Skip to main content

Cancer immunometabolism: advent, challenges, and perspective

Abstract

For decades, great strides have been made in the field of immunometabolism. A plethora of evidence ranging from basic mechanisms to clinical transformation has gradually embarked on immunometabolism to the center stage of innate and adaptive immunomodulation. Given this, we focus on changes in immunometabolism, a converging series of biochemical events that alters immune cell function, propose the immune roles played by diversified metabolic derivatives and enzymes, emphasize the key metabolism-related checkpoints in distinct immune cell types, and discuss the ongoing and upcoming realities of clinical treatment. It is expected that future research will reduce the current limitations of immunotherapy and provide a positive hand in immune responses to exert a broader therapeutic role.

Highlights

1. Attempting to delineate the complex and multidimensional interplays between metabolites (or metabolic enzymes) and predominant immune cell populations.

2. Metabolic checkpoints of immune cells are described and the contribution of these metabolic targets to determine the metabolic adaptations of distinct immune cells in specific tissue environments is emphasized.

3. Proposed cancer-immunometabolism subcycle, enriching the theoretical foundation of the cancer-immunology field.

4. Metabolism-induced disturbances in the acid-base balance of the tumor microenvironment have non-redundant effects on immunotherapy.

5. Pending challenges and clinical concerns in metabolic insights were addressed.

Introduction

As a pioneer in the quantitative study of cancer cell metabolism as well as photosynthesis and respiration, Otto Warburg and colleagues first unraveled the mystery of cancer’s ability to rapidly consume large amounts of glucose independent of oxygen for its growth and proliferation in the 1920s, a phenomenon also known as the Warburg effect [1, 2]. Indeed, various of solid tumors exhibit the Warburg effect while preserving mitochondrial respiration, which is an inefficient way to generate adenosine 5’-triphosphate (ATP), compared to oxidative phosphorylation (OXPHOS) [3]. Studies reported that the primary function of the Warburg effect may be to maintain high levels of glycolytic products, or even to enable “metabolic transformation” to support active anabolic reactions within the cell [4, 5]. Similarly, a consequence of oxidative metabolism is the production of reactive oxygen species (ROS), which could support tumorigenesis but require tight regulation of redox balance [6]. Of interest, tumors undergo dysregulation of multiple metabolic pathways improve the metabolic flexibility, which subsequently induces altered immune status and tumor progression [3, 7, 8]. For instance, certain metabolic processes are aberrantly enabled in cancer cells, including glutamate transport, rapid glutamine uptake, and fatty acid oxidation (FAO), which involve metabolites that act as immune mediators, resulting in reduced immunogenicity of cancer cells, immune escape, as well as state of localized immunosuppression in the tumor microenvironment and immunotherapy resistance [9,10,11,12]. However, the paucity of successful clinical data on metabolism-related therapies in cancer patients continues to attract researchers to initiate more in-depth studies.

The perception of metabolism by immune cells is closely linked to their fate decisions [13,14,15,16]. Nonetheless, it is inevitable that cancer cells are inherently competitive with immune cells in their demand for essential nutrients [17, 18]. The nutrient competition is depicted in Fig. 1. Chang and colleagues demonstrated metabolic competition between tumor cells and T cells in a mouse sarcoma model, which contributes to T cell dysfunction and tumor progression [19]. Hypoxia, one of the key drivers of tumor heterogeneity, mediates both metabolic reprogramming and immune escape [20]. Cholesterol metabolism produces important membrane components as well as metabolic derivatives with diverse biological functions [21]. Preclinical and clinical studies have shown that manipulation of cholesterol metabolism suppresses tumor growth and remodels the immune landscape [22, 23]. Specifically, the metabolic demands of immune cells largely affect the success of immunotherapy, which might be one of the principal reasons why many cancers remain resistant to immunotherapy and the long-term prognosis of patients cannot be guaranteed [17, 24]. Ultimately, if immunotherapy could be used early in the disease or as a link in combination therapy, the initiation of immune responses and transformation of the immunophenotype might be less restricted and perhaps more malleable.

Fig. 1
figure 1

Metabolic competition between tumor cells and immune cells. The availability of nutrients for metabolic processes is fundamental for cell survival, along with tumor cells and immune cells are no exception. Competitive uptake of nutrients by tumor cells in the tumor microenvironment may occur at all stages of immune cell life. Metabolite paucity tilts the energy balance in favour of the tumor cells (the negative direction), which in turn leads to further dysfunction of immune cells (such as naïve T cells, B cells, natural killer cells, macrophages, neutrophils, and dendritic cells, etc.)

