Skip to main content

The emerging role of glycolysis and immune evasion in ovarian cancer

Abstract

Ovarian cancer (OC) is one of the three most common malignant tumors of the female reproductive system, with the highest mortality rate among gynecologic malignancies. Like other tumors, OC cells undergo metabolic reprogramming phenomenon and convert glucose metabolism into “aerobic glycolysis” and generate a high concentration of lactate, i.e., the “Warburg effect”, which provides a large amount of energy and corresponding intermediary metabolites for their survival, reproduction and metastasis. Numerous studies have shown that targeted inhibition of aerobic glycolysis and lactate metabolism is a promising strategy to enhance the sensitivity of cancer cells to immunotherapy. Therefore, this review summarizes the metabolic features of glycolysis in OC cells and highlights how abnormal lactate concentration affects the differentiation, metabolism, and function of infiltrating immune cells, which contributes to immunosuppression, and how targeted inhibition of this phenomenon may be a potential strategy to enhance the therapeutic efficacy of OC.

Introduction

The metabolic activities of malignant tumor cells differ significantly from those of normal cells, and this special metabolic pattern is known as the phenomenon of tumor metabolic reprogramming [1]. Glucose as the most important material energy source of the cell, in normal cells is mainly through the oxidative phosphorylation metabolic pathway to consume glucose to produce energy, but in malignant tumor cells even in the oxygen-supplied sufficient environment, will be through a series of glycolysis under the action of glucose consumption of large quantities of glucose and the production of large quantities of lactate, so as to rapidly obtain energy, this metabolism is called aerobic glycolysis, namely “Warburg effect”, which is one of the first metabolic reprogramming discovered [2, 3]. Furthermore, this aerobic glycolysis can provide a large amount of energy as well as corresponding intermediate metabolites for metabolic processes such as tumor cell survival and reproduction [4, 5]. The end product of aerobic glycolysis is lactate, which acidifies the tumor microenvironment (TME) and can act as a multifunctional small molecule chemical to promote tumor development [6].

Ovarian cancer (OC) is a malignant neoplastic disease that occurs in the ovaries and is the most common malignant tumor in women, which is a major challenge that threatens the life and health of women [7]. Despite the development of early diagnosis and treatment methods, the overall incidence is still rising, and the five-year survival rate in Asian countries remains low at 59.6% [8]. OC has a unique TME, which is closely related to tumor proliferation, invasion and metastasis. A study showed that numerous cytokines in follicular fluid such as insulin-like growth factors (IGF), epidermal growth factor (EGF) and transforming growth factor-β (TGF-β) induce tumor cell proliferation and invasion [9]. Furthermore, OC has a high rate of peritoneal metastasis, and the fatty acid-rich peritoneal microenvironment composed of peritoneal fluid and peritoneum creates favorable conditions for the progression and metastasis of OC [10, 11], and alterations in glycolytic processes also create a facility for the development of peritoneal metastasis of OC, which distinguishes it from other modes of cancer metastasis [12]. Most OC originates from ovarian epithelial cells, and like other tumors, it exhibits the “Warburg effect”. OC cells create a localized hypoxic, hypoglycemic and acidified environment by consuming large amounts of glucose through aerobic glycolysis and generating the end product lactate. The acidified TME inhibits the activity of immune cells and helps OC cells evade the immune system, promoting their metastasis and invasion [11]. Further understanding of the effects of glycolysis and its metabolites on immune cell function in OC could provide better ideas for rational immunotherapy, leading to better efficacy.

Characteristic of aerobic glycolysis in OC

Enhancement of aerobic glycolysis in OC

Cancer cells follow the “Warburg effect”, and their metabolism is significantly different from that of normal cells. Fong et al. [13] used gas chromatography/mass spectrometry (GC/MS) and liquid chromatography tandem mass spectrometry (LC/MS/MS) methods to report the difference of metabolites between normal ovary and OC. Several intermediate metabolites and their derivatives in aerobic glycolysis, such as erythronate, fumarate, malate, etc. were found to be differential metabolites, which were significantly increased in cancer tissues. The altered glucose metabolism may serve as a crucial factor for distinguishing OC cells from normal cells. In another study, Cheng et al. [14] used LC-MS on serum from patients with OC and benign ovarian tumors and found that altered metabolites such as maltose, maltotriose, raffinose, and mannitol, which further caused disruption of the tricarboxylic acid (TCA) cycle, could differentiate benign tumors. Similarly, other researchers have observed abnormalities in glucose metabolism in OC [15, 16]. Loar P et al. [17] demonstrated that inhibition of glycolysis in OC cells enhances the effect of chemotherapeutic agents acting on cancer cells to cause apoptosis, and that the glycolytic phenotype of OC cells may be associated with chemotherapy resistance and could be targeted for therapeutic intervention. Wu et al. [18] further confirmed that Paris saponins VII (PSVII) inhibited glycolysis in OC cells, thus inhibiting cancer cell proliferation, inducing apoptosis and exerting its anticancer effect.

Proto-oncogenes and tumor suppressor genes can regulate cellular glucose metabolism. The development of cancer is closely related to the activation of proto-oncogenes and the inactivation of tumor suppressor genes, which cause corresponding changes in cellular glucose metabolism. Activation of the proto-oncogene c-MYC activates cellular glycolysis by transcriptase-upregulating the expression of key glycolytic enzymes, including glucose transporter 1(GLUT1), lactate dehydrogenase A (LDHA), and some key points of glucose metabolism such as hexokinase 2 (HK2) and pyruvate dehydrogenase kinase 1(PDK1) [19]. Human pituitary tumor transforming gene (PTTG), as a proto-oncogene, is involved in the occurrence, invasion and metastasis of OC. Wang et al. [20] found that the expression level of PTTG was synchronized with aerobic glycolysis in OC cells, and the expression of c-MYC and several key enzymes of aerobic glycolysis, such as GLUT1, LDHA and pyruvate Kinase M2 (PKM2), was down-regulated with the inhibition of PTTG, and oxidative phosphorylation was increased. Han et al. [21] showed that p53 regulates the intracellular transport of HK2, and in chemo-sensitive OC cells, phosphorylated p53 promotes the translocation of HK2 and apoptosis-inducing factor (AIF) from the mitochondria to the nucleus, which inhibits HK2-mediated glucose metabolism and induces apoptosis in cancer cells. The TP53 gene up-regulated TIGAR while down-regulating the expression of GLUT1 and some key enzymes of glycolysis, thereby reducing glucose consumption and lactate production in OC cells and inhibiting glycolysis [22]. Deletion of the tumor suppressor gene BRCA1 increases glycolytic activity by inducing the transcription factors MYC and STAT3 to activate HK2 expression [23]. In addition, studies have found that PI3K/AKT signaling activation plays an important role in tumor aerobic glycolysis, which can regulate the activity and expression of some glycolytic enzymes such as HK2, PFK, and PK. Oncogene ACTL6A not only mediates glucose utilization, lactate production and pyruvate levels by regulating PGK1, but also can be up-regulated by follicle-stimulating hormone through the PI3K/AKT pathway to promote glycolysis of OC [24]. In contrast to normal cells, cancer cells obtain their required energy mainly through glycolysis even under aerobic conditions. Although it is a much less efficient form of energy production than mitochondrial OXPHOS, the high level of glycolysis can provide intermediates and macromolecules for cancer cell biosynthesis, and these changes are considered to be the evolutionary advantages of cancer cells [25].

In addition, some studies have shown that cancer cells can alter TME to improve the fitness of cancer cells and their growth advantage through the lactate produced by glycolysis. Lactate derived from tumor cells can be released into the extracellular microenvironment in large quantities, resulting in a concentration of lactate 10 times that of the normal tissue microenvironment, and the pH value can be as low as 6.0-6.8 [26]. Battista et al. [27] detected that the lactate concentration of OC patients ranged from 17.51–37.16µmol/g, which was significantly higher than that of benign samples (5.43–6.54µmol/g). Due to the massive consumption of nutrients by tumor cells, when solid tumor tissues are still in the stage of insufficient vascular supply, they usually present a state of nutrient deficiency. Therefore, tumor cells require multiple mechanisms to maintain their own energy supply, including the phenomenon of “lactate exchange” between tumor cells in different tissue regions [28]. The TME shows high heterogeneity. According to the different oxygen levels in the extracellular environment, tumor cells can be divided into oxygen-rich tumor cells (oxidative type) or hypoxic tumor cells (glycolytic type). For both glycolytic and oxidized cancer cells, the use of alternative molecules such as lactate or glutamine as an energy source depends on the anatomical distribution of tumor cells, with glycolytic cancer cells located further away from blood vessels, whereas oxidized cancer cells are located around blood vessels [29]. Monocarboxylic transporters (MCTS) are crucial for transporting lactate and protons to the TME in cancer cells [30]. Specifically, tumor cells located deep in the TME, far from the oxygen supply in the blood, can utilize glycolysis and glutamine as energy sources and transport large amount of lactate to the extracellular environment by high expression of MCT4, while cancer cells located close to blood vessels with normal oxygen supply are more likely to take up lactate by high expression of MCT1 and energy is obtained through the oxidation of lactate in mitochondria and the TCA cycle [31,32,33]. Hypoxic cancer cells with a glycolytic phenotype promote the high expression of MCT4 by up-regulating hypoxia-inducible factor-1α (HIF-1α), which leads to rapid lactate transport [34]. Furthermore, inhibition of MCT1 impaired tumor-cell growth and glycolysis, reduced glucose transport, and decreased energy production and glutathione (GSH) levels. The intracellular hydrogen peroxide level increases with the decrease of GSH, resulting in mitochondrial damage and cell death [35]. Thus, lactate transport through the TME between different cell populations plays a crucial role in tumor cell growth and progression (Fig. 1). Lactate, the metabolite of tumor cells, accumulates in the TME, which can inhibit the function and metabolism of immune cells and induce tumor immunosuppressive microenvironment [36]. A growing body of research suggests that lactate and the acidic environment it creates have a broad and important role in the regulation of various natural immune cells in tumor tissues.

Fig. 1
figure 1

Regulation of lactate metabolism by oncogenes as well as other factors and the “lactate exchange” phenomenon between tumor cells. Hypoxic tumor cells (glycolytic type) can utilize glycolysis and glutamine as energy sources and transport large amounts of lactate produced to the TME through the high expression of MCT4, leaving it in a low pH acidic environment and inhibiting immune cell function and metabolism. Oxygen-rich tumor cells (oxidative type) preferred to take up lactate through high expression of MCT1 and gained energy through oxidation of lactate and TCA cycle

Key enzymes in aerobic glycolysis in OC

Aerobic glycolysis has three rate-limiting enzymes, hexokinase (HK), phosphofructokinase (PFK) and pyruvate kinases (PKs) (Fig. 2). Changes in the expression of these enzymes significantly influence the progression of OC.

Fig. 2
figure 2

Three rate-limiting enzymes in aerobic glycolysis. HK2, PFK1, and PKM2 are the three rate-limiting enzymes involved in glycolysis. HK2 catalyzes the conversion of glucose to G-6-P and binds to VDAC1 on the outer mitochondrial membrane, which allows preferential entry of ATP into the HK2-mediated glycolytic step via the open VDAC1. PFK1 catalyzes the conversion of F-6-P to F-1,6-BP, and its activity is regulated by the catalytic product F-2,6-BP of PFKFB3. PKM2 not only catalyzes PEP to pyruvate, but also translocates to the nucleus and acts as a protein kinase to phosphorylate histone H3 and activate various gene transcripts, such as c-MYC, STAT3, and NF-κB, which indirectly supports aerobic glycolysis and promotes the progression of tumor cell proliferation

Hexokinase 2 (HK2)

HK2, the first key rate-limiting enzyme of the “Warburg effect”, catalyzes the conversion of glucose to G-6-P, which is highly expressed in OC tissues and correlates directly with advanced and high-grade cancers and plays a role as an independent prognostic factor [37]. HK2 can bind to the Voltage-Dependent Anion Channel 1 (VDAC1), which inhibits the movement of the VDAC1 N-terminal helix and permits ATP to preferentially enter the HK2-mediated glycolytic step via the open VDAC1, thereby promoting aerobic glycolysis [38]. HK2 expression in OC is regulated by a variety of signaling pathways and transcription factors, including the PI3K/AKT signaling pathway, GPNMB/HIF-1α axis, STAT3 and miR-603 [39,40,41,42].

In OC, elevated HK2 expression can lead to chemoresistance. Yu et al. [43] demonstrated that mutations in the UBA structural domain of the autophagy receptor p62 altered the overall abundance and mitochondrial localization of HK2 in cells, which increased the phosphorylated ubiquitin form of parkin, stabilized mitochondrial autophagic activity, and ultimately induced cisplatin resistance. However, knockdown of HK2 could limit cisplatin-induced ERK1/2 phosphorylation and autophagy, which reduced cisplatin resistance in OC cells [44]. In addition, Li et al. [45] found that increased HK2 expression and binding to VDAC1 induced by the SBD structural domain of HSP70 maintained the integrity of the mitochondrial permeability transition pore (MPTP), which reduced apoptosis and promoted cisplatin resistance in OC. Based on its critical role in OC, HK2 is considered a promising metabolic target for the development of new OC therapeutic approaches.

Phosphofructokinase-1 (PFK1)

Phosphofructokinase-1 (PFK1) serves as the second rate-limiting enzyme in glycolysis and exhibits the most potent regulatory influence on glycolysis metabolism, which catalyzes the conversion of Fructose-6-P into Fructose-1,6-BP. There are three PFK1 isoforms existing in mammals, including PFK-L (liver), PFK-M (muscle), and PFK-P (platelet), and the varying expression levels of these isoforms in human tissues are potentially associated with energy metabolism requirements [46].

