Skip to main content

Targeting purine metabolism in ovarian cancer

Abstract

Purine, an abundant substrate in organisms, is a critical raw material for cell proliferation and an important factor for immune regulation. The purine de novo pathway and salvage pathway are tightly regulated by multiple enzymes, and dysfunction in these enzymes leads to excessive cell proliferation and immune imbalance that result in tumor progression. Maintaining the homeostasis of purine pools is an effective way to control cell growth and tumor evolution, and exploiting purine metabolism to suppress tumors suggests interesting directions for future research. In this review, we describe the process of purine metabolism and summarize the role and potential therapeutic effects of the major purine-metabolizing enzymes in ovarian cancer, including CD39, CD73, adenosine deaminase, adenylate kinase, hypoxanthine guanine phosphoribosyltransferase, inosine monophosphate dehydrogenase, purine nucleoside phosphorylase, dihydrofolate reductase and 5,10-methylenetetrahydrofolate reductase. Purinergic signaling is also described. We then provide an overview of the application of purine antimetabolites, comprising 6-thioguanine, 6-mercaptopurine, methotrexate, fludarabine and clopidogrel. Finally, we discuss the current challenges and future opportunities for targeting purine metabolism in the treatment-relevant cellular mechanisms of ovarian cancer.

Graphical Abstract

Introduction

Ovarian cancer (OC) is the seventh most common cancer and the fifth leading cause of cancer-related death among women worldwide, with a 5-year relative survival rate of 49% [1, 2]. Primary debulking surgery and adjuvant platinum-based chemotherapy are the first-line standard-of-care treatments [3]. More than 70% of patients experience will relapse after first-line treatment and acquire drug resistance, which highlights the need for novel treatment options.

As one of the most abundant components in organisms, purine, in addition to forming DNA and RNA, is involved in the stabilization of immune regulation and the formation of energy carriers and functions as an essential cofactor in biochemical reactions, thereby influencing the growth of both cancer and non-cancer cells [4, 5]. The metabolic enzymes implicated in purine metabolism cause imbalances in purine pools that interfere with cell proliferation, migration and death [6, 7]. Furthermore, various purine antimetabolites exert antitumor effects through multiple mechanisms, such as direct toxicity, interference with the tumor microenvironment (TME), inhibition of DNA synthesis and interference with DNA damage repair [8,9,10,11,12]. Disorders of extracellular ATP (eATP), extracellular adenosine (eADO) and subsequent purinergic signaling also delineate pro-oncogenic or anti-oncogenic outlines [13]. Overall, purine metabolism is closely related to tumor progression.

This review presents recent reports of major purine-metabolizing enzymes in purine synthetic pathways (Table 1), outlines the multiplicity of purinergic signaling in OC development, gives an overview of the application of purine antimetabolites in OC, and discusses potential therapeutic strategies to target purine metabolism in OC.

Table 1 Summary of the main molecular features of purine-metabolizing enzymes associated with OC

Purine metabolism pathways

The stabilization of purine pools is determined by the balance between the synthesis and degradation of purine nucleotides (Fig. 1). The salvage pathway and de novo pathway are two different pathways for the synthesis of purine nucleotides in mammals. The salvage pathway recycles the degraded purine bases or nucleosides via 5-phosphoribosyl-1-pyrophosphate (PRPP) and catalysis by adenine phosphoribosyltransferase (APRT) and hypoxanthine guanine phosphoribosyltransferase (HPRT). There are other enzymes involved in the recovery of purines. Generally, this simple pathway meets most of the cellular requirements of a large percentage of normal cells with very low energy consumption [44, 45]. Rapidly dividing cancer cells rely more heavily on the de novo pathway, the basis for replenishing purine pools, to meet high energy demands [46]. The de novo pathway is triggered by a dynamic complex called purinosome [47]. Purinosome accumulates in the vicinity of mitochondria and microtubules to accelerate purine nucleotide synthesis by catalyzing the important step PRPP to inosine monophosphate (IMP) [48,49,50]. Two important rate-limiting enzymes, adenylosuccinate synthase (ADSS) and IMP dehydrogenase (IMPDH), catalyze the conversion of intracellular IMP to succinyl-AMP and xanthosine monophosphate (XMP) and then to AMP and GMP which are gradually degraded to xanthine and eventually hydroxylated to uric acid (UA) under the action of xanthine oxidase (XO) [51].

Fig. 1
figure 1

De novo, salvage and degradation pathways of purine nucleotides under the regulation of purine-metabolizing enzymes. The de novo pathway converts PRPP to IMP and, ultimately, GMP and AMP that further involve in nucleotide synthesis. The salvage pathway recovers purine bases and purine nucleosides to generate purine nucleotides. The degraded purine base becomes Xan with eventual conversion to UA. Cyan: de novo pathway; red: salvage pathway; yellow: degradation pathway; gradient color: involved in multiple metabolic pathways; arrows: purine metabolic pathways; squares: purine-metabolizing enzymes involved in related pathways. R-5-P: ribose 5-phosphate; PRPP: 5-phosphoribosyl-1-pyrophosphate; Gln: glutamine; THF: Tetrahydrofolate; Asp: aspartate; Hyp: hypoxanthine; Ino: Inosine; IMP: inosine monophosphate; Xan: xanthine; XMP: xanthosine monophosphate; Gua: guanine; GMP: guanosine monophosphate; Ade: adenine; Ado: adenosine; AMP: ado monophosphate; SAMP: succinyl-AMP; UA: uric acid; PPAT: phosphoribosyl pyrophosphate amidotransferase; IMPDH: IMP dehydrogenase; GMPS: GMP synthase; ADSS: adenylosuccinate synthase; ADSL: adenylosuccinate lyase; HPRT: Hyp Gua phosphoribosyltransferase; APRT: Ade phosphoribosyltransferase; ADA: Ado deaminase; AK: adenylate kinase; PNP: purine nucleoside phosphorylase; XO: xanthine oxidase

Purine-metabolizing enzymes and mechanisms in OC

Ectonucleoside triphosphate diphosphohydrolase and ectosolic-5′-nucleotidase

Ectonucleoside triphosphate diphosphohydrolase (CD39, EC 3.6.1.5) and ectosolic-5′-nucleotidase (CD73, EC 3.1.3.5) are two membrane-bound ectonucleotidases. CD39 is the rate-limiting enzyme for the continuous dephosphorylation of eATP to extracellular AMP (eAMP) [52]. The 5'-nucleotidase activity of CD73 limits the rate of hydrolysis from eAMP to membrane-permeable eADO (Fig. 2) [53]. The eADO creates a highly immunosuppressive microenvironment by inhibiting the cytotoxicity of CD8 + T cells and NK cells while increasing activation of Treg cells and M2 macrophages [54,55,56,57]. This implies the balance between eATP and adenosine, which is jointly maintained by CD39 and CD73, is closely related to the immune-suppressive tumor microenvironment.

Fig. 2
figure 2

CD39 and CD73 in TME of OC. CD39 and CD73 localized on the surface of OC cells inhibit immune responses mediated by T cells, MDSC, and TAM in TME, and also induce cisplatin resistance. CD39 and CD73 dephosphorylate eATP to eAMP, ultimately converting it to eAdo. STAT3 induces cell surface acquiring CD39 in TME to promote immunosuppression. Metformin facilitates AMPKα phosphorylation and inhibits the HIF-α pathway to block the immunosuppression caused by high expression of CD39 and CD73 on MDSC. MDSC: myeloid-deriver suppressor cell; TAM: tumor-associated macrophage; TME: tumor microenvironment; eATP: extracellular ATP; eAMP: extracellular AMP; eAdo: extracellular Ado

Traditionally, CD39 is considered as a contactor between immune cells. Tumor-reactive T cells can be identified by the expression of CD39 alone or the co-expression with CD103 [58, 59]. In recent years, it has been shown that CD39 is widely expressed in multiple human cancers, and high-level CD39 is found in M2-polarized tumor-associated macrophages (TAMs) in OC tissues [10, 14, 15, 60, 61]. STAT3 induces the acquisition of CD39 on cell surface in TME to suppress T cell response [62]. Noteworthy, it is IL-27 that mediates the immunosuppression with CD39 involvement, whose polymorphisms is related to the susceptibility to OC [10, 63]. Furthermore, CD73 is expressed on the surface of OC, and high-level CD73 appears to be significantly associated with poor prognosis in patients with high-grade serous OC, which, as mentioned previously, is probably because that CD73 indirectly suppresses CD8 + T cells and promotes immune escape [16].

Surprisingly, the study by Hoon Kyu Oh et al. found that patients with CD73 overexpression present more frequently with low stage, moderate differentiation, no lymph node metastasis, and negative cytologic results, which means that more CD73 may lead to a better prognosis [17]. A study regarding ovarian tumor-initiating cells found that CD73 is essential for OC initiation and growth, and is able to regulate tumor-initiating cells at the transcriptional level to promote expression of epithelial–mesenchymal transition (EMT)-related genes [64]. Besides, blocking CD73 reverses drug resistance in cisplatin-resistant OC cells [65]. Studies on other human tumors have reported that AKT signaling plays a role when CD73 promotes tumor progression and metastasis [66,67,68,69]. However, more mechanistic studies are needed in OC.

Current researches have noted that the anticancer effect of CD39 or CD73 inhibitors depends overwhelmingly on relieving T cell targeted immunosuppression and additionally on myeloid-deriver suppressor cells (MDSCs), which suggest CD39 and CD73 may serve as potential therapeutic targets for OC treatment [70,71,72,73]. A study found the antitumor activity can be mediated by blocking CD39 via eATP-P2X7-ASC-NALP3-inflammasome-IL18 pathway [74]. Besides, CD39 inhibitor POM-1 partially relieves the immunosuppressive function of TAMs [10]. As we would expect, anti-CD73 antibody reverses the immunosuppressive environment developed by docetaxel in OC but the inhibition of CD73 alone rather than CD39 is less effective [75]. Thus, the protocol for co-suppressing CD39 and CD73 is expected. An important finding is that metformin, the first-line drug for type-2 diabetes, blocks CD39 and CD73 by increasing AMPKα phosphorylation and inhibiting HIF-α pathway to interrupt immunosuppression caused by MDSC as well as enhance the anti-tumor activity of CD8 + T cells [14]. In addition, a pivotal immunosuppressive factor programmed death-1 receptor (PD-1) encourages tumor spread through interfering protective immunity [76]. Anti-PD-1 combined with anti-CD39 or anti-CD73 demonstrates a more pronounced slowing of tumor growth [74, 77]. In any case, there is still a long way to go to clarify the specific roles and mechanisms of CD39 and CD73, and deep studies will provide new insights into tumor immune networks.

Adenosine deaminase

Adenosine deaminase (ADA, EC 3.5.4.4) catalyzes the irreversible hydrolytic deamination of adenosine and deoxyadenosine, the final products of which are inosine and deoxyinosine (Fig. 3) [18, 78, 79]. There are two isozymes of human ADA. ADA1 regulates adenosine concentration in the intracellular and interacts extracellularly with the adenosine receptor or dipeptidyl peptidase 4 (DDP4, EC 3.4.14.5) on the surface of immune cells [80, 81]. DDP4, expressed as a type II transmembrane protein, plays a pro-tumor role in a variety of human tumors and has the potential to act as a positive prognostic predictor [82]. It was reported that the interaction between ADA and DDP4 on the cell surface could regulate T cell activation and lymphocyte-epithelial cell adhesion [18, 83]. ADA2, with low abundance in humans, is mainly found in serum and its deficiency is associated with autoinflammatory diseases [84,85,86,87].

Fig. 3
figure 3

Role and mechanism of ADAR, ADA and its receptor DDP in OC. ADAR mediates A to I RNA editing to elicit CD8 T cell response and interferes with HMGA1 via miRNA Let-7d acting on OC apoptosis and chemotherapy sensitivity. ADA enhances the immune potency of TAM in TME with the capacity to convert Ado to inosine. DDP4, an important receptor for ADA, facilitates the migration, invasion and adhesion to mesothelial cells of OC. The DDP inhibitor Sitagliptin, increases caspase 3/7 activity to induce OC apoptosis on the one hand, and maintains the effect of paclitaxel on OC apoptosis via ERK and Akt pathways on the other hand

ADA level is significantly elevated in the serum and peritoneal fluid of OC patients and is positively correlated with pathological subtype and grade [19]. This may be a self-rescue measure in response to apoptosis caused by accumulation of toxic substrates adenosine and deoxyadenosine. D'Almeida SM et al. found that ADA, similar to CD39 inhibitors, partially blocked immunosuppression of TAMs [10]. At the same time, as a receptor of ADA, DDP4 is overexpressed strongly and the expression level is associated with FIGO stage or lymph node metastasis [88]. Besides, DDP4 increases the adhesion potency of OC to mesothelial cells with the participation of fibrin, and blocking DDP4 significantly inhibits cancer migration and invasiveness [89]. Sitagliptin, a selective DDP4 inhibitor, increases caspase 3/7 activity in OC by co-treatment with paclitaxel, and maintains apoptosis induction through the ERK and Akt signaling pathway [90, 91]. These studies imply that the application of ADA and DDP4 inhibitors seems to favor tumor-cell death. However, it is interesting to note that the addition of ADA reversed the decreased metastatic capacity of OC caused by adenosine [92]. Similarly, overexpression of DDP4 leads to enhanced chemosensitivity of OC cells to paclitaxel and reduced tumor cell invasiveness due to aberrant expression of E-cadherin, MMP-2 and TIMP [93,94,95]. These puzzling results suggest that ADA and DDP4 still need substantial and in-depth studies and their interaction could act as potential biomarkers or therapeutic targets in OC.

Adenosine deaminases acting on RNA(ADAR, EC 3.5.4.37)catalyzes the deamidation of adenosine on dsRNA to inosine, which means A-to-I RNA editing (Fig. 3) [96]. There are three members of the ADAR family: ADAR1 and ADAR2 are primarily used for RNA editing, and ADAR3 inhibits their editing activity by binding to substrates of the first two [97,98,99]. ADAR-triggered aberrant RNA editing regulation to promote tumor growth is widely observed in a variety of human solid tumors [100, 101]. The frequent over-editing of Cyclin I (CCNI) by ADAR1 in OC was found to produce peptide products that activate T-cell responses and specifically kill tumor cells [21]. Marek Cybulski et al. detected that the immune responses to CCNI in the nucleus and cytoplasm of epithelial OC are related to abnormal cell cycle but not to chemosensitivity [102]. Furthermore, ADAR1 is reported to induce tumor progression by impairing Let-7d in malignant tumors [103]. Studies have found that miRNA Let-7d plays a suppressive role in a variety of human tumors and its overexpression may inhibit HMGA1 by regulating the p53 signaling pathway to enhance chemosensitivity [104,105,106,107]. These studies provide us with inspiration for gene editing to regulate tumor progression. Overall, this is a promising therapeutic concept, but the specific mechanisms still need to be explored more deeply.

Adenylate kinase

Adenylate kinase (AK, EC 2.7.4.3) catalyzes the reversible transfer of phosphate group from ATP to AMP in purine synthesis pathway. AK-induced AMP signal changes affect adenosine pools status and energy information through metabolic sensors, which subsequently balance ATP level to regulate energy metabolism [108, 109]. AK family isozymes AK1-9 have been identified in human tissues [110], and existing studies have reported that AK1, 2, 4, 6 and 7 subtypes function in the regulation of tumor growth, metabolism, energy allocation and invasion [111,112,113,114]. So far, researches on AK in OC have been focused on AK4 and AK7.

AK4 is mainly expressed in kidney, heart and liver tissues, probably caused by the high-mitochondrial content. Significantly elevated level of AK4 is also found in some highly aggressive tumors [115,116,117]. Research has revealed that AK4 stabilizes the purine nucleotide pools and balances energy through the regulation of AMPK [118], and is able to interact with HIF-1 to regulate mitochondrial activity and enhance cellular hypoxia tolerance by promoting hypoxia [116, 119]. Cellular drug resistance is enhanced due to the progress in resistance to hypoxia. The interaction of hypoxia on AK4 level is different in diverse cell lines [120], however, it is clear that changes in AK4 level are able to protect cells from environmental stimuli, and blocking AK4 can resist the protective mechanism of cells against hypoxic stress [119]. AK4 can be found in oocytes, follicular epithelial cells and corpus luteum cells in normal ovary tissue [121]. AK4 is overexpressed in OC and the expression level is significantly correlated with tumor size and FIGO stage [26]. A meaningful finding is that upregulation of miRNA-3666 inhibits OC cell proliferation and migration by blocking the STAT3/AK4 axis and inducing apoptosis simultaneously [27]. Besides, AK4 favors tumor development and metastasis in an ATF3-dependent manner [122], which interestingly plays contradictory roles of either inducing apoptosis or promoting proliferation in different types of OC [123, 124].