With the proposed cross-cutting field of immunometabolism, the immune system underlying the metabolic landscape is being redefined by oncologists from multiple perspectives [25,26,27,28]. This review attempts to delineate the complex and multidimensional crosstalk between metabolites (or metabolic enzymes) and predominant immune cell populations, and highlights the contributions made by metabolic targets to the metabolic adaptations of immune cells in specific environments. Accordingly, clinical oncology treatment has progressed based on attempts to combine nutritional therapies with immunotherapy, yet there are still open questions.

Metabolic reprogramming for immune regulation

Accumulating evidence has led oncologists and immunologists to appreciate that metabolites and enzymes are important regulators of the immune system, which involved in energy circuits and signaling cascades [3, 29, 30]. Therefore, metabolic reprogramming caused by abnormal metabolites or metabolic enzymes produces a profound effect on the immune response.

Metabolites act as immune mediators

Metabolites have functions in the immune system independent of their traditional roles as biosynthesis and energy supply [31]. However, most studies to date have focused on the regulation of metabolic pathways during immune responses [32]. Notably, the discovery of metabolites and intermediates as novel signaling molecules is thought to produce a profound effect on immune regulation [33, 34].

Glucose

Glucose supply and glycolysis processes play an important role in the development and progression of tumors (Fig. 2) [35]. CD8+ T cell proliferation and cytokine production depend on enhanced glucose metabolism [36]. Glucose restriction can activate AMPK-coupled SENP1-Sirt3 signaling in mitochondria and promote T cell development [37]. Animal studies have indicated that diabetic mice exhibit larger breast tumors characterized by altered collagen structure, increased tumor-allowed M2 macrophage infiltration, and early spread metastasis [38]. Competitive glucose metabolism may also be a target to improve the efficacy of bladder cancer immunotherapy [39]. Notably, a recent study has focused on glucose-promoting tumor progression and immunotherapy resistance in a non-classical metabolism-dependent manner, directly in the form of signal transduction molecules [12]. Collectively, these studies not only show that glucose levels play an important role in the energy interaction between tumors and immune cells, but also highlight the role of glucose molecules as signaling molecules for immune regulation from a new perspective.

Fig. 2
figure 2

Immune-related intracellular energy metabolism and substance synthesis. The diagram shows the metabolic activities and synthetic reactions that occur in the cell under enzymatic reactions and correlate with or might cause immune changes (enzymes are labeled in purple, nutrients or metabolites are labeled in green, molecules or targets are labeled in yellow, and inhibitors are labeled in grey). For example, lipid uptake from the TME leads to elevated intracellular cholesterol concentrations, which in turn triggers ER stress (inducing CD8+ T cell dysfunction). The PI3K/AKT pathway, which is activated by growth factor signals, stimulates the mTOR family molecules, which in turn elicits vital activities such as protein synthesis, cell proliferation, and autophagy, and so on. The mTOR family molecules are also regulated by amino acids. FA synthesis is coordinated sequentially by several enzymes involving ACC1 (inhibition of ACC1 reduces TH17 cell differentiation but enhances the formation of memory CD4+ T cells). And C75 inhibits FASN which in turn diminishes FA synthesis. In addition, the induction of FA oxidation was associated with an increase in AMPK activity (AMPK also promotes the generation of memory CD8+ T cells). JHU083 inhibited glutaminase-mediated glutaminolysis. TME, tumor microenvironment; SREBP, sterol regulatory element binding protein; ER, endoplasmic reticulum; PI3K, phosphatidylinositol 3-kinase; FA, Fatty acid; ACC1, acetyl-CoA carboxylase 1; PDH, pyruvate dehydrogenase; αKG, α-ketoglutarate; FASN, fatty acid synthase; AMPK, AMP-activated protein kinase

Amino acids

Tumor growth and development depend on the intake of foreign amino acids, which affects the function of immune cells [40, 41]. Therefore, alterations in amino acid metabolism could be used not only as a clinical indicator of cancer progression but also as a therapeutic strategy.

Leucine (Leu) & Arginine (Arg)

As one of the branched-chain amino acids, leucine (Leu) acts as a nitrogen donor to produce biomolecules such as nucleotides, which are indispensable for the growth of cancer cells [42]. In a clinical study (NTR2121), a nutritional intervention with a high-leucine specific medical food rapidly increased the percentage of EPA and DHA in leukocyte phospholipids and lowered serum levels of the inflammatory mediator PGE2 within one week in cancer patients undergoing radiation therapy [43]. Leu restriction has now been shown to limit the response of premalignant B cells. It is now well established that limiting leucine then limits the response of pro-cancer B cells [44, 45]. Arginine (Arg) metabolism affects not only malignant cells but also the behavior of surrounding immune cells [46, 47]. Inhibition of Arg by CB-1158 blocked myelocyte-mediated immunosuppression in the tumor microenvironment (TME) [48]. Miret et al.’s data suggests Arg is an immunomodulatory target in KRASG12D genetically engineered mouse models, and inhibition of arginase attenuated tumor growth [49]. Therefore, it is promising to develop therapeutic strategies targeting immunomodulatory pathways controlled by Leu and/or Arg degradation.