PFK1 is active as a tetramer but less active as a monomer and dimer, and this structural transition in PFK1 largely influences glycolytic fluxes [47]. For example, downstream products of the PFK1 enzyme, including high concentrations of ATP, phosphoenolpyruvate (PEP), citrate, and lactate, can induce the dissociation of the PFK1 tetramer into a dimer, leading to the inhibition of PFK1 activity and providing negative feedback to the glycolytic process [48]. In contrast, Fructose-2,6-BP, a product of PFKFB3-catalyzed Fructose-6-P generation, was considered the most potent allosteric activator of PFK1, capable of stabilizing the PFK1 tetramer structure and counteracting the inhibition of PFK1 by high concentrations of ATP [49].

Cancer cells express different levels of PFK-2/FBPase-2 isozymes and regulate their relative kinase and bisphosphatase activities in a spatial or temporal manner according to metabolic needs [50]. PFKFB3 is widely overexpressed in OC, contributes to increased PFK1 activity, promotes aerobic glycolysis in OC cells, and is associated with poor patient prognosis [51]. Meanwhile, there is growing evidence that high levels of PFKFB3 are involved in chemoresistance in OC cells. Jiang et al. [52] demonstrated that PFKFB3 may enhance aerobic glycolysis and promote OC cell metastasis by regulating the apoptosis protein inhibitors and NF-κB signaling pathway. PFK-158 and 3-PO inhibit glycolysis by reducing glucose uptake, ATP production, and lactate metabolism through inhibition of PFKFB3 activity, which ameliorate resistance to carboplatin, paclitaxel and cisplatin in OC [52, 53]. In addition, knockdown of PFKP/PFKFB3 in OC-resistant cells also reduces levels of lactate and sensitizes the cells to cisplatin treatment, accompanied by increased poly ADP-ribose polymerase (PARP) cleavage [54]. It follows that PFKFB3 plays an important role in the regulation of glycolysis as well as growth and metastasis of OC. Therefore, targeting PFKFB3 to inhibit PFK1 activity could be a potential therapeutic option for OC. In addition, PFKFB2 also plays a role in OC development and chemoresistance. It was found that high CXCL14-expressing CAFs mediated the binding of LINC00092 to PFKFB2 in OC cells, thereby enhancing glycolysis of OC cells and the glycolytic phenotype of OC cells in turn contributed to the maintenance of local supportive functions of CAFs, enhancing the malignant behavior of the tumor [55]. Yang et al. [56] found that knockdown of PFKFB2, while increasing the rate of glycolysis, reduced the flow of intermediates through the pentose phosphate pathway in cancer cell lines with the wtTP53 gene, thereby decreasing NADPH levels, and that the accumulation of ROS after knockdown of PFKFB2 stimulated the phosphorylation of JNK, which induced G1 cell cycle arrest and apoptosis and reduced paclitaxel resistance. Similarly, PFKFB4 depletion enhances caspase 3/7 activity, increases ROS levels in OC cells during prolonged mitotic arrest, and promotes mitotic cell death after paclitaxel treatment [57]. Therefore, inhibitors of PFKFB2 or PFKFB4 in combination with paclitaxel may be an additional beneficial alternative for the treatment of OC.

Pyruvate kinases, type M2(PKM2)

Pyruvate kinase is the last rate-limiting enzyme in the regulation of glycolytic metabolism and serves to catalyze the conversion of PEP and ADP to pyruvate and ATP. In mammals, the PK family consists of four isoforms: liver-type PK (PKL), red blood cell PK (PKR) and PK muscle isozyme M1 and M2 (PKM1 and PKM2, respectively) [58]. Among them, PKM2 showed a high level of expression in OC cells and a negative correlation with prognosis [59].

PKM2 exists as an inactive monomer, a less active dimer and an active tetramer. Among them, PKM2 tetramer increases glycolytic flux and promotes high ATP production, whereas dimeric PKM2 not only directly promotes aerobic glycolysis by redirecting glucose-generated carbon to biogenic enzymes, but also translocases to the nucleus and acts as a protein kinase to phosphorylate histone H3 and activate the transcription of various genes such as c-MYC and STAT3, which indirectly supports aerobic glycolysis and promoting the progression of tumor cell proliferation [60,61,62,63].

With regard to the regulation of OC pathogenesis by PKM2, Sun et al. [64] found that PSMD14 mediated the K63-linked deubiquitylation on PKM2, resulting in an increase in its dimerization ratio and nuclear translocation, which promoted aerobic glycolysis in OC cells, thereby stimulating the proliferation and invasion of OC cells. PKM2 could be targeted and upregulated by AKT2, which mediated elevated STAT3 expression and activation of nuclear translocation of NF-κB p65, increasing migration and invasion of OC cells in vitro [65]. Similarly, PKM2 expression could be upregulated by follicle stimulating hormone (FSH), which promoted aerobic glycolysis, and stimulated OC cell proliferation and cisplatin resistance [66]. In addition, PKM2 overexpression also affected the expression levels of several tumor-associated genes, such as up-regulation of CCND1 expression and reduction of CDKN1A expression, which increased the number of S-phase cells and induced the proliferation and survival of OC cells [67]. Whereas compound 3 K specifically inhibited pyruvate kinase M2, induced AMPK activation, accompanied by mTOR inhibition, which induced autophagic death of OC cells [68]. Based on these facts, blocking the glycolytic pathway and targeting cancer cell metabolism using specific PKM2 inhibitors may be a promising strategy for the treatment of OC.

Regulatory mechanisms of aerobic glycolysis in OC

AMP-activated protein kinase (AMPK)

The AMPK is a highly conserved Ser/Thr kinase, which is composed of catalytic α subunits and regulatory β and γ subunits, and serves as a well-known energy status sensor and regulator of energy homeostasis, encompassing glucose, protein and fatty acid metabolism, as well as autophagy [69, 70]. During energetic stress, AMPK inhibits ATP consuming processes, such as fatty acid and protein synthesis, and cell proliferation, while promoting ATP preservation processes, including autophagy, glycolysis, TCA cycle and fatty acid oxidation [71]. AMPK is commonly activated during energetic stress by upstream kinases, including liver kinase B1 (LKB1) and Ca2+/calmodulin-dependent protein kinase kinase (CaMMKβ), which mediate the phosphorylation of threonine 172 (Thr172) in the catalytic α subunit of AMPK [72].

A growing number of studies have proved that AMPK plays an essential role in the proliferation, metastasis and invasion, and stress response in OC. Firstly, it has been found that transient receptor potential 7 (TRPM7) silencing activates the AMPK signaling pathway and promotes the ubiquitination degradation of HIF-1α, thereby inhibiting OC cell glycolysis and proliferation [73]; whereas inhibited AMPK activity is positively correlated with the increased aggressiveness of OC cells, thereby leading to a deteriorated prognosis for patients [74]. Secondly, it has also been suggested that AMPK may act as a tumor promoter related to the LKB1-AMPK pathway. For example, this pathway stabilizes the formation of epithelial OC cell spheroids, which promoted the development of metastasis and peritoneal lesions [75]. And under energy stress, lipolysis-stimulated lipoprotein receptor (LSR) also increases the phosphorylation levels of AMPKα and acetyl-CoA carboxylase (ACC) through this pathway, thereby promoting cancer cell survival and tumor progression [76]. Thirdly, AMPK also connects metabolic and signaling pathways, and its activation under high-energy stress maintains the link between glucose metabolism and cancer progression through GLUT3-mediated glucose energy homeostasis and regulation of the Hippo pathway [77].

In conclusion, AMPK represents a promising target for OC treatment. It has been discovered that natural AMPK activators, such as bitter melon extract, resveratrol, cordycepin, honokiol, or the novel AMPK-activating compound NT1014 may present new treatment options for OC by either inhibiting cell proliferation or promoting autophagy-induced cell death [78,79,80,81,82].

PI3K/AKT pathway

The phosphoinositide 3-kinases (PI3Ks) are lipid kinases that convert extracellular stimuli into intracellular signals through the production of phosphatidylinositol-3,4,5-trisphosphate (PIP3). Based on their different structures and specific substrates, PI3Ks are divided into three classes: class I (including class Ia and Ib), class II and class III [83]. Akt is a key serine/threonine kinase involved in the PI3K signaling pathway [84]. The PI3K/AKT signaling pathway broadly regulates normal cellular processes, but it is altered in cancer cells, promoting their proliferation, growth, motility, metabolism, angiogenesis, and inflammatory responses, and affecting the tumor microenvironment [85]. Thus, in OC, cell viability and EMT were similarly impaired after inhibition of the PI3K/AKT pathway by miR-22 [86]. PI3K/Akt pathway activation also contributes to cisplatin resistance in OC cells [87]. Using LY294002, a PI3K/Akt signaling pathway inhibitor could reverse cisplatin resistance in OC cells [88]. In addition, PTEN, as an oncogene, could induce dephosphorylation of PIP3, prevent Akt activation, and play a negative regulatory role in the PI3K/Akt signaling pathway, thus inhibiting tumor growth [89]. Compared with normal ovarian tissues, PTEN expression was absent in ovarian malignant tumor tissues, and PTEN expression in tissues was positively correlated with survival time and prognosis of patients [90].

The PI3K/AKT signaling pathway regulates glycolysis of OC through various mechanisms. Firstly, the PI3K/Akt signaling pathway induces the expression of GLUT1 and GLUT3, and mediates the translocation of GLUT4 from the cytoplasm to the cell membrane to enhance glucose uptake and transport [91]. Secondly, PI3K/Akt regulates the activity and expression of some glycolytic enzymes: (1) The oncogene ACTL6A accelerates OC glycolysis by up-regulating the PI3K/Akt pathway via FSH, which in turn regulates PGK1 to promote glucose utilization, lactate production, and pyruvate levels [24]. (2) The expression of HK2 can be up-regulated by the PI3K/AKT pathway, and the oncogene ARHGAP10 can inhibit OC cell glycolysis through this pathway [39]. (3) PKM2 can be regulated by epidermal growth factor (EGF) through the PI3K/AKT2 pathway, and the PI3K inhibitor LY294002 can block this effect [65]. Thirdly, PI3K/AKT indirectly affects glycolytic flux by interacting to regulate AMPK and HIF expression. For example, the PI3K/Akt pathway and the AMPK pathway can be synergistically affected by the enhanced p21 expression after creatine kinase B knockdown, resulting in glycolysis inhibition and OC apoptosis [92]. Additionally, knockdown of HIF-1α inhibited the PI3K/Akt/mTOR pathway and promoted autophagy in OC cells [93]. Finally, PI3K/AKT activates downstream regulators such as mTOR and VEGF to promote glycolysis and angiogenesis in OC. Wang et al. [94] found that FBN1 mediated the phosphorylation of vascular endothelial growth factor receptor 2 (VEGFR2) and activated the AKT pathway, which in turn induced the phosphorylation and nuclear translocation of STAT2, promoting glycolysis and angiogenesis. In addition, integrin alpha x (ITGAX) upregulates VEGFR2/VEGF-A expression through the PI3K/Akt pathway, which further promotes angiogenesis and OC progression [95].

Noncoding RNAs

Non-coding RNAs (ncRNAs) account for more than 90% of the human genome, which cannot be transcribed into proteins but are involved as functional RNAs in a variety of biological processes including development, proliferation, post-transcriptional modification, apoptosis, and cellular metabolism. ncRNAs can be mainly classified into three categories: long noncoding RNA (lncRNA), microRNA (miRNA), and circular RNA (circRNA), which can influence the “Warburg effect” by regulating the expression of glycolytic enzymes or glycolysis-related pathways [96, 97].

The three rate-limiting enzymes in aerobic glycolysis can all be regulated by various noncoding RNAs (Table 1). For example, the miR-654-3p was found to be inhibited by circular RNA pyridoxal kinase (circPDXK), which was highly expressed in tumor tissues and cells of OC patients, leading to upregulation of HK2 expression [98]. The miR-200 positively regulates HK2, which was associated with the fact that WTAP (WT1-associated protein) expression could be upregulated by HIF-1α and WTAP was able to interact with DGCR8 [99]. In addition, the DNMT3A/miR-603/HK2 axis was also involved in the “Warburg effect”, and the miR-603 expression was inhibited by DNMT3A, which promoted cell proliferation, migration, and invasion in OC [100]. Another key enzyme for glycolysis, PKM2, was found to be directly targeted by miR-324-5p, which negatively regulated glycolysis and lactate production in OC cells [101]. LINC00504 was up-regulated in OC tissues and was reported to inhibit miR-1244 to enhance the expression of glycolytic enzymes, including HK2, PKM2, and PDK1 [102]. By binding to PFKFB2, LINC00092 altered glycolysis and maintained local supportive functions of CAFs to promote metastasis of OC [55]. While the miR-206 directly targeted to regulate PFKFB3 expression and influenced glycolysis of OC [103]. In addition, lncRNA ceruloplasmin (NRCP) could function as an intermediate binding chaperone between STAT1 and RNA polymerase II, leading to increased expression of glucose-6-phosphate isomerase (G6PI) [104].

Table 1 Noncoding RNAs and their targets in aerobic glycolysis in OC

Other key proteins such as GLUT1 and LDHA are also regulated by ncRNAs in the OC, leading to changes in the levels of glucose uptake and lactate production. For example, the expression level of miR-1204 was positively correlated with that of GLUT1 in OC patients, and the overexpression of miR-1204 significantly promoted GLUT1 expression, enhanced glucose uptake and cell proliferation [105]. In addition, lncRNA HOXB-AS3 was abundantly expressed in EOC tissues, which was tightly associated with unfavorable overall survival status of the patients, whereas knockdown of HOXB-AS3 reduced LDHA expression via sponge miR-378a-3p, affecting the rate of acidification in the TME [106].