AK7 is mainly expressed in cilia-rich sites such as respiratory tract and fallopian tube [109, 110]. Ciliary structural abnormalities caused by AK7 damage often led to primary male infertility or primary ciliary dyskinesia [125,126,127]. Abnormal activation or loss of primary cilia will affect the progression and prognosis of diverse tumors. Study reported that cilia-related genes are lowly expressed in glioblastoma, breast cancer, OC and colon adenocarcinoma; in contrast, they are overexpressed in clear cell renal cell carcinoma, rectal adenocarcinoma, lung adenocarcinoma and lung squamous cell carcinoma [128]. It is well known that cilia play a key role in assisting transport and picking in the female reproductive system. Furthermore, some studies support the ability of cilia as angiopoietin (Ang) sensory organ to maintain the morphological and motor homeostasis of ovarian tissue [129, 130]. Recent studies support the standpoint that the majority of OC originate from the cilia-rich fallopian tube [131] and the Ang/Tie signaling pathway associated with cilia formation may be relevant to OC development and formation [129, 132] and the expression of Ang2 is closely related to angiogenesis in OC [133]. Thus, it is reasonable to speculate that AK7 may partially influence the ability of cilia to intervene in OC progression. Zhang et al. found that AK7 level in OC is significantly reduced by analysis of the TCGA database, the degree of which is positively correlated with tumor stage [25]. And this phenomenon is mainly related to conduction pathways such as EMT, TGF-b signaling and UV response. More importantly, OC patients with lower AK7 expression have a worse prognosis. These studies not only suggest that AK7 exhibits potential as a prognostic indicator for OC but the possibility of future treatment by elevating AK7 expression level or interfering with cilia activity.

Hypoxanthine guanine phosphoribosyltransferase

HPRT (EC 2.4.2.8) is one of the classical enzymes of the purine salvage pathway, which utilizes the transfer of ribose phosphate from PRPP to form IMP and GMP for DNA synthesis and repair [134]. As a housekeeping gene, HPRT is maintained at a low level in all somatic cells except the central nervous system [135]. HPRT deficiency leads to failure of the salvage for hypoxanthine and guanine, increasing purine degradation and UA production and triggering a series of diseases: partial deficiency causes gout-like symptoms, while complete deficiency results in Lesch-Nyhan syndrome [136]. Although other enzymes have complementary roles, these symptoms are unique to HPRT deficiency [134].

HPRT is closely associated with tumor development and is overexpressed in various malignancies. However, it has been reported that HPRT level is fluctuating in OC but stable in borderline ovarian tumor and normal ovarian tissue, making it a suitable reference gene for biochemical processes [33, 137]. HPRT performs a potential auxiliary role in DNA mismatch-repair. The reason for mismatch is that most DNA polymerases preferentially pair 7,8-dihydro-8-oxoguanine with adenine rather than cytosine during DNA oxidation [138]. The MutY DNA glycosylase homologue recognizes and repairs mismatches, whose level changes lead to opposite changes in the mutation rate of HPRT in OC cell line A2780 [139]. It implies that high-level HPRT mutation is associated with DNA mismatch activity. Nevertheless, HPRT mutation rate is reduced in OC with defects in DNA mismatch repair gene hMSH2 [140]. Also, HPRT mutation may be in connection with tumor risk and chemotherapy resistance in OC [141,142,143,144,145]. These contradictory results suggest that more research is needed to elucidate the relationship between HPRT mutation and DNA mismatch repair and effects on tumors. In general, current research on HPRT is focused on report gene, while tumor-related research on HPRT is relatively stagnant. In-depth investigations are still needed to explore its specific effects on tumors and therapeutic strategies.

Inosine monophosphate dehydrogenase

The catalytic oxidation of IMP to XMP by inosine monophosphate dehydrogenase (IMPDH, EC 1.1.1.205) in cytoplasm is the NAD + -dependent rate-limiting step in de novo synthesis of GTP [146]. GTP, a classical signaling molecule that regulates various cellular activities as well as an energy supplier for protein synthesis, is elevated in a variety of tumor cells [147]. Although HPRT recovers guanines in salvage pathway, it is unable to meet the demand of malignant tumor cells. Elevated IMPDH activity enhances the ability of cancer cells to synthesize GTP, which provides raw material to supply rapid proliferation.

Abnormally high level of IMPDH is found in multiple malignancies, including OC [28, 148, 149]. Among the two isoforms of human IMPDH, IMPDH2 rather than IMPDH1 is upregulated in a variety of human tumors [29, 149,150,151,152,153]. Researches have reported that high expression of IMPDH2 in OC is associated with different tumor types, lower survival rates and higher stage, which implies a poorer prognosis [29]. IMPDH2 activates PI3K/Akt and Wnt/β-catenin pathways to promote progression or drive EMT processes in diverse cancers [28, 148]. The abnormally increased activity makes IMPDH, especially IMPDH2, a new target for anti-cancer drug development.

Indeed, the use of IMPDH inhibitor as an antineoplastic agent has a long history [154]. A few studies have linked IMPDH inhibitor to apoptosis: downregulation of MEK/ERK pathway leads to Bcl-2 inhibition [155]; downregulation of the Src/PI3K/Akt pathway with inhibition of mTOR to activate Bax and Bak; inhibition of mTORC1 or c-Myc signaling reduces ribosomal RNA synthesis [147, 156, 157]; drive of mitochondria-dependent mechanisms causes apoptosis [158]. Katherine Y et al. found that the addition of thiazole-4-carboxamide adenine dinucleotide, an intracellular active metabolite of tiazofurin, is effective in inhibiting IMPDH activity in OC [28]. Utilizing the IMPDH inhibitor benzamide riboside (BR) makes it possible to observe apoptosis induced by cell cycle arrest due to c-Myc and downstream Cdc25A inhibition [28, 30, 159]. Unfortunately, BR-induced apoptosis is only seen in partial OC cell lines, and the specific regulatory mechanisms between IMPDH inhibitor and apoptosis remain unknown. Furthermore, the development and application of IMPDH inhibitor has been limited by unstable effects, adverse effects at high doses, and discrepancies in IMPDH levels in different tumors [154]. For now, it is still of much value to explore IMPDH such as immunosuppression or biomarkers and this may be an opportunity to reactivate it as an antitumor agent.

Purine nucleoside phosphorylase

Purine nucleoside phosphorylase (PNP, EC 2.4.2.1), which catalyzes the reversible phosphorylation of adenosine, guanosine and inosine, is an important enzyme in salvage and degradation pathway [32]. Among them, homologous E. coli PNP (ePNP) rather than human PNP is able to enrich purine pools with adenosine as a substrate [160]. Gene directed enzyme prodrug therapy (GDEPT), also known as suicide gene therapy, relies on transgenic methods to encode enzymes that convert non-toxic drug precursors into active toxic metabolites to target tumor cells injury [161]. The anti-tumor effect of selective expression of the suicide gene PNP has been observed in a variety of malignancy researches (Fig. 4) [162, 163]. PNP-GDEPT exerts anti-tumor activity in a unique way compared to others. It cleaves 6-methylpurine-2'-deoxyriboside and fludarabine phosphate to toxic purine analogs 6-methylpurine and 2-Fluoroadenine, which inhibits nucleic acid and protein synthesis [164]. These toxins are then released into extracellular matrix and remain active [32]. That is, toxic metabolites can spread to surroundings to achieve significant bystander killing effect even if only few tumor cells express ePNP, which suggests a potential chemotherapeutic advantage [165, 166].

Fig. 4
figure 4

Application of PNP-GDEPT in OC. PNP cleaved MePdR and Fludarabine phosphate to the toxic products MeP and 2-FA. Implementation of GDEPT using ePNP or adenovirus-mediated PNP is able to induce apoptosis in OC cells and exert the bystander effect. PNP-GDEPT acts synergistically with docetaxel and cisplatin and high-body-temperature environment enhance the expression efficiency of ePNP. GDEPT: gene directed enzyme prodrug therapy; ePNP: E. coli PNP; MePdR: 6-methylpurine-2’-deoxyriboside; MeP: 6-methylpurine; 2-FA: 2-Fluoroadenine

A few attempts of this approach have been tested in OC cells. V K Gadi et al. have successfully modeled ePNP bystander killing which exerts significant antitumor effects of OC in vitro and in vivo [166]. Other researchers reported that adenovirus-mediated PNP-GDEPT caused upregulation of Bax, Bik and Bok and downregulation of Bcl-2 [167]. Meanwhile, there was synergy between PNP-GDEPT and docetaxel or carboplatin on platinum-resistant OC cells, implying the potential for reversing drug resistance. Unfortunately, this approach exhibited a low success rate. In this regard, researchers combined PNP-GDEPT with high-body-temperature environment created by human telomerase reverse transcriptase and heat shock elements to improve PNP expression efficiency [168]. Thus, while novel evidence is currently limited, these investigations may provide a new strategy for GDEPT in OC.

Other purine-metabolizing enzyme

Dihydrofolate reductase (DHFR, EC 1.5.1.3) and 5,10-methylenetetrahydrofolate reductase (MTHFR, EC 1.5.1.20) are not conventional purine metabolizing enzymes as they control the metabolism of purines and pyrimidines indirectly by regulating folate synthesis (Fig. 5). Folate acts as a raw material and coenzyme as well as a methyl donor in the biosynthesis of nucleotides [169]. DHFR reduces dihydrofolate to tetrahydrofolate (THF) which subsequently acquires one-carbon units from amino acids to install methyl groups [170]. MTHFR continues to convert methylene-THF to methyl-THF for further participation in nucleic acid biosynthesis. In fact, methyl-THF, methylene-THF and formyl-THF are interconverted to provide one-carbon units for methionine cycle pathway, thymidylate synthesis pathway and purine synthesis pathway in turn [171]. Studies have reported that folate-mediated one-carbon metabolism plays an important supporting role for rapidly proliferating cells, especially tumor cells [169] and DNA methylation is expected to be used as an early diagnostic marker for various malignancies [172,173,174]. Overall, folate metabolism, the upstream of purine metabolism, serves an important role in tumor development.

Fig. 5
figure 5

Role of DHFR and MTHFR in OC. DHFR and MTHFR are involved in the formation of important one-carbon units for purine metabolism. DHFR promotes drug resistance and inhibits omentum metastasis, while resisting apoptosis caused by TMZ through AMPK pathway activation and mTOR pathway inhibition. Berberine, PTX, MTX, and some quinoxalines (453R&311S) have been found to act as DHFR inhibitors. MTHFR inhibits FBP expression and enhances drug sensitivity, which is inhibited by HOTAIR. TMZ: temozolomide; PTX: pemetrexed; MTX: methotrexate; 453R: 3-methyl-7-trifluoromethyl-2(R)-[3,4,5-trimethoxyanilino]-quinoxaline; 311S: 3-piperazinilmethyl-2[4(oxymethyl)-phenoxy]-quinoxaline; FBP: folate binding protein

Multiple studies have shown that the overexpression of DHFR is associated with platinum resistance in OC [36,37,38]. Interestingly, Jia Chen et al. reported that DHFR expression is increased in benign ovarian tumor but decreased in malignant OC with a significant correlation with omentum metastasis [39]. In this case, however, DHFR expression is still significantly higher in platinum-resistant patients than platinum-sensitive patients, highlighting the potential diagnostic value of DHFR in chemoresistance. Actually, the commonly used chemotherapy drug methotrexate (MTX) targets DHFR to block THF production to inhibit rapidly proliferating tumor cells [175]; however, given its toxicity and resistance, there has been a search for new DHFR inhibitors. Pemetrexed, a DHFR and thymidylate synthase inhibitor, activates AMPK and inhibits mTOR signaling pathway by synergizing with temozolomide (TMZ) to obstruct tumor growth [176]. Berberine disturbs folate-metabolizing enzymes including DHFR to reduce the viability of OC cells, especially platinum-resistant cells. In addition, there is evidence that some compounds with quinoxaline structures are not cross-resistant with cisplatin and show remarkably inhibition on OC growth, demonstrating the potential as DHFR inhibitors [37, 177].

As a rate-limiting enzyme for folate metabolism, MTHFR with low viability slows down the synthesis and repair of nucleotides and disrupts the methylation of homocysteine [178, 179]. Defective MTHFR in OC combined with overexpression of high-affinity folate-binding proteins lead to increased folate uptake, implying that low-activity MTHFR is a potential risk factor for OC [40]. Current studies have found that MTHFR single nucleotide polymorphisms (SNPs), particularly C677T polymorphism which affects enzyme activity, are likely to increase tumor risk [41, 180]. Many studies on the association of C677T with susceptibility and risk of OC in Asians rather than whites have been performed [181,182,183,184]. However, no effect of C677T polymorphism in Asians was found in a 2012 meta-analysis, possibly because of the limitation of sample size [42]. A possible association between high folic acid intake and low patient survival was reported in research by S C Dixon et al. [43]. However, an interesting finding is that a reduced risk of OC in Chinese with high folate intake was more pronounced in MTHFR 677CC mutation, but no such protection was observed in Australians [185, 186]. This discrepancy could be attributed to the limitations of the sample and ethnicity, emphasizing the need for more research. Recently investigators have examined the effect of MTHFR on the efficacy of 5-fluorouracil (5-FU) chemotherapy in OC patients. Silencing HOX transcript antisense RNA elevates the sensitivity to 5-FU due to decreased MTHFR methylation [187]. Besides, MTHFR SNP has been found correlated with prognosis or hematologic toxicity in 5-FU-treated patients in other human tumors [188,189,190]. Yet research on this link in OC has remained an under-explored domain and more in-depth explorations are needed.

Purinergic signaling in OC

Key agonist, eATP and eADO

The extracellular purine, mainly eATP and eADO, has active traffic with intracellular signals. In normal physiological environments, high eATP and low eADO levels are maintained through exocytosis or vesicles [191]. Nevertheless, the overall level is subject to strict limitations [192]. The eATP concentration increases dramatically with intracellular ATP in disordered TME (even up to two times the intracellular concentration), which may be due to active release from tumor cells or immune cells, or to abnormal metabolites. As previously described, CD39 and CD73 contribute to the progressive dephosphorylation of eATP to eADO. Continuous stimulation of ATP causes active phospholipase D to promote ATP degradation [193]. Meanwhile, hypoxia effectively stimulates adenosine release [194,195,196], possibly due to HIF-1α-induced adenosine kinase inhibition that impedes the conversion of ADO to AMP and thus promotes its efflux [197], or possibly as a consequence of retaliatory metabolism [196]. The cascade response results in abnormal levels of both intra- and extracellular ATP and ADO to regulate tumor development via purinergic signaling as autocrine or paracrine messengers [13].

Generally, eATP and eADO are considered as important signals to regulate immunity with pro-inflammatory and anti-inflammatory effects, respectively [198, 199]. The sharp increase in eADO hinders immune cell activation to exert pro-cancer effects and reduces survival of OC patients [200,201,202,203]. This effect is closely related to purinergic receptors, which will be described later (Fig. 6). For other malignant features of ovarian cancer, eATP and eADO also play see-saw roles. The eATP activates corresponding purinergic receptors and its co-expressed KCa3.1 channel, triggering subsequent complex electrical membrane responses that contribute to cancer migration [204]. Conversely, eADO supplementation or its analogs have been reported to inhibit the ability of OC to invasion or angiogenesis, partially dependent on the promotion of RhoGDI2 subsequently related gene expression [205, 206]. The effect of adenosine analog supplementation on OC chemoresistance likewise depends on the type of receptor activated [207]. Notably, low doses of adenosine were associated with G0/G1 phase arrest, whereas high concentrations of adenosine induced both early and late apoptosis in a dose-dependent manner by activating the Bcl-2/Bax and caspase-3 pathway in OC cells [208]. These complex effects shift our thinking to future studies of eATP/eADP ratios rather than single substance levels.

Fig. 6
figure 6

Purinergic signaling pathway in OC. Extra- and intracellular adenosine and ATP are key agonists. Purinergic receptors are expressed in a variety of cells in OC TME. Activation or antagonism of these receptors, as well as interaction with other signaling will ultimately affect the progression and malignant features of OC

Purinergic receptors and their therapeutic potential

Previous studies have shown that purinergic receptors are expressed in almost all immune cells, and an increasing number of studies have identified the presence of purinergic receptors in tumor cells [209]. The identified purinergic receptors include P1 receptors that mediate signals triggered by adenosine and P2 receptors that respond to adenine and uridine nucleotides. In general, P1 receptors and P2 receptors mediate pro-inflammatory and anti-inflammatory responses, respectively, which should actually be more dependent on specific receptor types and signaling transduction ratios [210].

P1 receptors

P1 receptors, also known as adenosine receptors, are divided into four subtypes, A1R, A2AR, A2BR and A3R, of which A2R is a key factor in protecting tissues from excessive immune responses and commonly expressed in a variety of human tumours [211,212,213,214]. A2AR is a high-affinity receptor involved in the regulation of T-cell function to promote immune evasion, and inhibits the secretion of defensive substances by neutrophils to attenuate the inflammatory response [54, 215]. A2BR appears to have low affinity and is more expressed in macrophages and dendritic cells. Another study found that activation of A2BR was able to interfere with TME via MDSCs to promote cancer growth [216]. Besides, the high-affinity receptors A1R and A3R are considered as immune-promoting adenosine receptors, possibly because of promoting IL-10 expression or inhibiting cAMP production [217, 218]. Therefore, adenosine receptors deserve to be another target of intervention in tumor immunotherapy and hold considerable promise.