Glutamate (Glu) & glutamine (Gln)

Glutamate (Glu) is a major excitatory neurotransmitter in the central nervous system (CNS) and also plays a critical function in tissue and cellular metabolism through the tricarboxylic acid cycle [50, 51]. Long et al. found that dysregulated Glu transport enhances T regulatory cell (Treg cell) proliferation, activation, and immunosuppressive functions, as well as promotes resistance to VEGF blockade of glioblastomas in vitro [9]. The use of a glutaminase antagonist, JHU083, effectively inhibited tumor growth in a variety of solid tumor models and significantly improved mouse survival [52]. Furthermore, activation of naïve T cells is associated with rapid Gln uptake [10]. Selective inhibition of Gln metabolism in tumor cells increases anti-tumor T lymphocyte activity in triple-negative breast cancer (TNBC) patients [53]. Thus, Glu and Gln metabolism are reprogrammed during tumorigenesis and are considered a promising target for cancer therapy.

Tryptophan (Trp) & Asparagine (Asn)

Substantial evidence suggests that tryptophan (Trp) and asparagine (Asn) metabolism are physiologically and pathologically involved in the progression and treatment of a wide range of diseases, including cancer [54,55,56]. Qin et al. found that IDO inhibitors could mediate tryptophanyl-tRNA synthetase (WARS) overexpression via accumulating Trp, which accelerates TRIP12 tryptophanylation and reduces surface PD-1 of mouse CD8+ T cells [57]. Thus, supplementation with exogenous Trp or use of IDO inhibitors to impede Trp catabolism may be beneficial for PD-1 blockade therapy. Likewise, it has been demonstrated that the Trp metabolizing enzyme tryptophan 2,3-dioxygenase (TDO) inhibits the anti-tumor activity of CD8 T cells in TNBC [58]. Additionally, Trp metabolism mediates impaired differentiation of myeloid cells infiltrated by IDH mutant gliomas, resulting in an immature immune phenotype [59]. Recent evidence suggests that Trp metabolites released by Lactobacillus tumefaciens locally promote interferon-gamma (IFN-γ)-producing CD8 T cells, thereby enhancing immune checkpoint inhibitor efficacy [60]. The intimate link between the kynurenine (Kyn) pathway of tryptophan metabolism and T cell function has been widely reported to date. As proof, IDO inhibitors enhance CD8+ T cell effects by accumulating tryptophan and or inhibiting Kyn production [57, 61]. A dietary model constructed by Siska et al. demonstrated that a high concentration (1 mM) of D-kyn could inhibit T cell proliferation via apoptosis manner [62]. Kyn derivatives 3-hydroxyanthranilic acid inhibits pro-inflammatory factors in several cell subsets including resident macrophages, proliferating macrophages, and plasmacytoid dendritic cells (pDCs) [63]. Nevertheless, in a study that enrolled 891 non-small cell lung cancer (NSCLC) samples, Bessede et al. noted that combined anti-PD-1/PD-L1 targeting of IDO1 might only be beneficial in patients with inflammatory tumors, and that the IDO1 pathway in NSCLC is driven by the immune system rather than tumor cells [64]. Correspondingly, the SRC family protein tyrosine kinase LCK is phosphorylated at tyrosine 394 and 505 upon binding to Asn, which subsequently increases T cell activation and anti-tumor effects [65]. Emerging evidence reveals that Asn restriction allows for increased metabolic capacity and anti-tumor function in CD8+ T cells in an NRF2-dependent manner of enhanced stress response [66]. To sum up, studies on amino acid metabolism and immunomodulation and immunotherapy or combination therapy based on amino acid metabolism still need to be further explored.