Non-coding RNAs can also regulate glycolysis levels in OC by targeting various signaling pathways involved in glycolysis. For example, downregulation of miR-3652 expression enabled overexpression of the CRTC2 gene, which functioned in the SIK/TORC pathway and was involved in the regulation of glucose production through the LKB1/AMPK/TORC2 pathway [107]. The miR-1180 activated Wnt signaling by targeting SFRP1 in OC cells to accelerate their proliferation and glycolysis [108]. The c-MYC is the target of LINC00629 and lncRNA CTSLP8 [109, 110], and the Hippo signaling pathway is the target of lncRNA LINC00857 [111], and PI3K/AKT is the target of circRHOBTB3 and miR-29b [112, 113]. Through directly binding to miR-375, LINC00662 was able to act on HIF-1α to promote glycolysis of OC and reduce apoptosis [114].

The role of glycolysis in peritoneal metastasis in OC

Altered glycolysis in cancer cells and related cells in the metastatic environment, such as mesothelial cells and cancer-associated fibroblasts(CAFs), can promote the process of peritoneal metastasis in OC.

Overexpressed in OC, especially in ascites and metastatic foci, HK2 can control lactate production and promote the process of OC metastasis via FAK/ERK/1/2 signaling pathway-mediated expression of MMP9/NANOG/SOX9 [115]. Knockdown of the HK2 gene in OC cells and ascites-derived tumor cells with the use of the glycolysis inhibitor 2-DG impeded lactate production, cell migration and invasion, and cell stemness properties, while decreasing FAK/ERK1/2 activation and metastasis and stemness-related genes [116]. Increased levels of fatty acid synthase correlate with high expression of P-cadherin, supporting enhanced ab initio lipogenesis in the metastatic microenvironment. P-calmodulin may enhance the activity of mesothelial cell glycolytic enzymes HK2, PGK1 through signaling cascades involving HIF or AMPK [117], thereby promoting glycolysis under metabolic stress or hypoxic conditions. Enriched lactate produced by the glycolytic process provides substrates for lipid synthesis and promotes ab initio synthesis of lipids in highly metastatic OC. Inhibition of lactate uptake by targeted inhibition of MCT1 could eliminate metabolic crosstalk and mesothelial metastasis [118]. Furthermore, PFKFB2 expression was significantly higher in metastatic OC tissues than in normal ovarian tissues and OC tissues without metastasis. It has been mentioned previously that binding of PFKFB2 to LINC00092 maintains the glycolytic phenotype of OC cells and enhances tumor aggressiveness through interactions with CAFs that highly express CXCL14 [55]. Schab et al. [119] found that DDR2 is highly expressed in OC stromal cells, which inhibits fructose 1,6-bisphosphatase and increases the activity of the key enzyme HK by targeting the AKT/SNAI1 pathway and that DDR2 increases the secretion of the collagen cross-linking agent LOXL2, which promotes the secretion of extracellular matrix proteins, and enhances the ability of cancer cells to spread in the peritoneal cavity.

Glycolysis mediates immunosuppression

Lactate produced by glycolysis in cancer cells is one of the important metabolites in the tumor microenvironment, which is closely related to the immune escape of tumor cells. It can regulate the metabolism of immune cells, inhibit the activation and proliferation of immune cells such as T cells, macrophages, natural killer cells (NK cells) and dendritic cells (DCs), and regulate the immune response of tumor cells as a signal molecule, which plays an essential role in immune surveillance and escape [120, 121]. This new research direction has a variety of potential clinical applications. Therefore, here we will focus on the research progress of the regulation of lactate on T cells, NK cells, Treg cells, tumor-associated macrophages (TAMs), which promotes immune escape of cancer cells (Fig. 3).

Fig. 3
figure 3

Impact of lactate accumulation on immune cells within the TME. Tumor cells produce excess lactate through aerobic glycolysis, causing acidosis and immunosuppression. Tumor-derived lactate regulates immune cell metabolism, inhibits the activation and proliferation of immune cells such as T cells, macrophages, and NK cells, and acts as a signaling molecule to regulate the immune response of tumor cells. In contrast, Treg cells and M2 phenotype macrophages are better able to maintain their immunosuppressive function in the acidic TME caused by lactate

T cells

T cell-mediated immune responses play an essential role in inhibiting tumor cell growth. T cells involved in tumor immunity include MHC class I antigen-restricted CD8+ cytotoxic T lymphocyte (CTL), MHC class II antigen-restricted CD4+ helper T cell (TH), tumor-promoting Tregs and etc. T cells are capable of sensing extracellular lactate levels, which elicits intracellular adjustments in signaling, cellular function, and homeostasis of the internal environment. Acidosis caused by excess extracellular lactate inhibits T cell-mediated immune responses, and tumor-specific CD8+ T cells are usually induced to an unresponsive state with reduced cytolytic activity and cytokine production at extracellular environmental pH values of 6.0 to 6.5 [122, 123]. In addition, lactate affects the metabolic aspects of T cells, which can inhibit the proliferation and cytokine production of human CTL with an efficiency of up to 95% and leading to a 50% reduction in cytotoxicity, which is attributed to the fact that high extracellular concentrations of lactate prevent the excretion of lactate from the cytoplasm of CTL, disrupting the metabolism of CTL [124]. In a high lactate environment, the lactate concentration difference formed inside and outside the cell prompts extracellular lactate to enter the cell by diffusion through MCT1 or other transporters on T cells as well as transporter-independent diffusion, leading to an intracellular low pH environment that prevents the upregulation of nuclear factor of activated T cells (NFAT) in T and NK cells, which leads to the reduction of interferon-γ (IFN-γ) and affects the activation of T cells and the participation in the process of tumor immunosurveillance. At the same time, intracellular acidification reduces the production of granzyme B by T cells and NK cells, which weakens the cytotoxicity and fails to target the apoptosis of tumor cells [125]. In addition, tumor cell-derived lactate impairs the function of natural killer T cells by inhibiting the mTOR signaling pathway and nuclear translocation of zinc finger proteins, resulting in reduced secretion of IFN-γ and IL-4 and significantly weakened anti-tumor effect [126]. The increased extracellular acidification rate (ECAR) of glycolytic phenotype OC stem cells attenuates T cell responses to TCR stimulation and reprograms T cell metabolism to inhibit T cell activation, proliferation, and action, as well as suppressing T cell responses by inhibiting ERK-AKT-mTOR signaling and activating the integrated stress response pathway [127]. In addition to its effects on common types of T cells, lactate promotes loss of the FAK family-interacting protein FIP200 in OC, driving apoptosis of naïve T cells and decreases cholesterol synthesis and IFN-γ release from invariant natural killer T cells (iNKT cells), which ultimately affects tumor immunity [128, 129].

Understanding the mechanisms by which the tumor microenvironment regulates T cells is of critical importance for the use of immunotherapy against cancer. However, there are still many unknowns in this complex microenvironment that are waiting to be explored.

NK cells

NK cells, as an effective component of tumor immune surveillance and control, have a favorable prognostic impact on OC [130]. NK cells, the main effectors of the innate immune system, produce antibody-dependent cytotoxicity as well as granzymes and perforins on target cells lacking MHC class I molecules when activated, as well as apoptotic signaling of target cells through the Fas/FasL and TRAIL/TRAILR pathways and secretion of a variety of cytokines including TNF-α and IFN-γ, which kill OC cells directly or indirectly participate in the immune response against OC development [131, 132]. NK cells can recognize stress-associated antigens through the NKG2D receptor, and this recognition is MHC-independent. MICA is a functional ligand that stimulates the NKG2D receptor, and in addition to MICA, the interaction of MICB and ULBP, which are part of the human NKG2D ligand, with the NKG2D receptor is important for cancer cell recognition and NK cell-mediated cytotoxicity [133]. Studies have shown that MICA/B and ULBP2 are expressed in most OC cells, but not in normal ovarian epithelium. Human OC over-expresses a variety of matrix metalloproteinases, such as MMP2, MMP9 and MMP14. MICA/B and ULBP2 on the membrane of OC cells can be cleaved by matrix metalloproteinases to produce soluble molecules, which systematically downregulate the expression of NKG2D receptor or directly block the function of NKG2D, thus impairing the immune surveillance of NK cells, resulting in the immune escape of OC cells [134,135,136]. The anti-stress natural product Ashwagandha is able to block tumor-induced NK cell inhibition by down-regulating the expression of GRP78 and the protein hydrolyzing agent ADAM10 on the surface of OC cells, thereby reducing MICA/B shedding [137]. In addition, OC-associated miR-20a was found to bind directly to the 3’-untranslated region of MICA/B mRNA, leading to the degradation of MICA/B and a decrease in the level of MICA/B proteins on the OC cell membrane, which prevented NK cells from exerting cytotoxicity against OC [138].

Lactate, as an important participant in suppressing tumor immunity, also has a significant impact on NK cells. The presence of lactate in the TME leads to polarization of the immunosuppressive phenotype of dendritic cells, thereby impairing the cytotoxicity of T and NK cells [139]. The key enzyme of lactate production, LDH, can spontaneously release its activity, a feature that is related to the dysfunction of NK cells. The acidic environment caused by lactate in tumor tissues can significantly inhibit the activity of NK cells and significantly reduce the release of cytotoxic molecules, such as perforin and granzymes, resulting in a reduction of their killing effect on tumor cells, and tumor-derived lactate can induce apoptosis of NK cells and reduce their infiltration in vivo [125, 139]. Large amounts of lactate can lower the pH in the tumor microenvironment, which in turn prevents NK cells from expelling cytoplasmic lactate down the lactate concentration gradient, leading to impaired metabolism, mitochondrial stress and apoptosis in NK cells [140]. Neutralization of the acidic environment in tumor tissues or inhibition of MCT4 in tumor cells to inhibit the release of lactate into the extracellular environment can restore the activity of NK cells in the TME to a certain extent [141]. In conclusion, it is possible that the inhibition of NK cell activity by lactate can be reversed by relevant treatment.

Treg cells

Treg cells can inhibit the abnormal over-immunization of immune cells against antigens, thus maintaining the immune homeostasis of the body, and they generally inhibit anti-tumor immunity and play a pro-tumorigenic role. For example, in vitro-induced CD8+ Treg cells inhibit the proliferation of naïve CD4+ T cells through TGF-β1 and IFN-γ mediation [142]. Alternatively, CD4+ T cells can be promoted for conversion to Treg cells by immunosuppressive regulatory B cells that produce IL-10 and express high levels of CD80 and CD86, thereby inhibiting CD4+ and CD8+ T cell proliferation [143].

Unlike other immune cells that show a negative response to lactate, Treg cells are able to increase their infiltration within tumors in an environment overwhelmed with lactate and upregulate the Treg cells activation marker CD25, and lactate released by tumor cells in the TME impairs cytotoxic T cell function and causes apoptosis, but Treg cells are highly resistant to this, and lactate can be used as an energy source for their uptake and utilization, as well as supporting more active metabolism of Treg cells [144, 145]. In the TME of low glucose and high lactate, forkhead frame protein 3 (FOXP3) in Treg cells, as a metabolic regulatory hub, has the ability to down-regulate the level of cellular glycolysis and induce OXPHOS by modulating MYC gene expression, thus increasing the ratio of NAD+ to NADH in the cells. High concentrations of lactate damage effector T cells through LDH-mediated NADH depletion, but Treg cells are highly resistant to it, which in turn promotes tumor development [146]. Compared to patients with benign ovarian tumors and healthy controls, OC patients had an increased CD8+ Treg cell subset with increased expression of CD25, CTLA-4, and Foxp3 and decreased expression of CD28, whereas CD8+ Treg cells induced in vitro by co-culture with OC cells similarly showed increased expression of CTLA-4 and Foxp3 and decreased expression of CD28 [142]. In addition, the large amount of lactate taken up by Treg cells can also be used for their own gene expression regulation and epigenetic modification. Lactate-modified MOESIN is able to promote the activation of the TGF-β-mediated SMAD3 signaling pathway and the transcription of the FOXP3 gene, which can enhance the function of Treg cells and drive the tumorigenesis [147]. Around highly glycolyzed tumor cells, Treg cells actively take up lactate via MCT1, which promotes NFAT1 into the nucleus, thus enhancing PD-1 expression, while PD-1 expression of effector T cells is inhibited, resulting in Treg cells expressing more PD-1 than effector T cells, leading to a decreased anti-PD-1 therapeutic effect [148]. In addition, mTOR and its downstream signaling pathway were significantly activated in CD4+ Treg of OC patients. mTOR signaling pathway could affect the glucose metabolism process of CD4+ Treg by regulating the levels of CD4+ Treg’s glucose uptake and glycolysis, and modulating the expression levels of glucose metabolism-related genes and proteins. Inhibition of mTOR signaling in CD4+ Tregs in OC cells growth environment resulted in a significant decrease in the expression levels of glucose metabolism-related genes and proteins, as well as a significant decrease in the levels of glucose uptake and glycolysis, and the simultaneous inhibition of mTOR signaling and activation of TLR8 signaling had a synergistic inhibitory effect on glucose metabolism and the immune-suppressing function of CD4+ Tregs. Activation of TLR8 signaling inhibited glucose metabolism in CD4+ Tregs by down-regulating mTOR signaling, thereby reversing the immunosuppressive function of these cells in the OC cells growth environment [149,150,151]. Gao et al. [152] found that miR-124 reduces lactate uptake by directly targeting MCT1, which ultimately impairs the immunosuppressive function of Treg cells, thereby slowing down the growth of OC cells and increasing their response to PD-1 blocking therapies. Wang et al. [153] demonstrated that oridonin could inhibit the metastasis of OC cells by inhibiting the mTOR signaling pathway and upregulating the level of FOXP3. Therefore, it is evident that targeting inhibition of MCT1, mTOR signaling and activation of TLR8 signaling in tumor tissues enriched with Treg cells may contribute to the improvement of immunotherapy efficacy of OC.