A2BR and A3R are the predominantly expressed adenosine receptors in OC [219]. It has been reported that activation of A2BR significantly increases cAMP levels in OC, while activation of A3R exerts the opposite effect, and the other two receptors do not affect cAMP levels, implying that A2BR and A3R are major functional adenosine receptors in OC. In addition, A2BR activation significantly inhibits the migration ability of OC and maintains the epithelioid phenotype [220]. Kaplan–Meier survival analyses showed that OC patients with low A2BR expression levels, especially with early-stage OC, had shorter OS. Interestingly, however, Hajiahmadi et al. found that either the A2BR agonist NECA or the A3R agonist IB-MECA dose-dependently inhibited OC cell viability in a manner that related to the loss of mitochondrial membrane potential to activate apoptosis [221, 222]. Although these findings reflect the potential OC-inhibitory effect of activating A2R and A3R, changes in cAMP, which link essential malignant features such as invasion, metastasis, apoptosis resistance and chemoresistance, are not mentioned [223,224,225]. The non-dominantly expressed adenosine receptor, A2AR, plays a role in assisting chimeric antigen receptor (CAR) T cells from hindrance by the immunosuppressive microenvironment [226]. Liu G et al. found that anti-mesothelin CAR T cells released more TNF-α and IFN-γ after suppressing A2AR, and extremely enhanced anti-OC efficacy, which provided a new perspective for OC treatment [227]. It should be emphasized that the complex regulatory mechanisms and effects are still not completely clear because different cell status, receptor types, and environmental conditions may cause influences, and further studies to identify the pro- or anti-cancer roles played by adenosine receptors and the specific mechanisms are urgent and necessary.

In fact, targeting adenosine receptors has shown anticancer therapeutic potential in OC. Theobromine, a non-selective adenosine receptor antagonist, inhibits OC angiogenic activity by reducing vascular endothelial growth factor production [228]. Further studies revealed that the inhibition of angiogenesis was associated with A2BR interaction of CD45 lymphocytes [229, 230]. However, blind application of the adenosine analogue ZM241385, an A2R antagonist, may increase the resistance to cisplatin [207], and the A1R antagonist PSB36 brought a sensitizing effect. Overall, activation of adenosine receptors can be anti- and pro-cancer, and the opposing efficacy of their dual action may bring valuable insights for future clinical therapeutic interventions with further studies.

P2 receptors

P2 receptors consist of two subfamilies, P2XRs and P2YRs. P2XRs are ATP-gated non-selective cationic conductance channels with seven isoforms (P2X1-7R) that are expressed in a variety of tumor cells, immune cells, and stromal cells [209]. The eATP is a recognized agonist, and some non-nucleotide compounds have been reported to modulate it as well [231,232,233]. P2XRs, especially P2X7R, are the ones that have been focused on for their multiple roles in mediating tumor growth, metastasis, invasion, drug resistance, and death, in addition to pro- and anti-inflammatory effects [234,235,236,237,238]. Overexpression of P2X7R in OC contributes to cell proliferation and viability [239, 240]. Vázquez-Cuevas FG et al. found that activation of P2X7R caused an increase in intracellular Ca(2+) concentration and phosphorylation of ERK and AKT, but did not cause apoptosis [241]. Notably, eATP, or actually mainly the activation of P2X7R, promotes NLPR3 inflammasome activation and assembly, followed by activation of caspase 1, which is dependent on ATP induction and subsequently K(+) efflux [75]. Meanwhile, P2X7R drives IL-1β maturation in response to activation of the NLRP3 inflammasome [242]. It was shown that excessive NLRP3 in OC promotes EMT [243, 244] and mediates gemcitabine resistance [245] through the Wnt/β-catenin signaling pathway, and is associated with poorer OS [246]. Once NLRP3 is absent, the expression of P2X7R will be promoted, which contributes to cancer growth as a negative feedback loop [247]. Co-localization of P2X7R/NLRP3 in the adipocyte plasma membrane of omental tissue of OC patients implies its possible contribution to tumor metastasis [248]. However, it is noteworthy that antagonizing P2X7R inhibits pyroptosis via NLRP3/caspase1 and promotes apoptosis mediated by Bcl-2/caspase9/caspase3 pathway at the same time [249]. In addition, P2X7R stimulation in alternative activated M2 macrophages has been reported to release abundant anti-inflammatory proteins, suggesting a contribution to the inflammation resolution [238]. The P2X7R/NLRP3 complex lacks detailed studies in OC, and the pro- and anti-inflammatory and even pro- and anti-cancer responses it mediates are incompletely understood as well. The balance between mutually inhibited pyroptosis and apoptosis may bring a fresh perspective to precisely fight OC.

P2YRs are G protein-coupled and contain a total of eight isoforms (P2Y1R, P2Y2R, P2Y4R, P2Y6R, P2Y11R, P2Y12R, P2Y13R and P2Y14R) in mammals. Compared to P2XRs, P2YRs are more sensitive to slight changes in local nucleotide or agonist concentrations [250]. These members are activated by different nucleotides, and only P2Y2R and P2Y11R are responsive to ATP [251]. The prominent contribution of P2YRs in tumor growth and metastasis cannot be ignored, which is associated with the regulation of intracellular Ca(2+) concentration and subsequently cAMP changes. In this regard, P2Y1R, P2Y2R, P2Y4R, P2Y6R bind to Gq-dominated protein subunits to activate PLCβ/IP3/DAG signaling pathway to increase intracellular Ca(2+) concentration, which may be affected by the acidic intracellular environment [251, 252]. P2Y12-14R couple to Gi/o to inhibit adenylate cyclase activity to reduce intracellular cAMP levels. As previously described, low levels of ATP stimulate P2Y2R, and activate co-localized KCa3.1 channels to promote migration of OC cells [204]. Martínez-Ramírez AS et al. suggested that the contribution of P2Y2R to OC cell migration may derive from an interaction with EGFR [253]. In addition, there is an increasing number of studies reporting the special status of platelets in OC. OC cells induce platelets activation, and in turn, platelets stimulate OC growth [254, 255]. Cho MS et al. revealed that activation of P2Y12R on platelets by ticagrelor contributes to OC growth, for which eADP secreted by OC cells is an important activator [256]. P2Y2R and P2Y12R may play as pivots in OC progression, however, more specific mechanisms need to be further investigated.

Altogether, the understanding of purinergic receptors in OC is currently incompletely clear. More in-depth studies need to be conducted to map a detailed purinergic signaling network in OC to provide novel insights for precise treatment.

Purine antimetabolites

Purine antimetabolites are one of the classical approaches to antitumor, which are chemical analogues of purine metabolism substrates and block purine metabolism by two mechanisms: mimic physiological substrates and compete for the same metabolic enzyme binding site to interfere with biochemical reaction rates resulting in reduced production of normal metabolites [257]; bind to the active site and generate inactive or even toxic metabolites which cause DNA damage and induce apoptosis [258]. Purine antimetabolites are mainly classified into thiopurines and purine deoxynucleoside analogues according to the structure and mechanism [257]. Current researches on OC focuses on 6-thioguanine (6-TG), 6-mercaptopurine (6-MP), MTX and fludarabine (Table 2). Besides, here we classify clopidogrel, which targets purinergic receptors, as an atypical purine antimetabolite.

Table 2 The basic information and effects of 6-thioguanine, 6-mercaptopurine, methotrexate, fludarabine and clopidogrel in OC

6-TG is a guanosine analogue that is converted by HPRT to toxic 6-thioguanine nucleotides, which exert pharmacological effects by binding to DNA [11]. It has been suggested that 6-TG may be an effective therapeutic agent for BRCA-deficient tumors, especially for platinum-resistant and PARP inhibitor-resistant tumors [259]. This is due to the fact that homologous recombination is reactivated in resistant cells without complete recovery from the damage caused by 6-TG. Unfortunately, long-term treatment is not recommended because of its hepatotoxicity [270]. 6-MP is a structural analogue of hypoxanthine with a similar mechanism as 6-TG and a lower toxicity [271]. After treatment with 6-MP and MTX for two months, 30% of OC patients with BRCA mutations showed stable disease status afterwards, and 14% showed longer-term clinical benefit [260]. Nevertheless, neutropenia and anemia are the most common adverse effects. A phase II clinical trial continues to evaluate the safety of 6-MP, which may provide an attack on BRCA-deficient OC [270]. The mechanism of MTX resistance to purine metabolism was described above. MTX was recently reported to induce apoptosis by increasing ROS, inducing DNA damage and modulating mitochondrial membrane potential [261]. Oral low-dose MTX and cyclophosphamide may serve as maintenance therapy after chemotherapy for patients with advanced OC [272]. Furthermore, Courtney A Penn et al. found that therapeutic combination of MTX and nanoparticle targeting TAM for OC exhibited superior activity over cisplatin alone [262]. At the same time, MTX ovarian toxicity cannot be ignored [273]. As for fludarabine, a fluorinated nucleotide analogue of vidarabine, is relatively resistant to ADA inactivation [274]. It is first converted to 9-β-D-arabino-furanosyl-2-fluoradenine (F-ara-A) that can be taken up by cells, and then to F-ara-A triphosphate which inhibits nucleotide reductase, DNA polymerase and DNA ligase, and ultimately causes impaired DNA synthesis and apoptosis [12]. Fludarabine was found to reduce OC migration and adhesion by inhibiting the FAK/STAT1 pathway [263] and was able to inhibit VEGF via Hif-1α and PI3K/AKT signaling pathways to arrest the progression of OC [264]. Moreover, it exerts synergistic effects with cisplatin, seemingly supporting its potential as a modulator of chemotherapeutic agents [265]. Two clinical trials utilize fludarabine as an immunosuppressant to deplete lymphocytes for allogeneic NK cell therapy in OC patients [275, 276]. There were also some important differences in other purine antimetabolites that contribute to the development of OC [277]. These conflicting roles indicate the need for more research and clinical trials on purine antimetabolites in OC. Additionally, interventions on purinergic signaling are the result of atypical purine antimetabolites, which lack sufficient evidence for therapeutic trials in OC. High-dose clopidogrel has been shown to cause high-level P2Y12 blockade [267]. A concern is that clopidogrel may increase the risk of toxicity when used with paclitaxel, as was found in a 60-year-old patient with OC [268]. A higher risk is also seen in the report by Park SH et al. [269]. It remains to be considered whether interventions on purinergic signaling pathways can provide benefits to OC patients.

Conclusions

Homeostasis of purine pools in vitro and in vivo is essential for the maintenance of healthy state and normal function of cells. The depletion of purine in normal cells can be saved by an increase in purine synthesis [278]. In some cases, excessive purine consumption is beyond the capacity of synthesis and is unable to be compensated. For example, excessive DNA synthesis in S1 phase leads to a dramatic increase in purine consumption, or a plethora of purine neurotransmitters are released during the active phase of certain neurons [279]. Imbalance of purine pools contributes to dysregulation of genomic stability and ultimately to metabolic disease or tumor development [280, 281]. Herein, we summarize the role of major purine-metabolizing enzymes, describe purinergic signaling pathways, outline the roles and mechanisms of partial purine antimetabolites and point out potential therapeutic strategies for targeting purine metabolism in OC.

In fact, purine-metabolizing enzymes also participate in other biological processes. For instance, CD39 and ADA are involved in immune regulation [18, 58, 59, 83]; IMPDH carries out the function of a transcription factor [282]; SAM Domain And HD Domain-Containing Protein-1 assists in regulating the cell cycle [283] and phosphoribosylaminoimidazole succinocarboxamide synthetase is correlated with DNA damage repair [284, 285]. It can help to identify new interventions in tumor process if we focus on purine metabolic processes and understand the mechanisms of action of the relevant enzymes as well as the interaction with other biological processes. Additionally, uncontrolled tumor proliferation due to excessive purine synthesis plays a role in chemoresistance, which may be sensitized by inhibitors of purine metabolism [286]. Purine antimetabolites and other anti-purine-metabolizing agents may provide an additional line of attack that would be a promising strategy to overcome tumor resistance in combination with conventional chemotherapy.

However, it cannot be denied that purine metabolism in OC has not been comprehensively studied and is still at a relatively superficial stage. So far, most studies have only briefly described purine-metabolizing enzyme changes and subsequent cell apoptosis, without exploring the relationship and mechanisms between the two in detail. Several studies even showed contradictory treatment results, probably caused by the lack of samples. Evidence above casts doubt on the credibility of the link between the two. Besides, it is difficult to determine whether adenosine disorders are caused by enzyme effect or an action of Adenosine itself, or are related to adenosine receptors. Likewise, it is a critical option to activate or antagonize purinergic receptors in OC. The interaction of purinergic signaling with hypoxia or NLRP3 inflammasomes is also a valuable reflection on how to intervene to obtain maximum benefit. Another limitation is that basic researches on OC treatment options targeting purine metabolizing enzymes or purinergic signaling pathways are currently stagnant and lack the support of clinical research results. Therefore, it is necessary to focus on the association of purine metabolism with other human tumors, which may open up new possibilities for OC research.

Recent studies have reported that multiple kinases interfere with tumor progression by affecting purine de novo pathway via regulation of important transcription factors or intervention with rate-limiting enzymes: DYRK3 regulates ATF4 transcriptional activity and inhibits PPAT to suppress hepatocellular carcinoma proliferation and metastasis [287]; UHMK1 modulates the NCOA3/ATF4 axis and may activates ATIC to promote gastric cancer development [288]; CLK3 stabilizes the USP13/Fbxl14/c-Myc axis to enhance cholangiocarcinoma aggressiveness [289]. These outcomes broaden the horizon, including targets beyond the enzymes that are directly involved in purine metabolism.

In any case, more researches are needed to understand the mechanisms of aberrant purine metabolism in OC. In-depth knowledge of purine metabolic processes may help to better understand cancer-related metabolic reprogramming and develop new inhibitors accordingly, which presents exciting opportunities for OC therapy. We will investigate this in more depth in the future and believe that this may become a promising novel strategy for OC therapy. 

Availability of data and materials

Not applicable.

Abbreviations

ADA:

Adenosine deaminase

ADAR:

Adenosine deaminases acting on RNA

ADSS:

Adenylosuccinate synthase

AK:

Adenylate kinase

Ang:

Angiopoietin

APRT:

Adenine phosphoribosyltransferase

BR:

Benzamide riboside

CAR:

Chimeric antigen receptor

CCNI:

Cyclin I

CD39:

Ectonucleoside triphosphate diphosphohydrolase

CD73:

Ectosolic-5′-nucleotidase

DDP4:

Dipeptidyl peptidase 4

DHFR:

Dihydrofolate reductase

eADO:

Extracellular adenosine

eAMP:

Extracellular AMP

eATP:

Extracellular ATP

EMT:

Epithelial–mesenchymal transition

ePNP:

E. coli PNP

F-ara-A:

9-β-D-arabino-furanosyl-2-fluoradenine

GDEPT:

Gene directed enzyme prodrug therapy

HPRT:

Hypoxanthine guanine phosphoribosyltransferase

IMP:

Inosine monophosphate

IMPDH:

Inosine monophosphate dehydrogenase

MDSC:

Myeloid-deriver suppressor cell

MTHFR:

5,10-Methylenetetrahydrofolate reductase

MTX:

Methotrexate

OC:

Ovarian cancer

PD-1:

Programmed death-1

PNP:

Purine nucleoside phosphorylase

PRPP:

5-Phosphoribosyl-1-pyrophosphate

SNP:

Single nucleotide polymorphisms

TAM:

Tumor-associated macrophage

THF:

Tetrahydrofolate

TME:

Tumor microenvironment

TMZ:

Temozolomide

UA:

Uric acid

XMP:

Xanthosine monophosphate

XO:

Xanthine oxidase

5-FU:

5-Fluorouracil

6-MP:

6-Mercaptopurine

6-TG:

6-Thioguanine

References

  1. Doherty JA, Peres LC, Wang C, Way GP, Greene CS, Schildkraut JM. Challenges and opportunities in studying the epidemiology of ovarian cancer subtypes. Curr Epidemiol Rep. 2017;4(3):211–20.

    Article  PubMed  PubMed Central  Google Scholar 

  2. Siegel RL, Miller KD, Fuchs HE, Jemal A. Cancer statistics, 2021. CA Cancer J Clin. 2021;71(1):7–33.

    Article  PubMed  Google Scholar 

  3. Coleman RL, Monk BJ, Sood AK, Herzog TJ. Latest research and treatment of advanced-stage epithelial ovarian cancer. Nat Rev Clin Oncol. 2013;10(4):211–24.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Hess JR, Greenberg NA. The role of nucleotides in the immune and gastrointestinal systems: potential clinical applications. Nutr Clin Pract. 2012;27(2):281–94.

