, genome stability, etc. [116,117,118]. Moreover, EMT has been shown to be associated with resistance to radiotherapy and chemotherapy and thus may contribute to unfavorable outcomes in patients. The interplay between hypoxia and EMT, as well as that between EMT and chemoresistance, has been vigorously studied and reviewed, so the details will not be reiterated herein [119,120,121,122,123,124,125].
Notably, in order to elucidate the linkage between EMT-TFs/EMT markers and chemoresistance, a vast majority of mechanistic studies utilized in vitro and xenograft models, and patient samples/databases have been used for correlational studies. Compelling evidence comes from two separate studies employing transgenic mouse models. Zheng et al. showed that the abrogation of EMT-TF (SNAIL or TWIST) enhances the sensitivity to the gemcitabine of primary PDAC tumor [126]. And Fischer et al., by distinguishing the epithelial/mesenchymal transcription activity of cells with transgenic fluorescent proteins, showed that cells that have undergone EMT in primary mammary adenocarcinoma are significantly less susceptible to cyclophosphamide treatment [127]. More recently, another group unraveled the mechanism responsible for EMT-associated chemoresistance in primary skin squamous cell carcinoma, showing that EMT cells express the small GTPase RHOJ to regulate nuclear actin dynamics and promote DNA repair, thereby inducing resistance to cisplatin and 5-FU treatment [128].
However, it should be noted that the experiments and analyses in these studies were performed without taking oxygen content into account; so, whether these mechanisms can function identically in hypoxic cells to induce chemoresistance remains to be verified. Also, whilst hypoxia is considered to be an important driver of EMT via HIF activation, studies have shown that treatment with chemotherapeutic agents could trigger EMT and subsequently promote resistance and metastasis in vitro and in vivo [129,130,131,132,133,134,135]. Clarifying the difference between hypoxia/HIF-induced EMT and chemotherapy-induced EMT may help reveal the relevant molecular mechanisms that hypoxic cancer cells exploited to induce EMT-mediated chemoresistance. Indeed, a continuous effort has been made to examine the mechanism of hypoxia contributing to EMT-mediated chemoresistance [136,137,138], and it has been reported that in hepatocellular carcinoma cells and colorectal cancer cells, the in vitro blockade of hypoxia-induced EMT with different small-molecule compounds could potentially attenuate HIF-mediated chemoresistance [139,140].

4. HIF-Independent Mechanisms behind Chemotherapy Resistance of Cancer Cells under Hypoxia

Whilst HIF functions as a robust transcription factor regulating multiple downstream targets upon hypoxia and induces the drug resistance of hypoxic cancer cells, other HIF-independent molecular pathways have also been reported to contribute to chemotherapy resistance under hypoxia (Figure 2).

4.1. HIF-Independent Mechanisms behind Chemotherapy Resistance

In response to DNA-damaging drugs, hypoxic cancer cells also rely on HIF-independent mechanisms to circumvent apoptosis and thereby develop resistance to chemotherapy. p53 signaling has been reported to affect the drug resistance of U2OS osteosarcoma cells under hypoxia. Treatment with DNA-damaging agents, including cisplatin, etoposide, doxorubicin, etc., can trigger p53 activation by phosphorylation at the Ser-15 residue under normoxia. But the phospho-p53 (Ser-15) level is significantly lowered under hypoxia despite drug treatment, and such a decrease is independent of HIF-1 activity. Correspondingly, the expression levels of p53 downstream targets like NOXA and PUMA are also suppressed under hypoxia after drug treatment, potentially mitigating the pro-apoptotic effect induced by these chemotherapeutic drugs [141]. In other studies using the Hep G2 hepatoma cell, it has been shown that hypoxia decreases p53 expression, resulting in the suppression of the expression of pro-apoptotic protein BAK1, and enhances the JNK-AP-1 pathway, resulting in the reduction in apoptosis upon etoposide treatment [142,143]. Moreover, the expression of the anti-apoptotic factor Pim-1 is enhanced under hypoxia in a HIF-independent manner, leading to decreased activities of caspase-9 and caspase-3 and the stabilization of mitochondrial transmembrane potential. Such suppression of the intrinsic apoptosis pathway results in a higher resistance to cisplatin treatment in hypoxic pancreatic cancer cells, and it was also shown that the forced expression of a dominant-negative form of Pim-1 can sensitize xenografted tumor to chemotherapy in vivo [144]. Other HIF-independent pathways that enhance resistance to chemotherapeutic agents by suppressing apoptotic signals under hypoxia have also been reported in A549 lung adenocarcinoma cells, including the cyclooxygenase-2 (COX-2)–prostaglandin E2 (PGE2) pathway and the Sphingosine Kinase 2 (SphK2)–Sphingosine-1-Phosphate (S1P)–MAPK pathway [145,146].
Additionally, hypoxia-responsive non-coding RNAs functioning independent of HIF have been shown to contribute to chemoresistance. For example, the hypoxia-inducible lncRNA HIF1A-AS2 regulates HMGA1 expression to suppress p53 family protein activities under hypoxic conditions. This was shown to downregulate Bax expression and promote cisplatin resistance in bladder cancer cells in vitro [147]. Also, circELP3 (hsa_circ_0001785) has been shown to be elevated under hypoxia in a HIF-independent manner, and its depletion by siRNA-mediated knockdown abolished the hypoxia-induced cisplatin resistance in bladder cancer cells in vitro. However, the detailed molecular mechanism like the target of this circRNA was not elucidated [148]. CircUBE2D2 (hsa_circ_0005728) has also been found to promote glycolysis and induce chemoresistance in hepatocellular carcinoma cells under hypoxia; this circRNA was increased in vitro after exposure to mild hypoxia (5% O2) and could sponge miR-889-3p to upregulate the expression of LDHA, through which it is suggested to induce sorafenib resistance [149]. Also, in vitro assays have shown that exosomes from hypoxic pancreatic cells have larger amounts of the lncRNA lncROR, and this suppresses the Hippo-YAP pathway in the recipient cells, rendering them gemcitabine-resistant by preventing cell cycle arrest and apoptosis [150]. And the exosomes from hypoxic stromal cells, like cancer-associated fibroblasts (CAFs), contain the hypoxia-responsive lncRNA H19, which activates the DNMT-miR-497 axis and induces paclitaxel resistance in recipient breast cancer cells [151].
CD133 expression has been reported to be upregulated via HIF-dependent and HIF-independent pathways under hypoxia, and both can render resistance against temozolomide and cisplatin to glioblastoma cells [152,153]. Notably, CD133 is commonly used as a surface marker of cancer stem cells. In separate studies using patient-derived samples of colon cancer or rectal cancer, it was found that CD133-positive cells are mainly located in the hypoxic fractions of tumors, but the expression level of CD133 is inversely correlated with that of HIF-1α in patients receiving preoperative chemoradiotherapy. Also, CD133-positive proliferative cells increased after 5-FU chemotherapy, whereas the amount of CD133-negative proliferative population remained unchanged, indicating the potential contribution of CD133 expression to therapy resistance and post-therapy recurrence in a HIF-independent manner [154,155].
Moreover, hypoxia has been reported to cause aneuploidy in Ewing sarcoma via the HIF-independent DDP4/Neuropeptide Y (NPY)–NPY-Y5 Receptor (Y5R)–Rho A axis. These hypoxic polyploid cells were shown to demonstrate higher resistance to doxorubicin treatment than normoxic cells, implying the role of hypoxia-induced polyploidy in chemotherapy failure [156,157,158].
Apart from conventional anticancer agents, hypoxia also induces resistance to targeted chemotherapeutic agents. For instance, Sorafenib, an inhibitor of the RAF-MEK-ERK pathway approved for hepatocellular carcinoma treatment, had significantly lower efficacy in inducing apoptosis under hypoxia due to the hypoxic activation of the compensatory Hippo-Yes-associated Protein (YAP) pathway [159,160]. Together, these findings signify the importance of HIF-independent mechanisms. However, since the molecular mechanisms remain largely unknown, it is critical to unveil them and utilize the acquired information to establish a means to overcome hypoxia-induced therapy resistance. Notably, the mechanisms of hypoxia-induced chemotherapy resistance can usually be triggered by multiple pathways. Those activated by HIF signaling, as detailed in the previous section, may also be induced via HIF-independent pathways to confer chemoresistance under hypoxia. For instance, ABCB1 expression, which facilitates drug efflux, can be enhanced by NRF2 and thus contributes to resistance against cisplatin and doxorubicin in hypoxic MCF7 and Hep G2 cells in vitro, respectively [161,162]. NRF2 has also been shown to upregulate the expression of antioxidant proteins GCLC and GCLM in a HIF-independent manner, thereby attenuating oxidative stress induced by cisplatin treatment under hypoxia [162]. In addition, the induction of NURP1 in hypoxic glioma cells has been shown to promote the KDM3A–TFEB axis to enhance autophagy-mediated temozolomide resistance both in vitro and in vivo [163].

4.2. Hypoxia-Associated Proteotoxicity, UPR, and Chemotherapy Resistance

Tumor hypoxia is a hostile microenvironment for cancer cells not only because it imposes oxidative stress and metabolic stress, but also due to the proteotoxicity it inflicts. Post-translational protein folding and isomerization in the endoplasmic reticulum (ER) have been shown to be oxygen-dependent; under hypoxia, protein maturation is impaired, and this induces ER stress as unfolded peptides/misfolded proteins accumulate and aggregate in the ER [164]. In order to resolve ER stress, signaling pathways of the unfolded protein response (UPR) are activated by GRP78 (glucose-regulated protein 78, also named BIP) to boost chaperone activity, induce autophagy, and promote ER-associated degradation (ERAD), all of which are directed towards the refolding and/or clearance of misfolded proteins/aggregates.
Notably, the fate of cells after UPR activation is highly context dependent. Depending on the severity and duration of hypoxia/ER stress, as well as on the status and crosstalk with other molecular responses, different branches of UPR pathways could be induced differently, either leading to the upregulation of the pro-survival gene profile, or when ER stress cannot be tolerated, to the initiation of the apoptosis program [165,166,167].
In general, UPR pathways are augmented under hypoxic conditions to help tumor cell survival and cancer progression. Thus, as a pro-survival mechanism to adapt hypoxia, it seems reasonable that UPR is adopted as yet another tactic by hypoxic cancer cells to withstand chemotherapeutic treatment. Indeed, there have been abundant experimental and clinical data demonstrating the aberrant expression of UPR pathway proteins (e.g., GRP78, XBP1(s), PERK, and ATF4) in various types of cancers, and the molecular mechanism by which UPR induces chemoresistance has also been intensively studied [167,168,169]. But, as mentioned above, the result of UPR activation depends heavily on the cellular context, and it has been pointed out that the temporal order of events (whether hypoxia drives UPR or UPR occurs before the hypoxic response) might also be one of the determinants [170]; hence, whether the findings can be extended to the situation in hypoxic cancer cells requires further examination.
More direct evidence of hypoxia-driven URP inducing chemoresistance came from two in vitro studies. The first study used a hypopharyngeal squamous carcinoma cell line and showed that the knockdown of GRP78, abrogating the hypoxic activation of UPR, leads to the increased expression of Bax (pro-apoptotic) and the reduced expression of Bcl-2 (anti-apoptotic) upon cisplatin treatment [171]. The second study showed that hyperoxia treatment (40% O2) can redirect UPR to pro-apoptotic signaling and sensitize glioblastoma cells to temozolomide [172]. In addition, recent results have also started to reveal the underlying molecular mechanisms in further detail; one group reported that hypoxia induces the IRE1α-XBP1(s) arm of URP to enhance 5-FU resistance in colorectal cancer cells, likely by unleashing autophagy suppressed by miR-34a [173].
Of note, although the majority of UPR pathways are often considered HIF-independent, accumulating evidence has shown the opposite [66,174,175,176,177,178]. This certainly adds further complexity to deciphering the interplay between hypoxia response and UPR, as well as the resulting effect on chemoresistance under hypoxia.
In addition to UPR, proteotoxic stress is known to activate heat shock transcription factor 1 (HSF1) as the stress-denatured proteins sequester heat shock proteins (HSPs) like HSP70 and HSP90. The liberated HSF1 can then upregulate the transcription of downstream genes, including HSP genes, which will potentially form a feedback loop. Under hypoxic conditions, it has been shown that the expressions of HSF1, HSP70, and HSP90 are increased, and the phosphorylation of HSF1, which is crucial for its transactivation activity, is also enhanced [179,180]. Indeed, several mechanisms by which HSPs contribute to therapy resistance have been proposed, and studies have also suggested that the overexpression of HSF1 may contribute to chemoresistance [181,182]. In particular, HSF1 has been shown to upregulate the expression of the MDR1 gene, and HSF1 overexpression indeed promoted efflux activity and induced resistance to doxorubicin and paclitaxel in melanoma cells in vitro [183,184]. Nonetheless, whether the HSF1-mediated induction of MDR1 also occurs under hypoxia is untested. Similarly, although there are results showing that heat stress induces the expression of superoxide dismutase (SOD), whose antioxidant activity has been implicated in drug resistance, the involvement of HSF1 as well as the linkage with hypoxia remain unknown [185,186,187,188]. Hence, while there are data indicative of the contribution of HSF1/HSPs, with the current information being fragmentary and the lack of concrete supporting evidence to date, their roles in hypoxic cancer cells acquiring chemoresistance are yet to be confirmed.
Regardless, taken all together, with the involvement of HIF-independent mechanisms being increasingly recognized, it is expected that such knowledge can provide us new opportunities and potential targets for overcoming hypoxia-induced chemotherapy resistance.