TAMs

Macrophages are one of the most infiltrated immune cell populations within the TME in OC and are mainly differentiated from monocytes, including M1 phenotype and M2 phenotype. Macrophages themselves have strong plasticity, and M1 phenotype macrophages produce inflammatory and pro-Th1 cytokines and low concentrations of IL-10, which have tumor-killing effects, while M2 phenotype polarized macrophages secrete a variety of anti-inflammatory factors and growth factors, which have immune-suppressive functions and play a key role in tumor development [154, 155]. Compared with specimens from benign ovarian lesions, patients with OC had increased densities of TAMs, more pronounced infiltration, and decreased M1/M2 ratios in their cancer tissues, thus exerting a role in promoting tumor growth, invasion, and metastasis, and suggesting a poor prognosis [156, 157]. Unlike other immune cells, the interaction between lactate and TAMs is reciprocal. On the one hand, there is evidence that the large amount of lactate in the TME significantly affects the polarization phenotype of macrophages, and that tumor cell-derived lactate induces polarization toward the M2 phenotype by stabilizing HIF-1α and promoting angiogenesis-induced production of VEGF, which phenotype demonstrates greater survival and adaptive capacity than M1 phenotype in the acidic environment of tumor tissues [158, 159]. Besides, OC-derived lactic acidosis promotes differentiation of monocytes into M2 phenotype macrophages with high levels of CD14 and CD163 markers and IL-1β, and certain OC cell lines generate IL-6 and M-CSF in addition to lactate to induce M2 phenotype macrophage polarization [160]. It was found that mTORC-related signaling pathway activation is essential in lactate-mediated M2 phenotype polarization of macrophages. Lactate activates the mTORC1 signaling pathway in TAMs, which in turn downregulates the expression of the lysosomal vesicle hydrolase subunit ATP6V0d2 through transcriptional regulation, and macrophage intracellular HIF-2α protein stability is increased as a result of lysosomal functional limitation, which ultimately maintains its own tumor-promoting M2 phenotype and supports tumor progression [161]. Chen et al. [162] found that the phosphorylation of mTOR, signal transduction and STAT3 could be inhibited by cardamonin, which exerted the effect of inhibiting M2 phenotype macrophage polarization and down-regulating the secretion of oncogenic factors by TAMs, and thus impeded the tumor-promoting function of TAMs on OC cells. The TAM surface receptor Gpr132 functions in response to lactate stimulation and induces its own M2 phenotype polarization, whereas the tumor-associated macrophage surface olfactory receptor Olfr78 has a similar effect. Olfr78 synergistically mediates the polarization of M2 phenotype macrophages with Gpr132, and Olfr78 functionally deficient mice exhibit the characteristics of tumor growth suppression, metastasis inhibition, and increased anti-tumor immune cell populations [163]. On the other hand, TAMs also affect tumor cell glycolytic metabolism as well as promote tumor progression. Colegio et al. [158] found that tumor-derived lactate was able to induce M2 phenotype polarization of macrophages via HIF-1α as well as arginase 1(Arg 1), while polarized macrophages could increase the utilization of lactate. In addition, lactate produced by tumors stimulates the production of IL-23 in TAM, leading to tumor growth by inducing the production of IL-17 and IL-22 [154]. Overall, the interactions between tumor cells and macrophages in the TME play a crucial role in the metabolic changes of macrophages, and lactate in the TME promotes their metabolic reprogramming, prompting macrophages to undergo an M2 phenotypic switch, in which the M1 phenotype relies mainly on aerobic glycolysis and fatty acid synthesis, whereas the M2 phenotype macrophage relies more on OXPHOS and fatty acid oxidation, and this phenotypic transition profoundly affects the TAMs immune response to tumors [154]. Additionally, it has been observed that OC cells are able to secrete lactate to promote the synthesis and secretion of IL-1β by TAMs, and the secreted IL-1β activates the nuclear factor kappa-B (NF-κB) signaling pathway in cancer cells to induce the up-regulation of the expression of PD-L1 and CCL2. IL-1β not only leads to the inhibitory effect of cancer cells on CD8+ T cells directly, but also recruits other components of immunosuppressive cells, including M2 phenotype macrophages, into the microenvironment, generating a vicious cycle leading to tumor immune escape. Targeting IL-1β can reverse the immunosuppressive phenotype of the TME, and its synergistic effect with PD-1-targeted immune checkpoint therapy can be achieved [164].

Endothelial cells and dendritic cells

Endothelial cells (ECs) in the TME are an integral part of the vasculature and are critical in transporting substances, angiogenesis and tumor metastasis. ECs in tumor tissues show significant heterogeneity and plasticity, which express lower levels of adhesion molecules, leading to impaired barrier function and increased levels of inhibitory immune checkpoint molecules which contribute to immunosuppression [165]. Like OC cells, ECs in cancerous tissues are more dependent on glycolysis for energy production than normal ECs, exhibiting a hyper glycolytic phenotype, the mechanism of action of which may be related to increased expression of GLUT1 and the glycolytic metabolizing enzyme 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 3 (PFKFB3) [103]. Hypoxia and pro-inflammatory cytokines in the TME upregulate PFKFB3. It has been demonstrated that inhibition of PFKFB3 reduces glycolysis in ECs in the TME to prevent the proliferation of tumor cells. High doses of 3-PO, a small molecule inhibitor of PFKFB3, can inhibit proliferation of cancer cells and reduce the growth of primary tumors, whereas a low dose of 3-PO induced tumor vascular barrier tightening and maturation, reducing cancer cell intravasation and metastasis [166]. Besides, lactate released into the microenvironment by tumor cells demonstrates the function of activating vascular ECs in tumor tissues and promoting angiogenesis. The uptake of lactate by ECs via MCT1 into the intracellular compartment for conversion to pyruvate results in the accumulation of NADH, which activates NADH oxidase, thereby inducing an increase in ROS and inhibition of prolyl hydroxylase domain 2 (PHD2). The increase in ROS and the inhibition of PHD2 lead to an increase in the degradation of human nuclear factor kappa-B inhibitor alpha (IκBα), which promotes the activity of NF-κB and activates the pro-angiogenic signaling IL-8 pathway to achieve the goal of promoting blood vessel growth [167].

Monocyte-derived dendritic cells exhibit an essential role in differentiation and activation of T cells, and normal functioning DCs must be available in tissues to serve as antigen-presenting cells, which recognize and process tumor antigens, form antigen-MHC complexes that are expressed on the surface of the cells for T-lymphocytes to recognize, promoting the maturation of naive T-lymphocytes. However, lactate-mediated signaling blocks the normal differentiation of DCs and affects their activation and antigen presentation, resulting in a tendency towards an immunosuppressive phenotype, in which DCs have lower levels of MHC II and co-stimulatory molecules, produce lower levels of IL-12 and higher levels of IL-10, and reduce the production of immune-activating cytokines such as IFN-γ, TNF-α, IL-1β and IL-6, which prevent the presentation of cancer cell-specific antigens to immune cells, negatively regulate the immune response, and result in immunosuppressive effects [168, 169]. Additionally, lactate induces intracellular Ca2+ mobilization by binding to GPR81 on the surface of plasmacytoid dendritic cells (pDCs) or transported to the cytoplasm of pDCs via MCT1, which participates in calcineurin phosphatase (CALN) signaling and affects cellular metabolism required for pDCs activation, and attenuates release of IFN-α by pDCs, affecting anti-tumor immune responses, while elevated intracellular lactate increases tryptophan metabolism in pDCs, leading to overproduction of L-kynurenine, which contributes to the expansion of FoxP3+CD4+ regulatory T cells, the main immunosuppressive immune cell subset in the TME [170]. Moreover, pDCs can secrete high levels of IFN-α through TLR, but the functional activity of pDCs in the acidic TME was altered, and instead, they highly expressed immunosuppressive factors such as IL-10, TGF-β, and indoleamine 2,3-dioxygenase (IDO), which induced the expansion of Tregs. Meanwhile, IL-10 enhances the secretion of serine protease granzyme B (GrB) from pDCs and inhibits the expansion of CD4+ T cells [171, 172]. The various effects of lactic acid accumulation on DCs play a negative role in the first stage of tumor cellular immunity, i.e., the antigen presentation process, thus impeding tumor immunity. Therefore, immunotherapy aimed at enhancing the antigen presentation function of DCs, inducing or enhancing the effective anti-tumor immune response against tumor antigens would be a very effective anticancer means.

Targeting glycolysis combined with immunotherapy in OC

TME is increasingly recognized as a key factor in tumor progression, and although researches on the effects of TME on immunotherapy and the mechanisms of its regulation are still in progress, its presence of negative regulation of immunotherapy cannot be ignored. In tumor tissues, the metabolites present in TME cause inhibition of immune cell function, especially contributing to the immune deficiency of tumor-infiltrating T lymphocytes. Therefore, interventions targeting glucose metabolism of cells in the TME to influence the composition of components in the TME become attractive adjunctive therapies to synergistically improve the anti-tumor immune response induced by the immune checkpoint inhibitor (ICI). Targeted understanding the effects of different metabolites on other immune cell components could also create more selective solution strategies for cancer immunotherapy (Table 2).

Table 2 Compounds and their targets and experimental phases in OC therapy

Reducing glycolytic metabolism in cancer cells by inhibiting key enzymes of glycolysis or focusing on the use of the competitive glucose analog 2-DG or the GLUT1 inhibitor BAY-876 is effective in inhibiting cancer cell proliferation, which may be able to support the formation of long-term memory CD8+ T cells [173, 174]. It was found that the glycolysis gating enzyme PDK1 in OC inhibits IFN-γ secretion by over-activating the PD-1/PD-L1 pathway through the c-Jun-NH2-kinase (JNK)-c-Jun pathway, resulting in the inhibition of CD8+ T cell function, and the knockdown of PDK1 in OC cells was able to block its induced expression of PD-L1 and promote the activation and infiltration of CD8+ T cells into tumor tissues to inhibit tumor growth, and the combination of PDK inhibitor dichloroacetic acid (DCA) and PD-L1 antibody increased IFN-γ secretion in monocyte-infiltrated tumor islets, attenuated tumor immune escape, and significantly enhanced the anti-tumor effect of immune checkpoint blockade [175]. Besides, Ho et al. [176] found that overexpression of the gluconeogenesis enzyme PCK1 enhanced the antitumoral response of T cells in glucose-deficient TME by inhibiting sarco/ER Ca2+-ATPase (SERCA) activity, and the results of this study support that glycolytic intermediates play an important role in the proliferation and function of effector T cells. Further, it has been shown that immune checkpoints including PD-1, PD-L1, and CTLA-4 act by partially inhibiting metabolic reprogramming and glycolysis of immune cells while increasing lipolysis and fatty acid β-oxidation (FAO) [177]. These studies reveal a link between cellular metabolism and checkpoint blockade, and thus approaches combined with alterations of glucose metabolism in cancer cell or immune cell may provide new opportunities for immunotherapy and improve the efficacy of checkpoint inhibitors in OC.

The aerobic glycolysis in cancer cells makes lactate prevalent in the TME, which is detrimental to antitumoral immunotherapy. GAO et al. [178] constructed a hollow MnO2 (HMnO2)-catalyzed nanosystem (PMLR) loaded with lactate oxidase (LOX) and glycolysis inhibitor (3-PO) and encapsulated in red blood cell membranes (mRBC). Owing to the long-circulating properties of mRBC, PMLR gradually accumulates at the tumor site, and the extracellular PMLR catalyzes the oxidative reaction of lactate via LOX to consume the lactate in TME, while intracellular PMLR releases glycolysis inhibitors to block the lactate source. Moreover, PMLR decreased the proportion of M2 phenotype macrophages in combination with PD-L1 therapy, and immune-activating cytokines, IFN-γ and IL-6, were detected, and the survival of experimental mice was prolonged compared with PD-L1 therapy alone, suggesting that lactate depletion can improve the efficacy of PD-L1 therapy. In addition, inhibition of MCT function to disrupt lactate transport between cells in tumor tissues also plays a role in suppressing tumor development. For lactate produced by high flux of glycolysis in cancer cells, an acidic environment-activatable nanomedicine, AZD3965, was able to reverse lactate-induced tumor immunosuppression by targeting on inhibiting MCT1 expression. Moreover, when oral AZD3965 is combined with anti-PD-1 therapy, increased T cell infiltration and decreased depletion of PD-1+ Tim3+ T lymphocyte effectively inhibits tumor growth and improves survival [179]. Additionally, given the important role of the mTOR pathway in lactate metabolism, combination therapy with mTOR inhibitors and lactate metabolism or transport inhibitors is considered to enhance anti-tumor immunity. Rapalogs (temsirolimus, everolimus), ATP-competitive mTOR inhibitors (MLN0128, PP242, AZD2014, AZD8055) and dual PI3K/mTOR inhibitors (NVP- BEZ235, LY3023414, etc.) have been applied to treat a variety of cancers, and the mTORC1/2 inhibitor onatasertib (ATG-008) demonstrates synergistic anti-tumor activity in combination with PD-1 antibody [180, 181].

To date, fewer studies have been conducted on the combined use of glycolytic or lactate inhibitors and immunotherapy aimed at OC, and further studies are expected to clarify the role of cancer-specific glycolytic inhibitors in combination with immunotherapy in the creation of new effective therapeutic regimens in OC.

Conclusion

A defining characteristic of OC is the alteration in metabolic processes, specifically aerobic glycolysis. This phenomenon provides valuable insights into the development of OC and the identification of biomarkers targeting metabolic aspects. The focus of this review is to examine how abnormal lactate concentrations impact tumor immune evasion through various pathways involving immune cell differentiation, metabolism, and function. By uncovering targets associated with aerobic glycolysis and lactate metabolism in regulating tumor immunity, new opportunities for immunotherapy arise. A comprehensive understanding of how metabolites produced during aerobic glycolysis influence immune cell function has the potential to greatly enhance the effectiveness of immunotherapy treatments. Although there are currently challenges hindering the clinical application of combining immunotherapy with inhibitors targeting glycolysis or lactate metabolism, further extensive research in these areas is expected to facilitate their utilization in improved combination regimens for OC treatment.

Data availability

No datasets were generated or analysed during the current study.