    Article  PubMed  Google Scholar 

  5. De Vitto H, Arachchige DB, Richardson BC, French JB. The intersection of purine and mitochondrial metabolism in cancer. Cells-Basel. 2021;10(10):2603.

    Article  CAS  Google Scholar 

  6. Camici M, Garcia-Gil M, Pesi R, Allegrini S, Tozzi MG. Purine-metabolising enzymes and apoptosis in cancer. Cancers (Basel). 2019;11(9):1354.

    Article  CAS  Google Scholar 

  7. Meuth M. The molecular basis of mutations induced by deoxyribonucleoside triphosphate pool imbalances in mammalian cells. Exp Cell Res. 1989;181(2):305–16.

    Article  PubMed  CAS  Google Scholar 

  8. Reilly NM, Novara L, Di Nicolantonio F, Bardelli A. Exploiting DNA repair defects in colorectal cancer. Mol Oncol. 2019;13(4):681–700.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. Karran P. Thiopurines, DNA damage, DNA repair and therapy-related cancer. Br Med Bull. 2006;79–80:153–70.

    Article  PubMed  CAS  Google Scholar 

  10. D’Almeida SM, Kauffenstein G, Roy C, Basset L, Papargyris L, Henrion D, Catros V, Ifrah N, Descamps P, Croue A, et al. The ecto-ATPDase CD39 is involved in the acquisition of the immunoregulatory phenotype by M-CSF-macrophages and ovarian cancer tumor-associated macrophages: Regulatory role of IL-27. Oncoimmunology. 2016;5(7):e1178025.

  11. Coulthard S, Hogarth L. The thiopurines: an update. Invest New Drugs. 2005;23(6):523–32.

    Article  PubMed  CAS  Google Scholar 

  12. Anderson VR, Perry CM. Fludarabine: a review of its use in non-Hodgkin’s lymphoma. Drugs. 2007;67(11):1633–55.

  13. Campos-Contreras A, Diaz-Munoz M, Vazquez-Cuevas FG. Purinergic signaling in the hallmarks of cancer. Cells-Basel. 2020;9(7):1612.

    Article  CAS  Google Scholar 

  14. Li L, Wang L, Li J, Fan Z, Yang L, Zhang Z, Zhang C, Yue D, Qin G, Zhang T, et al. Metformin-induced reduction of CD39 and CD73 blocks myeloid-derived suppressor cell activity in patients with ovarian cancer. Cancer Res. 2018;78(7):1779–91.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  15. Montalban DBI, Penski C, Schlahsa L, Stein RG, Diessner J, Wockel A, Dietl J, Lutz MB, Mittelbronn M, Wischhusen J, et al. Adenosine-generating ovarian cancer cells attract myeloid cells which differentiate into adenosine-generating tumor associated macrophages - a self-amplifying, CD39- and CD73-dependent mechanism for tumor immune escape. J Immunother Cancer. 2016;4:49.

    Article  Google Scholar 

  16. Turcotte M, Spring K, Pommey S, Chouinard G, Cousineau I, George J, Chen GM, Gendoo DM, Haibe-Kains B, Karn T, et al. CD73 is associated with poor prognosis in high-grade serous ovarian cancer. Cancer Res. 2015;75(21):4494–503.

    Article  PubMed  CAS  Google Scholar 

  17. Oh HK, Sin JI, Choi J, Park SH, Lee TS, Choi YS. Overexpression of CD73 in epithelial ovarian carcinoma is associated with better prognosis, lower stage, better differentiation and lower regulatory T cell infiltration. J Gynecol Oncol. 2012;23(4):274–81.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Gines S, Marino M, Mallol J, Canela EI, Morimoto C, Callebaut C, Hovanessian A, Casado V, Lluis C, Franco R. Regulation of epithelial and lymphocyte cell adhesion by adenosine deaminase-CD26 interaction. Biochem J. 2002;361(Pt 2):203–9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  19. Urunsak IF, Gulec UK, Paydas S, Seydaoglu G, Guzel AB, Vardar MA. Adenosine deaminase activity in patients with ovarian neoplasms. Arch Gynecol Obstet. 2012;286(1):155–9.

    Article  PubMed  CAS  Google Scholar 

  20. Zavialov AV, Gracia E, Glaichenhaus N, Franco R, Zavialov AV, Lauvau G. Human adenosine deaminase 2 induces differentiation of monocytes into macrophages and stimulates proliferation of T helper cells and macrophages. J Leukoc Biol. 2010;88(2):279–90.

    Article  PubMed  CAS  Google Scholar 

  21. Zhang M, Fritsche J, Roszik J, Williams LJ, Peng X, Chiu Y, Tsou CC, Hoffgaard F, Goldfinger V, Schoor O, et al. RNA editing derived epitopes function as cancer antigens to elicit immune responses. Nat Commun. 2018;9(1):3919.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Patterson JB, Samuel CE. Expression and regulation by interferon of a double-stranded-RNA-specific adenosine deaminase from human cells: evidence for two forms of the deaminase. Mol Cell Biol. 1995;15(10):5376–88.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Barraud P, Banerjee S, Mohamed WI, Jantsch MF, Allain FH. A bimodular nuclear localization signal assembled via an extended double-stranded RNA-binding domain acts as an RNA-sensing signal for transportin 1. Proc Natl Acad Sci U S A. 2014;111(18):E1852–61.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Tan TY, Sedmik J, Fitzgerald MP, Halevy RS, Keegan LP, Helbig I, Basel-Salmon L, Cohen L, Straussberg R, Chung WK, et al. Bi-allelic ADARB1 variants associated with microcephaly, intellectual disability, and seizures. Am J Hum Genet. 2020;106(4):467–83.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Zhang XY, Zhou LL, Jiao Y, Li YQ, Guan YN, Zhao YC, Zheng LW. Adenylate kinase 7 is a prognostic indicator of overall survival in ovarian cancer. Medicine (Baltimore). 2021;100(1):e24134.

    Article  CAS  Google Scholar 

  26. Huang M, Qin X, Wang Y, Mao F. Identification of AK4 as a novel therapeutic target for serous ovarian cancer. Oncol Lett. 2020;20(6):346.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  27. Tan H, Wu C, Huang B, Jin L, Jiang X. MiR-3666 serves as a tumor suppressor in ovarian carcinoma by down-regulating AK4 via targeting STAT3. Cancer Biomark. 2021;30(4):355–63.

    Article  PubMed  CAS  Google Scholar 

  28. Look KY, Sutton GP, Natsumeda Y, Eble JN, Stehman FB, Ehrlich CE, Olah E, Prajda N, Bosze P, Eckhardt S, et al. Inhibition by tiazofurin of inosine 5’-phosphate dehydrogenase (IMP DH) activity in extracts of ovarian carcinomas. Gynecol Oncol. 1992;47(1):66–70.

  29. Tian Y, Zhang J, Chen L, Zhang X. The expression and prognostic role of IMPDH2 in ovarian cancer. Ann Diagn Pathol. 2020;46:151511.

    Article  PubMed  Google Scholar 

  30. Hunakova L, Bies J, Sedlak J, Duraj J, Jakubikova J, Takacsova X, Novotny L, Chorvath B. Differential sensitivity of ovarian carcinoma cell lines to apoptosis induced by the IMPDH inhibitor benzamide riboside. Neoplasma. 2000;47(5):274–9.

    PubMed  CAS  Google Scholar 

  31. Duong-Ly KC, Kuo YM, Johnson MC, Cote JM, Kollman JM, Soboloff J, Rall GF, Andrews AJ, Peterson JR. T cell activation triggers reversible inosine-5’-monophosphate dehydrogenase assembly. J Cell Sci. 2018;131(17):jcs223289.

  32. Giuliani P, Zuccarini M, Buccella S, Pena-Altamira LE, Polazzi E, Virgili M, Monti B, Poli A, Rathbone MP, Di Iorio P, et al. Evidence for purine nucleoside phosphorylase (PNP) release from rat C6 glioma cells. J Neurochem. 2017;141(2):208–21.

    Article  PubMed  CAS  Google Scholar 

  33. Ofinran O, Bose U, Hay D, Abdul S, Tufatelli C, Khan R. Selection of suitable reference genes for gene expression studies in normal human ovarian tissues, borderline ovarian tumours and ovarian cancer. Mol Med Rep. 2016;14(6):5725–31.

    Article  PubMed  CAS  Google Scholar 

  34. Wang L, Wang Y, Han N, Wang X, Ruan M. HPRT promotes proliferation and metastasis in head and neck squamous cell carcinoma through direct interaction with STAT3. Exp Cell Res. 2021;399(1):112424.

    Article  PubMed  CAS  Google Scholar 

  35. Keough DT, Hockova D, Holy A, Naesens LM, Skinner-Adams TS, Jersey J, Guddat LW. Inhibition of hypoxanthine-guanine phosphoribosyltransferase by acyclic nucleoside phosphonates: a new class of antimalarial therapeutics. J Med Chem. 2009;52(14):4391–9.

    Article  PubMed  CAS  Google Scholar 

  36. Scanlon KJ, Kashani-Sabet M. Elevated expression of thymidylate synthase cycle genes in cisplatin-resistant human ovarian carcinoma A2780 cells. Proc Natl Acad Sci U S A. 1988;85(3):650–3.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Marverti G, Ligabue A, Paglietti G, Corona P, Piras S, Vitale G, Guerrieri D, Luciani R, Costi MP, Frassineti C, et al. Collateral sensitivity to novel thymidylate synthase inhibitors correlates with folate cycle enzymes impairment in cisplatin-resistant human ovarian cancer cells. Eur J Pharmacol. 2009;615(1–3):17–26.

    Article  PubMed  CAS  Google Scholar 

  38. Trent JM, Buick RN, Olson S, Horns RJ, Schimke RT. Cytologic evidence for gene amplification in methotrexate-resistant cells obtained from a patient with ovarian adenocarcinoma. J Clin Oncol. 1984;2(1):8–15.

    Article  PubMed  CAS  Google Scholar 

  39. Chen J, Li L. Aberrant Expression of Folate Metabolism Enzymes and Its Diagnosis and Survival Prediction in Ovarian Carcinoma. Anal Cell Pathol (Amst). 2019;2019:1438628.

    Google Scholar 

  40. Viel A, Dall’Agnese L, Simone F, Canzonieri V, Capozzi E, Visentin MC, Valle R, Boiocchi M. Loss of heterozygosity at the 5,10-methylenetetrahydrofolate reductase locus in human ovarian carcinomas. Br J Cancer. 1997;75(8):1105–10.

  41. Gong JM, Shen Y, Shan WW, He YX. The association between MTHFR polymorphism and cervical cancer. Sci Rep. 2018;8(1):7244.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Liu L, Liao SG, Wang YJ. MTHFR polymorphisms and ovarian cancer risk: a meta-analysis. Mol Biol Rep. 2012;39(11):9863–8.

    Article  PubMed  CAS  Google Scholar 

  43. Dixon SC, Ibiebele TI, Protani MM, Beesley J, Defazio A, Crandon AJ, Gard GB, Rome RM, Webb PM, Nagle CM. Dietary folate and related micronutrients, folate-metabolising genes, and ovarian cancer survival. Gynecol Oncol. 2014;132(3):566–72.

    Article  PubMed  CAS  Google Scholar 

  44. Natsumeda Y, Prajda N, Donohue JP, Glover JL, Weber G. Enzymic capacities of purine de Novo and salvage pathways for nucleotide synthesis in normal and neoplastic tissues. Cancer Res. 1984;44(6):2475–9.

    PubMed  CAS  Google Scholar 

  45. Yamaoka T, Kondo M, Honda S, Iwahana H, Moritani M, Ii S, Yoshimoto K, Itakura M. Amidophosphoribosyltransferase limits the rate of cell growth-linked de novo purine biosynthesis in the presence of constant capacity of salvage purine biosynthesis. J Biol Chem. 1997;272(28):17719–25.

    Article  PubMed  CAS  Google Scholar 

  46. Yin J, Ren W, Huang X, Deng J, Li T, Yin Y. Potential Mechanisms Connecting Purine Metabolism and Cancer Therapy. Front Immunol. 2018;9:1697.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. Pareek V, Pedley AM, Benkovic SJ. Human de novo purine biosynthesis. Crit Rev Biochem Mol Biol. 2021;56(1):1–16.

    Article  PubMed  CAS  Google Scholar 

  48. Pedley AM, Benkovic SJ. A New view into the regulation of purine metabolism: the purinosome. Trends Biochem Sci. 2017;42(2):141–54.

    Article  PubMed  CAS  Google Scholar 

  49. Zhao H, Chiaro CR, Zhang L, Smith PB, Chan CY, Pedley AM, Pugh RJ, French JB, Patterson AD, Benkovic SJ. Quantitative analysis of purine nucleotides indicates that purinosomes increase de novo purine biosynthesis. J Biol Chem. 2015;290(11):6705–13.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. An S, Kumar R, Sheets ED, Benkovic SJ. Reversible compartmentalization of de novo purine biosynthetic complexes in living cells. Science. 2008;320(5872):103–6.

    Article  PubMed  CAS  Google Scholar 

  51. Miyazono KI, Ishino S, Makita N, Ito T, Ishino Y, Tanokura M. Crystal structure of the novel lesion-specific endonuclease PfuEndoQ from Pyrococcus furiosus. Nucleic Acids Res. 2018;46(9):4807–18.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  52. Moesta AK, Li XY, Smyth MJ. Targeting CD39 in cancer. Nat Rev Immunol. 2020;20(12):739–55.

    Article  PubMed  CAS  Google Scholar 

  53. Regateiro FS, Cobbold SP, Waldmann H. CD73 and adenosine generation in the creation of regulatory microenvironments. Clin Exp Immunol. 2013;171(1):1–7.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  54. Ohta A, Gorelik E, Prasad SJ, Ronchese F, Lukashev D, Wong MK, Huang X, Caldwell S, Liu K, Smith P, et al. A2A adenosine receptor protects tumors from antitumor T cells. Proc Natl Acad Sci U S A. 2006;103(35):13132–7.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Young A, Ngiow SF, Gao Y, Patch AM, Barkauskas DS, Messaoudene M, Lin G, Coudert JD, Stannard KA, Zitvogel L, et al. A2AR adenosine signaling suppresses natural killer cell maturation in the tumor microenvironment. Cancer Res. 2018;78(4):1003–16.

    Article  PubMed  CAS  Google Scholar 

  56. Zheng X, Wang D. The adenosine A2A receptor agonist accelerates bone healing and adjusts Treg/Th17 cell balance through interleukin 6. Biomed Res Int. 2020;2020:2603873.

    PubMed  PubMed Central  Google Scholar 

  57. Csoka B, Selmeczy Z, Koscso B, Nemeth ZH, Pacher P, Murray PJ, Kepka-Lenhart D, Morris SJ, Gause WC, Leibovich SJ, et al. Adenosine promotes alternative macrophage activation via A2A and A2B receptors. Faseb J. 2012;26(1):376–86.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Kortekaas KE, Santegoets SJ, Sturm G, Ehsan I, van Egmond SL, Finotello F, Trajanoski Z, Welters M, van Poelgeest M, van der Burg SH. CD39 identifies the CD4(+) tumor-specific T-cell population in human cancer. Cancer Immunol Res. 2020;8(10):1311–21.

    Article  PubMed  CAS  Google Scholar 

  59. Duhen T, Duhen R, Montler R, Moses J, Moudgil T, de Miranda NF, Goodall CP, Blair TC, Fox BA, Mcdermott JE, et al. Co-expression of CD39 and CD103 identifies tumor-reactive CD8 T cells in human solid tumors. Nat Commun. 2018;9(1):2724.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  60. Ni X, Wan W, Ma J, Liu X, Zheng B, He Z, Yang W, Huang L. A Novel Prognostic Biomarker of Luminal Breast Cancer: High CD39 Expression Is Related to Poor Survival. Front Genet. 2021;12:682503.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  61. Giatromanolaki A, Kouroupi M, Pouliliou S, Mitrakas A, Hasan F, Pappa A, Koukourakis MI. Ectonucleotidase CD73 and CD39 expression in non-small cell lung cancer relates to hypoxia and immunosuppressive pathways. Life Sci. 2020;259:118389.

    Article  PubMed  CAS  Google Scholar 

  62. Mascanfroni ID, Yeste A, Vieira SM, Burns EJ, Patel B, Sloma I, Wu Y, Mayo L, Ben-Hamo R, Efroni S, et al. IL-27 acts on DCs to suppress the T cell response and autoimmunity by inducing expression of the immunoregulatory molecule CD39. Nat Immunol. 2013;14(10):1054–63.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  63. Zhang Z, Zhou B, Wu Y, Gao Q, Zhang K, Song Y, Zhang L, Xi M. Prognostic value of IL-27 polymorphisms and the susceptibility to epithelial ovarian cancer in a Chinese population. Immunogenetics. 2014;66(2):85–92.

    Article  PubMed  CAS  Google Scholar 

  64. Lupia M, Angiolini F, Bertalot G, Freddi S, Sachsenmeier KF, Chisci E, Kutryb-Zajac B, Confalonieri S, Smolenski RT, Giovannoni R, et al. CD73 regulates stemness and epithelial-mesenchymal transition in ovarian cancer-initiating cells. Stem Cell Rep. 2018;10(4):1412–25.