5. The Mechanisms of Therapy Resistance Acquisition through Hypoxia-Dependent Epigenetic Regulations: The Role of a Histone Acetyl Reader Protein, ATAD2

The hypoxic response of tumor cells is largely regulated by HIF-1, as mentioned above; however, in addition to HIF-1, epigenetic modifications such as chromatin condensation have also been thought to play a crucial role in regulating cell cycles and contribute significantly to the therapeutic resistance of hypoxic cancer cells [189]. A typical example of such epigenetic regulation is histone modification. In particular, the acetylation status of histone, which is mainly regulated by histone acetyltransferase (HAT) and histone deacetylase (HDAC), is known to control global transcription activity. HATs enzymatically facilitate the acetylation of various lysine residues in the N-terminal tail of histones, and this neutralizes the positive charge of lysine side chains, thus weakening the interaction between histones and DNA. This results in the destabilization of nucleosomes and the loosening of chromatin (euchromatinization) so that genomic DNA is more accessible to transcription machineries, thereby activating gene transcription. Conversely, HDACs remove the acetyl group from histone tails, leading to chromatin condensation and the suppression of gene transcription [190].
Indeed, it has been shown that hypoxia enhances HDAC activity in general [191]; two major ubiquitously expressed HDACs, HDAC1 and HDAC2, are reported to be rapidly activated in response to hypoxia [192]. At the chromatin level, hypoxia can induce global chromatin compaction and, correspondingly, decrease the amount of diffuse DNA within the nucleus [193]. Moreover, an ATAC-seq (Assay for Transposase-Accessible Chromatin with sequencing) analysis has indicated that chromatin accessibility to promoter regions was generally lowered upon both moderate and severe hypoxia treatments [194]. These results are in line with the global repression of gene transcription activity, as manifested by the decreases in the total RNA and mRNA synthesis of approximately 40–65% and 20–50%, respectively, under hypoxia [55]. Importantly, such transcription repression under hypoxia was retained even in the absence of HIF-1α or HIF-1β, supporting that the HIF-independent epigenetic regulation of gene expression would also substantially influence hypoxic responses and can thus be targeted for overcoming tumor hypoxia.
The mutation and dysregulation of HAT and HDAC have been found to contribute to the development, malignant progression, and recurrence of cancers and leukemia [4,195]. Particularly, it is known that the aberrant activation of HDAC can lead to chromatin aggregation and induce chemotherapy resistance. For example, the HDAC1/4 expression levels are significantly higher in patients with lung adenocarcinoma who are unresponsive to docetaxel-based chemotherapy, and they are correlated with shorter progression-free survival [196], whereas augmented HDAC2 expression has been shown to cause chromatin remodeling and cisplatin resistance in patient-derived ovarian cancer cells [197]. Therefore, in recent years, a variety of HDAC inhibitors have been developed, with several that have already been approved by the FDA for the treatment of hematological malignancies. Continuous effort is devoted to evaluate their applicability to treat solid tumors as well as to sensitize conventional chemotherapy [198].
Nonetheless, the progress towards clinical application for the treatment of solid tumors is rather limited to date [199,200], as adverse outcomes, which can be exemplified by the low efficacy accompanied by significant toxicity in urothelial cancer and exacerbated metastasis in breast cancer, have also been reported [201,202]. Thus, in order to overcome the anticancer drug resistance of hypoxic cancer cells, it is necessary to identify the key genes involved in chemotherapy resistance through hypoxia-dependent epigenetic regulation.
On top of the acetylation status of histones, which is determined by HATs and HDACs, the importance of histone acetyl readers, containing bromodomains (BRDs) that can recognize acetylated lysine residues, has also recently attracted meticulous attention [203]. A sequence analysis has identified over 40 BRD-containing proteins in the human proteome, and studies have found most of these BRD-containing proteins to be crucial to chromatin remodeling and gene transcription regulation [204]. The aberrant expression of BRD-containing protein has been implicated in inflammatory and autoimmune responses, neurological disorders, as well as hematological malignancies and cancers; thus, these proteins have emerged as attractive therapeutic targets for multiple diseases [205,206,207]. Indeed, approximately 15 different BRD inhibitors have already been used in clinical trials for the treatment of various cancers [208]. Yet, the regulation and function of BRD-containing proteins in the context of tumor hypoxia still remain largely unknown; further studies are therefore important to fully explore their potential as therapeutic targets.
ATAD2 is a protein with a bromodomain that binds to acetylated lysines 5 and 12 of histone H4 (H4K5ac and H4K12ac, respectively), and it has been reported to maintain histone hyperacetylation by preventing HDAC2-mediated deacetylation [209,210]. Recognition of these acetylated histones by ATAD2 has been reported to activate transcription factors such as E2F and Myc that positively regulate cell growth and the cell cycle [211,212].
Given the reports of ATAD2 functioning as a co-regulator of these oncogenic transcription factors, we recently analyzed the function of ATAD2 in the cell cycle progression and proliferation of cancer cells and its role in cancer chemoresistance [4]. ATAD2 protein levels were found to be significantly decreased under severe hypoxia (O2 < 0.1%), and this was accompanied by the reduction in acetylated histone H3 lysine 27 (H3K27ac) protein levels. The decrease in acetylated H3K27 was partially but significantly rescued by the forced expression of ATAD2, indicating the involvement of ATAD2 reduction in the global heterochromatinization of genome DNA. An FACS analysis subsequently revealed that severe hypoxic stress significantly decreases the proportion of cells proceeding from the early S phase to the late S phase. This phenomenon was significantly reversed by ATAD2 overexpression. These results indicate that ATAD2 works in cell cycle progression in the S phase and that a decrease in ATAD2 causes cell cycle retardation under severe hypoxia.
Cell cycle delay caused by ATAD2 reduction was confirmed to contribute to chemotherapy resistance. The knockdown of ATAD2 resulted in the resistance of cells towards a topoisomerase-inhibiting chemotherapeutic agent, camptothecin, even under normoxia. ATAD2-knockout HEK293 cells showed treatment resistance, which was recovered by cell cycle progression forced by ATAD2 reconstitution. Hypoxic stimulation was confirmed to induce resistance to camptothecin, and cells overexpressing ATAD2 were less likely to show resistance to chemotherapy induced by hypoxic stimulation. When administering other common chemotherapeutics, including platinum-based carboplatin and tubulin-stabilizing docetaxel, ATAD2-silenced cells exhibited a higher resistance to these drugs, indicating the generality of this phenomenon. Based on these results, we concluded that hypoxia induces chemotherapy resistance in cancer cells by decreasing the ATAD2 protein levels and delaying cell cycle progression, especially in the early S phase (Figure 3). In addition to this conclusion, our finding that the forced expression of ATAD2 increases the sensitivity of hypoxic cancer cells to anticancer drugs justifies the development of a novel strategy to inhibit the reduction in ATAD2 protein to sensitize hypoxic tumor cells to chemotherapy (Figure 3).

6. Oxygen-Dependent Regulatory Mechanism of ATAD2 Expression

How do ATAD2 protein levels decrease in cells exposed to hypoxic stimuli? To answer this question, a strategy to suppress chemotherapy resistance caused by the decreased ATAD2 protein levels under hypoxia should be suggested.
HIF-1α and its isoforms (HIFs) are stabilized and activated under hypoxia due to the inactivation of PHD and FIH-1, both of which belong to the Fe2+/2-OG-dependent dioxygenase (2-OGDD) superfamily. 2-OGDDs require oxygen for their activities, and their importance in hypoxic response has been receiving increasing attention [213]. Intriguingly, the stability of ATAD2 under normoxia was also found to be regulated by a protein belonging to the 2-OGDDs superfamily through the following results [4]: Treatment with an Fe2+ chelator, deferoxamine (DFO), or with a 2-OG analogue, dimethyloxalylglycine (DMOG), both of which inactivate 2-OGDDs, resulted in the decrease in ATAD2 protein levels even under normoxic conditions in various types of cells. And a proteasome inhibitor, MG-132, reversed the decrease in ATAD2 protein. Notably, the degradation of ATAD2 protein was observed in cell lines harboring mutant VHL, like clear cell renal cell carcinoma-derived cell lines, RCC4 and 786-O, upon hypoxic treatment, indicating that the regulation of ATAD2 protein was independent of the pVHL–HIF axis and its downstream genes [4]. Together, these results suggest that the chemotherapy resistance of hypoxic cancer cells due to ATAD2 protein degradation is triggered by the inactivation of 2-OGDD and mediated by the proteasome pathway independent of HIFs and pVHL (Figure 4). The further elucidation of this regulatory mechanism, particularly the identification of the responsible 2-OGDD(s), could thus lead to the establishment of new therapeutic strategies to overcome the chemotherapy resistance of hypoxic cancer cells by inhibiting ATAD2 proteolysis under hypoxic conditions and restoring cell cycle delay (Figure 5).

7. Summary and Perspectives

In a situation where the microenvironment in tumor tissues is highly heterogeneous and dynamic in time and space, individual cancer cells may acquire resistance to therapy by unique mechanisms. To overcome this problem, it is necessary to analyze how cancer cells adapt to their microenvironments and acquire resistance to therapy, and to develop therapeutic strategies to overcome resistance based on the knowledge obtained. It is known that the mechanisms behind the hypoxia response of cancer cells, which are mainly regulated by HIFs and their downstream genes, play pivotal roles in the chemoresistance of cancer cells under hypoxic conditions.
In the first part of this review, we summarized the major mechanisms triggered by HIF, each with examples of downstream pathways included, showing how HIF induces chemotherapy resistance in hypoxic cancer cells. From these results, it can be acknowledged that the inhibition of HIF would be a promising therapeutic strategy to not only suppress the HIF-induced malignant properties but also to overcome hypoxia-mediated resistance to therapy. Our group has previously reviewed a variety of HIF-targeting anticancer drugs as regards the diverse non-canonical mechanisms regulating HIF expression and activity [5]. Indeed, there has been a continuous effort to put HIF inhibition into clinical trials, with most studies using repurposed drugs that can indirectly inhibit HIF. Currently, almost all trials are Phase I or Phase II studies, and the results are rather limited [214]. In many cases, HIF inhibitors are administered in combination with other anticancer drugs; however, due to drug–drug interactions and differences in cancer types, it is usually difficult to predict clinical outcomes from other trials. For example, Ganetespib, a small-molecule drug that can potentially impede HIF stabilization by targeting HSP90, has been used in a Phase I study together with the antiangiogenic agent ziv-aflibercept to treat patients with advanced carcinoma or sarcoma; however, due to severe toxicity and adverse events, the trial was terminated (ClinicalTrials.gov ID: NCT02192541) [215]. In another Phase I trial targeting patients with a specific type of sarcoma (Malignant Peripheral Nerve Sheath Tumor, MPNST), the combined treatment of Ganetespib and mTOR inhibitor Sirolimus was tolerable, and a partial response was observed in one patient; however, when the study proceeded to Phase II, no responses could be observed despite there being manageable toxicity (ClinicalTrials.gov ID: NCT02008877) [216]. Similarly, in a study using Ganetespib and docetaxel in patients with advanced lung adenocarcinoma, a Phase II trial showed that treatment prolonged progression-free survival and overall survival [217], but the subsequent Phase III trial was terminated, likely because of the lack of supporting data from the interim analysis (ClinicalTrial.gov ID NCT01798485). Regardless, there are still on-going trials (e.g., ClinicalTrials.gov ID: NCT01560416) examining the possibilities of using Ganetespib with other therapeutic agents, as is the case with other HIF inhibitors.
Notably, recent trials have also been exploring the clinical value of HIF-2α inhibition. Promising results from the administration of Belzutifan, a specific oral HIF-2α inhibitor, to patients with VHL germline alteration-associated cancers prompted the approval of its use for the treatment of VHL-associated renal cell carcinoma, central nervous system hemangioblastomas, and pancreatic neuroendocrine tumors [218]. This encouraged trials with a broader application of Belzutifan to other solid tumors (ClinicalTrials.gov ID: NCT02974738), as well as the use of other HIF-2α inhibitors (e.g., ClinicalTrials.gov ID: NCT04895748, NCT05119335, NCT05935748) [219].
With continual clinical trials and studies alongside the enhancing quality of preclinical research, it is envisioned that the HIF inhibitor can be adopted synergistically with other drugs to sensitize chemotherapy treatment in the future.
On the other hand, factors other than HIFs have recently been found to contribute to cellular hypoxic responses as well, and we introduced how some of these HIF-independent mechanisms engender chemoresistance in hypoxic cancer cells. In addition, hypoxia-induced biological responses, like EMT and UPR, involving both HIF-dependent and HIF-independent pathways, have been shown to promote the acquisition of resistance to chemotherapeutic drugs due to the coinciding molecular bases. Although our understanding of their molecular mechanisms has gradually advanced, there remains a lot of unknown parts. The complicated crosstalk therein, along with the resulting context dependence, also adds difficulties to utilize them as therapeutic targets. Hence, basic studies to uncover relevant molecular details would be paramount.
We also briefly discussed the role of hypoxia-responsive non-coding RNAs in promoting chemoresistance. Whilst supportive evidence is accumulating, huge difficulties can still be expected to convert our current knowledge into clinical application, at least in the near future. First, the incomplete understanding of molecular mechanisms would increase the likelihood of unexpected off-target/side effects. Many miRNAs have multiple target mRNAs, and one mRNA may also be regulated by multiple miRNAs; similarly, one lncRNA or circRNA can also regulate multiple miRNAs [220]. Simplistically inhibiting the target RNA molecule may likely result in a disruption to the regulation of other biological processes. Also, most non-coding RNAs originate from the corresponding “parent” genes; targeting non-coding RNA may potentially affect the expressions of the parent genes and other genes containing homologous sequences. Moreover, unlike targeting proteins/enzymes, very little is currently known about the inhibition of specific RNA using small molecules. Although there have been attempts to apply oligonucleotides, sometimes in combination with nanoparticle carriers, most of them were carried out in in vitro studies and did not take oxygen status into account; their applicability to in vivo models or to patients, as well as to hypoxic cancer tissues, is yet to be investigated [221]. Potential obstacles, like unexpected immune responses, should also be considered, and it can be perceived that effective drug delivery, particularly to hypoxic cancer cells, and dosage control would require proper optimization as well. Undoubtedly, the involvement of hypoxia-responsive non-coding RNAs in drug resistance is increasingly being recognized; further studies surveying molecular details and how to utilize them as therapeutic targets in practice would warrant their emerging potential in overcoming chemoresistance in hypoxia.
Given the complexity of the molecular pathways altered under hypoxic conditions, other therapeutic agents have also been tested/developed in an attempt to annihilate hypoxic cancer cells from alternative approaches. For example, metformin and atovaquone are two inhibitors of the mitochondrial electron transport chain and are expected to alleviate hypoxia by reducing cellular oxygen consumption; trials evaluating their oxygenating effect in sensitizing radiotherapy and chemotherapy have been launched (ClinicalTrials.gov ID: NCT04275713, NCT02628080, NCT04648033). Moreover, there were also some hypoxia-activated prodrugs in trials, like CP-506 with carboplatin for triple-negative breast cancer and ovarian cancer (ClinicalTrials.gov ID: NCT04954599), and evofosfamide (TH-302) for recurrent bevacizumab (anti-VEGF antibody)-refractory glioblastoma [222].
Finally, in the latter part of this review, we focused on the contribution of hypoxia-dependent epigenetic regulation and reported a specific example wherein the results demonstrate that the therapeutic resistance of hypoxic cancer cells can be induced by cell cycle retardation because of the degradation of the ATAD2 protein in a 2-OGDD-dependent but HIF-independent manner. Based on this discovery, we opened the potential to develop new therapeutic strategies to overcome the chemotherapy resistance of hypoxic cancer cells by inhibiting ATAD2 degradation under hypoxic conditions and restoring cell cycle delay. While further efforts will be needed to uncover the detailed molecular mechanism and identify the exact 2-OGDD(s) involved, these results exemplify the importance of 2-OGDD proteins, apart from regulating the HIF pathway, in the hypoxic response, shedding light on their potential role in combating the chemotherapy resistance of hypoxic cancer cells.
Table 1. List of long non-coding RNAs contributing to chemoresistance under hypoxia in relation to HIF.
Table 1. List of long non-coding RNAs contributing to chemoresistance under hypoxia in relation to HIF.
NameTarget/EffectResistanceCancer Type, ModelRef.
HIF downstreamLINC03000-201 (lncMat2B, ENST00000486913.3)↓ ROS production
↓ DNA damage
CisplatinBreast cancer, in vitro[223]
PVT1miR-140-3p/ATG5
↓ autophagy
CisplatinLung cancer, in vitro and in vivo[60]
lncRNA-CBSLRYTHDF2/CBS
↓ ferroptosis
CisplatinGastric cancer, in vivo[224]
ANRILmiR-328/ABCG2, MDR1CisplatinRetinoblastoma, in vitro[225]
LUCAT1Interaction with PTBP1
↓ DNA damage
5-Fluorouracil, Camptothecin, Doxorubicin and OxaliplatinColorectal cancer, in vitro, in vivo, and patient cohorts[226]
NLUCAT1
(HIF-2α-dependent)
↓ ROS productionCisplatinLung adenocarcinoma, in vitro[227]
HIF upstreamPVT1miR-194-5p/HIF1A
↑ proliferation
CisplatinOral SCC, in vitro[228]
HIF1A-AS1Interaction with YB1
↑ HIF-1α (positive feedback)
↑ glycolysis
GemcitabinePancreatic cancer, in vitro, in vivo, and patient cohorts[229]
HIF1A-AS2↑ HIF-1α
↑ autophagy
DoxorubicinSmall cell lung cancer, in vitro[230]
NORADmiR-495-3p/HIF-1α
↑ vasculogenic mimicry
5-FluorouracilColorectal cancer, in vitro[231]
Potentially HIF-dependentHOTAIRmiR-1277-5p/ZEB1
↑ EMT
OxaliplatinColorectal cancer, in vitro and in vivo[232]
lncRNA-EMSmiR-758-3p/WTAPCisplatinEsophageal cancer, in vitro and in vivo[233]