References

  1. Ohshima K, Morii E. Metabolic reprogramming of Cancer cells during Tumor Progression and Metastasis. Metabolites. 2021;11(1):null.

    Article  Google Scholar 

  2. Alkhathami AG, Sahib AS, Al Fayi MS, Fadhil AA, Jawad MA, Shafik SA, Sultan SJ, Almulla AF, Shen M. Glycolysis in human cancers: Emphasis circRNA/glycolysis axis and nanoparticles in glycolysis regulation in cancer therapy. Environ res 2023, 234(null):116007.

  3. Chang X, Liu X, Wang H, Yang X, Gu Y. Glycolysis in the progression of pancreatic cancer. Am J Cancer Res. 2022;12(2):861–72.

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Bouillaud F, Hammad N, Schwartz L. Warburg Effect, glutamine, Succinate, Alanine, when oxygen matters. Biology (Basel). 2021;10(10):null.

    Google Scholar 

  5. Huang Q, Tan Y, Yin P, Ye G, Gao P, Lu X, Wang H, Xu G. Metabolic characterization of hepatocellular carcinoma using nontargeted tissue metabolomics. Cancer res. 2013;73(16):4992–5002.

    Article  CAS  PubMed  Google Scholar 

  6. Gao Y, Zhou H, Liu G, Wu J, Yuan Y, Shang A. Tumor Microenvironment: Lactic Acid Promotes Tumor Development. J immunol res 2022, 2022(null):3119375.

  7. Sehouli J, Grabowski JP. Surgery in recurrent ovarian cancer. Cancer-am cancer soc. 2019;125(Suppl 24):4598–601.

    Google Scholar 

  8. Maleki Z, Vali M, Nikbakht HA, Hassanipour S, Kouhi A, Sedighi S, Farokhi R, Ghaem H. Survival rate of ovarian cancer in Asian countries: a systematic review and meta-analysis. BMC Cancer. 2023;23(1):558.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Micek HM, Visetsouk MR, Fleszar AJ, Kreeger PK. The many microenvironments of ovarian Cancer. Adv exp med biol. 2020;1296(null):199–213.

    Article  CAS  PubMed  Google Scholar 

  10. Duarte Mendes A, Freitas AR, Vicente R, Vitorino M, Vaz Batista M, Silva M, Braga S. Adipocyte microenvironment in ovarian Cancer: a critical contributor? Int J Mol Sci. 2023;24(23):null.

    Article  Google Scholar 

  11. Lin Y, Zhou X, Ni Y, Zhao X, Liang X. Metabolic reprogramming of the tumor immune microenvironment in ovarian cancer: A novel orientation for immunotherapy. Front Immunol 2022, 13(null):1030831.

  12. Edwards S, Pyne E, Lui L, Guinan J, Frisard M, Schmelz E. Abstract 808: aggregation alters the metabolism of serous ovarian cancer cells. Cancer res. 2019;79(13Supple):808–808.

    Article  Google Scholar 

  13. Fong MY, McDunn J, Kakar SS. Identification of metabolites in the normal ovary and their transformation in primary and metastatic ovarian cancer. PLoS ONE. 2011;6(5):e19963.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Cheng Y, Li L, Zhu B, Liu F, Wang Y, Gu X, Yan C. Expanded metabolomics approach to profiling endogenous carbohydrates in the serum of ovarian cancer patients. J sep sci. 2016;39(2):316–23.

    Article  CAS  PubMed  Google Scholar 

  15. Vermeersch KA, Wang L, McDonald JF, Styczynski MP. Distinct metabolic responses of an ovarian cancer stem cell line. BMC Syst Biol. 2014;8(null):134.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Compton SLE, Grieco JP, Gollamudi B, Bae E, Van Mullekom JH, Schmelz EM. Metabolic reprogramming of Ovarian Cancer spheroids during Adhesion. Cancers (Basel). 2022;14(6):null.

    Article  Google Scholar 

  17. Loar P, Wahl H, Kshirsagar M, Gossner G, Griffith K, Liu JR. Inhibition of glycolysis enhances cisplatin-induced apoptosis in ovarian cancer cells. Am j Obstet Gynecol. 2010;202(4):e371371–378.

    Article  Google Scholar 

  18. Wu Z, Yuan C, Zhang Z, Wang M, Xu M, Chen Z, Tian J, Cao W, Wang Z. Paris saponins VII inhibits glycolysis of ovarian cancer via the RORC/ACK1 signaling pathway. Biochem Pharmacol. 2023;213null:115597.

    Article  Google Scholar 

  19. S SXX HHGG, RF LLJJ, XD XDZZ RFTT. M MWW: LncRNA IDH1-AS1 links the functions of c-Myc and HIF1α via IDH1 to regulate the Warburg effect. Proc Natl Acad Sci USA 2018(7):E1465–74.

  20. X XWW WWDD. X XLL, J JLL, D DLL, L LYY, L LQQ, A AYY, Q QXX, H HLL: PTTG regulates the metabolic switch of ovarian cancer cells via the c-myc pathway. Oncotarget 2015(38):40959–69.

  21. Han CY, Patten DA, Kim SI, Lim JJ, Chan DW, Siu MKY, Han Y, Carmona E, Parks RJ, Lee C, et al. Nuclear HKII-P-p53 (Ser15) Interaction is a prognostic biomarker for chemoresponsiveness and glycolytic regulation in epithelial ovarian Cancer. Cancers (Basel). 2021;13(14):null.

    Article  Google Scholar 

  22. Zhao Y, Zhang L, Wu Y, Dai Q, Zhou Y, Li Z, Yang L, Guo Q, Lu N. Selective anti-tumor activity of wogonin targeting the Warburg effect through stablizing p53. Pharmacol res. 2018;135(null):49–59.

    Article  CAS  PubMed  Google Scholar 

  23. Chiyoda T, Hart PC, Eckert MA, McGregor SM, Lastra RR, Hamamoto R, Nakamura Y, Yamada SD, Olopade OI, Lengyel E, et al. Loss of BRCA1 in the cells of origin of Ovarian Cancer induces glycolysis: a window of opportunity for Ovarian Cancer Chemoprevention. Cancer prev res. 2017;10(4):255–66.

    Article  CAS  Google Scholar 

  24. Zhang J, Zhang J, Wei Y, Li Q, Wang Q. ACTL6A regulates follicle-stimulating hormone-driven glycolysis in ovarian cancer cells via PGK1. Cell Death Dis. 2019;10(11):811.

    Article  PubMed  PubMed Central  Google Scholar 

  25. Kobayashi H. Recent advances in understanding the metabolic plasticity of ovarian cancer: a systematic review. Heliyon. 2022;8(11):e11487.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Corbet C, Feron O. Tumour acidosis: from the passenger to the driver’s seat. Nat rev cancer. 2017;17(10):577–93.

    Article  CAS  PubMed  Google Scholar 

  27. Battista MJ, Goetze K, Schmidt M, Cotarelo C, Weyer-Elberich V, Hasenburg A, Mueller-Klieser W, Walenta S. Feasibility of induced metabolic bioluminescence imaging in advanced ovarian cancer patients: first results of a pilot study. J cancer res clin. 2016;142(9):1909–16.

    Article  CAS  Google Scholar 

  28. Li X, Yang Y, Zhang B, Lin X, Fu X, An Y, Zou Y, Wang JX, Wang Z, Yu T. Lactate metabolism in human health and disease. Signal Transduct tar. 2022;7(1):305.

    Article  CAS  Google Scholar 

  29. CT CTHH, B BFF QQYY, N NL-CL-C EEJJ. J JKK, L LJJ, B BKK, R RSS, L LLL: Metabolic heterogeneity in human lung tumors. Cell 2016(4):681–94.

  30. Hatami H, Sajedi A, Mir SM, Memar MY. Importance of lactate dehydrogenase (LDH) and monocarboxylate transporters (MCTs) in cancer cells. Health Sci Rep. 2023;6(1):e996.

    Article  PubMed  Google Scholar 

  31. Chatterjee P, Bhowmik D, Roy SS. A systemic analysis of monocarboxylate transporters in ovarian cancer and possible therapeutic interventions. Channels. 2023;17(1):2273008.

    Article  PubMed  PubMed Central  Google Scholar 

  32. L LWW MMRR, M MD-VD-V, K KTT NNPP, U UM-OM-O. J JCC,: Metabolic coupling and the Reverse Warburg Effect in cancer: implications for novel biomarker and anticancer agent development. Semin Oncol 2017(3):198–203.

  33. Payen VL, Mina E, Van Hée VF, Porporato PE, Sonveaux P. Monocarboxylate transporters in cancer. Mol Metab. 2020;33(null):48–66.

    Article  CAS  PubMed  Google Scholar 

  34. Yamaguchi A, Mukai Y, Sakuma T, Narumi K, Furugen A, Yamada Y, Kobayashi M. Monocarboxylate transporter 4 involves in energy metabolism and drug sensitivity in hypoxia. Sci Rep. 2023;13(1):1501.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Doherty JR, Yang C, Scott KE, Cameron MD, Fallahi M, Li W, Hall MA, Amelio AL, Mishra JK, Li F, et al. Blocking lactate export by inhibiting the Myc target MCT1 disables glycolysis and glutathione synthesis. Cancer res. 2014;74(3):908–20.

    Article  CAS  PubMed  Google Scholar 

  36. Gao T, Lin YQ, Ye HY, Lin WM. miR-124 delivered by BM-MSCs-derived exosomes targets MCT1 of tumor-infiltrating Treg cells and improves ovarian cancer immunotherapy. Neoplasma 2023, null(null):null.

  37. Siu MKY, Jiang YX, Wang JJ, Leung THY, Han CY, Tsang BK, Cheung ANY, Ngan HYS, Chan KKL. Hexokinase 2 regulates ovarian Cancer Cell Migration, Invasion and Stemness via FAK/ERK1/2/MMP9/NANOG/SOX9 signaling cascades. Cancers (Basel). 2019;11(6):null.

    Article  Google Scholar 

  38. Zhang D, Yip YM, Li L. In silico construction of HK2-VDAC1 complex and investigating the HK2 binding-induced molecular gating mechanism of VDAC1. Mitochondrion. 2016;30(null):222–8.

    Article  CAS  PubMed  Google Scholar 

  39. Wu K, Gong W, Sun H, Li W, Chen L, Duan Y, Zhu J, Zhang H, Ke H. SMAD4 inhibits glycolysis in ovarian cancer through PI3K/AKT/HK2 signaling pathway by activating ARHGAP10. Cancer Rep (Hoboken) 2024, null(null):e1976.

  40. Lu J, Wang L, Chen W, Wang Y, Zhen S, Chen H, Cheng J, Zhou Y, Li X, Zhao L. miR-603 targeted hexokinase-2 to inhibit the malignancy of ovarian cancer cells. Arch Biochem Biophys. 2019;661(null):1–9.

    Article  PubMed  Google Scholar 

  41. Tuo X, Zhou Y, Yang X, Ma S, Liu D, Zhang X, Hou H, Wang R, Li X, Zhao L. miR-532-3p suppresses proliferation and invasion of ovarian cancer cells via GPNMB/HIF-1α/HK2 axis. Pathol res pract 2022, 237(null):154032.

  42. Zou J, Wang Y, Liu M, Huang X, Zheng W, Gao Q, Wang H. Euxanthone inhibits glycolysis and triggers mitochondria-mediated apoptosis by targeting hexokinase 2 in epithelial ovarian cancer. Cell Biochem Funct. 2018;36(6):303–11.

    Article  CAS  PubMed  Google Scholar 

  43. Yu S, Yan X, Tian R, Xu L, Zhao Y, Sun L, Su J. An experimentally Induced Mutation in the UBA Domain of p62 changes the sensitivity of cisplatin by Up-Regulating HK2 localisation on the Mitochondria and increasing Mitophagy in A2780 Ovarian Cancer cells. Int J Mol Sci. 2021;22(8):null.

    Article  Google Scholar 

  44. Zhang XY, Zhang M, Cong Q, Zhang MX, Zhang MY, Lu YY, Xu CJ. Hexokinase 2 confers resistance to cisplatin in ovarian cancer cells by enhancing cisplatin-induced autophagy. Int j Biochem cell b. 2018;95(null):9–16.

    Article  CAS  Google Scholar 

  45. Li Y, Kang J, Fu J, Luo H, Liu Y, Li Y, Sun L. PGC1α promotes Cisplatin Resistance in Ovarian Cancer by regulating the HSP70/HK2/VDAC1 signaling pathway. Int J Mol Sci. 2021;22(5):null.

    Article  Google Scholar 

  46. LV MMCC. LVAA: Hitting the Sweet Spot: how glucose metabolism is orchestrated in space and time by Phosphofructokinase-1. Cancers 2023(1).

  47. Fernandes PM, Kinkead J, McNae I, Michels PAM, Walkinshaw MD. Biochemical and transcript level differences between the three human phosphofructokinases show optimisation of each isoform for specific metabolic niches. Biochem j. 2020;477(22):4425–41.

    Article  CAS  PubMed  Google Scholar 

  48. Li L, Li L, Li W, Chen T, Bin Z, Zhao L, Wang H, Wang X, Xu L, Liu X, et al. TAp73-induced phosphofructokinase-1 transcription promotes the Warburg effect and enhances cell proliferation. Nat Commun. 2018;9(1):4683.

    Article  PubMed  PubMed Central  Google Scholar 

  49. Zancan P, Marinho-Carvalho MM, Faber-Barata J, Dellias JM, Sola-Penna M. ATP and fructose-2,6-bisphosphate regulate skeletal muscle 6-phosphofructo-1-kinase by altering its quaternary structure. IUBMB Life. 2008;60(8):526–33.

    Article  CAS  PubMed  Google Scholar 

  50. Lu Z. Abstract IA08: kinase-mediated modulation of paclitaxel sensitivity in ovarian cancer. Clin cancer res. 2018;24(15Supple):Ia08–08.