    Article  CAS  Google Scholar 

  65. Nevedomskaya E, Perryman R, Solanki S, Syed N, Mayboroda OA, Keun HC. A systems oncology approach identifies NT5E as a key metabolic regulator in tumor cells and modulator of platinum sensitivity. J Proteome Res. 2016;15(1):280–90.

    Article  PubMed  CAS  Google Scholar 

  66. Ma XL, Shen MN, Hu B, Wang BL, Yang WJ, Lv LH, Wang H, Zhou Y, Jin AL, Sun YF, et al. CD73 promotes hepatocellular carcinoma progression and metastasis via activating PI3K/AKT signaling by inducing Rap1-mediated membrane localization of P110beta and predicts poor prognosis. J Hematol Oncol. 2019;12(1):37.

    Article  PubMed  PubMed Central  Google Scholar 

  67. Ma XL, Hu B, Tang WG, Xie SH, Ren N, Guo L, Lu RQ. CD73 sustained cancer-stem-cell traits by promoting SOX9 expression and stability in hepatocellular carcinoma. J Hematol Oncol. 2020;13(1):11.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  68. Zhou L, Jia S, Chen Y, Wang W, Wu Z, Yu W, Zhang M, Ding G, Cao L. The distinct role of CD73 in the progression of pancreatic cancer. J Mol Med (Berl). 2019;97(6):803–15.

    Article  CAS  Google Scholar 

  69. Yu J, Wang X, Lu Q, Wang J, Li L, Liao X, Zhu W, Lv L, Zhi X, Yu J, et al. Extracellular 5’-nucleotidase (CD73) promotes human breast cancer cells growth through AKT/GSK-3beta/beta-catenin/cyclinD1 signaling pathway. Int J Cancer. 2018;142(5):959–67.

  70. Hausler SF, Montalban DBI, Strohschein J, Chandran PA, Engel JB, Honig A, Ossadnik M, Horn E, Fischer B, Krockenberger M, et al. Ectonucleotidases CD39 and CD73 on OvCA cells are potent adenosine-generating enzymes responsible for adenosine receptor 2A-dependent suppression of T cell function and NK cell cytotoxicity. Cancer Immunol Immunother. 2011;60(10):1405–18.

    Article  PubMed  CAS  Google Scholar 

  71. Fang F, Yu M, Cavanagh MM, Hutter SJ, Qi Q, Ye Z, Le Saux S, Sultan W, Turgano E, Dekker CL, et al. Expression of CD39 on activated t cells impairs their survival in older individuals. Cell Rep. 2016;14(5):1218–31.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  72. Li J, Wang L, Chen X, Li L, Li Y, Ping Y, Huang L, Yue D, Zhang Z, Wang F, et al. CD39/CD73 upregulation on myeloid-derived suppressor cells via TGF-beta-mTOR-HIF-1 signaling in patients with non-small cell lung cancer. Oncoimmunology. 2017;6(6):e1320011.

    Article  PubMed  PubMed Central  Google Scholar 

  73. Limagne E, Euvrard R, Thibaudin M, Rebe C, Derangere V, Chevriaux A, Boidot R, Vegran F, Bonnefoy N, Vincent J, et al. Accumulation of MDSC and Th17 cells in patients with metastatic colorectal cancer predicts the efficacy of a FOLFOX-bevacizumab drug treatment regimen. Cancer Res. 2016;76(18):5241–52.

    Article  PubMed  CAS  Google Scholar 

  74. Li XY, Moesta AK, Xiao C, Nakamura K, Casey M, Zhang H, Madore J, Lepletier A, Aguilera AR, Sundarrajan A, et al. Targeting CD39 in cancer reveals an extracellular ATP- and inflammasome-driven tumor immunity. Cancer Discov. 2019;9(12):1754–73.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  75. Li H, Lv M, Qiao B, Li X. Blockade pf CD73/adenosine axis improves the therapeutic efficacy of docetaxel in epithelial ovarian cancer. Arch Gynecol Obstet. 2019;299(6):1737–46.

    Article  PubMed  CAS  Google Scholar 

  76. Salmaninejad A, Khoramshahi V, Azani A, Soltaninejad E, Aslani S, Zamani MR, Zal M, Nesaei A, Hosseini SM. PD-1 and cancer: molecular mechanisms and polymorphisms. Immunogenetics. 2018;70(2):73–86.

    Article  PubMed  CAS  Google Scholar 

  77. Allard B, Pommey S, Smyth MJ, Stagg J. Targeting CD73 enhances the antitumor activity of anti-PD-1 and anti-CTLA-4 mAbs. Clin Cancer Res. 2013;19(20):5626–35.

    Article  PubMed  CAS  Google Scholar 

  78. Lindley ER, Pisoni RL. Demonstration of adenosine deaminase activity in human fibroblast lysosomes. Biochem J. 1993;290(Pt 2):457–62.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  79. Eltzschig HK, Faigle M, Knapp S, Karhausen J, Ibla J, Rosenberger P, Odegard KC, Laussen PC, Thompson LF, Colgan SP. Endothelial catabolism of extracellular adenosine during hypoxia: the role of surface adenosine deaminase and CD26. Blood. 2006;108(5):1602–10.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  80. Franco R, Pacheco R, Gatell JM, Gallart T, Lluis C. Enzymatic and extraenzymatic role of adenosine deaminase 1 in T-cell-dendritic cell contacts and in alterations of the immune function. Crit Rev Immunol. 2007;27(6):495–509.

    Article  PubMed  CAS  Google Scholar 

  81. Zhong J, Rao X, Deiuliis J, Braunstein Z, Narula V, Hazey J, Mikami D, Needleman B, Satoskar AR, Rajagopalan S. A potential role for dendritic cell/macrophage-expressing DPP4 in obesity-induced visceral inflammation. Diabetes. 2013;62(1):149–57.

    Article  PubMed  CAS  Google Scholar 

  82. Enz N, Vliegen G, De Meester I, Jungraithmayr W. CD26/DPP4 - a potential biomarker and target for cancer therapy. Pharmacol Ther. 2019;198:135–59.

    Article  PubMed  CAS  Google Scholar 

  83. Martinez-Navio JM, Casanova V, Pacheco R, Naval-Macabuhay I, Climent N, Garcia F, Gatell JM, Mallol J, Gallart T, Lluis C, et al. Adenosine deaminase potentiates the generation of effector, memory, and regulatory CD4+ T cells. J Leukoc Biol. 2011;89(1):127–36.

    Article  PubMed  CAS  Google Scholar 

  84. Gakis C. Adenosine deaminase (ADA) isoenzymes ADA1 and ADA2: diagnostic and biological role. Eur Respir J. 1996;9(4):632–3.

    Article  PubMed  CAS  Google Scholar 

  85. Zavialov AV, Engstrom A. Human ADA2 belongs to a new family of growth factors with adenosine deaminase activity. Biochem J. 2005;391(Pt 1):51–7.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  86. Zhou Q, Yang D, Ombrello AK, Zavialov AV, Toro C, Zavialov AV, Stone DL, Chae JJ, Rosenzweig SD, Bishop K, et al. Early-onset stroke and vasculopathy associated with mutations in ADA2. N Engl J Med. 2014;370(10):911–20.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  87. Caorsi R, Penco F, Grossi A, Insalaco A, Omenetti A, Alessio M, Conti G, Marchetti F, Picco P, Tommasini A, et al. ADA2 deficiency (DADA2) as an unrecognised cause of early onset polyarteritis nodosa and stroke: a multicentre national study. Ann Rheum Dis. 2017;76(10):1648–56.

    Article  PubMed  CAS  Google Scholar 

  88. Zhang M, Xu L, Wang X, Sun B, Ding J. Expression levels of seprase/FAPalpha and DPPIV/CD26 in epithelial ovarian carcinoma. Oncol Lett. 2015;10(1):34–42.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  89. Kikkawa F, Kajiyama H, Ino K, Shibata K, Mizutani S. Increased adhesion potency of ovarian carcinoma cells to mesothelial cells by overexpression of dipeptidyl peptidase IV. Int J Cancer. 2003;105(6):779–83.

    Article  PubMed  CAS  Google Scholar 

  90. El-Kashef DH, Serrya MS. Sitagliptin ameliorates thioacetamide-induced acute liver injury via modulating TLR4/NF-KB signaling pathway in mice. Life Sci. 2019;228:266–73.

    Article  PubMed  CAS  Google Scholar 

  91. Kosowska A, Garczorz W, Klych-Ratuszny A, Aghdam M, Kimsa-Furdzik M, Simka-Lampa K, Francuz T. Sitagliptin Modulates the Response of Ovarian Cancer Cells to Chemotherapeutic Agents. Int J Mol Sci. 2020;21(23):8976.

    Article  PubMed Central  CAS  Google Scholar 

  92. Martinez-Ramirez AS, Diaz-Munoz M, Battastini AM, Campos-Contreras A, Olvera A, Bergamin L, Glaser T, Jacintho MC, Ulrich H, Vazquez-Cuevas FG. Cellular Migration Ability Is Modulated by Extracellular Purines in Ovarian Carcinoma SKOV-3 Cells. J Cell Biochem. 2017;118(12):4468–78.

    Article  PubMed  CAS  Google Scholar 

  93. Kajiyama H, Shibata K, Ino K, Mizutani S, Nawa A, Kikkawa F. The expression of dipeptidyl peptidase IV (DPPIV/CD26) is associated with enhanced chemosensitivity to paclitaxel in epithelial ovarian carcinoma cells. Cancer Sci. 2010;101(2):347–54.

    Article  PubMed  CAS  Google Scholar 

  94. Kajiyama H, Shibata K, Terauchi M, Ino K, Nawa A, Kikkawa F. Involvement of DPPIV/CD26 in epithelial morphology and suppressed invasive ability in ovarian carcinoma cells. Ann N Y Acad Sci. 2006;1086:233–40.

    Article  PubMed  CAS  Google Scholar 

  95. Kajiyama H, Kikkawa F, Khin E, Shibata K, Ino K, Mizutani S. Dipeptidyl peptidase IV overexpression induces up-regulation of E-cadherin and tissue inhibitors of matrix metalloproteinases, resulting in decreased invasive potential in ovarian carcinoma cells. Cancer Res. 2003;63(9):2278–83.

    PubMed  CAS  Google Scholar 

  96. Palladino MJ, Keegan LP, O’Connell MA, Reenan RA. dADAR, a Drosophila double-stranded RNA-specific adenosine deaminase is highly developmentally regulated and is itself a target for RNA editing. RNA. 2000;6(7):1004–18.

  97. Wang C, Zou J, Ma X, Wang E, Peng G. Mechanisms and implications of ADAR-mediated RNA editing in cancer. Cancer Lett. 2017;411:27–34.

    Article  PubMed  CAS  Google Scholar 

  98. Nishikura K. Editor meets silencer: crosstalk between RNA editing and RNA interference. Nat Rev Mol Cell Biol. 2006;7(12):919–31.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  99. Chen CX, Cho DS, Wang Q, Lai F, Carter KC, Nishikura K. A third member of the RNA-specific adenosine deaminase gene family, ADAR3, contains both single- and double-stranded RNA binding domains. RNA. 2000;6(5):755–67.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  100. Peng X, Xu X, Wang Y, Hawke DH, Yu S, Han L, Zhou Z, Mojumdar K, Jeong KJ, Labrie M, et al. A-to-I RNA editing contributes to proteomic diversity in cancer. Cancer Cell. 2018;33(5):817–28.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  101. Zhang Y, Qian H, Xu J, Gao W. ADAR, the carcinogenesis mechanisms of ADAR and related clinical applications. Ann Transl Med. 2019;7(22):686.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  102. Cybulski M, Jarosz B, Nowakowski A, Jeleniewicz W, Seroczynski P, Mazurek-Kociubowska M. Cyclin I correlates with VEGFR-2 and cell proliferation in human epithelial ovarian cancer. Gynecol Oncol. 2012;127(1):217–22.

    Article  PubMed  CAS  Google Scholar 

  103. Zipeto MA, Court AC, Sadarangani A, Delos SN, Balaian L, Chun HJ, Pineda G, Morris SR, Mason CN, Geron I, et al. ADAR1 activation drives leukemia stem cell self-renewal by impairing Let-7 biogenesis. Cell Stem Cell. 2016;19(2):177–91.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  104. Yang ZY, Wang Y, Liu Q, Wu M. microRNA cluster MC-let-7a-1~let-7d promotes autophagy and apoptosis of glioma cells by down-regulating STAT3. Cns Neurosci Ther. 2020;26(3):319–31.

    Article  PubMed  CAS  Google Scholar 

  105. Garcia-Vazquez R, Gallardo RD, Ruiz-Garcia E, Meneses GA, Hernandez DLCO, Astudillo-De LVH, Isla-Ortiz D, Marchat LA, Salinas-Vera YM, Carlos-Reyes A, et al. let-7d-3p is associated with apoptosis and response to neoadjuvant chemotherapy in ovarian cancer. Oncol Rep. 2018;39(6):3086–94.

    PubMed  CAS  Google Scholar 

  106. Jiang J, Liu HL, Tao L, Lin XY, Yang YD, Tan SW, Wu B. Let7d inhibits colorectal cancer cell proliferation through the CST1/p65 pathway. Int J Oncol. 2018;53(2):781–90.

    PubMed  CAS  Google Scholar 

  107. Chen YN, Ren CC, Yang L, Nai MM, Xu YM, Zhang F, Liu Y. MicroRNA let7d5p rescues ovarian cancer cell apoptosis and restores chemosensitivity by regulating the p53 signaling pathway via HMGA1. Int J Oncol. 2019;54(5):1771–84.

    PubMed  CAS  Google Scholar 

  108. Carrasco AJ, Dzeja PP, Alekseev AE, Pucar D, Zingman LV, Abraham MR, Hodgson D, Bienengraeber M, Puceat M, Janssen E, et al. Adenylate kinase phosphotransfer communicates cellular energetic signals to ATP-sensitive potassium channels. Proc Natl Acad Sci U S A. 2001;98(13):7623–8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  109. Dzeja P, Terzic A. Adenylate kinase and AMP signaling networks: metabolic monitoring, signal communication and body energy sensing. Int J Mol Sci. 2009;10(4):1729–72.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  110. Panayiotou C, Solaroli N, Karlsson A. The many isoforms of human adenylate kinases. Int J Biochem Cell Biol. 2014;49:75–83.

    Article  PubMed  CAS  Google Scholar 

  111. Jan YH, Lai TC, Yang CJ, Huang MS, Hsiao M. A co-expressed gene status of adenylate kinase 1/4 reveals prognostic gene signature associated with prognosis and sensitivity to EGFR targeted therapy in lung adenocarcinoma. Sci Rep. 2019;9(1):12329.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  112. Vasseur S, Malicet C, Calvo EL, Dagorn JC, Iovanna JL. Gene expression profiling of tumours derived from rasV12/E1A-transformed mouse embryonic fibroblasts to identify genes required for tumour development. Mol Cancer. 2005;4(1):4.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  113. Lanning NJ, Looyenga BD, Kauffman AL, Niemi NM, Sudderth J, Deberardinis RJ, Mackeigan JP. A mitochondrial RNAi screen defines cellular bioenergetic determinants and identifies an adenylate kinase as a key regulator of ATP levels. Cell Rep. 2014;7(3):907–17.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  114. Bai D, Zhang J, Li T, Hang R, Liu Y, Tian Y, Huang D, Qu L, Cao X, Ji J, et al. The ATPase hCINAP regulates 18S rRNA processing and is essential for embryogenesis and tumour growth. Nat Commun. 2016;7:12310.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  115. Zhang J, Yin YT, Wu CH, Qiu RL, Jiang WJ, Deng XG, Li ZX. AK4 Promotes the Progression of HER2-Positive Breast Cancer by Facilitating Cell Proliferation and Invasion. Dis Markers. 2019;2019:8186091.

    PubMed  PubMed Central  Google Scholar 

  116. Jan YH, Lai TC, Yang CJ, Lin YF, Huang MS, Hsiao M. Adenylate kinase 4 modulates oxidative stress and stabilizes HIF-1alpha to drive lung adenocarcinoma metastasis. J Hematol Oncol. 2019;12(1):12.

    Article  PubMed  PubMed Central  Google Scholar 

  117. Xin F, Yao DW, Fan L, Liu JH, Liu XD. Adenylate kinase 4 promotes bladder cancer cell proliferation and invasion. Clin Exp Med. 2019;19(4):525–34.

    Article  PubMed  CAS  Google Scholar 

  118. Liu R, Strom AL, Zhai J, Gal J, Bao S, Gong W, Zhu H. Enzymatically inactive adenylate kinase 4 interacts with mitochondrial ADP/ATP translocase. Int J Biochem Cell Biol. 2009;41(6):1371–80.