Author Contributions

P.W.T.L. wrote the first draft and revised it; L.R.K. wrote the first draft; T.H. conducted the ATAD2 research and wrote the draft on the part; and H.H. and M.K. planned the overall concept, brushed it up, and revised the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by funds from the AMED-CREST (21gm1110010s0303, 22gm1110010s0304, 23gm1110010s0305) and the Promotion of Cancer Research and Therapeutic Evolution (P-PROMOTE: 22ama221417h0001, 23ama221417h0002) to H.H. and from the AMED-PRIME (22gm6710010h0001, 23gm6710010h0002, 23gm6710010h9902) to T.Y. from the Japan Agency for Medical Research and development (AMED); by funds from the Core-to-Core Program (JPJSCCA20200009) to H.H., KAKENHI (21KK0144, 23H02855, and 23K18274 to H.H., 21K07727 to M.K., and 20KK0339, 21H03597, 22H05585, and 23H04274 to T.Y.), and the Grant-in-Aid for JSPS Fellows (23KJ1316) to L.R.K. from the Japan Society for the Promotion of Science (JSPS); and by funds from the research grant programs of the Princess Takamatsu Cancer Research Fund, Uehara Memorial Foundation, Takeda Science Foundation, Ichiro Kanehara Foundation for the Promotion of Medical Sciences and Medical Care, Kobayashi Foundation for Cancer Research, Yasuda Medical Foundation, the Foundation for Promotion of Cancer Research, Suzuken Memorial Foundation, and Daiichi Sankyo Foundation of Life Science to H.H. and from the Takeda Science Foundation, Sumitomo Foundation, and Asian Young Scientist Fellowship to T.Y. This study was conducted through the CORE Programs of the Radiation Biology Center, Kyoto University and the Joint Usage Program of the Institute for Integrated Radiation and Nuclear Science, Kyoto University. P.W.T.L. was supported by the Japanese Government (MEXT) scholarship program. L.R.K. was supported by JSPS as a DC1 fellow (23KJ1316).

Conflicts of Interest

The authors declare no competing interests.