Lipids

Lipid-rich lung-resident mesenchymal cells (MCs) are known to promote lung metastasis of breast cancer. Lipid-loaded MCs transport lipids to tumor cells and natural killer (NK) cells via exosome-like vesicles, leading to enhanced tumor cell survival and proliferation as well as NK cell dysfunction [67]. Accordingly, lipid droplets are intracellular lipid reservoirs that are utilized by effector memory CD4+ T cells in nutrient-deficient environments [68]. Cholesterol metabolism plays a crucial role in regulating anti-tumor immune responses by acting on various immune cells involved in innate and adaptive immune responses [69, 70]. In addition, caloric restriction decreases total cholesterol and triglyceride levels, stimulates cancer immune surveillance, and reduces the migration of immunosuppressive regulatory T cells to tumors [71]. Cholesterol in TME induces dysfunctional CD8+ T cells by triggering endoplasmic reticulum (ER) stress, manifested by certain co-inhibitory molecule expression and impaired effector function [72]. In addition to serving as a fuel source for energy production, fatty acid (FA) primarily serves as structural components of membrane matrices and important secondary messengers (Fig. 2) [73, 74]. FA synthesis is coordinated by several enzymes involving acetyl-CoA carboxylase 1 (ACC1), and inhibition of ACC1 decreases TH17 cell differentiation but enhances the formation of memory CD4+ T cells [75,76,77]. Induction of FAO is associated with an elevation of AMP-activated protein kinase (AMPK) activity, while AMPK promotes the generation of central memory CD8+ T cells [78, 79]. Grajchen et al. found that enzyme-catalyzed desaturation of FAs is an important determinant of Treg differentiation and autoimmunity [80]. FAs play a double-edged role in the immunomodulation of the body (Fig. 3) [11, 81]. In addition, studies on lipid and lipoprotein transport pathways have provided options for improving prelipidic routes of administration for oral administration and therapy, with the promise of involvement in immunotherapy through the lymphatic system [82, 83]. Thus, the importance of altered cholesterol and FA metabolism in cancer should receive new attention.

Fig. 3
figure 3

Double-edged swords in cancer immunometabolism. Implications of FA synthesis (1) and catabolism (2) on tumor progression. 1) Upregulation of SREBP activity in Treg cells synergizes with FASN to promote FA synthesis, which in turn activates the PI3K pathway and facilitates the maturation of Treg cells. Specific deletion of SCAP (an essential factor for SREBP activity) by Treg cells enhances anti-PD-1 immunotherapy; 2) Leptin downregulates CD8+ T cell effector function through activation of STAT3-FAO and inhibition of glycolysis. Ablating T cell STAT3 or treatment with perhexiline (FTO inhibitor) in obese mice spontaneously developing breast tumor reduces FAO, increases glycolysis and CD8T effector cell functions, leading to inhibition of breast tumor development. Additionally, the effects of lactate on cancer and immune cells in TME can be complex and difficult to decipher, which is further confounded by acid protons (byproducts of glycolysis). 3) Tumor-derived lactate is an inhibitor of CD8+ T cell cytotoxicity. Cytotoxic T cells shunt succinate out of the TCA circulation to promote autocrine signalling via the succinate receptor (SUCNR1). Moreover, cytotoxic T cells rely on PC to replenish succinate. Lactate decreases PC activity, and similarly, inhibition of PDH restores PC activity, succinate secretion, and SUCNR1 activation; 4) Lactate increases CD8+ T cell stemness and enhances anti-tumor immunity. Subcutaneous injection of lactate in mice transplanted with MC38 tumors leads to CD8+ T cell-dependent tumor growth inhibition. Mechanistically, lactate inhibits histone deacetylase activity, leading to increased acetylation of the Tcf7 super-enhancer site, H3K27, which results in increased Tcf7 gene expression. FA, fatty acid; SREBP, sterol regulatory element binding protein; FASN, fatty acid synthase; PI3K, phosphatidylinositol 3-kinase; FAO, fatty acid oxidation; TME, tumor microenvironment; PC, pyruvate carboxylase; PDH, pyruvate dehydrogenase

Other alternative metabolites

Citrate

Citrate produces a profound effect on the immune and inflammatory responses engaged in both primary and adaptive immune cells and is also thought to play a crucial role in cancer metabolism [15, 84, 85]. Studies have shown that citrate could induce macrophages to rapidly secrete pro-inflammatory cytokines, which in turn facilitates the destruction of cancer stem cells (CSCs) [86]. Inhibition of citric acid carriers could cause peripheral macrophage inactivation and reduce cerebral thrombosis [84]. In fact, carcinogenic signaling pathways, such as HIF-1α and RAS/PI3K/AKT, may cause resistance by enhancing the aerobic glycolysis of cancer cells, known as the “Warburg effect” [86,87,88]. However, most drugs are weakly alkaline molecules, and this metabolism that promotes the development and aggressiveness of cancer cells can also induce increased extracellular acidity to weaken the penetration of compounds into cancer cells and even lead to the occurrence of multi-drug resistance events. Interestingly, citrate-rich organs, such as the liver, brain, and bone, are also common sites for metastasis of various malignant tumors, and it is possible that high citrate forms a good metastatic niche for the growth of secondary tumors and improves the survival rate of colonized cancer cells [89, 90]. Therefore, it might be possible to enhance the local infiltration of tumor chemotherapy drugs by reducing extracellular acidity strategies to achieve effective therapeutic concentrations and improve prognosis in clinical or preclinical trials.