    Article  Google Scholar 

  51. Antoun S, Atallah D, Tahtouh R, Moubarak M, Alaeddine N, Hilal G. TP53 gene mutations differently regulate ovarian cancer metabolism: ex vivo and in vitro studies. Cancer Res. 2018;78(13Supplement):2413–2413.

    Article  Google Scholar 

  52. YX Y-XJJ, MKY MKYSS, JJ J-JWW, THY THYLL, DW DWCC, ANY ANYCC, HYS HYSNN, KKL KKLCC.: PFKFB3 regulates Chemoresistance, Metastasis and Stemness via IAP Proteins and the NF-κB signaling pathway in Ovarian Cancer. Front Oncol 2022:748403.

  53. Mondal S, Roy D, Kalogera E, Khurana A, Tapolsky GH, Telang S, Chesney J, Shridhar V. Inhibition of PFKFB3/glycolysis overcomes chemoresistance in ovarian cancers. Cancer Res. 2015;75(15Supplement):2706–2706.

    Article  Google Scholar 

  54. Wang J, Siu MK, Jiang Y, Leung TH, NGAN HY, Cheung AN, Chan KK. Abstract LB-269: PFKP/PFKFB3 modulate metabolic switch and chemoresistance to cisplatin in ovarian cancer. Cancer Research 2017, 77(13_Supplement):LB-269-LB-269.

  55. Zhao L, Ji G, Le X, Wang C, Xu L, Feng M, Zhang Y, Yang H, Xuan Y, Yang Y, et al. Long noncoding RNA LINC00092 acts in Cancer-Associated fibroblasts to Drive Glycolysis and Progression of Ovarian Cancer. Cancer res. 2017;77(6):1369–82.

    Article  CAS  PubMed  Google Scholar 

  56. Yang H, Shu Z, Jiang Y, Mao W, Pang L, Redwood A, Jeter-Jones SL, Jennings NB, Ornelas A, Zhou J, et al. 6-Phosphofructo-2-Kinase/Fructose-2,6-Biphosphatase-2 regulates TP53-Dependent paclitaxel sensitivity in ovarian and breast cancers. Clin cancer res. 2019;25(18):5702–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Taylor C, Mannion D, Miranda F, Karaminejadranjbar M, Herrero-Gonzalez S, Hellner K, Zheng Y, Bartholomeusz G, Bast RC, Ahmed AA. Loss of PFKFB4 induces cell death in mitotically arrested ovarian cancer cells. Oncotarget. 2017;8(11):17960–80.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Dong G, Mao Q, Xia W, Xu Y, Wang J, Xu L, Jiang F. PKM2 and cancer: the function of PKM2 beyond glycolysis. Oncol lett. 2016;11(3):1980–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Papadaki C, Manolakou S, Lagoudaki E, Pontikakis S, Ierodiakonou D, Vogiatzoglou K, Messaritakis I, Trypaki M, Giannikaki L, Sfakianaki M, et al. Correlation of PKM2 and CD44 protein expression with poor prognosis in platinum-treated epithelial ovarian Cancer: a retrospective study. Cancers (Basel). 2020;12(4):null.

    Article  Google Scholar 

  60. Wong N, Ojo D, Yan J, Tang D. PKM2 contributes to cancer metabolism. Cancer Lett. 2015;356(2 Pt A):184–91.

    Article  CAS  PubMed  Google Scholar 

  61. Hosios AM, Fiske BP, Gui DY, Vander Heiden MG. Lack of evidence for PKM2 protein kinase activity. Mol cell. 2015;59(5):850–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. XGG X, HWW H, JJYY JJ, XLL X, Z-RLL ZR. Pyruvate kinase M2 regulates gene transcription by acting as a protein kinase. Mol Cell. 2012;5:598–609.

    Google Scholar 

  63. KE KEKK, ZM ZMDD, ZW ZWDD, YS Y-SLL:. SAICAR induces protein kinase activity of PKM2 that is necessary for sustained proliferative signaling of cancer cells. Mol Cell 2014(5):700–9.

  64. TSS T, ZLL Z, FBB F, QYY Q. Deubiquitinase PSMD14 promotes ovarian cancer progression by decreasing enzymatic activity of PKM2. Mol Oncol. 2021;12:3639–58.

    Google Scholar 

  65. Zheng B, Geng L, Zeng L, Liu F, Huang Q. AKT2 contributes to increase ovarian cancer cell migration and invasion through the AKT2-PKM2-STAT3/NF-κB axis. Cell Signal. 2018;45(null):122–31.

    Article  CAS  PubMed  Google Scholar 

  66. Li S, Ji X, Wang R, Miao Y. Follicle-stimulating hormone promoted pyruvate kinase isozyme type M2-induced glycolysis and proliferation of ovarian cancer cells. Arch Gynecol Obstet. 2019;299(5):1443–51.

    Article  CAS  PubMed  Google Scholar 

  67. Zheng B, Liu F, Zeng L, Geng L, Ouyang X, Wang K, Huang Q. Overexpression of pyruvate kinase type M2 (PKM2) promotes ovarian Cancer Cell Growth and Survival Via Regulation of Cell Cycle Progression related with upregulated CCND1 and downregulated CDKN1A expression. Med Sci Monit. 2018;24(null):3103–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Park JH, Kundu A, Lee SH, Jiang C, Lee SH, Kim YS, Kyung SY, Park SH, Kim HS. Specific pyruvate kinase M2 inhibitor, compound 3K, induces autophagic cell death through disruption of the glycolysis pathway in Ovarian Cancer cells. Int J Biol Sci. 2021;17(8):1895–908.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Yuan W, Fang W, Zhang R, Lyu H, Xiao S, Guo D, Ali DW, Michalak M, Chen XZ, Zhou C, et al. Therapeutic strategies targeting AMPK-dependent autophagy in cancer cells. Bba-mol cell res. 2023;1870(7):119537.

    CAS  Google Scholar 

  70. Keerthana CK, Rayginia TP, Shifana SC, Anto NP, Kalimuthu K, Isakov N, Anto RJ. The role of AMPK in cancer metabolism and its impact on the immunomodulation of the tumor microenvironment. Front Immunol 2023, 14(null):1114582.

  71. Hsu CC, Peng D, Cai Z, Lin HK. AMPK signaling and its targeting in cancer progression and treatment. Semin cancer biol. 2022;85(null):52–68.

    Article  CAS  PubMed  Google Scholar 

  72. Monteverde T, Muthalagu N, Port J, Murphy DJ. Evidence of cancer-promoting roles for AMPK and related kinases. Febs j. 2015;282(24):4658–71.

    Article  CAS  PubMed  Google Scholar 

  73. Chen Y, Liu L, Xia L, Wu N, Wang Y, Li H, Chen X, Zhang X, Liu Z, Zhu M, et al. TRPM7 silencing modulates glucose metabolic reprogramming to inhibit the growth of ovarian cancer by enhancing AMPK activation to promote HIF-1α degradation. J Exp Clin Cancer Res. 2022;41(1):44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Li C, Liu VW, Chiu PM, Yao KM, Ngan HY, Chan DW. Reduced expression of AMPK-β1 during tumor progression enhances the oncogenic capacity of advanced ovarian cancer. Mol Cancer. 2014;13(null):49.

    Article  PubMed  PubMed Central  Google Scholar 

  75. Peart T, Ramos Valdes Y, Correa RJ, Fazio E, Bertrand M, McGee J, Préfontaine M, Sugimoto A, DiMattia GE, Shepherd TG. Intact LKB1 activity is required for survival of dormant ovarian cancer spheroids. Oncotarget. 2015;6(26):22424–38.

    Article  PubMed  PubMed Central  Google Scholar 

  76. Takahashi Y, Serada S, Ohkawara T, Fujimoto M, Hiramatsu K, Ueda Y, Kimura T, Takemori H, Naka T. LSR promotes epithelial ovarian cancer cell survival under energy stress through the LKB1-AMPK pathway. Biochem Bioph res co. 2021;537(null):93–9.

    Article  CAS  Google Scholar 

  77. Wang W, Xiao ZD, Li X, Aziz KE, Gan B, Johnson RL, Chen J. AMPK modulates Hippo pathway activity to regulate energy homeostasis. Nat cell biol. 2015;17(4):490–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Lee JS, Sul JY, Park JB, Lee MS, Cha EY, Ko YB. Honokiol induces apoptosis and suppresses migration and invasion of ovarian carcinoma cells via AMPK/mTOR signaling pathway. Int j mol med. 2019;43(5):1969–78.

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Liu Y, Tong L, Luo Y, Li X, Chen G, Wang Y. Resveratrol inhibits the proliferation and induces the apoptosis in ovarian cancer cells via inhibiting glycolysis and targeting AMPK/mTOR signaling pathway. J cell Biochem. 2018;119(7):6162–72.

    Article  CAS  PubMed  Google Scholar 

  80. Yoon SY, Lindroth AM, Kwon S, Park SJ, Park YJ. Adenosine derivatives from Cordyceps exert antitumor effects against ovarian cancer cells through ENT1-mediated transport, induction of AMPK signaling, and consequent autophagic cell death. Biomed pharmacother 2022, 153(null):113491.

  81. Yung MM, Ross FA, Hardie DG, Leung TH, Zhan J, Ngan HY, Chan DW. Bitter Melon (Momordica charantia) Extract inhibits tumorigenicity and overcomes cisplatin-resistance in ovarian Cancer cells through Targeting AMPK Signaling Cascade. Integr cancer ther. 2016;15(3):376–89.

    Article  CAS  PubMed  Google Scholar 

  82. Zhang L, Han J, Jackson AL, Clark LN, Kilgore J, Guo H, Livingston N, Batchelor K, Yin Y, Gilliam TP, et al. NT1014, a novel biguanide, inhibits ovarian cancer growth in vitro and in vivo. J Hematol Oncol. 2016;9(1):91.

    Article  PubMed  PubMed Central  Google Scholar 

  83. Yang J, Nie J, Ma X, Wei Y, Peng Y, Wei X. Targeting PI3K in cancer: mechanisms and advances in clinical trials. Mol Cancer. 2019;18(1):26.

    Article  PubMed  PubMed Central  Google Scholar 

  84. Revathidevi S, Munirajan AK. Akt in cancer: Mediator and more. Semin cancer biol. 2019;59(null):80–91.

    Article  CAS  PubMed  Google Scholar 

  85. Trigueiros B, Santos IJS, Pimenta FP, Ávila AR. A long way to go: a scenario for clinical trials of PI3K inhibitors in treating Cancer. Cancer Control. 2024;31(null):10732748241238047.

    Article  PubMed  PubMed Central  Google Scholar 

  86. 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  CAS  PubMed  Google Scholar 

  87. Wu D, Lu P, Mi X, Miao J. Downregulation of miR-503 contributes to the development of drug resistance in ovarian cancer by targeting PI3K p85. Arch Gynecol Obstet. 2018;297(3):699–707.

    Article  CAS  PubMed  Google Scholar 

  88. Jiang L, Liu Y, Liu M, Zheng Y, Chen L, Shan F, Ji J, Cao Y, Kai H, Kang X. REG3A promotes proliferation and DDP resistance of ovarian cancer cells by activating the PI3K/Akt signaling pathway. Environ Toxicol. 2024;39(1):85–96.

    Article  CAS  PubMed  Google Scholar 

  89. Cai J, Xu L, Tang H, Yang Q, Yi X, Fang Y, Zhu Y, Wang Z. The role of the PTEN/PI3K/Akt pathway on prognosis in epithelial ovarian cancer: a meta-analysis. Oncologist. 2014;19(5):528–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Martins FC, Couturier DL, Paterson A, Karnezis AN, Chow C, Nazeran TM, Odunsi A, Gentry-Maharaj A, Vrvilo A, Hein A, et al. Clinical and pathological associations of PTEN expression in ovarian cancer: a multicentre study from the Ovarian Tumour Tissue Analysis Consortium. Brit j cancer. 2020;123(5):793–802.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Barron CC, Bilan PJ, Tsakiridis T, Tsiani E. Facilitative glucose transporters: implications for cancer detection, prognosis and treatment. Metabolism. 2016;65(2):124–39.

    Article  CAS  PubMed  Google Scholar 

  92. Li XH, Chen XJ, Ou WB, Zhang Q, Lv ZR, Zhan Y, Ma L, Huang T, Yan YB, Zhou HM. Knockdown of creatine kinase B inhibits ovarian cancer progression by decreasing glycolysis. Int j Biochem cell b. 2013;45(5):979–86.

    Article  Google Scholar 

  93. Huang J, Gao L, Li B, Liu C, Hong S, Min J, Hong L. Knockdown of Hypoxia-Inducible factor 1α (HIF-1α) promotes Autophagy and inhibits phosphatidylinositol 3-Kinase (PI3K)/AKT/Mammalian target of Rapamycin (mTOR) signaling pathway in Ovarian Cancer cells. Med Sci Monit. 2019;25(null):4250–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Wang Z, Chen W, Zuo L, Xu M, Wu Y, Huang J, Zhang X, Li Y, Wang J, Chen J, et al. The Fibrillin-1/VEGFR2/STAT2 signaling axis promotes chemoresistance via modulating glycolysis and angiogenesis in ovarian cancer organoids and cells. Cancer Commun (Lond). 2022;42(3):245–65.

    Article  CAS  PubMed  Google Scholar 

  95. Wang J, Yang L, Liang F, Chen Y, Yang G. Integrin alpha x stimulates cancer angiogenesis through PI3K/Akt signaling-mediated VEGFR2/VEGF-A overexpression in blood vessel endothelial cells. J cell Biochem. 2019;120(2):1807–18.

    Article  CAS  PubMed  Google Scholar 

  96. Peng B, Li J, Yan Y, Liu Y, Liang Q, Liu W, Thakur A, Zhang K, Xu Z, Wang J et al. Non-coding RNAs: The recently accentuated molecules in the regulation of cell autophagy for ovarian cancer pathogenesis and therapeutic response. Front Pharmacol 2023, 14(null):1162045.