References

  1. Yeom, C.J.; Goto, Y.; Zhu, Y.; Hiraoka, M.; Harada, H. Microenvironments and cellular characteristics in the micro tumor cords of malignant solid tumors. Int. J. Mol. Sci. 2012, 13, 13949–13965. [Google Scholar] [CrossRef]
  2. Walsh, J.C.; Lebedev, A.; Aten, E.; Madsen, K.; Marciano, L.; Kolb, H.C. The clinical importance of assessing tumor hypoxia: Relationship of tumor hypoxia to prognosis and therapeutic opportunities. Antioxid. Redox Signal. 2014, 21, 1516–1554. [Google Scholar] [CrossRef]
  3. Osinsky, S.; Zavelevich, M.; Vaupel, P. Tumor hypoxia and malignant progression. Exp. Oncol. 2009, 31, 80–86. [Google Scholar]
  4. Haitani, T.; Kobayashi, M.; Koyasu, S.; Akamatsu, S.; Suwa, T.; Onodera, Y.; Nam, J.-M.; Nguyen, P.T.L.; Menju, T.; Date, H.; et al. Proteolysis of a histone acetyl reader, atad2, induces chemoresistance of cancer cells under severe hypoxia by inhibiting cell cycle progression in s phase. Cancer Lett. 2022, 528, 76–84. [Google Scholar] [CrossRef]
  5. Shirai, Y.; Chow, C.C.T.; Kambe, G.; Suwa, T.; Kobayashi, M.; Takahashi, I.; Harada, H.; Nam, J.-M. An overview of the recent development of anticancer agents targeting the hif-1 transcription factor. Cancers 2021, 13, 2813. [Google Scholar] [CrossRef]
  6. Yoshimura, M.; Itasaka, S.; Harada, H.; Hiraoka, M. Microenvironment and radiation therapy. Biomed. Res. Int. 2013, 2013, 685308. [Google Scholar] [CrossRef]
  7. Mellor, H.R.; Callaghan, R. Resistance to chemotherapy in cancer: A complex and integrated cellular response. Pharmacology 2008, 81, 275–300. [Google Scholar] [CrossRef]
  8. Yeldag, G.; Rice, A.; Del Río Hernández, A. Chemoresistance and the self-maintaining tumor microenvironment. Cancers 2018, 10, 471. [Google Scholar] [CrossRef]
  9. Weniger, M.; Honselmann, K.C.; Liss, A.S. The extracellular matrix and pancreatic cancer: A complex relationship. Cancers 2018, 10, 316. [Google Scholar] [CrossRef] [PubMed]
  10. Masamune, A.; Kikuta, K.; Watanabe, T.; Satoh, K.; Hirota, M.; Shimosegawa, T. Hypoxia stimulates pancreatic stellate cells to induce fibrosis and angiogenesis in pancreatic cancer. Am. J. Physiol. Gastrointest. Liver Physiol. 2008, 295, G709–G717. [Google Scholar] [CrossRef] [PubMed]
  11. Goda, N.; Ryan, H.E.; Khadivi, B.; McNulty, W.; Rickert, R.C.; Johnson, R.S. Hypoxia-inducible factor 1alpha is essential for cell cycle arrest during hypoxia. Mol. Cell. Biol. 2003, 23, 359–369. [Google Scholar] [CrossRef]
  12. Harada, H. How can we overcome tumor hypoxia in radiation therapy? J. Radiat. Res. 2011, 52, 545–556. [Google Scholar] [CrossRef]
  13. Kizaka-Kondoh, S.; Inoue, M.; Harada, H.; Hiraoka, M. Tumor hypoxia: A target for selective cancer therapy. Cancer Sci. 2003, 94, 1021–1028. [Google Scholar] [CrossRef]
  14. Kizaka-Kondoh, S.; Tanaka, S.; Harada, H.; Hiraoka, M. The hif-1-active microenvironment: An environmental target for cancer therapy. Adv. Drug Deliv. Rev. 2009, 61, 623–632. [Google Scholar] [CrossRef] [PubMed]
  15. Brown, J.M.; Wilson, W.R. Exploiting tumour hypoxia in cancer treatment. Nat. Rev. Cancer 2004, 4, 437–447. [Google Scholar] [CrossRef]
  16. Begg, K.; Tavassoli, M. Inside the hypoxic tumour: Reprogramming of the ddr and radioresistance. Cell Death Discov. 2020, 6, 77. [Google Scholar] [CrossRef] [PubMed]
  17. Wang, G.L.; Semenza, G.L. Purification and characterization of hypoxia-inducible factor 1. J. Biol. Chem. 1995, 270, 1230–1237. [Google Scholar] [CrossRef]
  18. Hirota, K.; Semenza, G.L. Regulation of hypoxia-inducible factor 1 by prolyl and asparaginyl hydroxylases. Biochem. Biophys. Res. Commun. 2005, 338, 610–616. [Google Scholar] [CrossRef] [PubMed]
  19. Epstein, A.C.; Gleadle, J.M.; McNeill, L.A.; Hewitson, K.S.; O’Rourke, J.; Mole, D.R.; Mukherji, M.; Metzen, E.; Wilson, M.I.; Dhanda, A.; et al. C. Elegans egl-9 and mammalian homologs define a family of dioxygenases that regulate hif by prolyl hydroxylation. Cell 2001, 107, 43–54. [Google Scholar] [CrossRef]
  20. Ivan, M.; Kondo, K.; Yang, H.; Kim, W.; Valiando, J.; Ohh, M.; Salic, A.; Asara, J.M.; Lane, W.S.; Kaelin, W.G., Jr. Hifalpha targeted for vhl-mediated destruction by proline hydroxylation: Implications for o2 sensing. Science 2001, 292, 464–468. [Google Scholar] [CrossRef]
  21. Jaakkola, P.; Mole, D.R.; Tian, Y.-M.; Wilson, M.I.; Gielbert, J.; Gaskell, S.J.; von Kriegsheim, A.; Hebestreit, H.F.; Mukherji, M.; Schofield, C.J.; et al. Targeting of hif-alpha to the von hippel-lindau ubiquitylation complex by o2-regulated prolyl hydroxylation. Science 2001, 292, 468–472. [Google Scholar] [CrossRef]
  22. Rosenberg, N.; Gervais, P. Evaluation of the sequelae of occupational asthma. Rev. Mal. Respir. 1989, 6, 35–38. [Google Scholar]
  23. Kaelin, W.G., Jr. The von hippel-lindau tumour suppressor protein: O2 sensing and cancer. Nat. Rev. Cancer 2008, 8, 865–873. [Google Scholar] [CrossRef]
  24. Tanimoto, K.; Makino, Y.; Pereira, T.; Poellinger, L. Mechanism of regulation of the hypoxia-inducible factor-1 alpha by the von hippel-lindau tumor suppressor protein. EMBO J. 2000, 19, 4298–4309. [Google Scholar] [CrossRef]
  25. Ohh, M.; Park, C.W.; Ivan, M.; Hoffman, M.A.; Kim, T.-Y.; Huang, L.E.; Pavletich, N.; Chau, V.; Kaelin, W.G. Ubiquitination of hypoxia-inducible factor requires direct binding to the beta-domain of the von hippel-lindau protein. Nat. Cell Biol. 2000, 2, 423–427. [Google Scholar] [CrossRef]
  26. Maxwell, P.H.; Wiesener, M.S.; Chang, G.-W.; Clifford, S.C.; Vaux, E.C.; Cockman, M.E.; Wykoff, C.C.; Pugh, C.W.; Maher, E.R.; Ratcliffe, P.J. The tumour suppressor protein vhl targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature 1999, 399, 271–275. [Google Scholar] [CrossRef]
  27. Arany, Z.; Huang, L.E.; Eckner, R.; Bhattacharya, S.; Jiang, C.; Goldberg, M.A.; Bunn, H.F.; Livingston, D.M. An essential role for p300/cbp in the cellular response to hypoxia. Proc. Natl. Acad. Sci. USA 1996, 93, 12969–12973. [Google Scholar] [CrossRef]
  28. Lando, D.; Peet, D.J.; Whelan, D.A.; Gorman, J.J.; Whitelaw, M.L. Asparagine hydroxylation of the hif transactivation domain a hypoxic switch. Science 2002, 295, 858–861. [Google Scholar] [CrossRef]
  29. Mahon, P.C.; Hirota, K.; Semenza, G.L. Fih-1: A novel protein that interacts with hif-1alpha and vhl to mediate repression of hif-1 transcriptional activity. Genes Dev. 2001, 15, 2675–2686. [Google Scholar] [CrossRef]
  30. Koyasu, S.; Kobayashi, M.; Goto, Y.; Hiraoka, M.; Harada, H. Regulatory mechanisms of hypoxia-inducible factor 1 activity: Two decades of knowledge. Cancer Sci. 2018, 109, 560–571. [Google Scholar] [CrossRef]
  31. Yeom, C.J.; Zeng, L.; Goto, Y.; Morinibu, A.; Zhu, Y.; Shinomiya, K.; Kobayashi, M.; Itasaka, S.; Yoshimura, M.; Hur, C.-G.; et al. Ly6e: A conductor of malignant tumor growth through modulation of the pten/pi3k/akt/hif-1 axis. Oncotarget 2016, 7, 65837–65848. [Google Scholar] [CrossRef] [PubMed]
  32. Zeng, L.; Morinibu, A.; Kobayashi, M.; Zhu, Y.; Wang, X.; Goto, Y.; Yeom, C.J.; Zhao, T.; Hirota, K.; Shinomiya, K.; et al. Aberrant idh3alpha expression promotes malignant tumor growth by inducing hif-1-mediated metabolic reprogramming and angiogenesis. Oncogene 2015, 34, 4758–4766. [Google Scholar] [CrossRef]
  33. Harada, H.; Kizaka-Kondoh, S.; Li, G.; Itasaka, S.; Shibuya, K.; Inoue, M.; Hiraoka, M. Significance of hif-1-active cells in angiogenesis and radioresistance. Oncogene 2007, 26, 7508–7516. [Google Scholar] [CrossRef]
  34. Liu, Y.; Cox, S.R.; Morita, T.; Kourembanas, S. Hypoxia regulates vascular endothelial growth factor gene expression in endothelial cells. Identification of a 5′ enhancer. Circ. Res. 1995, 77, 638–643. [Google Scholar] [CrossRef] [PubMed]
  35. Goto, Y.; Zeng, L.; Yeom, C.J.; Zhu, Y.; Morinibu, A.; Shinomiya, K.; Kobayashi, M.; Hirota, K.; Itasaka, S.; Yoshimura, M.; et al. Uchl1 provides diagnostic and antimetastatic strategies due to its deubiquitinating effect on hif-1alpha. Nat. Commun. 2015, 6, 6153. [Google Scholar] [CrossRef]
  36. Semenza, G.L. Molecular mechanisms mediating metastasis of hypoxic breast cancer cells. Trends Mol. Med. 2012, 18, 534–543. [Google Scholar] [CrossRef] [PubMed]
  37. Zhong, H.; De Marzo, A.M.; Laughner, E.; Lim, M.; Hilton, D.A.; Zagzag, D.; Buechler, P.; Isaacs, W.B.; Semenza, G.L.; Simons, J.W. Overexpression of hypoxia-inducible factor 1alpha in common human cancers and their metastases. Cancer Res. 1999, 59, 5830–5835. [Google Scholar]
  38. Semenza, G.L. Mitochondrial autophagy: Life and breath of the cell. Autophagy 2008, 4, 534–536. [Google Scholar] [CrossRef] [PubMed]
  39. Comerford, K.M.; Wallace, T.J.; Karhausen, J.; Louis, N.A.; Montalto, M.C.; Colgan, S.P. Hypoxia-inducible factor-1-dependent regulation of the multidrug resistance (mdr1) gene. Cancer Res. 2002, 62, 3387–3394. [Google Scholar]
  40. He, M.; Wu, H.; Jiang, Q.; Liu, Y.; Han, L.; Yan, Y.; Wei, B.; Liu, F.; Deng, X.; Chen, H.; et al. Hypoxia-inducible factor-2alpha directly promotes bcrp expression and mediates the resistance of ovarian cancer stem cells to adriamycin. Mol. Oncol. 2019, 13, 403–421. [Google Scholar] [CrossRef]
  41. Pinzón-Daza, M.L.; Cuellar-Saenz, Y.; Nualart, F.; Ondo-Mendez, A.; Del Riesgo, L.; Castillo-Rivera, F.; Garzón, R. Oxidative stress promotes doxorubicin-induced pgp and bcrp expression in colon cancer cells under hypoxic conditions. J. Cell. Biochem. 2017, 118, 1868–1878. [Google Scholar] [CrossRef]
  42. Wang, H.; Wu, X.; Hudkins, K.; Mikheev, A.; Zhang, H.; Gupta, A.; Unadkat, J.D.; Mao, Q. Expression of the breast cancer resistance protein (bcrp1/abcg2) in tissues from pregnant mice: Effects of pregnancy and correlations with nuclear receptors. Am. J. Physiol. Endocrinol. Metab. 2006, 291, E1295–E1304. [Google Scholar] [CrossRef] [PubMed]
  43. Hyun, J.-Y.; Chun, Y.-S.; Kim, T.-Y.; Kim, H.-L.; Kim, M.-S.; Park, J.-W. Hypoxia-inducible factor 1alpha-mediated resistance to phenolic anticancer. Chemotherapy 2004, 50, 119–126. [Google Scholar] [CrossRef]
  44. Chen, J.; Ding, Z.; Peng, Y.; Pan, F.; Li, J.; Zou, L.; Zhang, Y.; Liang, H. Hif-1alpha inhibition reverses multidrug resistance in colon cancer cells via downregulation of mdr1/p-glycoprotein. PLoS ONE 2014, 9, e98882. [Google Scholar] [CrossRef] [PubMed]
  45. Wu, X.-Y.; Fu, Z.-X.; Wang, X.-H. Effect of hypoxia-inducible factor 1-alpha on survivin in colorectal cancer. Mol. Med. Rep. 2010, 3, 409–415. [Google Scholar] [CrossRef] [PubMed]
  46. Sun, X.; Dong, X.; Lin, L.; Jiang, X.; Wei, Z.; Zhai, B.; Sun, B.; Zhang, Q.; Wang, X.; Jiang, H.; et al. Up-regulation of survivin by akt and hypoxia-inducible factor 1alpha contributes to cisplatin resistance in gastric cancer. FEBS J. 2014, 281, 115–128. [Google Scholar] [CrossRef]
  47. Li, H.; Sun, X.; Li, J.; Liu, W.; Pan, G.; Mao, A.; Liu, J.; Zhang, Q.; Rao, L.; Xie, X.; et al. Hypoxia induces docetaxel resistance in triple-negative breast cancer via the hif-1alpha/mir-494/survivin signaling pathway. Neoplasia 2022, 32, 100821. [Google Scholar] [CrossRef] [PubMed]
  48. Zichittella, C.; Barreca, M.M.; Cordaro, A.; Corrado, C.; Alessandro, R.; Conigliaro, A. Mir-675-5p supports hypoxia-induced drug resistance in colorectal cancer cells. BMC Cancer 2022, 22, 567. [Google Scholar] [CrossRef]
  49. Feng, L.; Shen, F.; Zhou, J.; Li, Y.; Jiang, R.; Chen, Y. Hypoxia-induced up-regulation of mir-27a promotes paclitaxel resistance in ovarian cancer. Biosci. Rep. 2020, 40, BSR20192457. [Google Scholar] [CrossRef]
  50. Xu, K.; Zhan, Y.; Yuan, Z.; Qiu, Y.; Wang, H.; Fan, G.; Wang, J.; Li, W.; Cao, Y.; Shen, X.; et al. Hypoxia induces drug resistance in colorectal cancer through the hif-1alpha/mir-338-5p/il-6 feedback loop. Mol. Ther. 2019, 27, 1810–1824. [Google Scholar] [CrossRef]
  51. Lu, H.; Samanta, D.; Xiang, L.; Zhang, H.; Hu, H.; Chen, I.; Bullen, J.W.; Semenza, G.L. Chemotherapy triggers hif-1-dependent glutathione synthesis and copper chelation that induces the breast cancer stem cell phenotype. Proc. Natl. Acad. Sci. USA 2015, 112, E4600–E4609. [Google Scholar] [CrossRef] [PubMed]
  52. Roncuzzi, L.; Pancotti, F.; Baldini, N. Involvement of hif-1alpha activation in the doxorubicin resistance of human osteosarcoma cells. Oncol. Rep. 2014, 32, 389–394. [Google Scholar] [CrossRef] [PubMed]