Other mitochondria and TCA cycle metabolites

Most of the intermediates in the TCA cycle are important raw materials for the synthesis of the three nutrients (glucoses, lipids, and proteins), so the TCA cycle is also considered to be the hub of the metabolic connection of the three nutrients [91, 92]. The TCA cycle and oxidative phosphorylation process in mitochondria could produce a variety of metabolites and energy substances including ATP, NADH, α-ketoglutaric acid, etc., and further cross-link with immune cells (Fig. 2) [93,94,95]. For instance, as a by-product of dehydrogenase reaction in the electron transport chain, NADH is involved in regulating the energy metabolism of immune cells [94]. And α-ketoglutarate (αKG) is an intermediate product involved in energy metabolism and synthesis of T cell proliferation and inflammation [95, 96].

PGE2 and lactate

Conventional type 1 dendritic cell (cDC1) is an important anti-tumor immune cell, which can present tumor antigens and secrete IL-12 and other cytokines that promote T cell activation and effector function. Studies have confirmed that PGE2 can directly inhibit the survival of NK cells and the production of chemokines, and reduce the reactivity of cDC1 to chemokines, thereby blocking the recruitment of cDC1 [97]. Using a variety of mouse tumor models, researchers have found that the endogenous cyclooxygenase-2 / prostaglandin E2 (COX-2/PGE2) pathway in tumor cells inhibits NK cell infiltration and IFN-γ production, thereby promoting tumor evasion from immune surveillance [98]. Watson et al. reported that lactate treatment prevented the damaging effects on the function and stability of Treg cells under high glucose conditions [99]. In addition, lactic acid accumulation in the tissue microenvironment limits the function of immune cells, yet activated immune cells need lactic acid to perform their functions [100,101,102]. Briefly, lactate reduces pyruvate carboxylase (PC) activity, succinate secretion and SUCNR1 activation, which in turn inhibits autocrine signaling in cytotoxic T cells [102]. Therefore, lactate produces a double-edged effect on the immune process, suggesting a complex link between tumor immunity and metabolite regulation (Fig. 3) [100, 102].

Metabolic enzymes act as immune mediators

Inevitably, enzymatic reactions catalyzed by metabolic enzymes play a non-negligible role in the synthesis or catabolism of metabolites, especially in pathological states [103]. Changes resulting from metabolic enzyme abnormalities may be efficient and specific. Therefore, unlike nutritional therapies related to metabolite deficiencies, specific blocking agents are also used for treatments or study against metabolic enzyme abnormalities.

Glycolytic rate-limiting enzyme

Hexokinase 2(HK2) is one of the key protein kinases in the glycolysis pathway, which is mainly located in the mitochondrial outer membrane and could regulate the permeability of mitochondrial membrane [104, 105]. HK2 is usually induced to catalyze glucose metabolism in cancer cells and is highly expressed in various tumors, including prostate cancer, liver cancer, gastric cancer (GC), glioblastoma, and breast cancer [105,106,107,108,109,110]. As a sensor in the first step of catalytic gluconeogenesis, HK2 could exert regulatory effects independently of downstream glycolysis reactions. To illustrate, glucose is involved in inducing upregulation of programmed cell death ligand 1 (PD-L1) expression in glioblastoma via HK2 in a dose-dependent manner, and that this induction process is independent of oxygen availability [108]. HK2 also acts as an A-kinase anchoring protein (AKAP) to increase the stability of GSK3 targets, mediating SNAIL glycosylation to promote epithelial-mesenchymal transformation (EMT) in mouse models of BC metastasis, which is independent of glucokinase activity of HK2 [109]. Identification based on the HK subtype showed that most of the non-tumor tissues expressed only HK1, while most of the tumor tissues expressed both HK1 and HK2 [110]. Therefore, the investigation of HK2 and its regulation of the tumor immune microenvironment may remain unclarified.

Another supervisor of the glycolytic pathway, phosphofructokinase 1(PFK1) could catalyze the irreversible conversion of fructose-6 phosphate (F6P) and ATP to fructose-1, 6-diphosphate (F1,6BP) and ADP [111]. PFK1 accelerates glycolysis and the formation of a local acidic microenvironment in tumors [112]. Additionally, extracellular acidity promotes invasion, immunosuppression, and therapeutic resistance [113,114,115]. Activation of phosphofructokinase-1 liver type (PFKL) and inhibition of the pentose phosphate pathway suppresses NOX2-dependent oxidative burst in neutrophils [116]. Transforming growth factor-β (TGF-β) increased PFKL expression and activity during macrophage activation, promoting glycolysis but inhibiting pro-inflammatory cytokine production [117]. Importantly, numerous in vitro studies confirm that administration of high concentrations of citrate, a potent physiological inhibitor of PFK1 and PFK2, reduces ATP production, induces apoptosis, and sensitizes cells to cisplatin treatment [86, 118, 119]. Therefore, an in-depth understanding of the role of PFK1 in the maintenance of immune homeostasis and disease progression is essential to help probe the overall regulation and mutual collaboration between cellular metabolic activities and immune regulation.