    Article  PubMed  CAS  Google Scholar 

  119. Fujisawa K, Terai S, Takami T, Yamamoto N, Yamasaki T, Matsumoto T, Yamaguchi K, Owada Y, Nishina H, Noma T, et al. Modulation of anti-cancer drug sensitivity through the regulation of mitochondrial activity by adenylate kinase 4. J Exp Clin Cancer Res. 2016;35:48.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  120. Kong F, Binas B, Moon JH, Kang SS, Kim HJ. Differential expression of adenylate kinase 4 in the context of disparate stress response strategies of HEK293 and HepG2 cells. Arch Biochem Biophys. 2013;533(1–2):11–7.

    Article  PubMed  CAS  Google Scholar 

  121. Miyoshi K, Akazawa Y, Horiguchi T, Noma T. Localization of adenylate kinase 4 in mouse tissues. Acta Histochem Cytochem. 2009;42(2):55–64.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  122. Jan YH, Tsai HY, Yang CJ, Huang MS, Yang YF, Lai TC, Lee CH, Jeng YM, Huang CY, Su JL, et al. Adenylate kinase-4 is a marker of poor clinical outcomes that promotes metastasis of lung cancer by downregulating the transcription factor ATF3. Cancer Res. 2012;72(19):5119–29.

    Article  PubMed  CAS  Google Scholar 

  123. Syed V, Mukherjee K, Lyons-Weiler J, Lau KM, Mashima T, Tsuruo T, Ho SM. Identification of ATF-3, caveolin-1, DLC-1, and NM23-H2 as putative antitumorigenic, progesterone-regulated genes for ovarian cancer cells by gene profiling. Oncogene. 2005;24(10):1774–87.

    Article  PubMed  CAS  Google Scholar 

  124. Murph MM, Liu W, Yu S, Lu Y, Hall H, Hennessy BT, Lahad J, Schaner M, Helland A, Kristensen G, et al. Lysophosphatidic acid-induced transcriptional profile represents serous epithelial ovarian carcinoma and worsened prognosis. PLoS ONE. 2009;4(5):e5583.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  125. Lores P, Coutton C, El KE, Stouvenel L, Givelet M, Thomas L, Rode B, Schmitt A, Louis B, Sakheli Z, et al. Homozygous missense mutation L673P in adenylate kinase 7 (AK7) leads to primary male infertility and multiple morphological anomalies of the flagella but not to primary ciliary dyskinesia. Hum Mol Genet. 2018;27(7):1196–211.

    Article  PubMed  CAS  Google Scholar 

  126. Mata M, Lluch-Estelles J, Armengot M, Sarrion I, Carda C, Cortijo J. New adenylate kinase 7 (AK7) mutation in primary ciliary dyskinesia. Am J Rhinol Allergy. 2012;26(4):260–4.

    Article  PubMed  Google Scholar 

  127. Milara J, Armengot M, Mata M, Morcillo EJ, Cortijo J. Role of adenylate kinase type 7 expression on cilia motility: possible link in primary ciliary dyskinesia. Am J Rhinol Allergy. 2010;24(3):181–5.

    Article  PubMed  Google Scholar 

  128. Shpak M, Goldberg MM, Cowperthwaite MC. Cilia gene expression patterns in cancer. Cancer Genomics Proteomics. 2014;11(1):13–24.

    PubMed  Google Scholar 

  129. Teilmann SC, Christensen ST. Localization of the angiopoietin receptors Tie-1 and Tie-2 on the primary cilia in the female reproductive organs. Cell Biol Int. 2005;29(5):340–6.

    Article  PubMed  CAS  Google Scholar 

  130. Dominska K. Involvement of ACE2/Ang-(1–7)/MAS1 axis in the regulation of ovarian function in mammals. Int J Mol Sci. 2020;21(13):4572.

    Article  PubMed Central  CAS  Google Scholar 

  131. Lohmussaar K, Kopper O, Korving J, Begthel H, Vreuls C, van Es JH, Clevers H. Assessing the origin of high-grade serous ovarian cancer using CRISPR-modification of mouse organoids. Nat Commun. 2020;11(1):2660.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  132. Hata K, Udagawa J, Fujiwaki R, Nakayama K, Otani H, Miyazaki K. Expression of angiopoietin-1, angiopoietin-2, and Tie2 genes in normal ovary with corpus luteum and in ovarian cancer. Oncology. 2002;62(4):340–8.

    Article  PubMed  CAS  Google Scholar 

  133. Lin Z, Liu Y, Sun Y, He X. Expression of Ets-1, Ang-2 and maspin in ovarian cancer and their role in tumor angiogenesis. J Exp Clin Cancer Res. 2011;30:31.

    Article  PubMed  PubMed Central  Google Scholar 

  134. Townsend MH, Robison RA, O’Neill KL. A review of HPRT and its emerging role in cancer. Med Oncol. 2018;35(6):89.

  135. Stout JT, Caskey CT. HPRT: gene structure, expression, and mutation. Annu Rev Genet. 1985;19:127–48.

    Article  PubMed  CAS  Google Scholar 

  136. Mateos EA, Puig JG. Purine metabolism in Lesch-Nyhan syndrome versus Kelley-Seegmiller syndrome. J Inherit Metab Dis. 1994;17(1):138–42.

    Article  PubMed  CAS  Google Scholar 

  137. Zoref-Shani E, Lavie R, Bromberg Y, Beery E, Sidi Y, Sperling O, Nordenberg J. Effects of differentiation-inducing agents on purine nucleotide metabolism in an ovarian cancer cell line. J Cancer Res Clin Oncol. 1994;120(12):717–22.

    Article  PubMed  CAS  Google Scholar 

  138. Shibutani S, Takeshita M, Grollman AP. Insertion of specific bases during DNA synthesis past the oxidation-damaged base 8-oxodG. Nature. 1991;349(6308):431–4.

    Article  PubMed  CAS  Google Scholar 

  139. Dorn J, Ferrari E, Imhof R, Ziegler N, Hubscher U. Regulation of human MutYH DNA glycosylase by the E3 ubiquitin ligase mule. J Biol Chem. 2014;289(10):7049–58.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  140. Richards B, Zhang H, Phear G, Meuth M. Conditional mutator phenotypes in hMSH2-deficient tumor cell lines. Science. 1997;277(5331):1523–6.

    Article  PubMed  CAS  Google Scholar 

  141. Wang SS, O’Neill JP, Qian GS, Zhu YR, Wang JB, Armenian H, Zarba A, Wang JS, Kensler TW, Cariello NF, et al. Elevated HPRT mutation frequencies in aflatoxin-exposed residents of daxin, Qidong county, People’s Republic of China. Carcinogenesis. 1999;20(11):2181–4.

  142. Watanabe N, Okochi E, Mochizuki M, Sugimura T, Ushijima T. The presence of single nucleotide instability in human breast cancer cell lines. Cancer Res. 2001;61(21):7739–42.

    PubMed  CAS  Google Scholar 

  143. Chang YJ, Tseng CY, Lin PY, Chuang YC, Chao MW. Acute exposure to DEHP metabolite, MEHP cause genotoxicity, mutagenesis and carcinogenicity in mammalian Chinese hamster ovary cells. Carcinogenesis. 2017;38(3):336–45.

    Article  PubMed  CAS  Google Scholar 

  144. Kamiyama H, Rauenzahn S, Shim JS, Karikari CA, Feldmann G, Hua L, Kamiyama M, Schuler FW, Lin MT, Beaty RM, et al. Personalized chemotherapy profiling using cancer cell lines from selectable mice. Clin Cancer Res. 2013;19(5):1139–46.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  145. Gercel-Taylor C, Scobee JJ, Taylor DD. Effect of chemotherapy on the mutation frequency of ovarian cancer cells at the HPRT locus. Anticancer Res. 2005;25(3B):2113–7.

    PubMed  CAS  Google Scholar 

  146. Isakovic AM, Dulovic M, Markovic I, Kravic-Stevovic T, Bumbasirevic V, Trajkovic V, Isakovic A. Autophagy suppression sensitizes glioma cells to IMP dehydrogenase inhibition-induced apoptotic death. Exp Cell Res. 2017;350(1):32–40.

    Article  PubMed  CAS  Google Scholar 

  147. Kofuji S, Sasaki AT. GTP metabolic reprogramming by IMPDH2: unlocking cancer cells’ fuelling mechanism. J Biochem. 2020;168(4):319–28.

  148. Huang F, Ni M, Chalishazar MD, Huffman KE, Kim J, Cai L, Shi X, Cai F, Zacharias LG, Ireland AS, et al. Inosine Monophosphate Dehydrogenase Dependence in a Subset of Small Cell Lung Cancers. Cell Metab. 2018;28(3):369–82.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  149. Duan S, Huang W, Liu X, Liu X, Chen N, Xu Q, Hu Y, Song W, Zhou J. IMPDH2 promotes colorectal cancer progression through activation of the PI3K/AKT/mTOR and PI3K/AKT/FOXO1 signaling pathways. J Exp Clin Cancer Res. 2018;37(1):304.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  150. Xu H, Ma H, Zha L, Li Q, Yang G, Pan H, Fei X, Xu X, Xing C, Zhang L. IMPDH2 promotes cell proliferation and epithelial-mesenchymal transition of non-small cell lung cancer by activating the Wnt/beta-catenin signaling pathway. Oncol Lett. 2020;20(5):219.

    PubMed  PubMed Central  CAS  Google Scholar 

  151. Kofuji S, Hirayama A, Eberhardt AO, Kawaguchi R, Sugiura Y, Sampetrean O, Ikeda Y, Warren M, Sakamoto N, Kitahara S, et al. IMP dehydrogenase-2 drives aberrant nucleolar activity and promotes tumorigenesis in glioblastoma. Nat Cell Biol. 2019;21(8):1003–14.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  152. He Y, Zheng Z, Xu Y, Weng H, Gao Y, Qin K, Rong J, Chen C, Yun M, Zhang J, et al. Over-expression of IMPDH2 is associated with tumor progression and poor prognosis in hepatocellular carcinoma. Am J Cancer Res. 2018;8(8):1604–14.

    PubMed  PubMed Central  CAS  Google Scholar 

  153. Xu Y, Zheng Z, Gao Y, Duan S, Chen C, Rong J, Wang K, Yun M, Weng H, Ye S, et al. High expression of IMPDH2 is associated with aggressive features and poor prognosis of primary nasopharyngeal carcinoma. Sci Rep. 2017;7(1):745.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  154. Naffouje R, Grover P, Yu H, Sendilnathan A, Wolfe K, Majd N, Smith EP, Takeuchi K, Senda T, Kofuji S, et al. Anti-Tumor Potential of IMP Dehydrogenase Inhibitors: A Century-Long Story. Cancers (Basel). 2019;11(9):1346.

    Article  CAS  Google Scholar 

  155. Floryk D, Thompson TC. Antiproliferative effects of AVN944, a novel inosine 5-monophosphate dehydrogenase inhibitor, in prostate cancer cells. Int J Cancer. 2008;123(10):2294–302.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  156. Valvezan AJ, Mcnamara MC, Miller SK, Torrence ME, Asara JM, Henske EP, Manning BD. IMPDH inhibitors for antitumor therapy in tuberous sclerosis complex. JCI Insight. 2020;5(7):e135071.

    Article  PubMed Central  Google Scholar 

  157. Huang F, Huffman KE, Wang Z, Wang X, Li K, Cai F, Yang C, Cai L, Shih TS, Zacharias LG, et al. Guanosine triphosphate links MYC-dependent metabolic and ribosome programs in small-cell lung cancer. J Clin Invest. 2021;131(1):e139929.

    Article  PubMed Central  CAS  Google Scholar 

  158. Meli M, Tolomeo M, Grifantini M, Franchetti P, Cappellacci L, Simoni D, Invidiata FP, Aiello S, Dusonchet L. The synergistic apoptotic effects of thiophenfurin, an inosine monophosphate dehydrogenase inhibitor, in combination with retinoids in HL60 cells. Oncol Rep. 2007;17(1):185–92.

    PubMed  CAS  Google Scholar 

  159. Grusch M, Rosenberger G, Fuhrmann G, Braun K, Titscher B, Szekeres T, Fritzer-Skekeres M, Oberhuber G, Krohn K, Hengstschlaeger M, et al. Benzamide riboside induces apoptosis independent of Cdc25A expression in human ovarian carcinoma N.1 cells. Cell Death Differ. 1999;6(8):736–44.

    Article  PubMed  CAS  Google Scholar 

  160. Bennett EM, Li C, Allan PW, Parker WB, Ealick SE. Structural basis for substrate specificity of Escherichia coli purine nucleoside phosphorylase. J Biol Chem. 2003;278(47):47110–8.

    Article  PubMed  Google Scholar 

  161. Zhang J, Kale V, Chen M. Gene-directed enzyme prodrug therapy. Aaps J. 2015;17(1):102–10.

    Article  PubMed  CAS  Google Scholar 

  162. Fu W, Lan H, Li S, Han X, Gao T, Ren D. Synergistic antitumor efficacy of suicide/ePNP gene and 6-methylpurine 2’-deoxyriboside via Salmonella against murine tumors. Cancer Gene Ther. 2008;15(7):474–84.

  163. Deharvengt S, Wack S, Uhring M, Aprahamian M, Hajri A. Suicide gene/prodrug therapy for pancreatic adenocarcinoma by E. coli purine nucleoside phosphorylase and 6-methylpurine 2’-deoxyriboside. Pancreas. 2004;28(2):E54–64.

  164. Zhang Y, Parker WB, Sorscher EJ, Ealick SE. PNP anticancer gene therapy. Curr Top Med Chem. 2005;5(13):1259–74.

    Article  PubMed  CAS  Google Scholar 

  165. Sorscher EJ, Peng S, Bebok Z, Allan PW, Bennett LJ, Parker WB. Tumor cell bystander killing in colonic carcinoma utilizing the Escherichia coli DeoD gene to generate toxic purines. Gene Ther. 1994;1(4):233–8.

    PubMed  CAS  Google Scholar 

  166. Gadi VK, Alexander SD, Kudlow JE, Allan P, Parker WB, Sorscher EJ. In vivo sensitization of ovarian tumors to chemotherapy by expression of E. coli purine nucleoside phosphorylase in a small fraction of cells. Gene Ther. 2000;7(20):1738–43.

    Article  PubMed  CAS  Google Scholar 

  167. Singh PP, Joshi S, Russell PJ, Nair S, Khatri A. Purine Nucleoside Phosphorylase mediated molecular chemotherapy and conventional chemotherapy: a tangible union against chemoresistant cancer. BMC Cancer. 2011;11:368.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  168. Wang X, Sun L, Sun X, Yu J, Wang K, Wu Y, Gao Q, Zheng J. Antitumor effects of a dual-specific lentiviral vector carrying the Escherichia coli purine nucleoside phosphorylase gene. Int J Oncol. 2017;50(5):1612–22.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  169. Yang C, Zhang J, Liao M, Yang Y, Wang Y, Yuan Y, Ouyang L. Folate-mediated one-carbon metabolism: a targeting strategy in cancer therapy. Drug Discov Today. 2021;26(3):817–25.

    Article  PubMed  CAS  Google Scholar 

  170. Duff MJ, Desai N, Craig MA, Agarwal PK, Howell EE. Crowders steal dihydrofolate reductase ligands through quinary interactions. Biochemistry-Us. 2019;58(9):1198–213.

    Article  CAS  Google Scholar 

  171. Tibbetts AS, Appling DR. Compartmentalization of Mammalian folate-mediated one-carbon metabolism. Annu Rev Nutr. 2010;30:57–81.

    Article  PubMed  CAS  Google Scholar 

  172. Tang Q, Cheng J, Cao X, Surowy H, Burwinkel B. Blood-based DNA methylation as biomarker for breast cancer: a systematic review. Clin Epigenetics. 2016;8:115.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  173. Tahara T, Arisawa T. DNA methylation as a molecular biomarker in gastric cancer. Epigenomics-Uk. 2015;7(3):475–86.

    Article  CAS  Google Scholar 

  174. Pan Y, Liu G, Zhou F, Su B, Li Y. DNA methylation profiles in cancer diagnosis and therapeutics. Clin Exp Med. 2018;18(1):1–14.

    Article  PubMed  CAS  Google Scholar 

  175. Goldman ID, Matherly LH. The cellular pharmacology of methotrexate. Pharmacol Ther. 1985;28(1):77–102.

    Article  PubMed  CAS  Google Scholar 

  176. Zhao M, Tan B, Dai X, Shao Y, He Q, Yang B, Wang J, Weng Q. DHFR/TYMS are positive regulators of glioma cell growth and modulate chemo-sensitivity to temozolomide. Eur J Pharmacol. 2019;863:172665.

    Article  PubMed  CAS  Google Scholar 

  177. Marverti G, Ligabue A, Lombardi P, Ferrari S, Monti MG, Frassineti C, Costi MP. Modulation of the expression of folate cycle enzymes and polyamine metabolism by berberine in cisplatin-sensitive and -resistant human ovarian cancer cells. Int J Oncol. 2013;43(4):1269–80.