  53. Wen, W.; Ding, J.; Sun, W.; Wu, K.; Ning, B.; Gong, W.; He, G.; Huang, S.; Ding, X.; Yin, P.; et al. Suppression of cyclin d1 by hypoxia-inducible factor-1 via direct mechanism inhibits the proliferation and 5-fluorouracil-induced apoptosis of a549 cells. Cancer Res. 2010, 70, 2010–2019. [Google Scholar] [CrossRef] [PubMed]
  54. Sullivan, R.; Paré, G.C.; Frederiksen, L.J.; Semenza, G.L.; Graham, C.H. Hypoxia-induced resistance to anticancer drugs is associated with decreased senescence and requires hypoxia-inducible factor-1 activity. Mol. Cancer Ther. 2008, 7, 1961–1973. [Google Scholar] [CrossRef]
  55. Johnson, A.B.; Denko, N.; Barton, M.C. Hypoxia induces a novel signature of chromatin modifications and global repression of transcription. Mutat. Res. 2008, 640, 174–179. [Google Scholar] [CrossRef] [PubMed]
  56. Liu, Y.-F.; Luo, D.; Li, X.M.; Li, Z.-Q.; Yu, X.; Zhu, H.-W. Pvt1 knockdown inhibits autophagy and improves gemcitabine sensitivity by regulating the mir-143/hif-1alpha/vmp1 axis in pancreatic cancer. Pancreas 2021, 50, 227–234. [Google Scholar] [CrossRef]
  57. Wu, H.-M.; Jiang, Z.-F.; Ding, P.-S.; Shao, L.-J.; Liu, R.-Y. Hypoxia-induced autophagy mediates cisplatin resistance in lung cancer cells. Sci. Rep. 2015, 5, 12291. [Google Scholar] [CrossRef]
  58. Liang, C.; Dong, Z.; Cai, X.; Shen, J.; Xu, Y.; Zhang, M.; Li, H.; Yu, W.; Chen, W. Hypoxia induces sorafenib resistance mediated by autophagy via activating foxo3a in hepatocellular carcinoma. Cell Death Dis. 2020, 11, 1017. [Google Scholar] [CrossRef]
  59. Yang, X.; Yin, H.; Zhang, Y.; Li, X.; Tong, H.; Zeng, Y.; Wang, Q.; He, W. Hypoxia-induced autophagy promotes gemcitabine resistance in human bladder cancer cells through hypoxia-inducible factor 1alpha activation. Int. J. Oncol. 2018, 53, 215–224. [Google Scholar] [CrossRef]
  60. Wang, J.; Dong, Z.; Sheng, Z.; Cai, Y. Hypoxia-induced pvt1 promotes lung cancer chemoresistance to cisplatin by autophagy via pvt1/mir-140-3p/atg5 axis. Cell Death Discov. 2022, 8, 104. [Google Scholar] [CrossRef]
  61. Huang, S.; Qi, P.; Zhang, T.; Li, F.; He, X. The hif-1alpha/mir-224-3p/atg5 axis affects cell mobility and chemosensitivity by regulating hypoxia-induced protective autophagy in glioblastoma and astrocytoma. Oncol. Rep. 2019, 41, 1759–1768. [Google Scholar] [CrossRef] [PubMed]
  62. Ma, J.; Weng, L.; Jia, Y.; Liu, B.; Wu, S.; Xue, L.; Yin, X.; Mao, A.; Wang, Z.; Shang, M. Ptbp3 promotes malignancy and hypoxia-induced chemoresistance in pancreatic cancer cells by atg12 up-regulation. J. Cell. Mol. Med. 2020, 24, 2917–2930. [Google Scholar] [CrossRef] [PubMed]
  63. Chavez-Dominguez, R.; Perez-Medina, M.; Lopez-Gonzalez, J.S.; Galicia-Velasco, M.; Aguilar-Cazares, D. The double-edge sword of autophagy in cancer: From tumor suppression to pro-tumor activity. Front. Oncol. 2020, 10, 578418. [Google Scholar] [CrossRef]
  64. Sannigrahi, M.; Singh, V.; Sharma, R.; Panda, N.; Khullar, M. Role of autophagy in head and neck cancer and therapeutic resistance. Oral Dis. 2015, 21, 283–291. [Google Scholar] [CrossRef]
  65. Sullivan, R.; Graham, C.H. Hypoxia prevents etoposide-induced DNA damage in cancer cells through a mechanism involving hypoxia-inducible factor 1. Mol. Cancer Ther. 2009, 8, 1702–1713. [Google Scholar] [CrossRef]
  66. Yan, Y.; He, M.; Zhao, L.; Wu, H.; Zhao, Y.; Han, L.; Wei, B.; Ye, D.; Lv, X.; Wang, Y.; et al. A novel hif-2alpha targeted inhibitor suppresses hypoxia-induced breast cancer stemness via sod2-mtros-pdi/gpr78-upr(er) axis. Cell Death Differ. 2022, 29, 1769–1789. [Google Scholar] [CrossRef]
  67. Yan, Y.; Liu, F.; Han, L.; Zhao, L.; Chen, J.; I Olopade, O.; He, M.; Wei, M. Hif-2alpha 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, 256. [Google Scholar] [CrossRef] [PubMed]
  68. Zhang, G.; Tao, X.; Ji, B.; Gong, J. Hypoxia-driven m2-polarized macrophages facilitate cancer aggressiveness and temozolomide resistance in glioblastoma. Oxidative Med. Cell. Longev. 2022, 2022, 1614336. [Google Scholar] [CrossRef]
  69. Zheng, H.; Yu, S.; Zhu, C.; Guo, T.; Liu, F.; Xu, Y. Hif1alpha promotes tumor chemoresistance via recruiting gdf15-producing tams in colorectal cancer. Exp. Cell Res. 2021, 398, 112394. [Google Scholar] [CrossRef]
  70. Yu, S.; Li, Q.; Yu, Y.; Cui, Y.; Li, W.; Liu, T.; Liu, F. Activated hif1alpha of tumor cells promotes chemoresistance development via recruiting gdf15-producing tumor-associated macrophages in gastric cancer. Cancer Immunol. Immunother. 2020, 69, 1973–1987. [Google Scholar] [CrossRef]
  71. Eptaminitaki, G.C.; Stellas, D.; Bonavida, B.; Baritaki, S. Long non-coding rnas (lncrnas) signaling in cancer chemoresistance: From prediction to druggability. Drug Resist. Updat. 2022, 65, 100866. [Google Scholar] [CrossRef] [PubMed]
  72. Liu, X.-Y.; Zhang, Q.; Guo, J.; Zhang, P.; Liu, H.; Tian, Z.-B.; Zhang, C.-P.; Li, X.-Y. The role of circular rnas in the drug resistance of cancers. Front. Oncol. 2021, 11, 790589. [Google Scholar] [CrossRef] [PubMed]
  73. Papatsirou, M.; I Artemaki, P.; Scorilas, A.; Kontos, C.K. The role of circular rnas in therapy resistance of patients with solid tumors. Pers. Med. 2020, 17, 469–490. [Google Scholar] [CrossRef] [PubMed]
  74. Qu, Y.; Tan, H.-Y.; Chan, Y.-T.; Jiang, H.; Wang, N.; Wang, D. The functional role of long noncoding rna in resistance to anticancer treatment. Ther. Adv. Med. Oncol. 2020, 12, 1758835920927850. [Google Scholar] [CrossRef]
  75. Cui, C.; Yang, J.; Li, X.; Liu, D.; Fu, L.; Wang, X. Functions and mechanisms of circular rnas in cancer radiotherapy and chemotherapy resistance. Mol. Cancer 2020, 19, 58. [Google Scholar] [CrossRef] [PubMed]
  76. Pucci, P.; Rescigno, P.; Sumanasuriya, S.; de Bono, J.; Crea, F. Hypoxia and noncoding rnas in taxane resistance. Trends Pharmacol. Sci. 2018, 39, 695–709. [Google Scholar] [CrossRef] [PubMed]
  77. Ebrahimi, A.; Bakhshaei Shahrebabaki, P.; Fouladi, H.; Mansoori Derakhshan, S. The impact of micrornas on the resistance of breast cancer subtypes to chemotherapy. Pathol. Res. Pract. 2023, 249, 154702. [Google Scholar] [CrossRef] [PubMed]
  78. Funamizu, N.; Honjo, M.; Tamura, K.; Sakamoto, K.; Ogawa, K.; Takada, Y. Micrornas associated with gemcitabine resistance via emt, tme, and drug metabolism in pancreatic cancer. Cancers 2023, 15, 1230. [Google Scholar] [CrossRef]
  79. Doghish, A.S.; Ismail, A.; Elrebehy, M.A.; Elbadry, A.M.; Mahmoud, H.H.; Farouk, S.M.; Abu Serea, G.A.; Elghany, R.A.A.; El-Halwany, K.K.; Alsawah, A.O.; et al. A study of mirnas as cornerstone in lung cancer pathogenesis and therapeutic resistance: A focus on signaling pathways interplay. Pathol. Res. Pract. 2022, 237, 154053. [Google Scholar] [CrossRef]
  80. Karami Fath, M.; Azargoonjahromi, A.; Kiani, A.; Jalalifar, F.; Osati, P.; Akbari Oryani, M.; Shakeri, F.; Nasirzadeh, F.; Khalesi, B.; Nabi-Afjadi, M.; et al. The role of epigenetic modifications in drug resistance and treatment of breast cancer. Cell. Mol. Biol. Lett. 2022, 27, 52. [Google Scholar] [CrossRef]
  81. Zeng, Z.; Chen, Y.; Geng, X.; Zhang, Y.; Wen, X.; Yan, Q.; Wang, T.; Ling, C.; Xu, Y.; Duan, J.; et al. Ncrnas: Multi-angle participation in the regulation of glioma chemotherapy resistance (review). Int. J. Oncol. 2022, 60, 76. [Google Scholar] [CrossRef]
  82. Ashrafizadeh, M.; Zarrabi, A.; Hushmandi, K.; Hashemi, F.; Moghadam, E.R.; Owrang, M.; Hashemi, F.; Makvandi, P.; Goharrizi, M.A.S.B.; Najafi, M.; et al. Lung cancer cells and their sensitivity/resistance to cisplatin chemotherapy: Role of micrornas and upstream mediators. Cell. Signal. 2021, 78, 109871. [Google Scholar] [CrossRef]
  83. Lin, Z.; Lu, S.; Xie, X.; Yi, X.; Huang, H. Noncoding rnas in drug-resistant pancreatic cancer: A review. Biomed. Pharmacother. 2020, 131, 110768. [Google Scholar] [CrossRef]
  84. Yete, S.; Saranath, D. Micrornas in oral cancer: Biomarkers with clinical potential. Oral Oncol. 2020, 110, 105002. [Google Scholar] [CrossRef]
  85. Ashrafizadeh, M.; Zarrabi, A.; Hushmandi, K.; Hashemi, F.; Hashemi, F.; Samarghandian, S.; Najafi, M. Micrornas in cancer therapy: Their involvement in oxaliplatin sensitivity/resistance of cancer cells with a focus on colorectal cancer. Life Sci. 2020, 256, 117973. [Google Scholar] [CrossRef]
  86. Si, W.; Shen, J.; Zheng, H.; Fan, W. The role and mechanisms of action of micrornas in cancer drug resistance. Clin. Epigenetics 2019, 11, 25. [Google Scholar] [CrossRef] [PubMed]
  87. Garrido-Cano, I.; Pattanayak, B.; Adam-Artigues, A.; Lameirinhas, A.; Torres-Ruiz, S.; Tormo, E.; Cervera, R.; Eroles, P. Micrornas as a clue to overcome breast cancer treatment resistance. Cancer Metastasis Rev. 2022, 41, 77–105. [Google Scholar] [CrossRef] [PubMed]
  88. Dong, B.; Li, S.; Zhu, S.; Yi, M.; Luo, S.; Wu, K. Mirna-mediated emt and cscs in cancer chemoresistance. Exp. Hematol. Oncol. 2021, 10, 12. [Google Scholar] [CrossRef]
  89. Pan, G.; Liu, Y.; Shang, L.; Zhou, F.; Yang, S. Emt-associated micrornas and their roles in cancer stemness and drug resistance. Cancer Commun. 2021, 41, 199–217. [Google Scholar] [CrossRef] [PubMed]
  90. Jing, Y.; Liang, W.; Liu, J.; Zhang, L.; Wei, J.; Yang, J.; Zhang, Y.; Huang, Z. Autophagy-mediating micrornas in cancer chemoresistance. Cell Biol. Toxicol. 2020, 36, 517–536. [Google Scholar] [CrossRef]
  91. Zeng, X.; Wang, H.-Y.; Bai, S.-Y.; Pu, K.; Wang, Y.-P.; Zhou, Y.-N. The roles of micrornas in multidrug-resistance mechanisms in gastric cancer. Curr. Mol. Med. 2020, 20, 667–674. [Google Scholar] [CrossRef] [PubMed]
  92. Liu, H.; Chen, C.; Zeng, J.; Zhao, Z.; Hu, Q. Microrna-210-3p is transcriptionally upregulated by hypoxia induction and thus promoting emt and chemoresistance in glioma cells. PLoS ONE 2021, 16, e0253522. [Google Scholar] [CrossRef] [PubMed]
  93. Jawad, S.F.; Altalbawy, F.M.A.; Hussein, R.M.; Fadhil, A.A.; Jawad, M.A.; Zabibah, R.S.; Taraki, T.Y.; Mohan, C.D.; Rangappa, K.S. The strict regulation of hif-1alpha by non-coding rnas: New insight towards proliferation, metastasis, and therapeutic resistance strategies. Cancer Metastasis Rev. 2024, 43, 5–27. [Google Scholar] [CrossRef] [PubMed]
  94. Xia, M.; Sheng, L.; Qu, W.; Xue, X.; Chen, H.; Zheng, G.; Chen, W. Mir-194-5p enhances the sensitivity of nonsmall-cell lung cancer to doxorubicin through targeted inhibition of hypoxia-inducible factor-1. World J. Surg. Oncol. 2021, 19, 174. [Google Scholar] [CrossRef] [PubMed]
  95. Luo, G.; Xia, X.; Wang, X.; Zhang, K.; Cao, J.; Jiang, T.; Zhao, Q.; Qiu, Z. Mir-301a plays a pivotal role in hypoxia-induced gemcitabine resistance in pancreatic cancer. Exp. Cell Res. 2018, 369, 120–128. [Google Scholar] [CrossRef] [PubMed]
  96. Jasim, S.A.; Majeed, A.A.; Uinarni, H.; Alshuhri, M.; Alzahrani, A.A.; Ibrahim, A.A.; Alawadi, A.; Abed Al-Abadi, N.K.; Mustafa, Y.F.; Ahmed, B.A. Long non-coding rna (lncrna) pvt1 in drug resistance of cancers: Focus on pathological mechanisms. Pathol. Res. Pract. 2024, 254, 155119. [Google Scholar] [CrossRef] [PubMed]
  97. Chen, H.; Zhang, M.; Deng, Y. Long noncoding rnas in taxane resistance of breast cancer. Int. J. Mol. Sci. 2023, 24, 12253. [Google Scholar] [CrossRef] [PubMed]
  98. Hou, J.; Zhang, G.; Wang, X.; Wang, Y.; Wang, K. Functions and mechanisms of lncrna malat1 in cancer chemotherapy resistance. Biomark. Res. 2023, 11, 23. [Google Scholar] [CrossRef]
  99. Zhu, C.; Wang, X.; Wang, Y.; Wang, K. Functions and underlying mechanisms of lncrna hotair in cancer chemotherapy resistance. Cell Death Discov. 2022, 8, 383. [Google Scholar] [CrossRef]
  100. Yao, W.; Li, S.; Liu, R.; Jiang, M.; Gao, L.; Lu, Y.; Liang, X.; Zhang, H. Long non-coding rna pvt1: A promising chemotherapy and radiotherapy sensitizer. Front. Oncol. 2022, 12, 959208. [Google Scholar] [CrossRef]
  101. Ashrafizaveh, S.; Ashrafizadeh, M.; Zarrabi, A.; Husmandi, K.; Zabolian, A.; Shahinozzaman; Aref, A.R.; Hamblin, M.R.; Nabavi, N.; Crea, F.; et al. Long non-coding rnas in the doxorubicin resistance of cancer cells. Cancer Lett. 2021, 508, 104–114. [Google Scholar] [CrossRef]
  102. Taheri, M.; Shoorei, H.; Anamag, F.T.; Ghafouri-Fard, S.; Dinger, M.E. Lncrnas and mirnas participate in determination of sensitivity of cancer cells to cisplatin. Exp. Mol. Pathol. 2021, 123, 104602. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, H.; Guan, Z.; He, K.; Qian, J.; Cao, J.; Teng, L. Lncrna uca1 in anti-cancer drug resistance. Oncotarget 2017, 8, 64638–64650. [Google Scholar] [CrossRef]