Fatty acid synthase (FASN)

Fatty acid synthase (FASN) meets the energy requirements of tumor cells during growth and proliferation by de novo synthesizing of FAs, and promotes various malignant phenotypes in tumors [120]. FASN also connects to cellular metabolism and tumor immunomodulation [121]. As proof, a pan-cancer analysis performed by Zhang et al. showed that the expression level of FASN was significantly negatively correlated with the immune infiltration in 35 tumors and immunotherapeutic targets (including PD-1, PD-L1 and CTLA-4, etc.) in 15 tumors [122]. Synthesis of new FAs mediated by FASN contributes to the functional maturation of Treg cells, and loss of FASN in Treg cells inhibits tumor growth [11]. In addition to reducing fat accumulation in hepatocytes, inhibition of FASN directly suppresses immune cells and stellate cells [123]. Phosphatidylinositol 3-kinase alpha (PI3Kα)-specific inhibitor CYH33 promotes the FA metabolism in TME and ultimately enhances the immune response in combination with the FASN inhibitor C75 [124]. This CYH33-driven process involves preferential M1 polarization of macrophages and increased activity of CD8+ T cells. The above suggests that targeting FASN may improve immunotherapy by altering the local immune microenvironment of tumors.

Lysosomal acid lipase (LAL) & ATP citrate lyase (Acly)

Lysosomal acid lipase (LAL), encoded by the lipase A gene, is the only lysosomal enzyme responsible for catalyzing the hydrolysis of cholesteryl esters and triglycerides at acidic pH [125, 126]. Huang et al. reported that LAL is involved in and ultimately determines M2 activation in macrophages [127]. LAL deficiency was reported to cause systemic expansion and infiltration of myeloid-derived suppressor cells (MDSCs) in multiple organs [128]. In the blood of LAL-deficient (Lal-/-) mice, an increase in CD11c+ cells were observed to be accompanied by an upregulation of PD-L1 expression, which may also cause value-added tumors in the bone marrow of mice [125, 129]. Another study observed suppressed immune rejection and allowed human lung cancer cell growth in Lal-/- mice [130]. Explicitly, O’Sullivan et al. revealed that memory T cells rely on cell-intrinsically expressed LAL to mobilize FA to support FAO and memory T cell development [131].

ATP citrate lyase (Acly), which converts citrate to acetyl-CoA in the cytoplasm, is one of the major enzymes catalyzing the formation of cytosolic acetyl-CoA [90, 132]. Acetyl-CoA synthesized by Acly plays an essential role in mitochondrial metabolic processes such as acetylation and lipid synthesis of various proteins [133,134,135]. Pharmacological analyses report that intracellular acetyl-CoA enhances the therapeutic effect of CD8+ T cells [85]. Toll-like receptor signaling re-mediates macrophage metabolism and promotes histone acetylation via Acly [136]. Acetyl-CoA production was dependent on the glucose transporters GLUT3 and Acly is a promising metabolic checkpoint for alleviating Th17 cell-mediated disease [137]. Furthermore, Acly-dependent histone acetylation promotes hematopoietic stem cell differentiation to CD48+ progenitors [90]. Notably, Acly was actively degraded during the differentiation of in vitro-derived Treg cells, leading to downregulation of FA synthesis to support Treg cell generation [138]. The above results highlight the global impact of LAL or Acly deficiency on metabolic homeostasis and immune cell function, and play a profound role in the metabolic regulation of cellular immunity.

Isocitrate dehydrogenase (IDH)

Isocitrate dehydrogenase (IDH) 1 and 2 (IDH1 and IDH2) are the most frequently mutated metabolic genes in human cancers [139,140,141]. IDH1 could play a key role in lipogenesis and maintenance of redox homeostasis in mammalian hepatocytes [142]. It has been shown that gain-of-function mutations in IDH in human cancers lead to the production of d-2-hydroxyglutarate (d-2HG), a metabolite that promotes tumorigenesis through epigenetic alterations and can alter T-cell metabolism and impair CD8+ T-cell function [143]. Defective metabolism of IDH was identified in M1 macrophages [144]. In addition, enzymatic properties of mutant IDH1 inhibited IFNγ-TET2 signaling and promoted immune escape and tumor viability in cholangiocarcinoma [145]. IDH mutations are highly correlated with the degree of intra-tumor heterogeneity, and IDH mutations produce a paracrine metabolite, (R)-2-hydroxyglutarate, which can be involved in shaping the tumor immune microenvironment [146, 147]. A high proportion of tumor-associated macrophage subpopulations mediating antigen presentation was found in IDH-mutated grade 4 astrocytomas [148]. In a mouse glioma model, treatment with mutant IDH1 reduced levels of the chemokine CXCL10 and inhibited T cell aggregation at the tumor site [149]. In conclusion, information about the association between IDH and immune cell regulation remains to be explored.