  97. Shankaraiah RC, Veronese A, Sabbioni S, Negrini M. Non-coding RNAs in the reprogramming of glucose metabolism in cancer. Cancer Lett. 2018;419(null):167–74.

    Article  CAS  PubMed  Google Scholar 

  98. Hou L, Wang W, Zhai J, Zhao H. Circular RNA pyridoxal kinase (circPDXK) involves in the progression of ovarian cancer and glycolysis via regulating mir-654-3p and hexokinase II. Appl biol chem. 2022;65(1):null.

    Article  Google Scholar 

  99. Lyu Y, Zhang Y, Wang Y, Luo Y, Ding H, Li P, Ni G. HIF-1α Regulated WTAP Overexpression Promoting the Warburg Effect of Ovarian Cancer by m6A-Dependent Manner. J immunol res 2022, 2022(null):6130806.

  100. Lu J, Zhen S, Tuo X, Chang S, Yang X, Zhou Y, Chen W, Zhao L, Li X. Downregulation of DNMT3A attenuates the Warburg Effect, Proliferation, and Invasion via promoting the inhibition of miR-603 on HK2 in Ovarian Cancer. Technol cancer res t. 2022;21null:15330338221110668.

    Article  Google Scholar 

  101. Zheng X, Zhou Y, Chen W, Chen L, Lu J, He F, Li X, Zhao L. Ginsenoside 20(S)-Rg3 prevents PKM2-Targeting mir-324-5p from H19 sponging to antagonize the Warburg Effect in Ovarian Cancer cells. Cell Physiol Biochem. 2018;51(3):1340–53.

    Article  CAS  PubMed  Google Scholar 

  102. Liu Y, He X, Chen Y, Cao D. Long non-coding RNA LINC00504 regulates the Warburg effect in ovarian cancer through inhibition of miR-1244. Mol cell Biochem. 2020;464(1–2):39–50.

    Article  CAS  PubMed  Google Scholar 

  103. Boscaro C, Baggio C, Carotti M, Sandonà D, Trevisi L, Cignarella A, Bolego C. Targeting of PFKFB3 with miR-206 but not mir-26b inhibits ovarian cancer cell proliferation and migration involving FAK downregulation. Faseb j. 2022;36(3):e22140.

    Article  CAS  PubMed  Google Scholar 

  104. Rupaimoole R, Lee J, Haemmerle M, Ling H, Previs RA, Pradeep S, Wu SY, Ivan C, Ferracin M, Dennison JB, et al. Long noncoding RNA ceruloplasmin promotes Cancer Growth by altering Glycolysis. Cell Rep. 2015;13(11):2395–402.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Xu J, Gu X, Yang X, Meng Y. MiR-1204 promotes ovarian squamous cell carcinoma growth by increasing glucose uptake. Biosci Biotech Bioch. 2019;83(1):123–8.

    Article  Google Scholar 

  106. Xu S, Jia G, Zhang H, Wang L, Cong Y, Lv M, Xu J, Ruan H, Jia X, Xu P, et al. LncRNA HOXB-AS3 promotes growth, invasion and migration of epithelial ovarian cancer by altering glycolysis. Life sci. 2021;264null:118636.

    Article  Google Scholar 

  107. Imran K, Iqbal MJ, Abid R, Ahmad MM, Calina D, Sharifi-Rad J, Cho WC. Cellular signaling modulated by miRNA-3652 in ovarian cancer: unveiling mechanistic pathways for future therapeutic strategies. Cell Commun Signal. 2023;21(1):289.

    Article  PubMed  PubMed Central  Google Scholar 

  108. Hu J, Zhao W, Huang Y, Wang Z, Jiang T, Wang L. MiR-1180 from bone marrow MSCs promotes cell proliferation and glycolysis in ovarian cancer cells via SFRP1/Wnt pathway. Cancer Cell Int. 2019;19null:66.

    Article  Google Scholar 

  109. Liu J, Zhu Y, Wang H, Han C, Wang Y, Tang R. LINC00629, a HOXB4-downregulated long noncoding RNA, inhibits glycolysis and ovarian cancer progression by destabilizing c-Myc. Cancer sci. 2024;115(3):804–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Li X, Zhang Y, Wang X, Lin F, Cheng X, Wang Z, Wang X. Long non-coding RNA CTSLP8 mediates ovarian cancer progression and chemotherapy resistance by modulating cellular glycolysis and regulating c-Myc expression through PKM2. Cell biol Toxicol. 2022;38(6):1027–45.

    Article  CAS  PubMed  Google Scholar 

  111. Lin X, Feng D, Li P, Lv Y. LncRNA LINC00857 regulates the progression and glycolysis in ovarian cancer by modulating the Hippo signaling pathway. Cancer Med. 2020;9(21):8122–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Sun Y, Liu Y, Chen H, Tan Y. Circular RNA circRHOBTB3 inhibits ovarian cancer progression through PI3K/AKT signaling pathway. Panminerva med. 2024;66(1):36–46.

    Article  PubMed  Google Scholar 

  113. Teng Y, Zhang Y, Qu K, Yang X, Fu J, Chen W, Li X. MicroRNA-29B (mir-29b) regulates the Warburg effect in ovarian cancer by targeting AKT2 and AKT3. Oncotarget. 2015;6(38):40799–814.

    Article  PubMed  PubMed Central  Google Scholar 

  114. Tao LM, Gong YF, Yang HM, Pei JH, Zhao XJ, Liu SS. LINC00662 promotes glycolysis and cell survival by regulating miR- 375/HIF-1α axis in ovarian cancer. J biol reg Homeos Ag. 2020;34(3):467–77.

    CAS  Google Scholar 

  115. Li C, Liu FY, Shen Y, Tian Y, Han FJ. Research progress on the mechanism of glycolysis in ovarian cancer. Front Immunol. 2023;14(null):1284853.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Siu MKY, Jiang YX, Wang JJ, Leung THY, Han CY, Tsang BK, Cheung ANY, Ngan HYS, Chan KKL. Hexokinase 2 regulates ovarian Cancer Cell Migration, Invasion and Stemness via FAK/ERK1/2/MMP9/NANOG/SOX9 signaling cascades. Cancers (Basel) 2019, 11(6).

  117. Si M, Wang Q, Li Y, Lin H, Luo D, Zhao W, Dou X, Liu J, Zhang H, Huang Y et al. Inhibition of hyperglycolysis in mesothelial cells prevents peritoneal fibrosis. Sci Transl Med 2019, 11(495).

  118. Ma J, To SKY, Fung KSW, Wang K, Zhang J, Ngan AHW, Yung S, Chan TM, Wong CCL, Ip PPC, et al. P-cadherin mechanoactivates tumor-mesothelium metabolic coupling to promote ovarian cancer metastasis. Cell Rep. 2024;44(1):115096.

    Article  PubMed  Google Scholar 

  119. Schab AM, Greenwade MM, Stock E, Lomonosova E, Cho K, Grither WR, Noia H, Wilke D, Mullen MM, Hagemann AR, et al. Stromal DDR2 promotes ovarian Cancer Metastasis through Regulation of Metabolism and Secretion of Extracellular Matrix proteins. Mol cancer res. 2023;21(11):1234–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. B MMJJWWCC, BCC MMXX. G GCC: Tumor-derived Lactate creates a Favorable Niche for Tumor via Supplying Energy Source for Tumor and modulating the Tumor Microenvironment. Front cell Dev Biology 2022:808859.

  121. DC DCHH, LA LASS. The Tumor Microenvironment innately modulates Cancer Progression. Cancer Res 2019(18):4557–66.

  122. Tu VY, Ayari A, O’Connor RS. Beyond the Lactate Paradox: how Lactate and Acidity Impact T Cell therapies against Cancer. Antibodies (Basel). 2021;10(3):null.

    Google Scholar 

  123. Aktar N, Yueting C, Abbas M, Zafar H, Paiva-Santos AC, Zhang Q, Chen T, Ahmed M, Raza F, Zhou X. Understanding of Immune Escape Mechanisms and Advances in Cancer Immunotherapy. J oncol 2022, 2022(null):8901326.

  124. K KFF PPHH. S SVV, N NMM, J JAA, M MEE, E EGG, S SSS, G GRR, S SHH: Inhibitory effect of tumor cell-derived lactic acid on human T cells. Blood 2007(9):3812–9.

  125. Brand A, Singer K, Koehl GE, Kolitzus M, Schoenhammer G, Thiel A, Matos C, Bruss C, Klobuch S, Peter K, et al. LDHA-Associated Lactic Acid Production blunts Tumor Immunosurveillance by T and NK Cells. Cell Metab. 2016;24(5):657–71.

    Article  CAS  PubMed  Google Scholar 

  126. D DXX SSZZ. L LBB: Lactic acid in tumor microenvironments causes dysfunction of NKT cells by interfering with mTOR signaling. Sci China Life Sci 2016(12):1290–6.

  127. Qian F, Liao J, Miliotto A, Collins K, Odunsi K. Abstract A35: ovarian cancer stem cells subvert tumor-specific T cells by disrupting T cells’ metabolic fitness. Clin cancer res. 2018;24(15Supple):A35–35.

    Article  Google Scholar 

  128. Xia H, Wang W, Crespo J, Kryczek I, Li W, Wei S, Bian Z, Maj T, He M, Liu RJ, et al. Suppression of FIP200 and autophagy by tumor-derived lactate promotes naïve T cell apoptosis and affects tumor immunity. Sci Immunol. 2017;2(17):null.

    Article  Google Scholar 

  129. Fu S, He K, Tian C, Sun H, Zhu C, Bai S, Liu J, Wu Q, Xie D, Yue T, et al. Impaired lipid biosynthesis hinders anti-tumor efficacy of intratumoral iNKT cells. Nat Commun. 2020;11(1):438.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Henriksen JR, Nederby L, Donskov F, Waldstrøm M, Adimi P, Jakobsen A, Dahl Steffensen K. Blood natural killer cells during treatment in recurrent ovarian cancer. Acta Oncol. 2020;59(11):1365–73.

    Article  CAS  PubMed  Google Scholar 

  131. Fan Z, Han D, Fan X, Zhao L. Ovarian cancer treatment and natural killer cell-based immunotherapy. Front Immunol. 2023;14(null):1308143.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Kumar V, Mukhopadhyay A. Monitoring natural killer cell function in human ovarian Cancer cells of Ascitic Fluid. Bio Protoc. 2018;8(24):e3124.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Tan G, Spillane KM, Maher J. The role and regulation of the NKG2D/NKG2D ligand system in Cancer. Biology (Basel). 2023;12(8):null.

    Google Scholar 

  134. K KLL MMMM, H HYY JJHHNNMMAASS, K KYY TTBB, I IKK. S SFF,: Clinical significance of the NKG2D ligands, MICA/B and ULBP2 in ovarian cancer: high expression of ULBP2 is an indicator of poor prognosis. Cancer Immunol Immunotherapy: CII 2009(5):641–52.

  135. S S, L P, W J: Expression levels of matrix metalloproteinases in ascites of patients with ovarian cancer and the relationship with pathological characteristics of ovarian cancer. Dcc 2019, 6(2):12.

  136. Vos MC, van der Wurff AAM, van Kuppevelt TH, Massuger L. The role of MMP-14 in ovarian cancer: a systematic review. J Ovarian Res. 2021;14(1):101.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Barua A, Bitterman P, Sharma S, Basu S, Bahr J. Abstract A48: ovarian tumor-induced immunosuppression of NK cells and its prevention by dietary supplementation of herbal withaferin A (Ashwagandha). Clin cancer res. 2016;22(2Supple):A48–48.

    Article  Google Scholar 

  138. Xie J, Liu M, Li Y, Nie Y, Mi Q, Zhao S. Ovarian tumor-associated microRNA-20a decreases natural killer cell cytotoxicity by downregulating MICA/B expression. Cell mol Immunol. 2014;11(5):495–502.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Jedlička M, Feglarová T, Janstová L, Hortová-Kohoutková M, Frič J. Lactate from the tumor microenvironment - A key obstacle in NK cell-based immunotherapies. Front Immunol 2022, 13(null):932055.

  140. Terrén I, Orrantia A, Vitallé J, Zenarruzabeitia O, Borrego F. NK Cell Metabolism and Tumor Microenvironment. Front Immunol. 2019;10(null):2278.

    Article  PubMed  PubMed Central  Google Scholar 

  141. Pötzl J, Roser D, Bankel L, Hömberg N, Geishauser A, Brenner CD, Weigand M, Röcken M, Mocikat R. Reversal of tumor acidosis by systemic buffering reactivates NK cells to express IFN-γ and induces NK cell-dependent lymphoma control without other immunotherapies. Int j cancer. 2017;140(9):2125–33.

    Article  PubMed  Google Scholar 

  142. Zhang S, Ke X, Zeng S, Wu M, Lou J, Wu L, Huang P, Huang L, Wang F, Pan S. Analysis of CD8 + Treg cells in patients with ovarian cancer: a possible mechanism for immune impairment. Cell mol Immunol. 2015;12(5):580–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Tokunaga R, Naseem M, Lo JH, Battaglin F, Soni S, Puccini A, Berger MD, Zhang W, Baba H, Lenz HJ. B cell and B cell-related pathways for novel cancer treatments. Cancer Treat rev. 2019;73(null):10–9.

    Article  CAS  PubMed  Google Scholar 

  144. Calcinotto A, Filipazzi P, Grioni M, Iero M, De Milito A, Ricupito A, Cova A, Canese R, Jachetti E, Rossetti M, et al. Modulation of microenvironment acidity reverses anergy in human and murine tumor-infiltrating T lymphocytes. Cancer res. 2012;72(11):2746–56.