  104. Jiao, B.; Liu, S.; Zhao, H.; Zhuang, Y.; Ma, S.; Lin, C.; Hu, J.; Liu, X. Hypoxia-responsive circrnas: A novel but important participant in non-coding rnas ushered toward tumor hypoxia. Cell Death Dis. 2022, 13, 666. [Google Scholar] [CrossRef]
  105. Wang, X.; Zhang, J.; Cao, G.; Hua, J.; Shan, G.; Lin, W. Emerging roles of circular rnas in gastric cancer metastasis and drug resistance. J. Exp. Clin. Cancer Res. 2022, 41, 218. [Google Scholar] [CrossRef]
  106. Mu, Q.; Lv, Y.; Luo, C.; Liu, X.; Huang, C.; Xiu, Y.; Tang, L. Research progress on the functions and mechanism of circrna in cisplatin resistance in tumors. Front. Pharmacol. 2021, 12, 709324. [Google Scholar] [CrossRef]
  107. Ameli-Mojarad, M.; Ameli-Mojarad, M.; Hadizadeh, M.; Young, C.; Babini, H.; Nazemalhosseini-Mojarad, E.; Bonab, M.A. The effective function of circular rna in colorectal cancer. Cancer Cell Int. 2021, 21, 496. [Google Scholar] [CrossRef] [PubMed]
  108. Xu, G.; Li, M.; Wu, J.; Qin, C.; Tao, Y.; He, H. Circular rna circnrip1 sponges microrna-138-5p to maintain hypoxia-induced resistance to 5-fluorouracil through hif-1alpha-dependent glucose metabolism in gastric carcinoma. Cancer Manag. Res. 2020, 12, 2789–2802. [Google Scholar] [CrossRef]
  109. Xu, M.; Liu, X.; Zhou, X.; Qin, Y.; Yang, L.; Wen, S.; Qiu, Y.; Chen, S.; Tang, R.; Guo, Y.; et al. Hypoxia-induced circstt3a enhances serine synthesis and promotes h3k4me3 modification to facilitate breast cancer stem cell formation. Pharmacol. Res. 2023, 197, 106964. [Google Scholar] [CrossRef]
  110. Zeng, Z.; Zhao, Y.; Chen, Q.; Zhu, S.; Niu, Y.; Ye, Z.; Hu, P.; Chen, D.; Xu, P.; Chen, J.; et al. Hypoxic exosomal hif-1alpha-stabilizing circznf91 promotes chemoresistance of normoxic pancreatic cancer cells via enhancing glycolysis. Oncogene 2021, 40, 5505–5517. [Google Scholar] [CrossRef]
  111. Deng, K.; Zou, F.; Xu, J.; Xu, D.; Luo, Z. Cancer-associated fibroblasts promote stemness maintenance and gemcitabine resistance via hif-1alpha/mir-21 axis under hypoxic conditions in pancreatic cancer. Mol. Carcinog. 2024, 63, 524–537. [Google Scholar] [CrossRef] [PubMed]
  112. Tam, S.Y.; Wu, V.W.C.; Law, H.K.W. Hypoxia-induced epithelial-mesenchymal transition in cancers: Hif-1alpha and beyond. Front. Oncol. 2020, 10, 486. [Google Scholar] [CrossRef] [PubMed]
  113. Zhang, Q.; Lou, Y.; Zhang, J.; Fu, Q.; Wei, T.; Sun, X.; Chen, Q.; Yang, J.; Bai, X.; Liang, T. Hypoxia-inducible factor-2alpha promotes tumor progression and has crosstalk with wnt/beta-catenin signaling in pancreatic cancer. Mol. Cancer 2017, 16, 119. [Google Scholar] [CrossRef] [PubMed]
  114. Kohnoh, T.; Hashimoto, N.; Ando, A.; Sakamoto, K.; Miyazaki, S.; Aoyama, D.; Kusunose, M.; Kimura, M.; Omote, N.; Imaizumi, K.; et al. Hypoxia-induced modulation of pten activity and emt phenotypes in lung cancers. Cancer Cell Int. 2016, 16, 33. [Google Scholar] [CrossRef] [PubMed]
  115. Jiang, J.; Tang, Y.-L.; Liang, X.-H. Emt: A new vision of hypoxia promoting cancer progression. Cancer Biol. Ther. 2011, 11, 714–723. [Google Scholar] [CrossRef] [PubMed]
  116. Brabletz, S.; Schuhwerk, H.; Brabletz, T.; Stemmler, M.P. Dynamic emt: A multi-tool for tumor progression. EMBO J. 2021, 40, e108647. [Google Scholar] [CrossRef] [PubMed]
  117. Gaponova, A.V.; Rodin, S.; Mazina, A.A.; Volchkov, P.V. Epithelial-mesenchymal transition: Role in cancer progression and the perspectives of antitumor treatment. Acta Naturae 2020, 12, 4–23. [Google Scholar] [CrossRef] [PubMed]
  118. Ribatti, D.; Tamma, R.; Annese, T. Epithelial-mesenchymal transition in cancer: A historical overview. Transl. Oncol. 2020, 13, 100773. [Google Scholar] [CrossRef] [PubMed]
  119. Hashemi, M.; Arani, H.Z.; Orouei, S.; Fallah, S.; Ghorbani, A.; Khaledabadi, M.; Kakavand, A.; Tavakolpournegari, A.; Saebfar, H.; Heidari, H.; et al. Emt mechanism in breast cancer metastasis and drug resistance: Revisiting molecular interactions and biological functions. Biomed. Pharmacother. 2022, 155, 113774. [Google Scholar] [CrossRef]
  120. Palamaris, K.; Felekouras, E.; Sakellariou, S. Epithelial to mesenchymal transition: Key regulator of pancreatic ductal adenocarcinoma progression and chemoresistance. Cancers 2021, 13, 5532. [Google Scholar] [CrossRef]
  121. Sha, J.; Bai, Y.; Ngo, H.X.; Okui, T.; Kanno, T. Overview of evidence-based chemotherapy for oral cancer: Focus on drug resistance related to the epithelial-mesenchymal transition. Biomolecules 2021, 11, 893. [Google Scholar] [CrossRef]
  122. Ashrafizadeh, M.; Mirzaei, S.; Hashemi, F.; Zarrabi, A.; Zabolian, A.; Saleki, H.; Sharifzadeh, S.O.; Soleymani, L.; Daneshi, S.; Hushmandi, K.; et al. New insight towards development of paclitaxel and docetaxel resistance in cancer cells: Emt as a novel molecular mechanism and therapeutic possibilities. Biomed. Pharmacother. 2021, 141, 111824. [Google Scholar] [CrossRef] [PubMed]
  123. De Las Rivas, J.; Brozovic, A.; Izraely, S.; Casas-Pais, A.; Witz, I.P.; Figueroa, A. Cancer drug resistance induced by emt: Novel therapeutic strategies. Arch. Toxicol. 2021, 95, 2279–2297. [Google Scholar] [CrossRef]
  124. Ashrafizadeh, M.; Zarrabi, A.; Hushmandi, K.; Kalantari, M.; Mohammadinejad, R.; Javaheri, T.; Sethi, G. Association of the epithelial-mesenchymal transition (emt) with cisplatin resistance. Int. J. Mol. Sci. 2020, 21, 4002. [Google Scholar] [CrossRef] [PubMed]
  125. Dudás, J.; Ladányi, A.; Ingruber, J.; Steinbichler, T.B.; Riechelmann, H. Epithelial to mesenchymal transition: A mechanism that fuels cancer radio/chemoresistance. Cells 2020, 9, 428. [Google Scholar] [CrossRef] [PubMed]
  126. Zheng, X.; Carstens, J.L.; Kim, J.; Scheible, M.; Kaye, J.; Sugimoto, H.; Wu, C.-C.; LeBleu, V.S.; Kalluri, R. Epithelial-to-mesenchymal transition is dispensable for metastasis but induces chemoresistance in pancreatic cancer. Nature 2015, 527, 525–530. [Google Scholar] [CrossRef]
  127. Fischer, K.R.; Durrans, A.; Lee, S.; Sheng, J.; Li, F.; Wong, S.T.C.; Choi, H.; El Rayes, T.; Ryu, S.; Troeger, J.; et al. Epithelial-to-mesenchymal transition is not required for lung metastasis but contributes to chemoresistance. Nature 2015, 527, 472–476. [Google Scholar] [CrossRef]
  128. Debaugnies, M.; Rodríguez-Acebes, S.; Blondeau, J.; Parent, M.-A.; Zocco, M.; Song, Y.; de Maertelaer, V.; Moers, V.; Latil, M.; Dubois, C.; et al. Rhoj controls emt-associated resistance to chemotherapy. Nature 2023, 616, 168–175. [Google Scholar] [CrossRef]
  129. Petrella, G.; Corsi, F.; Ciufolini, G.; Germini, S.; Capradossi, F.; Pelliccia, A.; Torino, F.; Ghibelli, L.; Cicero, D.O. Metabolic reprogramming of castration-resistant prostate cancer cells as a response to chemotherapy. Metabolites 2022, 13, 65. [Google Scholar] [CrossRef] [PubMed]
  130. Chen, X.; Zhou, Z.; Zhang, Z.; Zhao, C.; Li, J.; Jiang, J.; Huang, B.; Qin, Y. Puerarin inhibits emt induced by oxaliplatin via targeting carbonic anhydrase xii. Front. Pharmacol. 2022, 13, 969422. [Google Scholar] [CrossRef]
  131. Boulding, T.; McCuaig, R.D.; Tan, A.; Hardy, K.; Wu, F.; Dunn, J.; Kalimutho, M.; Sutton, C.R.; Forwood, J.K.; Bert, A.G.; et al. Lsd1 activation promotes inducible emt programs and modulates the tumour microenvironment in breast cancer. Sci. Rep. 2018, 8, 73. [Google Scholar] [CrossRef]
  132. Xie, S.; Fan, S.; Zhang, S.; Chen, W.; Li, Q.; Pan, G.; Zhang, H.; Wang, W.; Weng, B.; Zhang, Z.; et al. Sox8 regulates cancer stem-like properties and cisplatin-induced emt in tongue squamous cell carcinoma by acting on the wnt/beta-catenin pathway. Int. J. Cancer 2018, 142, 1252–1265. [Google Scholar] [CrossRef]
  133. Zheng, H.C. The molecular mechanisms of chemoresistance in cancers. Oncotarget 2017, 8, 59950–59964. [Google Scholar] [CrossRef]
  134. Li, P.; Zhang, X.; Wang, H.; Wang, L.; Liu, T.; Du, L.; Yang, Y.; Wang, C. Malat1 is associated with poor response to oxaliplatin-based chemotherapy in colorectal cancer patients and promotes chemoresistance through ezh2. Mol. Cancer Ther. 2017, 16, 739–751. [Google Scholar] [CrossRef]
  135. Li, Q.-Q.; Chen, Z.-Q.; Cao, X.-X.; Xu, J.-W.; Chen, Y.-Y.; Wang, W.-J.; Chen, Q.; Tang, F.; Liu, X.-P.; Xu, Z.-D. Involvement of nf-kappab/mir-448 regulatory feedback loop in chemotherapy-induced epithelial-mesenchymal transition of breast cancer cells. Cell Death Differ. 2011, 18, 16–25. [Google Scholar] [CrossRef] [PubMed]
  136. Liu, J.; Gao, L.; Zhan, N.; Xu, P.; Yang, J.; Yuan, F.; Xu, Y.; Cai, Q.; Geng, R.; Chen, Q. Hypoxia induced ferritin light chain (ftl) promoted epithelia mesenchymal transition and chemoresistance of glioma. J. Exp. Clin. Cancer Res. 2020, 39, 137. [Google Scholar] [CrossRef] [PubMed]
  137. Ni, J.; Zhou, S.; Yuan, W.; Cen, F.; Yan, Q. Mechanism of mir-210 involved in epithelial-mesenchymal transition of pancreatic cancer cells under hypoxia. J. Recept. Signal Transduct. Res. 2019, 39, 399–406. [Google Scholar] [CrossRef]
  138. Okumura, Y.; Noda, T.; Eguchi, H.; Sakamoto, T.; Iwagami, Y.; Yamada, D.; Asaoka, T.; Wada, H.; Kawamoto, K.; Gotoh, K.; et al. Hypoxia-induced plod2 is a key regulator in epithelial-mesenchymal transition and chemoresistance in biliary tract cancer. Ann. Surg. Oncol. 2018, 25, 3728–3737. [Google Scholar] [CrossRef] [PubMed]
  139. Wu, C.-E.; Zhuang, Y.-W.; Zhou, J.-Y.; Liu, S.-L.; Wang, R.-P.; Shu, P. Cinnamaldehyde enhances apoptotic effect of oxaliplatin and reverses epithelial-mesenchymal transition and stemnness in hypoxic colorectal cancer cells. Exp. Cell Res. 2019, 383, 111500. [Google Scholar] [CrossRef]
  140. Zhou, Q.-Y.; Tu, C.-Y.; Shao, C.-X.; Wang, W.-K.; Zhu, J.-D.; Cai, Y.; Mao, J.-Y.; Chen, W. Gc7 blocks epithelial-mesenchymal transition and reverses hypoxia-induced chemotherapy resistance in hepatocellular carcinoma cells. Am. J. Transl. Res. 2017, 9, 2608–2617. [Google Scholar]
  141. Adamski, J.; Price, A.; Dive, C.; Makin, G. Hypoxia-induced cytotoxic drug resistance in osteosarcoma is independent of hif-1alpha. PLoS ONE 2013, 8, e65304. [Google Scholar] [CrossRef] [PubMed]
  142. Cosse, J.-P.; Ronvaux, M.; Ninane, N.; Raes, M.J.; Michiels, C. Hypoxia-induced decrease in p53 protein level and increase in c-jun DNA binding activity results in cancer cell resistance to etoposide. Neoplasia 2009, 11, 976–986. [Google Scholar] [CrossRef] [PubMed]
  143. Piret, J.-P.; Cosse, J.-P.; Ninane, N.; Raes, M.; Michiels, C. Hypoxia protects hepg2 cells against etoposide-induced apoptosis via a hif-1-independent pathway. Exp. Cell Res. 2006, 312, 2908–2920. [Google Scholar] [CrossRef] [PubMed]
  144. Chen, J.; Kobayashi, M.; Darmanin, S.; Qiao, Y.; Gully, C.; Zhao, R.; Yeung, S.C.; Lee, M.H. Pim-1 plays a pivotal role in hypoxia-induced chemoresistance. Oncogene 2009, 28, 2581–2592. [Google Scholar] [CrossRef] [PubMed]
  145. Schnitzer, S.E.; Weigert, A.; Zhou, J.; Brüne, B. Hypoxia enhances sphingosine kinase 2 activity and provokes sphingosine-1-phosphate-mediated chemoresistance in a549 lung cancer cells. Mol. Cancer Res. 2009, 7, 393–401. [Google Scholar] [CrossRef]
  146. Schnitzer, S.E.; Schmid, T.; Zhou, J.; Brune, B. Hypoxia and hif-1alpha protect a549 cells from drug-induced apoptosis. Cell Death Differ. 2006, 13, 1611–1613. [Google Scholar] [CrossRef]
  147. Chen, X.; Liu, M.; Meng, F.; Sun, B.; Jin, X.; Jia, C. The long noncoding rna hif1a-as2 facilitates cisplatin resistance in bladder cancer. J. Cell Biochem. 2019, 120, 243–252. [Google Scholar] [CrossRef]
  148. Su, Y.; Yang, W.; Jiang, N.; Shi, J.; Chen, L.; Zhong, G.; Bi, J.; Dong, W.; Wang, Q.; Wang, C.; et al. Hypoxia-elevated circelp3 contributes to bladder cancer progression and cisplatin resistance. Int. J. Biol. Sci. 2019, 15, 441–452. [Google Scholar] [CrossRef]
  149. Huang, H.; Peng, J.; Yi, S.; Ding, C.; Ji, W.; Huang, Q.; Zeng, S. Circular rna circube2d2 functions as an oncogenic factor in hepatocellular carcinoma sorafenib resistance and glycolysis. Am. J. Transl. Res. 2021, 13, 6076–6086. [Google Scholar]
  150. Wang, H.; Min, J.; Xu, C.; Liu, Y.; Yu, Z.; Gong, A.; Xu, M. Hypoxia-elicited exosomes promote the chemoresistance of pancreatic cancer cells by transferring lncror via hippo signaling. J. Cancer 2023, 14, 1075–1087. [Google Scholar] [CrossRef]
  151. Tao, S.; Wang, J.; Li, F.; Shi, B.; Ren, Q.; Zhuang, Y.; Qian, X. Extracellular vesicles released by hypoxia-induced tumor-associated fibroblasts impart chemoresistance to breast cancer cells via long noncoding rna h19 delivery. FASEB J. 2024, 38, e23165. [Google Scholar] [CrossRef] [PubMed]