Hyaluronidase (HAase)

Hyaluronidase (HAase) on the cell surface hydrolyzes hyaluronic acid (HA) improves fluid permeability in tissues, and is a potent modulator of ER stress resistance [150]. It is generally accepted that HAase significantly improves the efficiency of percutaneous drug delivery and assists local anesthesia in reducing operative pain [151, 152]. HAase plays a pro-cancer role in a variety of cancers [153]. To illustrate, HAase accumulates at sites of inflammation in the body and accelerates the degradation of HA, which is present in high levels in the skin, thereby modulating tumor cell invasion and angiogenesis and protecting against immune cell attack [154]. In addition, HA stimulates the expression of various immune cells at the site of injury [155, 156]. Blair et al. reported that degradation of HA in combination with anti-PD-1 antibody and focal adhesion kinase inhibitor reduced granulocytes [157]. Of interest, Liu et al. developed a nanosystem that evidently increased HAase activity, which synergized with light irradiation could reduce HIF-1α expression and infiltration of immunosuppressive cells in breast cancer model [158]. In terms of clinical translation, HAase-mediated cascade degradation of the stromal barrier and immune cell penetration by microneedles enable efficient anti-tumor therapies [159]. Therefore, HAase is considered a potential target that can modulate immune cell metabolism and mediate immunotherapy that cannot be ignored.

Metabolic checkpoints: trigger targets for changing metabolic manners in distinct immune cell populations

Briefly, the regulatory role of metabolites (or metabolic enzymes) in immune responses was addressed above from the metabolic perspective. Therefore, this section focuses on the molecular targets and signaling pathways that alter the metabolic patterns of these immune cell populations. Referring to the immunologists’ nomenclature of immune checkpoints (targets that regulate autoimmune responses), we roughly named these targets that can change the metabolic pattern of immune cells as “metabolic checkpoints”. We emphasize the impact of metabolic checkpoints on the functional differentiation and fate determination of immune cells, aiming to reveal how metabolic networks mediate the function of specific subtypes of immune cells. Additionally, identifying the metabolic adaptability of different immune cells in specific tissue environments helps us to understand how these cells defend against pathogens and tumors, and how they maintain tissue health at barrier sites [13, 160]. Ultimately, we summarized multiple immunometabolism axes consisting of metabolites, metabolic checkpoints, and immune cell subpopulations (Table 1).

Table 1 Evolving roles alters the immunometabolism of cancer cell survival and growth

T cell

Proverbially, naïve CD8+ T cells differentiate into memory CD8+ T cells through the stages of initial activation, expansion, and sorting of the immune response, a process that is tightly regulated by cell-surface receptors, soluble factors, and transcriptional programs and associated with metabolic reprogramming [168]. As proof, signaling roles for the metabolism of lipid-derived molecules such as prostaglandins in T cell responses have been reported [169]. Scavenger receptor CD36 uptakes lipids and promotes Treg cell function, but inhibits the killing effect of CD8+ T cells in the TME [161, 170]. Interestingly, CD28 co-stimulatory signaling upregulates mammalian target of rapamycin (mTOR) complex 1 (mTORC1), which is activated in a T cell receptor (TCR)-dependent manner, suggesting that the CD28 molecule impacts lipid anabolism and immune responses of effector T cells [171, 172]. The involvement of mTORC1 in the downregulation of CXCR4 and inhibition of bone marrow infiltration of CAR-T cells and elimination of acute myeloid leukemia (AML) may provide a potential reason for the limited efficacy of cellular therapies in AML [173]. However, Werter et al. observed in patients with metastatic renal cell carcinoma (RCC) treated with everolimus that cyclophosphamide attenuates mTOR-mediated regulatory T-cell expansion without affecting clinical outcomes [174]. Likewise, Braun et al. reported that no immune infiltration phenotype was observed to correlate with clinical benefit in 66 patients with advanced RCC (clear cell histology, mTOR inhibition group) [175]. IL-15 signaling drives upregulation of CPT1a expression to promote FAO, whereas IL-7 signaling induces glycerol uptake to promote triacylglycerol synthesis and FAO, thus supporting the longevity of memory CD8+ T cells [167, 176]. Short-chain fatty acids (SCFAs) reconfigure metabolism to allow activated T cells to take up and oxidize more FAs, thereby transforming them into memory CD8+ T cells with long-term viability [177]. Additionally, cholesterol depletion promotes the generation of IL-9-producing CD8+ T cells (Tc9 cells, potent anti-tumor immune inducers) by modulating the activity of the transcription factor liver X receptor (LXR) [163]. Mediated by a metabolite (oxysterol 7α,25-dihydroxycholesterol), EBI2 enhances T follicular helper cell fate by promoting interaction with IL-2-quenched dendritic cells [178]. The LCA derivative 3-oxoLCA inhibits TH17 cell differentiation by binding to the TH17 cell-specific transcription factor RORγt (retinoic acid receptor-associated orphan receptor γt) [179]. In an obesity-associated breast tumor model, STAT3 activation induces FAO in CD8+ T cells and impairs CD8+ T cell effector function [81]. In addition, stimulated TCRs activate PI3K-Akt and the mTOR signaling pathway, which subsequently induces FA and mevalonate synthesis, whereas posttranslational modifications dependent on mevalonate metabolism are essential for Treg cell activation and the establishment of immune tolerance [180, 181]. In short, these studies adequately portray the flexible metabolic plasticity and subpopulation plasticity of T cells.