    Article  PubMed  CAS  Google Scholar 

  178. Eroglu A, Akar N. Methylenetetrahydrofolate reductase C677T polymorphism in breast cancer risk. Breast Cancer Res Treat. 2010;122(3):897–8.

    Article  PubMed  Google Scholar 

  179. Frosst P, Blom HJ, Milos R, Goyette P, Sheppard CA, Matthews RG, Boers GJ, den Heijer M, Kluijtmans LA, van den Heuvel LP, et al. A candidate genetic risk factor for vascular disease: a common mutation in methylenetetrahydrofolate reductase. Nat Genet. 1995;10(1):111–3.

    Article  PubMed  CAS  Google Scholar 

  180. Pepe C, Guidugli L, Sensi E, Aretini P, D’Andrea E, Montagna M, Manoukian S, Ottini L, Radice P, Viel A, et al. Methyl group metabolism gene polymorphisms as modifier of breast cancer risk in Italian BRCA1/2 carriers. Breast Cancer Res Treat. 2007;103(1):29–36.

  181. He L, Shen Y. MTHFR C677T polymorphism and breast, ovarian cancer risk: a meta-analysis of 19,260 patients and 26,364 controls. Onco Targets Ther. 2017;10:227–38.

    Article  PubMed  PubMed Central  Google Scholar 

  182. Pu D, Jiang SW, Wu J. Association between MTHFR gene polymorphism and the risk of ovarian cancer: a meta-analysis of the literature. Curr Pharm Des. 2014;20(11):1632–8.

    Article  PubMed  CAS  Google Scholar 

  183. Terry KL, Tworoger SS, Goode EL, Gates MA, Titus-Ernstoff L, Kelemen LE, Sellers TA, Hankinson SE, Cramer DW. MTHFR polymorphisms in relation to ovarian cancer risk. Gynecol Oncol. 2010;119(2):319–24.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  184. Ma C, Liu Y, Zhang W, Liu P. The association between MTHFR C677T polymorphism and ovarian cancer risk: a meta-analysis of 18,628 individuals. Mol Biol Rep. 2013;40(3):2061–8.

    Article  PubMed  CAS  Google Scholar 

  185. Zhang L, Liu W, Hao Q, Bao L, Wang K. Folate intake and methylenetetrahydrofolate reductase gene polymorphisms as predictive and prognostic biomarkers for ovarian cancer risk. Int J Mol Sci. 2012;13(4):4009–20.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  186. Webb PM, Ibiebele TI, Hughes MC, Beesley J, van der Pols JC, Chen X, Nagle CM, Bain CJ, Chenevix-Trench G. Folate and related micronutrients, folate-metabolising genes and risk of ovarian cancer. Eur J Clin Nutr. 2011;65(10):1133–40.

    Article  PubMed  CAS  Google Scholar 

  187. Zhang S, Zheng F, Zhang L, Huang Z, Huang X, Pan Z, Chen S, Xu C, Jiang Y, Gu S, et al. LncRNA HOTAIR-mediated MTHFR methylation inhibits 5-fluorouracil sensitivity in esophageal cancer cells. J Exp Clin Cancer Res. 2020;39(1):131.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  188. Yeh CC, Lai CY, Chang SN, Hsieh LL, Tang R, Sung FC, Lin YK. Polymorphisms of MTHFR C677T and A1298C associated with survival in patients with colorectal cancer treated with 5-fluorouracil-based chemotherapy. Int J Clin Oncol. 2017;22(3):484–93.

    Article  PubMed  CAS  Google Scholar 

  189. Wang Z, Chen JQ, Liu JL, Qin XG, Huang Y. Polymorphisms in ERCC1, GSTs, TS and MTHFR predict clinical outcomes of gastric cancer patients treated with platinum/5-Fu-based chemotherapy: a systematic review. Bmc Gastroenterol. 2012;12:137.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  190. Tang C, Yu S, Jiang H, Li W, Xu X, Cheng X, Peng K, Chen E, Cui Y, Liu T. A Meta-Analysis: Methylenetetrahydrofolate Reductase C677T Polymorphism in Gastric Cancer Patients Treated with 5-Fu Based Chemotherapy Predicts Serious Hematologic Toxicity but Not Prognosis. J Cancer. 2018;9(6):1057–66.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  191. Di Virgilio F, Adinolfi E. Extracellular purines, purinergic receptors and tumor growth. Oncogene. 2017;36(3):293–303.

    Article  PubMed  CAS  Google Scholar 

  192. Pellegatti P, Raffaghello L, Bianchi G, Piccardi F, Pistoia V, Di Virgilio F. Increased level of extracellular ATP at tumor sites: in vivo imaging with plasma membrane luciferase. PLoS ONE. 2008;3(7):e2599.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  193. Schultze-Mosgau A, Katzur AC, Arora KK, Stojilkovic SS, Diedrich K, Ortmann O. Characterization of calcium-mobilizing, purinergic P2Y(2) receptors in human ovarian cancer cells. Mol Hum Reprod. 2000;6(5):435–42.

    Article  PubMed  CAS  Google Scholar 

  194. Saito H, Nishimura M, Shinano H, Makita H, Tsujino I, Shibuya E, Sato F, Miyamoto K, Kawakami Y. Plasma concentration of adenosine during normoxia and moderate hypoxia in humans. Am J Respir Crit Care Med. 1999;159(3):1014–8.

    Article  PubMed  CAS  Google Scholar 

  195. Poth JM, Brodsky K, Ehrentraut H, Grenz A, Eltzschig HK. Transcriptional control of adenosine signaling by hypoxia-inducible transcription factors during ischemic or inflammatory disease. J Mol Med (Berl). 2013;91(2):183–93.

    Article  CAS  Google Scholar 

  196. Kusek J, Yang Q, Witek M, Gruber CW, Nanoff C, Freissmuth M. Chaperoning of the A1-adenosine receptor by endogenous adenosine - an extension of the retaliatory metabolite concept. Mol Pharmacol. 2015;87(1):39–51.

    Article  PubMed  CAS  Google Scholar 

  197. Morote-Garcia JC, Rosenberger P, Kuhlicke J, Eltzschig HK. HIF-1-dependent repression of adenosine kinase attenuates hypoxia-induced vascular leak. Blood. 2008;111(12):5571–80.

    Article  PubMed  CAS  Google Scholar 

  198. Allard B, Beavis PA, Darcy PK, Stagg J. Immunosuppressive activities of adenosine in cancer. Curr Opin Pharmacol. 2016;29:7–16.

    Article  PubMed  CAS  Google Scholar 

  199. Linden J, Koch-Nolte F, Dahl G. Purine Release, Metabolism, and Signaling in the Inflammatory Response. Annu Rev Immunol. 2019;37:325–47.

    Article  PubMed  CAS  Google Scholar 

  200. Fredholm BB. Adenosine–a physiological or pathophysiological agent? J Mol Med (Berl). 2014;92(3):201–6.

    Article  CAS  Google Scholar 

  201. Vijayan D, Young A, Teng M, Smyth MJ. Targeting immunosuppressive adenosine in cancer. Nat Rev Cancer. 2017;17(12):709–24.

    Article  PubMed  CAS  Google Scholar 

  202. Ohta A. A Metabolic Immune Checkpoint: Adenosine in Tumor Microenvironment. Front Immunol. 2016;7:109.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  203. Gaudreau PO, Allard B, Turcotte M, Stagg J. CD73-adenosine reduces immune responses and survival in ovarian cancer patients. Oncoimmunology. 2016;5(5):e1127496.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  204. Robles-Martinez L, Garay E, Martel-Gallegos MG, Cisneros-Mejorado A, Perez-Montiel D, Lara A, Arellano RO. Kca3.1 activation via P2y2 purinergic receptors promotes human ovarian cancer cell (Skov-3) migration. Sci Rep. 2017;7(1):4340.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  205. Xia B, Wang J. Adenosine inhibits ovarian cancer growth through regulating rhogdi2 protein expression. Drug Des Devel Ther. 2019;13:3837–44.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  206. Zhu F, Liu XG, Liang NC. Inhibitory effect of adenosine analogues on invasion of human ovarian cancer cell line HO-8910PM]. Ai Zheng. 2004;23(12):1646–50.

    PubMed  CAS  Google Scholar 

  207. Bednarska-Szczepaniak K, Krzyzanowski D, Klink M, Nowak M. Adenosine analogues as opposite modulators of the cisplatin resistance of ovarian cancer cells. Anticancer Agents Med Chem. 2019;19(4):473–86.

    Article  PubMed  CAS  Google Scholar 

  208. Shirali S, Aghaei M, Shabani M, Fathi M, Sohrabi M, Moeinifard M. Adenosine induces cell cycle arrest and apoptosis via cyclinD1/Cdk4 and Bcl-2/Bax pathways in human ovarian cancer cell line OVCAR-3. Tumour Biol. 2013;34(2):1085–95.

    Article  PubMed  CAS  Google Scholar 

  209. Di Virgilio F, Sarti AC, Falzoni S, De Marchi E, Adinolfi E. Extracellular ATP and P2 purinergic signalling in the tumour microenvironment. Nat Rev Cancer. 2018;18(10):601–18.

    Article  PubMed  CAS  Google Scholar 

  210. Woods LT, Forti KM, Shanbhag VC, Camden JM, Weisman GA. P2Y receptors for extracellular nucleotides: Contributions to cancer progression and therapeutic implications. Biochem Pharmacol. 2021;187:114406.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  211. Braganhol E, Wink MR, Lenz G, Battastini A. Purinergic Signaling in Glioma Progression. Adv Exp Med Biol. 2020;1202:87–108.

    Article  PubMed  CAS  Google Scholar 

  212. Kasama H, Sakamoto Y, Kasamatsu A, Okamoto A, Koyama T, Minakawa Y, Ogawara K, Yokoe H, Shiiba M, Tanzawa H, et al. Adenosine A2b receptor promotes progression of human oral cancer. BMC Cancer. 2015;15:563.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  213. Lan J, Lu H, Samanta D, Salman S, Lu Y, Semenza GL. Hypoxia-inducible factor 1-dependent expression of adenosine receptor 2B promotes breast cancer stem cell enrichment. Proc Natl Acad Sci U S A. 2018;115(41):E9640–8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  214. Zhou Y, Chu X, Deng F, Tong L, Tong G, Yi Y, Liu J, Tang J, Tang Y, Xia Y, et al. The adenosine A2b receptor promotes tumor progression of bladder urothelial carcinoma by enhancing MAPK signaling pathway. Oncotarget. 2017;8(30):48755–68.

    Article  PubMed  PubMed Central  Google Scholar 

  215. Cekic C, Linden J. Purinergic regulation of the immune system. Nat Rev Immunol. 2016;16(3):177–92.

    Article  PubMed  CAS  Google Scholar 

  216. Morello S, Miele L. Targeting the adenosine A2b receptor in the tumor microenvironment overcomes local immunosuppression by myeloid-derived suppressor cells. Oncoimmunology. 2014;3:e27989.

    Article  PubMed  PubMed Central  Google Scholar 

  217. Winerdal M, Winerdal ME, Wang YQ, Fredholm BB, Winqvist O, Aden U. Adenosine A1 receptors contribute to immune regulation after neonatal hypoxic ischemic brain injury. Purinergic Signal. 2016;12(1):89–101.

    Article  PubMed  CAS  Google Scholar 

  218. Butler M, Sanmugalingam D, Burton VJ, Wilson T, Pearson R, Watson RP, Smith P, Parkinson SJ. Impairment of adenosine A3 receptor activity disrupts neutrophil migratory capacity and impacts innate immune function in vivo. Eur J Immunol. 2012;42(12):3358–68.

    Article  PubMed  CAS  Google Scholar 

  219. Hajiahmadi S, Panjehpour M, Aghaei M, Mousavi S. Molecular expression of adenosine receptors in OVCAR-3, Caov-4 and SKOV-3 human ovarian cancer cell lines. Res Pharm Sci. 2015;10(1):43–51.

    PubMed  PubMed Central  CAS  Google Scholar 

  220. Campos-Contreras A, Gonzalez-Gallardo A, Diaz-Munoz M, Vazquez-Cuevas FG. Adenosine Receptor A2B Negatively Regulates Cell Migration in Ovarian Carcinoma Cells. Int J Mol Sci. 2022;23(9):4585.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  221. Hajiahmadi S, Panjehpour M, Aghaei M, Shabani M. Activation of A2b adenosine receptor regulates ovarian cancer cell growth: involvement of Bax/Bcl-2 and caspase-3. Biochem Cell Biol. 2015;93(4):321–9.

    Article  PubMed  CAS  Google Scholar 

  222. Abedi H, Aghaei M, Panjehpour M, Hajiahmadi S. Mitochondrial and caspase pathways are involved in the induction of apoptosis by IB-MECA in ovarian cancer cell lines. Tumour Biol. 2014;35(11):11027–39.

    Article  PubMed  CAS  Google Scholar 

  223. Jiang K, Yao G, Hu L, Yan Y, Liu J, Shi J, Chang Y, Zhang Y, Liang D, Shen D, et al. MOB2 suppresses GBM cell migration and invasion via regulation of FAK/Akt and cAMP/PKA signaling. Cell Death Dis. 2020;11(4):230.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  224. Yue W, Ma J, Xiao Y, Wang P, Gu X, Xie B, Li M. The Apoptotic Resistance of BRCA1-Deficient Ovarian Cancer Cells is Mediated by cAMP. Front Cell Dev Biol. 2022;10:889656.

    Article  PubMed  PubMed Central  Google Scholar 

  225. Zhu P, Wang L, Xu P, Tan Q, Wang Y, Feng G, Yuan L. GANT61 elevates chemosensitivity to cisplatin through regulating the Hedgehog, AMPK and cAMP pathways in ovarian cancer. Future Med Chem. 2022;14(7):479–500.

    Article  PubMed  CAS  Google Scholar 

  226. Martinez M, Moon EK. CAR T cells for solid tumors: new strategies for finding, infiltrating, and surviving in the tumor microenvironment. Front Immunol. 2019;10:128.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  227. Liu G, Zhang Q, Liu G, Li D, Zhang L, Gu Z, Tian H, Zhang Y, Tian X. Disruption of adenosine 2A receptor improves the anti-tumor function of anti-mesothelin CAR T cells both in vitro and in vivo. Exp Cell Res. 2021;409(1):112886.

    Article  PubMed  CAS  Google Scholar 

  228. Barcz E, Sommer E, Sokolnicka I, Gawrychowski K, Roszkowska-Purska K, Janik P, Skopinska-Rozewska E. The influence of theobromine on angiogenic activity and proangiogenic cytokines production of human ovarian cancer cells. Oncol Rep. 1998;5(2):517–20.

    PubMed  CAS  Google Scholar 

  229. Barcz E, Sommer E, Janik P, Marianowski L, Skopinska-Rozewska E. Adenosine receptor antagonism causes inhibition of angiogenic activity of human ovarian cancer cells. Oncol Rep. 2000;7(6):1285–91.

    PubMed  CAS  Google Scholar 

  230. Ryzhov S, Novitskiy SV, Zaynagetdinov R, Goldstein AE, Carbone DP, Biaggioni I, Dikov MM, Feoktistov I. Host A(2B) adenosine receptors promote carcinoma growth. Neoplasia. 2008;10(9):987–95.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  231. Di Virgilio F, Giuliani AL, Vultaggio-Poma V, Falzoni S, Sarti AC. Non-nucleotide Agonists Triggering P2X7 Receptor Activation and Pore Formation. Front Pharmacol. 2018;9:39.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  232. Tomasinsig L, Pizzirani C, Skerlavaj B, Pellegatti P, Gulinelli S, Tossi A, Di Virgilio F, Zanetti M. The human cathelicidin LL-37 modulates the activities of the P2X7 receptor in a structure-dependent manner. J Biol Chem. 2008;283(45):30471–81.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  233. Norenberg W, Hempel C, Urban N, Sobottka H, Illes P, Schaefer M. Clemastine potentiates the human P2X7 receptor by sensitizing it to lower ATP concentrations. J Biol Chem. 2011;286(13):11067–81.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  234. Zhang Y, Cheng H, Li W, Wu H, Yang Y. Highly-expressed P2X7 receptor promotes growth and metastasis of human HOS/MNNG osteosarcoma cells via PI3K/Akt/GSK3beta/beta-catenin and mTOR/HIF1alpha/VEGF signaling. Int J Cancer. 2019;145(4):1068–82.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  235. Brisson L, Chadet S, Lopez-Charcas O, Jelassi B, Ternant D, Chamouton J, Lerondel S, Le Pape A, Couillin I, Gombault A, et al. P2X7 receptor promotes mouse mammary cancer cell invasiveness and tumour progression, and is a target for anticancer treatment. Cancers (Basel). 2020;12(9):2342.

    Article  CAS  Google Scholar 

  236. Arnaud-Sampaio VF, Rabelo I, Ulrich H, Lameu C. The P2X7 receptor in the maintenance of cancer stem cells, chemoresistance and metastasis. Stem Cell Rev Rep. 2020;16(2):288–300.