  152. Musah-Eroje, A.; Watson, S. A novel 3d in vitro model of glioblastoma reveals resistance to temozolomide which was potentiated by hypoxia. J. Neurooncol 2019, 142, 231–240. [Google Scholar] [CrossRef] [PubMed]
  153. Ahmed, E.M.; Bandopadhyay, G.; Coyle, B.; Grabowska, A. A hif-independent, cd133-mediated mechanism of cisplatin resistance in glioblastoma cells. Cell. Oncol. 2018, 41, 319–328. [Google Scholar] [CrossRef] [PubMed]
  154. Mao, Q.; Zhang, Y.; Fu, X.; Xue, J.; Guo, W.; Meng, M.; Zhou, Z.; Mo, X.; Lu, Y. A tumor hypoxic niche protects human colon cancer stem cells from chemotherapy. J. Cancer Res. Clin. Oncol. 2013, 139, 211–222. [Google Scholar] [CrossRef] [PubMed]
  155. Saigusa, S.; Tanaka, K.; Toiyama, Y.; Yokoe, T.; Okugawa, Y.; Koike, Y.; Fujikawa, H.; Inoue, Y.; Miki, C.; Kusunoki, M. Clinical significance of cd133 and hypoxia inducible factor-1alpha gene expression in rectal cancer after preoperative chemoradiotherapy. Clin. Oncol. 2011, 23, 323–332. [Google Scholar] [CrossRef]
  156. Lu, C.; Mahajan, A.; Hong, S.-H.; Galli, S.; Zhu, S.; Tilan, J.U.; Abualsaud, N.; Adnani, M.; Chung, S.; Elmansy, N.; et al. Hypoxia-activated neuropeptide y/y5 receptor/rhoa pathway triggers chromosomal instability and bone metastasis in ewing sarcoma. Nat. Commun. 2022, 13, 2323. [Google Scholar] [CrossRef] [PubMed]
  157. Moffitt, L.R.; Bilandzic, M.; Wilson, A.L.; Chen, Y.; Gorrell, M.D.; Oehler, M.K.; Plebanski, M.; Stephens, A.N. Hypoxia regulates dpp4 expression, proteolytic inactivation, and shedding from ovarian cancer cells. Int. J. Mol. Sci. 2020, 21, 8110. [Google Scholar] [CrossRef] [PubMed]
  158. Tilan, J.U.; Lu, C.; Galli, S.; Izycka-Swieszewska, E.; Earnest, J.P.; Shabbir, A.; Everhart, L.M.; Wang, S.; Martin, S.; Horton, M.; et al. Hypoxia shifts activity of neuropeptide y in ewing sarcoma from growth-inhibitory to growth-promoting effects. Oncotarget 2013, 4, 2487–2501. [Google Scholar] [CrossRef] [PubMed]
  159. Zhou, T.-Y.; Zhuang, L.-H.; Hu, Y.; Zhou, Y.-L.; Lin, W.-K.; Wang, D.-D.; Wan, Z.-Q.; Chang, L.-L.; Chen, Y.; Ying, M.-D.; et al. Inactivation of hypoxia-induced yap by statins overcomes hypoxic resistance tosorafenib in hepatocellular carcinoma cells. Sci. Rep. 2016, 6, 30483. [Google Scholar] [CrossRef]
  160. Lang, L. Fda approves sorafenib for patients with inoperable liver cancer. Gastroenterology 2008, 134, 379. [Google Scholar] [CrossRef]
  161. Xia, X.; Wang, Q.; Ye, T.; Liu, Y.; Liu, D.; Song, S.; Zheng, C. Nrf2/abcb1-mediated efflux and parp1-mediated dampening of DNA damage contribute to doxorubicin resistance in chronic hypoxic hepg2 cells. Fundam. Clin. Pharmacol. 2020, 34, 41–50. [Google Scholar] [CrossRef]
  162. Syu, J.-P.; Chi, J.-T.; Kung, H.-N. Nrf2 is the key to chemotherapy resistance in mcf7 breast cancer cells under hypoxia. Oncotarget 2016, 7, 14659–14672. [Google Scholar] [CrossRef]
  163. Li, T.; Fu, X.; Wang, J.; Shang, W.; Wang, X.; Zhang, L.; Li, J. Mechanism of nurp1 in temozolomide resistance in hypoxia-treated glioma cells via the kdm3a/tfeb axis. Oncol. Res. 2023, 31, 345–359. [Google Scholar] [CrossRef]
  164. Koritzinsky, M.; Levitin, F.; van den Beucken, T.; Rumantir, R.A.; Harding, N.J.; Chu, K.C.; Boutros, P.C.; Braakman, I.; Wouters, B.G. Two phases of disulfide bond formation have differing requirements for oxygen. J. Cell Biol. 2013, 203, 615–627. [Google Scholar] [CrossRef]
  165. Bartoszewska, S.; Collawn, J.F.; Bartoszewski, R. The role of the hypoxia-related unfolded protein response (upr) in the tumor microenvironment. Cancers 2022, 14, 4870. [Google Scholar] [CrossRef]
  166. Chipurupalli, S.; Kannan, E.; Tergaonkar, V.; D’andrea, R.; Robinson, N. Hypoxia induced er stress response as an adaptive mechanism in cancer. Int. J. Mol. Sci. 2019, 20, 749. [Google Scholar] [CrossRef]
  167. Avril, T.; Vauléon, E.; Chevet, E. Endoplasmic reticulum stress signaling and chemotherapy resistance in solid cancers. Oncogenesis 2017, 6, e373. [Google Scholar] [CrossRef]
  168. Benedetti, R.; Romeo, M.A.; Arena, A.; Montani, M.S.G.; Di Renzo, L.; D’orazi, G.; Cirone, M. Atf6 prevents DNA damage and cell death in colon cancer cells undergoing er stress. Cell Death Discov. 2022, 8, 295. [Google Scholar] [CrossRef]
  169. Cho, J.; Min, H.-Y.; Pei, H.; Wei, X.; Sim, J.Y.; Park, S.-H.; Hwang, S.J.; Lee, H.-J.; Hong, S.; Shin, Y.K.; et al. The atf6-egf pathway mediates the awakening of slow-cycling chemoresistant cells and tumor recurrence by stimulating tumor angiogenesis. Cancers 2020, 12, 1772. [Google Scholar] [CrossRef] [PubMed]
  170. Akman, M.; Belisario, D.C.; Salaroglio, I.C.; Kopecka, J.; Donadelli, M.; De Smaele, E.; Riganti, C. Hypoxia, endoplasmic reticulum stress and chemoresistance: Dangerous liaisons. J. Exp. Clin. Cancer Res. 2021, 40, 28. [Google Scholar] [CrossRef] [PubMed]
  171. Pi, L.; Li, X.; Song, Q.; Shen, Y.; Lu, X.; DI, B. Knockdown of glucose-regulated protein 78 abrogates chemoresistance of hypopharyngeal carcinoma cells to cisplatin induced by unfolded protein in response to severe hypoxia. Oncol. Lett. 2014, 7, 685–692. [Google Scholar] [CrossRef]
  172. Lee, D.; Sun, S.; Ho, A.S.; Kiang, K.M.; Zhang, X.Q.; Xu, F.F.; Leung, G.K. Hyperoxia resensitizes chemoresistant glioblastoma cells to temozolomide through unfolded protein response. Anticancer. Res. 2014, 34, 2957–2966. [Google Scholar]
  173. Bouznad, N.; Rokavec, M.; Öner, M.G.; Hermeking, H. Mir-34a and ire1a/xbp-1(s) form a double-negative feedback loop to regulate hypoxia-induced emt, metastasis, chemo-resistance and autophagy. Cancers 2023, 15, 1143. [Google Scholar] [CrossRef]
  174. Moszyńska, A.; Collawn, J.F.; Bartoszewski, R. Ire1 endoribonuclease activity modulates hypoxic hif-1alpha signaling in human endothelial cells. Biomolecules 2020, 10, 895. [Google Scholar] [CrossRef]
  175. Lee, J.H.; Yoon, Y.M.; Lee, S.H. Hypoxic preconditioning promotes the bioactivities of mesenchymal stem cells via the hif-1alpha-grp78-akt axis. Int. J. Mol. Sci. 2017, 18, 1320. [Google Scholar] [CrossRef]
  176. Li, Z.; Wang, Y.; Newton, I.P.; Zhang, L.; Ji, P.; Li, Z. Grp78 is implicated in the modulation of tumor aerobic glycolysis by promoting autophagic degradation of ikkbeta. Cell. Signal. 2015, 27, 1237–1245. [Google Scholar] [CrossRef]
  177. Chen, X.; Iliopoulos, D.; Zhang, Q.; Tang, Q.; Greenblatt, M.B.; Hatziapostolou, M.; Lim, E.; Tam, W.L.; Ni, M.; Chen, Y.; et al. Xbp1 promotes triple-negative breast cancer by controlling the hif1alpha pathway. Nature 2014, 508, 103–107. [Google Scholar] [CrossRef]
  178. Miharada, K.; Karlsson, G.; Rehn, M.; Rörby, E.; Siva, K.; Cammenga, J.; Karlsson, S. Hematopoietic stem cells are regulated by cripto, as an intermediary of hif-1alpha in the hypoxic bone marrow niche. Ann. N. Y. Acad. Sci. 2012, 1266, 55–62. [Google Scholar] [CrossRef]
  179. Kang, M.J.; Jung, S.M.; Kim, M.J.; Bae, J.H.; Kim, H.B.; Kim, J.Y.; Park, S.J.; Song, H.S.; Kim, D.W.; Kang, C.D.; et al. DNA-dependent protein kinase is involved in heat shock protein-mediated accumulation of hypoxia-inducible factor-1alpha in hypoxic preconditioned hepg2 cells. FEBS J. 2008, 275, 5969–5981. [Google Scholar] [CrossRef] [PubMed]
  180. Mivechi, N.F.; Koong, A.C.; Giaccia, A.J.; Hahn, G.M. Analysis of hsf-1 phosphorylation in a549 cells treated with a variety of stresses. Int. J. Hyperth. 1994, 10, 371–379. [Google Scholar] [CrossRef] [PubMed]
  181. Cyran, A.M.; Zhitkovich, A. Heat shock proteins and hsf1 in cancer. Front. Oncol. 2022, 12, 860320. [Google Scholar] [CrossRef]
  182. Landriscina, M.; Maddalena, F.; Laudiero, G.; Esposito, F. Adaptation to oxidative stress, chemoresistance, and cell survival. Antioxid. Redox Signal 2009, 11, 2701–2716. [Google Scholar] [CrossRef]
  183. Vydra, N.; Toma, A.; Glowala-Kosinska, M.; Gogler-Piglowska, A.; Widlak, W. Overexpression of heat shock transcription factor 1 enhances the resistance of melanoma cells to doxorubicin and paclitaxel. BMC Cancer 2013, 13, 504. [Google Scholar] [CrossRef]
  184. Vilaboa, N.E.; Galán, A.; Troyano, A.; de Blas, E.; Aller, P. Regulation of multidrug resistance 1 (mdr1)/p-glycoprotein gene expression and activity by heat-shock transcription factor 1 (hsf1). J. Biol. Chem. 2000, 275, 24970–24976. [Google Scholar] [CrossRef]
  185. Yoo, H.J.; Im, C.-N.; Youn, D.-Y.; Yun, H.H.; Lee, J.-H. Bis is induced by oxidative stress via activation of hsf1. Korean J. Physiol. Pharmacol. 2014, 18, 403–409. [Google Scholar] [CrossRef]
  186. Shanker, M.; Willcutts, D.; Roth, J.A.; Ramesh, R. Drug resistance in lung cancer. Lung Cancer 2010, 1, 23–36. [Google Scholar]
  187. Banerjee Mustafi, S.; Chakraborty, P.K.; Dey, R.S.; Raha, S. Heat stress upregulates chaperone heat shock protein 70 and antioxidant manganese superoxide dismutase through reactive oxygen species (ros), p38mapk, and akt. Cell Stress Chaperones 2009, 14, 579–589. [Google Scholar] [CrossRef]
  188. Yoo, H.Y.; Chang, M.S.; Rho, H.M. The activation of the rat copper/zinc superoxide dismutase gene by hydrogen peroxide through the hydrogen peroxide-responsive element and by paraquat and heat shock through the same heat shock element. J. Biol. Chem. 1999, 274, 23887–23892. [Google Scholar] [CrossRef]
  189. Egger, G.; Liang, G.; Aparicio, A.; Jones, P.A. Epigenetics in human disease and prospects for epigenetic therapy. Nature 2004, 429, 457–463. [Google Scholar] [CrossRef]
  190. Marmorstein, R.; Zhou, M.-M. Writers and readers of histone acetylation: Structure, mechanism, and inhibition. Cold Spring Harb. Perspect. Biol. 2014, 6, a018762. [Google Scholar] [CrossRef]
  191. Kim, M.S.; Kwon, H.J.; Lee, Y.M.; Baek, J.H.; Jang, J.-E.; Lee, S.-W.; Moon, E.-J.; Kim, H.-S.; Lee, S.-K.; Chung, H.Y.; et al. Histone deacetylases induce angiogenesis by negative regulation of tumor suppressor genes. Nat. Med. 2001, 7, 437–443. [Google Scholar] [CrossRef] [PubMed]
  192. Pluemsampant, S.; Safronova, O.S.; Nakahama, K.; Morita, I. Protein kinase ck2 is a key activator of histone deacetylase in hypoxia-associated tumors. Int. J. Cancer 2008, 122, 333–341. [Google Scholar] [CrossRef] [PubMed]
  193. Kirmes, I.; Szczurek, A.; Prakash, K.; Charapitsa, I.; Heiser, C.; Musheev, M.; Schock, F.; Fornalczyk, K.; Ma, D.; Birk, U.; et al. A transient ischemic environment induces reversible compaction of chromatin. Genome Biol. 2015, 16, 246. [Google Scholar] [CrossRef] [PubMed]
  194. Leszczynska, K.B.; Dzwigonska, M.; Estephan, H.; Moehlenbrink, J.; Bowler, E.; Giaccia, A.J.; Mieczkowski, J.; Kaminska, B.; Hammond, E.M. Hypoxia-mediated regulation of ddx5 through decreased chromatin accessibility and post-translational targeting restricts r-loop accumulation. Mol. Oncol. 2023, 17, 1173–1191. [Google Scholar] [CrossRef] [PubMed]
  195. Gujral, P.; Mahajan, V.; Lissaman, A.C.; Ponnampalam, A.P. Histone acetylation and the role of histone deacetylases in normal cyclic endometrium. Reprod. Biol. Endocrinol. 2020, 18, 84. [Google Scholar] [CrossRef] [PubMed]
  196. Chen, D.-Q.; Pan, B.-Z.; Huang, J.-Y.; Zhang, K.; Cui, S.-Y.; De, W.; Wang, R.; Chen, L.-B. Hdac 1/4-mediated silencing of microrna-200b promotes chemoresistance in human lung adenocarcinoma cells. Oncotarget 2014, 5, 3333–3349. [Google Scholar] [CrossRef] [PubMed]
  197. Huang, R.; Langdon, S.P.; Tse, M.; Mullen, P.; Um, I.H.; Faratian, D.; Harrison, D.J. The role of hdac2 in chromatin remodelling and response to chemotherapy in ovarian cancer. Oncotarget 2016, 7, 4695–4711. [Google Scholar] [CrossRef] [PubMed]
  198. Taddei, A.; Roche, D.; Bickmore, W.A.; Almouzni, G. The effects of histone deacetylase inhibitors on heterochromatin: Implications for anticancer therapy? EMBO Rep. 2005, 6, 520–524. [Google Scholar] [CrossRef] [PubMed]
  199. Mehmood, S.A.; Sahu, K.K.; Sengupta, S.; Partap, S.; Karpoormath, R.; Kumar, B.; Kumar, D. Recent advancement of hdac inhibitors against breast cancer. Med. Oncol. 2023, 40, 201. [Google Scholar] [CrossRef]
  200. Falkenberg, K.J.; Johnstone, R.W. Histone deacetylases and their inhibitors in cancer, neurological diseases and immune disorders. Nat. Rev. Drug Discov. 2014, 13, 673–691. [Google Scholar] [CrossRef]
  201. Hu, Z.; Wei, F.; Su, Y.; Wang, Y.; Shen, Y.; Fang, Y.; Ding, J.; Chen, Y. Histone deacetylase inhibitors promote breast cancer metastasis by elevating nedd9 expression. Signal Transduct. Target. Ther. 2023, 8, 11. [Google Scholar] [CrossRef] [PubMed]