    Article  CAS  PubMed  Google Scholar 

  145. Watson MJ, Vignali PDA, Mullett SJ, Overacre-Delgoffe AE, Peralta RM, Grebinoski S, Menk AV, Rittenhouse NL, DePeaux K, Whetstone RD, et al. Metabolic support of tumour-infiltrating regulatory T cells by lactic acid. Nature. 2021;591(7851):645–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Angelin A, Gil-de-Gómez L, Dahiya S, Jiao J, Guo L, Levine MH, Wang Z, Quinn WJ, Kopinski PK, Wang L, et al. Foxp3 reprograms T cell metabolism to function in Low-Glucose, high-lactate environments. Cell Metab. 2017;25(6):1282–e12931287.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Gu J, Zhou J, Chen Q, Xu X, Gao J, Li X, Shao Q, Zhou B, Zhou H, Wei S, et al. Tumor metabolite lactate promotes tumorigenesis by modulating MOESIN lactylation and enhancing TGF-β signaling in regulatory T cells. Cell Rep. 2022;39(12):110986.

    Article  CAS  PubMed  Google Scholar 

  148. Kumagai S, Koyama S, Itahashi K, Tanegashima T, Lin YT, Togashi Y, Kamada T, Irie T, Okumura G, Kono H, et al. Lactic acid promotes PD-1 expression in regulatory T cells in highly glycolytic tumor microenvironments. Cancer Cell. 2022;40(2):201–e218209.

    Article  CAS  PubMed  Google Scholar 

  149. Shang W, Xu R, Xu T, Wu M, Xu J, Wang F. Ovarian Cancer cells promote glycolysis metabolism and TLR8-Mediated metabolic control of human CD4 + T cells. Front Oncol. 2020;10(null):570899.

    Article  PubMed  PubMed Central  Google Scholar 

  150. Wu M, Fu X, Xu R, Liu S, Li R, Xu J, Shang W, Chen X, Wang T, Wang F. Glucose metabolism and function of CD4 + tregs are regulated by the TLR8/mTOR signal in an environment of SKOV3 cell growth. Cancer Med. 2023;12(15):16310–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Xu R, Wu M, Liu S, Shang W, Li R, Xu J, Huang L, Wang F. Glucose metabolism characteristics and TLR8-mediated metabolic control of CD4 + Treg cells in ovarian cancer cells microenvironment. Cell Death Dis. 2021;12(1):22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Gao T, Lin YQ, Ye HY, Lin WM. miR-124 delivered by BM-MSCs-derived exosomes targets MCT1 of tumor-infiltrating Treg cells and improves ovarian cancer immunotherapy. Neoplasma. 2023;70(6):713–21.

    Article  CAS  PubMed  Google Scholar 

  153. Wang Y, Zhu Z. Oridonin inhibits metastasis of human ovarian cancer cells by suppressing the mTOR pathway. Arch med sci. 2019;15(4):1017–27.

    Article  CAS  PubMed  Google Scholar 

  154. Kumar S, Mittal S, Gupta P, Singh M, Chaluvally-Raghavan P, Pradeep S. Metabolic reprogramming in Tumor-Associated macrophages in the ovarian Tumor Microenvironment. Cancers (Basel). 2022;14(21):null.

    Article  Google Scholar 

  155. Sica A, Mantovani A. Macrophage plasticity and polarization: in vivo veritas. J clin Invest. 2012;122(3):787–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Kepuladze S, Devadze R, Gvenetadze A, Burkadze G. Distribution of tumor-associated macrophages and M1/M2 polarization in different types and grades of ovarian tumors. Ijpo. 2022;9(4):318–21.

    Article  Google Scholar 

  157. Vankerckhoven A, Wouters R, Mathivet T, Ceusters J, Baert T, Van Hoylandt A, Gerhardt H, Vergote I, Coosemans A. Opposite macrophage polarization in different subsets of Ovarian Cancer: Observation from a pilot study. Cells. 2020;9(2):null.

    Article  Google Scholar 

  158. Colegio OR, Chu NQ, Szabo AL, Chu T, Rhebergen AM, Jairam V, Cyrus N, Brokowski CE, Eisenbarth SC, Phillips GM, et al. Functional polarization of tumour-associated macrophages by tumour-derived lactic acid. Nature. 2014;513(7519):559–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Lian X, Yang K, Li R, Li M, Zuo J, Zheng B, Wang W, Wang P, Zhou S. Immunometabolic rewiring in tumorigenesis and anti-tumor immunotherapy. Mol Cancer. 2022;21(1):27.

    Article  PubMed  PubMed Central  Google Scholar 

  160. Paolini L, Adam C, Beauvillain C, Preisser L, Blanchard S, Pignon P, Seegers V, Chevalier LM, Campone M, Wernert R, et al. Lactic acidosis together with GM-CSF and M-CSF induces human macrophages toward an inflammatory Protumor phenotype. Cancer Immunol res. 2020;8(3):383–95.

    Article  CAS  PubMed  Google Scholar 

  161. Liu N, Luo J, Kuang D, Xu S, Duan Y, Xia Y, Wei Z, Xie X, Yin B, Chen F, et al. Lactate inhibits ATP6V0d2 expression in tumor-associated macrophages to promote HIF-2α-mediated tumor progression. J clin Invest. 2019;129(2):631–46.

    Article  PubMed  PubMed Central  Google Scholar 

  162. Chen H, Huang S, Niu P, Zhu Y, Zhou J, Jiang L, Li D, Shi D. Cardamonin suppresses pro-tumor function of macrophages by decreasing M2 polarization on ovarian cancer cells via mTOR inhibition. Mol Ther Oncolytics. 2022;26(null):175–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Vadevoo SMP, Gunassekaran GR, Lee C, Lee N, Lee J, Chae S, Park JY, Koo J, Lee B. The macrophage odorant receptor Olfr78 mediates the lactate-induced M2 phenotype of tumor-associated macrophages. P natl acad sci USA. 2021;118(37):null.

    Article  Google Scholar 

  164. Xu C, Xia Y, Zhang BW, Drokow EK, Li HY, Xu S, Wang Z, Wang SY, Jin P, Fang T, et al. Macrophages facilitate tumor cell PD-L1 expression via an IL-1β-centered loop to attenuate immune checkpoint blockade. MedComm (2020). 2023;4(2):e242.

    Article  CAS  PubMed  Google Scholar 

  165. de Visser KE, Joyce JA. The evolving tumor microenvironment: from cancer initiation to metastatic outgrowth. Cancer Cell. 2023;41(3):374–403.

    Article  PubMed  Google Scholar 

  166. Conradi LC, Brajic A, Cantelmo AR, Bouché A, Kalucka J, Pircher A, Brüning U, Teuwen LA, Vinckier S, Ghesquière B, et al. Tumor vessel disintegration by maximum tolerable PFKFB3 blockade. Angiogenesis. 2017;20(4):599–613.

    Article  CAS  PubMed  Google Scholar 

  167. Végran F, Boidot R, Michiels C, Sonveaux P, Feron O. Lactate influx through the endothelial cell monocarboxylate transporter MCT1 supports an NF-κB/IL-8 pathway that drives tumor angiogenesis. Cancer res. 2011;71(7):2550–60.

    Article  PubMed  Google Scholar 

  168. Hosonuma M, Yoshimura K. Association between pH regulation of the tumor microenvironment and immunological state. Front Oncol 2023, 13(null):1175563.

  169. Sangsuwan R, Thuamsang B, Pacifici N, Allen R, Han H, Miakicheva S, Lewis JS. Lactate exposure promotes immunosuppressive phenotypes in Innate Immune cells. Cell mol Bioeng. 2020;13(5):541–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Raychaudhuri D, Bhattacharya R, Sinha BP, Liu CSC, Ghosh AR, Rahaman O, Bandopadhyay P, Sarif J, D’Rozario R, Paul S, et al. Lactate induces pro-tumor reprogramming in Intratumoral Plasmacytoid dendritic cells. Front Immunol. 2019;10(null):1878.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Demoulin S, Herfs M, Delvenne P, Hubert P. Tumor microenvironment converts plasmacytoid dendritic cells into immunosuppressive/tolerogenic cells: insight into the molecular mechanisms. J Leukoc biol. 2013;93(3):343–52.

    Article  CAS  PubMed  Google Scholar 

  172. Jahrsdörfer B, Vollmer A, Blackwell SE, Maier J, Sontheimer K, Beyer T, Mandel B, Lunov O, Tron K, Nienhaus GU, et al. Granzyme B produced by human plasmacytoid dendritic cells suppresses T-cell expansion. Blood. 2010;115(6):1156–65.

    Article  PubMed  PubMed Central  Google Scholar 

  173. Kouidhi S, Ben Ayed F, Benammar Elgaaied A. Targeting Tumor metabolism: a New Challenge to Improve Immunotherapy. Front Immunol. 2018;9(null):353.

    Article  PubMed  PubMed Central  Google Scholar 

  174. Ma Y, Wang W, Idowu MO, Oh U, Wang XY, Temkin SM, Fang X. Ovarian Cancer relies on glucose transporter 1 to fuel glycolysis and Growth: Anti-tumor Activity of BAY-876. Cancers (Basel). 2018;11(1):null.

    Article  Google Scholar 

  175. Wang JJ, Siu MK, Jiang YX, Leung TH, Chan DW, Cheng RR, Cheung AN, Ngan HY, Chan KK. Aberrant upregulation of PDK1 in ovarian cancer cells impairs CD8 + T cell function and survival through elevation of PD-L1. Oncoimmunology. 2019;8(11):e1659092.

    Article  PubMed  PubMed Central  Google Scholar 

  176. Ho PC, Bihuniak JD, Macintyre AN, Staron M, Liu X, Amezquita R, Tsui YC, Cui G, Micevic G, Perales JC, et al. Phosphoenolpyruvate is a metabolic checkpoint of anti-tumor T cell responses. Cell. 2015;162(6):1217–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Patsoukis N, Bardhan K, Chatterjee P, Sari D, Liu B, Bell LN, Karoly ED, Freeman GJ, Petkova V, Seth P et al. PD-1 alters T-cell metabolic reprogramming by inhibiting glycolysis and promoting lipolysis and fatty acid oxidation. Nat Commun 2015, 6(null):6692.

  178. Gao F, Tang Y, Liu WL, Zou MZ, Huang C, Liu CJ, Zhang XZ. Intra/Extracellular lactic acid exhaustion for synergistic metabolic therapy and immunotherapy of tumors. Adv Mater. 2019;31(51):e1904639.

    Article  PubMed  Google Scholar 

  179. Huang T, Feng Q, Wang Z, Li W, Sun Z, Wilhelm J, Huang G, Vo T, Sumer BD, Gao J. Tumor-targeted inhibition of Monocarboxylate Transporter 1 improves T-Cell immunotherapy of solid tumors. Adv Healthc Mater. 2021;10(4):e2000549.

    Article  PubMed  Google Scholar 

  180. Hua H, Kong Q, Zhang H, Wang J, Luo T, Jiang Y. Targeting mTOR for cancer therapy. J Hematol Oncol. 2019;12(1):71.

    Article  PubMed  PubMed Central  Google Scholar 

  181. Shu P, Li X, Yuan L, Xie H, Zhou Q, Wang Y, Li Q, Zheng L. A phase 1/2 study of onatasertib, a dual TORC1/2 inhibitor, combined with the PD-1 antibody toripalimab in patients with advanced solid tumors (TORCH-2). J clin Oncol. 2022;40(16suppl):2610–2610.

    Article  Google Scholar 

  182. Davis S, Leal A, Messersmith W, Lieu C, Lam E, Corr B, et al. A phase Ib study of the combination of alisertib (Aurora A kinase inhibitor) and MLN0128 (dual TORC1/2 Inhibitor) in patients with advanced solid tumors, final expansion cohort data. J Clin Oncol. 2022;40(16_suppl):3112–12.

  183. Musa F, Alard A, David-West G, Curtin JP, Blank SV, Schneider RJ. Dual mTORC1/2 Inhibition as a Novel Strategy for the Resensitization and Treatment of Platinum-Resistant Ovarian Cancer. Mol Cancer Ther. 2016;15(7):1557–67.

  184. Pi R, Yang Y, Hu X, Li H, Shi H, Liu Y, et al. Dual mTORC1/2 inhibitor AZD2014 diminishes myeloid-derived suppressor cells accumulation in ovarian cancer and delays tumor growth. Cancer Lett. 2021;523(null):72–81.

  185. Fabi F, Adam P, Parent S, Tardif L, Cadrin M, Asselin E. Pharmacologic inhibition of Akt in combination with chemotherapeutic agents effectively induces apoptosis in ovarian and endometrial cancer cell lines. Mol Oncol. 2021;15(8):2106–19.

Download references

Acknowledgements

Not applicable.

Funding

This study was funded by Key Medical Discipline of Hangzhou City (2021–21); Key Medical Discipline of Zhejiang Province (2018–2–3); Key Laboratory of Clinical Cancer Pharmacology and Toxicology Research of Zhejiang Province (2020E10021); Zhejiang Province Medical and Health Science and Technology Program (2023KY933); Zhejiang Traditional Chinese Medicine Science and Technology Project (2023ZL565); Zhejiang Traditional Chinese Medicine Science and Technology Project (2022ZA139).

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed to the study concept and design; Bowen Jin, Junjie Pan and Zehua Miao did literature search; All authors performed critical review and interpretation of the findings; Bowen Jin drafted the article; Bowen Jin, Junjie Pan and Zehua Miao revised and edited the manuscript; All authors contributed to the final approval of the article.

Corresponding authors

Correspondence to Zheng Niu or Qiaoping Xu.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

The Author confirms: that the work described has not been published before; that it is not under consideration for publication elsewhere; that its publica- tion has been approved by all co-authors.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note

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

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/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jin, B., Miao, Z., Pan, J. et al. The emerging role of glycolysis and immune evasion in ovarian cancer. Cancer Cell Int 25, 78 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12935-025-03698-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12935-025-03698-x

Keywords