    Article  PubMed  Google Scholar 

  237. Hofman P, Cherfils-Vicini J, Bazin M, Ilie M, Juhel T, Hebuterne X, Gilson E, Schmid-Alliana A, Boyer O, Adriouch S, et al. Genetic and pharmacological inactivation of the purinergic P2RX7 receptor dampens inflammation but increases tumor incidence in a mouse model of colitis-associated cancer. Cancer Res. 2015;75(5):835–45.

    Article  PubMed  CAS  Google Scholar 

  238. de Torre-Minguela C, Barbera-Cremades M, Gomez AI, Martin-Sanchez F, Pelegrin P. Macrophage activation and polarization modify P2X7 receptor secretome influencing the inflammatory process. Sci Rep. 2016;6:22586.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  239. Vazquez-Cuevas FG, Cruz-Rico A, Garay E, Garcia-Carranca A, Perez-Montiel D, Juarez B, Arellano RO. Differential expression of the P2X7 receptor in ovarian surface epithelium during the oestrous cycle in the mouse. Reprod Fertil Dev. 2013;25(7):971–84.

    Article  PubMed  CAS  Google Scholar 

  240. Adinolfi E, Raffaghello L, Giuliani AL, Cavazzini L, Capece M, Chiozzi P, Bianchi G, Kroemer G, Pistoia V, Di Virgilio F. Expression of P2X7 receptor increases in vivo tumor growth. Cancer Res. 2012;72(12):2957–69.

    Article  PubMed  CAS  Google Scholar 

  241. Vazquez-Cuevas FG, Martinez-Ramirez AS, Robles-Martinez L, Garay E, Garcia-Carranca A, Perez-Montiel D, Castaneda-Garcia C, Arellano RO. Paracrine stimulation of P2X7 receptor by ATP activates a proliferative pathway in ovarian carcinoma cells. J Cell Biochem. 2014;115(11):1955–66.

    PubMed  CAS  Google Scholar 

  242. Ferrari D, Chiozzi P, Falzoni S, Dal Susino M, Melchiorri L, Baricordi OR, Di Virgilio F. Extracellular ATP triggers IL-1 beta release by activating the purinergic P2Z receptor of human macrophages. J Immunol. 1997;159(3):1451–8.

    PubMed  CAS  Google Scholar 

  243. Luborsky J, Barua A, Edassery S, Bahr JM, Edassery SL. Inflammasome expression is higher in ovarian tumors than in normal ovary. PLoS ONE. 2020;15(1):e227081.

    Article  CAS  Google Scholar 

  244. Wu H, Liu J, Zhang Y, Li Q, Wang Q, Gu Z. miR-22 suppresses cell viability and EMT of ovarian cancer cells via NLRP3 and inhibits PI3K/AKT signaling pathway. Clin Transl Oncol. 2021;23(2):257–64.

    Article  PubMed  CAS  Google Scholar 

  245. Alrashed MM, Alharbi H, Alshehry AS, Ahmad M, Aloahd MS. MiR-624-5p enhances NLRP3 augmented gemcitabine resistance via EMT/IL-1beta/Wnt/beta-catenin signaling pathway in ovarian cancer. J Reprod Immunol. 2022;150:103488.

    Article  PubMed  CAS  Google Scholar 

  246. Ding Y, Yan Y, Dong Y, Xu J, Su W, Shi W, Zou Q, Yang X. NLRP3 promotes immune escape by regulating immune checkpoints: A pan-cancer analysis. Int Immunopharmacol. 2022;104:108512.

    Article  PubMed  CAS  Google Scholar 

  247. Salaro E, Rambaldi A, Falzoni S, Amoroso FS, Franceschini A, Sarti AC, Bonora M, Cavazzini F, Rigolin GM, Ciccone M, et al. Involvement of the P2X7-NLRP3 axis in leukemic cell proliferation and death. Sci Rep. 2016;6:26280.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  248. Solini A, Cobuccio L, Rossi C, Parolini F, Biancalana E, Cosio S, Chiarugi M, Gadducci A. Molecular characterization of peritoneal involvement in primary colon and ovary neoplasm: the possible clinical meaning of the p2x7 receptor-inflammasome complex. Eur Surg Res. 2021:1–9.

  249. Zhang Y, Li F, Wang L, Lou Y. A438079 affects colorectal cancer cell proliferation, migration, apoptosis, and pyroptosis by inhibiting the P2X7 receptor. Biochem Biophys Res Commun. 2021;558:147–53.

    Article  PubMed  CAS  Google Scholar 

  250. Jacobson KA, Muller CE. Medicinal chemistry of adenosine, P2Y and P2X receptors. Neuropharmacology. 2016;104:31–49.

    Article  PubMed  CAS  Google Scholar 

  251. Erb L, Weisman GA. Coupling of P2Y receptors to G proteins and other signaling pathways. Wiley Interdiscip Rev Membr Transp Signal. 2012;1(6):789–803.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  252. Wei WC, Jacobs B, Becker EB, Glitsch MD. Reciprocal regulation of two G protein-coupled receptors sensing extracellular concentrations of Ca2+ and H. Proc Natl Acad Sci U S A. 2015;112(34):10738–43.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  253. Martinez-Ramirez AS, Garay E, Garcia-Carranca A, Vazquez-Cuevas FG. The P2RY2 receptor induces carcinoma cell migration and EMT through cross-talk with epidermal growth factor receptor. J Cell Biochem. 2016;117(4):1016–26.

    Article  PubMed  CAS  Google Scholar 

  254. Egan K, Crowley D, Smyth P, O’Toole S, Spillane C, Martin C, Gallagher M, Canney A, Norris L, Conlon N, et al. Platelet adhesion and degranulation induce pro-survival and pro-angiogenic signalling in ovarian cancer cells. PLoS ONE. 2011;6(10):e26125.

  255. Davis AN, Afshar-Kharghan V, Sood AK. Platelet effects on ovarian cancer. Semin Oncol. 2014;41(3):378–84.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  256. Cho MS, Noh K, Haemmerle M, Li D, Park H, Hu Q, Hisamatsu T, Mitamura T, Mak S, Kunapuli S, et al. Role of ADP receptors on platelets in the growth of ovarian cancer. Blood. 2017;130(10):1235–42.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  257. Parker WB. Enzymology of purine and pyrimidine antimetabolites used in the treatment of cancer. Chem Rev. 2009;109(7):2880–93.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  258. Lee SA, Jung M. The nucleoside analog sangivamycin induces apoptotic cell death in breast carcinoma MCF7/adriamycin-resistant cells via protein kinase Cdelta and JNK activation. J Biol Chem. 2007;282(20):15271–83.

    Article  PubMed  CAS  Google Scholar 

  259. Issaeva N, Thomas HD, Djureinovic T, Jaspers JE, Stoimenov I, Kyle S, Pedley N, Gottipati P, Zur R, Sleeth K, et al. 6-thioguanine selectively kills BRCA2-defective tumors and overcomes PARP inhibitor resistance. Cancer Res. 2010;70(15):6268–76.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  260. Roberts C, Strauss VY, Kopijasz S, Gourley C, Hall M, Montes A, Abraham J, Clamp A, Kennedy R, Banerjee S, et al. Results of a phase II clinical trial of 6-mercaptopurine (6MP) and methotrexate in patients with BRCA-defective tumours. Br J Cancer. 2020;122(4):483–90.

    Article  PubMed  CAS  Google Scholar 

  261. Albasher G, Alkahtane AA, Alarifi S, Ali D, Alessia MS, Almeer RS, Abdel-Daim MM, Al-Sultan NK, Al-Qahtani AA, Ali H, et al. Methotrexate-induced apoptosis in human ovarian adenocarcinoma SKOV-3 cells via ROS-mediated bax/bcl-2-cyt-c release cascading. Onco Targets Ther. 2019;12:21–30.

    Article  PubMed  CAS  Google Scholar 

  262. Penn CA, Yang K, Zong H, Lim JY, Cole A, Yang D, Baker J, Goonewardena SN, Buckanovich RJ. Therapeutic impact of nanoparticle therapy targeting tumor-associated macrophages. Mol Cancer Ther. 2018;17(1):96–106.

    Article  PubMed  CAS  Google Scholar 

  263. Zhang L, Zou W. Inhibition of integrin beta1 decreases the malignancy of ovarian cancer cells and potentiates anticancer therapy via the FAK/STAT1 signaling pathway. Mol Med Rep. 2015;12(6):7869–76.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  264. Fang J, Cao Z, Chen YC, Reed E, Jiang BH. 9-beta-D-arabinofuranosyl-2-fluoroadenine inhibits expression of vascular endothelial growth factor through hypoxia-inducible factor-1 in human ovarian cancer cells. Mol Pharmacol. 2004;66(1):178–86.

    Article  PubMed  CAS  Google Scholar 

  265. Zaffaroni N, Orlandi L, Gornati D, De Marco C, Vaglini M, Silvestrini R. Fludarabine as a modulator of cisplatin activity in human tumour primary cultures and established cell lines. Eur J Cancer. 1996;32A(10):1766–73.

    Article  PubMed  CAS  Google Scholar 

  266. Xue J, Bi X, Wu G, Meng D, Fang J. Fludarabine reduces survivability of HepG2 cells through VEGF under hypoxia. Arch Biochem Biophys. 2007;468(1):100–6.

    Article  PubMed  CAS  Google Scholar 

  267. Frelinger AR, Michelson AD, Wiviott SD, Trenk D, Neumann FJ, Miller DL, Jakubowski JA, Costigan TM, Mccabe CH, Antman EM, et al. Intrinsic platelet reactivity before P2Y12 blockade contributes to residual platelet reactivity despite high-level P2Y12 blockade by prasugrel or high-dose clopidogrel. Results from PRINCIPLE-TIMI 44. Thromb Haemost. 2011;106(2):219–26.

    PubMed  CAS  Google Scholar 

  268. Bergmann TK, Filppula AM, Launiainen T, Nielsen F, Backman J, Brosen K. Neurotoxicity and low paclitaxel clearance associated with concomitant clopidogrel therapy in a 60-year-old Caucasian woman with ovarian carcinoma. Br J Clin Pharmacol. 2016;81(2):313–5.

    Article  PubMed  CAS  Google Scholar 

  269. Park SH, Byon JS, Lee SW, Lee SJ, Jin DK, Shin WY. Coronary artery thrombosis associated with Paclitaxel in advanced ovarian cancer. Korean Circ J. 2009;39(3):124–7.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  270. Nicum S, Roberts C, Boyle L, Kopijasz S, Gourley C, Hall M, Montes A, Poole C, Collins L, Schuh A, et al. A phase II clinical trial of 6-mercaptopurine (6MP) and methotrexate in patients with BRCA defective tumours: a study protocol. BMC Cancer. 2014;14:983.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  271. Estlin EJ. Continuing therapy for childhood acute lymphoblastic leukaemia: clinical and cellular pharmacology of methotrexate, 6-mercaptopurine and 6-thioguanine. Cancer Treat Rev. 2001;27(6):351–63.

    Article  PubMed  CAS  Google Scholar 

  272. El-Husseiny K, Motawei H, Ali MS. Continuous low-dose oral cyclophosphamide and methotrexate as maintenance therapy in patients with advanced ovarian carcinoma after complete clinical response to platinum and paclitaxel chemotherapy. Int J Gynecol Cancer. 2016;26(3):437–42.

    Article  PubMed  Google Scholar 

  273. Hortu I, Ozceltik G, Ergenoglu AM, Yigitturk G, Atasoy O, Erbas O. Protective effect of oxytocin on a methotrexate-induced ovarian toxicity model. Arch Gynecol Obstet. 2020;301(5):1317–24.

    Article  PubMed  CAS  Google Scholar 

  274. O’Brien S, Kantarjian H, Keating MJ. Purine analogs in chronic lymphocytic leukemia and Waldenstrom’s macroglobulinemia. Ann Oncol. 1996;7(Suppl 6):S27–33.

  275. Hoogstad-Van EJ, Bekkers R, Ottevanger N, Schaap N, Hobo W, Jansen JH, Massuger L, Dolstra H. Intraperitoneal infusion of ex vivo-cultured allogeneic NK cells in recurrent ovarian carcinoma patients (a phase I study). Medicine (Baltimore). 2019;98(5):e14290.

    Article  CAS  Google Scholar 

  276. Geller MA, Cooley S, Judson PL, Ghebre R, Carson LF, Argenta PA, Jonson AL, Panoskaltsis-Mortari A, Curtsinger J, Mckenna D, et al. A phase II study of allogeneic natural killer cell therapy to treat patients with recurrent ovarian and breast cancer. Cytotherapy. 2011;13(1):98–107.

    Article  PubMed  CAS  Google Scholar 

  277. Teneriello M, Farley J, Parker M, O’Connor D, Shaver T, Park R, Barnhill D. Management of advanced ovarian epithelial cancer in the renal transplant patient. Gynecol Oncol. 1993;50(3):374–8.

  278. Fu R, Sutcliffe D, Zhao H, Huang X, Schretlen DJ, Benkovic S, Jinnah HA. Clinical severity in Lesch-Nyhan disease: the role of residual enzyme and compensatory pathways. Mol Genet Metab. 2015;114(1):55–61.

    Article  PubMed  CAS  Google Scholar 

  279. Abbracchio MP, Burnstock G, Verkhratsky A, Zimmermann H. Purinergic signalling in the nervous system: an overview. Trends Neurosci. 2009;32(1):19–29.

    Article  PubMed  CAS  Google Scholar 

  280. Chatzidoukaki O, Goulielmaki E, Schumacher B, Garinis GA. DNA damage response and metabolic reprogramming in health and disease. Trends Genet. 2020;36(10):777–91.

    Article  PubMed  CAS  Google Scholar 

  281. Mathews CK. DNA precursor metabolism and genomic stability. Faseb J. 2006;20(9):1300–14.

    Article  PubMed  CAS  Google Scholar 

  282. Kozhevnikova EN, van der Knaap JA, Pindyurin AV, Ozgur Z, van Ijcken WF, Moshkin YM, Verrijzer CP. Metabolic enzyme IMPDH is also a transcription factor regulated by cellular state. Mol Cell. 2012;47(1):133–9.

    Article  PubMed  CAS  Google Scholar 

  283. Coquel F, Silva MJ, Techer H, Zadorozhny K, Sharma S, Nieminuszczy J, Mettling C, Dardillac E, Barthe A, Schmitz AL, et al. SAMHD1 acts at stalled replication forks to prevent interferon induction. Nature. 2018;557(7703):57–61.

    Article  PubMed  CAS  Google Scholar 

  284. Du B, Zhang Z, Di W, Xu W, Yang L, Zhang S, He G, Yang R, Wang M. PAICS is related to glioma grade and can promote glioma growth and migration. J Cell Mol Med. 2021;25(16):7720–33.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  285. Huang N, Xu C, Deng L, Li X, Bian Z, Zhang Y, Long S, Chen Y, Zhen N, Li G, et al. PAICS contributes to gastric carcinogenesis and participates in DNA damage response by interacting with histone deacetylase 1/2. Cell Death Dis. 2020;11(7):507.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  286. Fulda S. Tumor resistance to apoptosis. Int J Cancer. 2009;124(3):511–5.

    Article  PubMed  CAS  Google Scholar 

  287. Ma F, Zhu Y, Liu X, Zhou Q, Hong X, Qu C, Feng X, Zhang Y, Ding Q, Zhao J, et al. Dual-specificity tyrosine phosphorylation-regulated kinase 3 loss activates purine metabolism and promotes hepatocellular carcinoma progression. Hepatology. 2019;70(5):1785–803.

    Article  PubMed  CAS  Google Scholar 

  288. Feng X, Ma D, Zhao J, Song Y, Zhu Y, Zhou Q, Ma F, Liu X, Zhong M, Liu Y, et al. UHMK1 promotes gastric cancer progression through reprogramming nucleotide metabolism. Embo J. 2020;39(5):e102541.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  289. Zhou Q, Lin M, Feng X, Ma F, Zhu Y, Liu X, Qu C, Sui H, Sun B, Zhu A, et al. Targeting CLK3 inhibits the progression of cholangiocarcinoma by reprogramming nucleotide metabolism. J Exp Med. 2020;217(8):e20191779.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

Grant sponsor: Hubei Province's Outstanding Medical Academic Leader Programme and the Fundamental Research Funds for the Central Universities; Grant numbers: 2042021kf0125.

Author information

Authors and Affiliations

Authors

Contributions

 J Liu and S Hong provided ideas and wrote articles. J Yang, X Zhang and L Hong reviewed and proofread the article. Y Wang, H Wang and J Peng checked and modified the figures and tables. J Liu and S Hong contributed equally to this work. The author(s) read and approved the final manuscript.

Corresponding author

Correspondence to Li Hong.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

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

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, J., Hong, S., Yang, J. et al. Targeting purine metabolism in ovarian cancer. J Ovarian Res 15, 93 (2022). https://doi.org/10.1186/s13048-022-01022-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13048-022-01022-z

Keywords