  202. Quinn, D.I.; Tsao-Wei, D.D.; Twardowski, P.; Aparicio, A.M.; Frankel, P.; Chatta, G.; Wright, J.J.; Groshen, S.G.; Khoo, S.; Lenz, H.-J.; et al. Phase ii study of the histone deacetylase inhibitor vorinostat (suberoylanilide hydroxamic acid; saha) in recurrent or metastatic transitional cell carcinoma of the urothelium—An nci-ctep sponsored: California cancer consortium trial, nci 6879. Investig. New Drugs 2021, 39, 812–820. [Google Scholar] [CrossRef] [PubMed]
  203. Fujisawa, T.; Filippakopoulos, P. Functions of bromodomain-containing proteins and their roles in homeostasis and cancer. Nat. Rev. Mol. Cell Biol. 2017, 18, 246–262. [Google Scholar] [CrossRef]
  204. Filippakopoulos, P.; Picaud, S.; Mangos, M.; Keates, T.; Lambert, J.-P.; Barsyte-Lovejoy, D.; Felletar, I.; Volkmer, R.; Müller, S.; Pawson, T.; et al. Histone recognition and large-scale structural analysis of the human bromodomain family. Cell 2012, 149, 214–231. [Google Scholar] [CrossRef] [PubMed]
  205. Pan, Z.; Zhao, Y.; Wang, X.; Xie, X.; Liu, M.; Zhang, K.; Wang, L.; Bai, D.; Foster, L.J.; Shu, R.; et al. Targeting bromodomain-containing proteins: Research advances of drug discovery. Mol. Biomed. 2023, 4, 13. [Google Scholar] [CrossRef]
  206. Zaware, N.; Zhou, M.-M. Bromodomain biology and drug discovery. Nat. Struct. Mol. Biol. 2019, 26, 870–879. [Google Scholar] [CrossRef]
  207. Muller, S.; Filippakopoulos, P.; Knapp, S. Bromodomains as therapeutic targets. Expert. Rev. Mol. Med. 2011, 13, e29. [Google Scholar] [CrossRef] [PubMed]
  208. Gajjela, B.K.; Zhou, M.-M. Bromodomain inhibitors and therapeutic applications. Curr. Opin. Chem. Biol. 2023, 75, 102323. [Google Scholar] [CrossRef] [PubMed]
  209. Lazarchuk, P.; Hernandez-Villanueva, J.; Pavlova, M.N.; Federation, A.; MacCoss, M.; Sidorova, J.M. Mutual balance of histone deacetylases 1 and 2 and the acetyl reader atad2 regulates the level of acetylation of histone h4 on nascent chromatin of human cells. Mol. Cell. Biol. 2020, 40, e00421-19. [Google Scholar] [CrossRef]
  210. Koo, S.J.; Fernández-Montalván, A.E.; Badock, V.; Ott, C.J.; Holton, S.J.; von Ahsen, O.; Toedling, J.; Vittori, S.; Bradner, J.E.; Gorjánácz, M. Atad2 is an epigenetic reader of newly synthesized histone marks during DNA replication. Oncotarget 2016, 7, 70323–70335. [Google Scholar] [CrossRef]
  211. Revenko, A.S.; Kalashnikova, E.V.; Gemo, A.T.; Zou, J.X.; Chen, H.-W. Chromatin loading of e2f-mll complex by cancer-associated coregulator ancca via reading a specific histone mark. Mol. Cell. Biol. 2010, 30, 5260–5272. [Google Scholar] [CrossRef]
  212. Ciró, M.; Prosperini, E.; Quarto, M.; Grazini, U.; Walfridsson, J.; McBlane, F.; Nucifero, P.; Pacchiana, G.; Capra, M.; Christensen, J.; et al. Atad2 is a novel cofactor for myc, overexpressed and amplified in aggressive tumors. Cancer Res. 2009, 69, 8491–8498. [Google Scholar] [CrossRef]
  213. Losman, J.-A.; Koivunen, P.; Kaelin, W.G., Jr. 2-oxoglutarate-dependent dioxygenases in cancer. Nat. Rev. Cancer 2020, 20, 710–726. [Google Scholar] [CrossRef]
  214. Kao, T.-W.; Bai, G.-H.; Wang, T.-L.; Shih, I.-M.; Chuang, C.-M.; Lo, C.-L.; Tsai, M.-C.; Chiu, L.-Y.; Lin, C.-C.; Shen, Y.-A. Novel cancer treatment paradigm targeting hypoxia-induced factor in conjunction with current therapies to overcome resistance. J. Exp. Clin. Cancer Res. 2023, 42, 171. [Google Scholar] [CrossRef]
  215. Meehan, R.; Kummar, S.; Do, K.; Coyne, G.O.; Juwara, L.; Zlott, J.; Rubinstein, L.; Doroshow, J.H.; Chen, A.P. A phase i study of ganetespib and ziv-aflibercept in patients with advanced carcinomas and sarcomas. Oncologist 2018, 23, 1269-e125. [Google Scholar] [CrossRef]
  216. Kim, A.; Lu, Y.; Okuno, S.H.; Reinke, D.; Maertens, O.; Perentesis, J.; Basu, M.; Wolters, P.L.; De Raedt, T.; Chawla, S.; et al. Targeting refractory sarcomas and malignant peripheral nerve sheath tumors in a phase i/ii study of sirolimus in combination with ganetespib (sarc023). Sarcoma 2020, 2020, 5784876. [Google Scholar] [CrossRef]
  217. Ramalingam, S.; Goss, G.; Rosell, R.; Schmid-Bindert, G.; Zaric, B.; Andric, Z.; Bondarenko, I.; Komov, D.; Ceric, T.; Khuri, F.; et al. A randomized phase ii study of ganetespib, a heat shock protein 90 inhibitor, in combination with docetaxel in second-line therapy of advanced non-small cell lung cancer (galaxy-1). Ann. Oncol. 2015, 26, 1741–1748. [Google Scholar] [CrossRef]
  218. Fallah, J.; Brave, M.H.; Weinstock, C.; Mehta, G.U.; Bradford, D.; Gittleman, H.; Bloomquist, E.W.; Charlab, R.; Hamed, S.S.; Miller, C.P.; et al. Fda approval summary: Belzutifan for von hippel-lindau disease-associated tumors. Clin. Cancer Res. 2022, 28, 4843–4848. [Google Scholar] [CrossRef]
  219. Strowd, R.; Ellingson, B.; Raymond, C.; Yao, J.; Wen, P.Y.; Ahluwalia, M.; Piotrowski, A.; Desai, A.; Clarke, J.L.; Lieberman, F.S.; et al. Activity of a first-in-class oral hif2-alpha inhibitor, pt2385, in patients with first recurrence of glioblastoma. J. Neuro-Oncol. 2023, 165, 101–112. [Google Scholar] [CrossRef]
  220. Xu, P.; Wu, Q.; Yu, J.; Rao, Y.; Kou, Z.; Fang, G.; Shi, X.; Liu, W.; Han, H. A systematic way to infer the regulation relations of mirnas on target genes and critical mirnas in cancers. Front. Genet. 2020, 11, 278. [Google Scholar] [CrossRef]
  221. He, Z.; Zhu, Q. Circular rnas: Emerging roles and new insights in human cancers. Biomed. Pharmacother. 2023, 165, 115217. [Google Scholar] [CrossRef]
  222. Brenner, A.J.; Floyd, J.; Fichtel, L.; Michalek, J.; Kanakia, K.P.; Huang, S.; Reardon, D.; Wen, P.Y.; Lee, E.Q. Phase 2 trial of hypoxia activated evofosfamide (th302) for treatment of recurrent bevacizumab-refractory glioblastoma. Sci. Rep. 2021, 11, 2306. [Google Scholar] [CrossRef]
  223. García-Venzor, A.; Mandujano-Tinoco, E.A.; Ruiz-Silvestre, A.; Sánchez, J.M.; Lizarraga, F.; Zampedri, C.; Melendez-Zajgla, J.; Maldonado, V. Lncmat2b regulated by severe hypoxia induces cisplatin resistance by increasing DNA damage repair and tumor-initiating population in breast cancer cells. Carcinogenesis 2020, 41, 1485–1497. [Google Scholar] [CrossRef]
  224. Yang, H.; Hu, Y.; Weng, M.; Liu, X.; Wan, P.; Hu, Y.; Ma, M.; Zhang, Y.; Xia, H.; Lv, K. Hypoxia inducible lncrna-cbslr modulates ferroptosis through m6a-ythdf2-dependent modulation of cbs in gastric cancer. J. Adv. Res. 2022, 37, 91–106. [Google Scholar] [CrossRef]
  225. Yin, X.; Liao, Y.; Xiong, W.; Zhang, Y.; Zhou, Y.; Yang, Y. Hypoxia-induced lncrna anril promotes cisplatin resistance in retinoblastoma cells through regulating abcg2 expression. Clin. Exp. Pharmacol. Physiol. 2020, 47, 1049–1057. [Google Scholar] [CrossRef]
  226. Huan, L.; Guo, T.; Wu, Y.; Xu, L.; Huang, S.; Xu, Y.; Liang, L.; He, X. Hypoxia induced lucat1/ptbp1 axis modulates cancer cell viability and chemotherapy response. Mol. Cancer 2020, 19, 11. [Google Scholar] [CrossRef]
  227. Moreno Leon, L.; Gautier, M.; Allan, R.; Ilié, M.; Nottet, N.; Pons, N.; Paquet, A.; Lebrigand, K.; Truchi, M.; Fassy, J.; et al. The nuclear hypoxia-regulated nlucat1 long non-coding rna contributes to an aggressive phenotype in lung adenocarcinoma through regulation of oxidative stress. Oncogene 2019, 38, 7146–7165. [Google Scholar] [CrossRef]
  228. Wang, F.; Ji, X.; Wang, J.; Ma, X.; Yang, Y.; Zuo, J.; Cui, J. Lncrna pvt1 enhances proliferation and cisplatin resistance via regulating mir-194-5p/hif1a axis in oral squamous cell carcinoma. Onco Targets Ther. 2020, 13, 243–252. [Google Scholar] [CrossRef]
  229. Xu, F.; Huang, M.; Chen, Q.; Niu, Y.; Hu, Y.; Hu, P.; Chen, D.; He, C.; Huang, K.; Zeng, Z.; et al. Lncrna hif1a-as1 promotes gemcitabine resistance of pancreatic cancer by enhancing glycolysis through modulating the akt/yb1/hif1alpha pathway. Cancer Res. 2021, 81, 5678–5691. [Google Scholar] [CrossRef]
  230. Güçlü, E.; Güneş, C.E.; Kurar, E.; Vural, H. Knockdown of lncrna hif1a-as2 increases drug sensitivity of sclc cells in association with autophagy. Med. Oncol. 2021, 38, 113. [Google Scholar] [CrossRef]
  231. Zhang, L.; Wu, H.; Zhang, Y.; Xiao, X.; Chu, F. Induction of lncrna norad accounts for hypoxia-induced chemoresistance and vasculogenic mimicry in colorectal cancer by sponging the mir-495-3p/ hypoxia-inducible factor-1alpha (hif-1alpha). Bioengineered 2022, 13, 950–962. [Google Scholar] [CrossRef]
  232. Weng, X.; Liu, H.; Ruan, J.; Du, M.; Wang, L.; Mao, J.; Cai, Y.; Lu, X.; Chen, W.; Huang, Y.; et al. Hotair/mir-1277-5p/zeb1 axis mediates hypoxia-induced oxaliplatin resistance via regulating epithelial-mesenchymal transition in colorectal cancer. Cell Death Discov. 2022, 8, 310. [Google Scholar] [CrossRef]
  233. Zhu, Z.-J.; Pang, Y.; Jin, G.; Zhang, H.-Y.; Wang, W.-H.; Liu, J.-W.; Tuo, G.-X.; Wu, P.; Yang, Y.; Wang, Z.-Q.; et al. Hypoxia induces chemoresistance of esophageal cancer cells to cisplatin through regulating the lncrna-ems/mir-758-3p/wtap axis. Aging 2021, 13, 17155–17176. [Google Scholar] [CrossRef]
Figure 1. Mechanisms of hypoxia-induced chemotherapy resistance mediated by HIF.
Figure 1. Mechanisms of hypoxia-induced chemotherapy resistance mediated by HIF.
Cancers 16 01729 g001
Figure 2. Mechanisms of hypoxia-induced chemotherapy resistance independent of HIF.
Figure 2. Mechanisms of hypoxia-induced chemotherapy resistance independent of HIF.
Cancers 16 01729 g002
Figure 3. ATAD2 degradation, cell cycle delay, and chemotherapy resistance under hypoxic conditions. Under hypoxic conditions, ATAD2 protein degradation delays the cell cycle from the early to late S phase and induces chemotherapy resistance.
Figure 3. ATAD2 degradation, cell cycle delay, and chemotherapy resistance under hypoxic conditions. Under hypoxic conditions, ATAD2 protein degradation delays the cell cycle from the early to late S phase and induces chemotherapy resistance.
Cancers 16 01729 g003
Figure 4. The 2-OGDD-dependent regulation of ATAD2 proteolysis. Under normoxic conditions, the ATAD2 protein is hydroxylated and stabilized by 2-OGDD. On the other hand, under hypoxic conditions, the activity of oxygen-requiring 2-OGDD is reduced, so the hydroxylation of the ATAD2 protein does not occur, resulting in the degradation of the ATAD2 protein by the proteasome pathway. Note: At present, there is also a possibility that a factor regulating ATAD2 stability receives hydroxylation by 2-OGDD.
Figure 4. The 2-OGDD-dependent regulation of ATAD2 proteolysis. Under normoxic conditions, the ATAD2 protein is hydroxylated and stabilized by 2-OGDD. On the other hand, under hypoxic conditions, the activity of oxygen-requiring 2-OGDD is reduced, so the hydroxylation of the ATAD2 protein does not occur, resulting in the degradation of the ATAD2 protein by the proteasome pathway. Note: At present, there is also a possibility that a factor regulating ATAD2 stability receives hydroxylation by 2-OGDD.
Cancers 16 01729 g004
Figure 5. Future perspective: Therapeutic strategies must be developed to overcome the chemoresistance of hypoxic cancer cells via ATAD2 degradation. If the degradation mechanism of the ATAD2 protein can be elucidated and inhibited, a novel therapeutic strategy to overcome the chemotherapy resistance of hypoxic cancer cells can be established.
Figure 5. Future perspective: Therapeutic strategies must be developed to overcome the chemoresistance of hypoxic cancer cells via ATAD2 degradation. If the degradation mechanism of the ATAD2 protein can be elucidated and inhibited, a novel therapeutic strategy to overcome the chemotherapy resistance of hypoxic cancer cells can be established.
Cancers 16 01729 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lee, P.W.T.; Koseki, L.R.; Haitani, T.; Harada, H.; Kobayashi, M. Hypoxia-Inducible Factor-Dependent and Independent Mechanisms Underlying Chemoresistance of Hypoxic Cancer Cells. Cancers 2024, 16, 1729. https://doi.org/10.3390/cancers16091729

AMA Style

Lee PWT, Koseki LR, Haitani T, Harada H, Kobayashi M. Hypoxia-Inducible Factor-Dependent and Independent Mechanisms Underlying Chemoresistance of Hypoxic Cancer Cells. Cancers. 2024; 16(9):1729. https://doi.org/10.3390/cancers16091729

Chicago/Turabian Style

Lee, Peter Wai Tik, Lina Rochelle Koseki, Takao Haitani, Hiroshi Harada, and Minoru Kobayashi. 2024. "Hypoxia-Inducible Factor-Dependent and Independent Mechanisms Underlying Chemoresistance of Hypoxic Cancer Cells" Cancers 16, no. 9: 1729. https://doi.org/10.3390/cancers16091729

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

Article Metrics

Back to TopTop