B cell

B cells are important components of adaptive immunity and the relationship between fate determination of B cells and glucose or glutamine metabolic pathways has received much attention [14, 182]. Metabolic reprogramming of activated B cells has been reported to require the involvement of the sterol regulatory binding protein (SREBP) pathway [183, 184]. CD37 inhibits the FA transporter FATP1 through molecular interactions, which subsequently leads to the inhibition of FA metabolism in aggressive B cell lymphomas [162]. In addition, the findings of Cheng et al. linked cellular metabolism to B cell antigen receptor signaling reveal that fumaric acid inhibits B-cell activation and function by directly inactivating the tyrosine kinase LYN [185]. Naïve B cells treated with 25-hydroxycholesterol inhibit IL-2-mediated B cell proliferation, leading to a significant reduction in IgA [164]. 24-hydroxycholesterol is involved in angiogenesis and in the development of pancreatic neuroendocrine tumors [186]. Oxysterol gradient, produced in lymphoid stromal cells, binds to the upregulated EBI2 receptor on the surface of B cells and promotes the movement of B cells in and out of follicles in response to antigenic stimulation [187]. Sphingosine 1-phosphate (S1P) is a metabolic intermediate of sphingomyelin that functions as a multi-effector lipid mediator in tissues such as the circulatory, nervous, and lymphatic systems [188, 189]. S1P/S1P1 signaling has been reported to help guide the release of nascent immature B cells from the bone marrow into the bloodstream [190, 191]. Significant reductions in germinal center responses, antibody production, mitochondrial mobilization, and OXPHOS have been demonstrated in CD36-deficient B cells [192]. Therefore, understanding B cell metabolic patterns is expected to provide therapeutic targets for B cell-associated immune processes.

Macrophage

In vitro studies performed in the context of pro- or anti-inflammatory activation highlighted the metabolic plasticity of macrophages [15]. Evidence suggests that M1 macrophages prefer to receive activation signals via glycolysis, whereas M2 macrophages favor mitochondrial metabolism and FAO [127, 193]. CD36 molecules, which act as signaling receptors and FA transporters, are known to regulate the metabolism and fate of immune cells, particularly macrophages and T cells [194]. In addition, macrophage Acly deficiency stabilizes atherosclerotic plaques [195]. Previously reported Tissue-resident macrophages (TRMs) are diverse cell families, which generally present long-lived and self-renewing [196, 197]. TRMs are exposed to and adapt to many tissue-specific growth factors and actively participate in cellular metabolism to maintain tissues and organism balance [198, 199]. For example, adipose tissue TRMs contribute to metabolic processes such as insulin sensitivity, adipogenesis, and adaptive thermogenesis [200]. Lack of peroxisome proliferator-activated receptor-γ (PPARγ) targeting or disturbed lipid metabolism in alveolar macrophages leads to pathologic accumulation of surface-active substances in the lungs [201]. Similarly, lowering systemic cholesterol levels with PPARγ agonists, lung X receptor agonists, or statins reduces the pathologic changes of proteolytic disease in mice [202]. Hence, the above reports confirm the importance of variation in metabolism-related targets for functional alterations in macrophages, as well as macrophage regulation of metabolic processes.

Dendritic cell

Differentiation of human monocytes to dendritic cells (DCs) is accompanied by increased expression of PPARγ, a key transcription factor controlling lipid metabolism [203, 204]. The deoxycholic acid derivative 3β-hydroxydeoxycholic acid (isoDCA) inhibits NR1H4 transcriptional activity in DCs and subsequently induces Foxp3 expression [16]. MYC is a transcription factor that promotes the expression of genes encoding proteins in the glycolytic pathway [205]. However, MYC expression is downregulated during DC development with the emergence of MYCL expression in conventional DCs (cDCs) progenitor cells [206]. Resting GM-CSF-induced bone marrow-derived DCs (BMDCs) differ from activated DCs in their weaker ability to interact with and activate T cells. BMDCs have been shown to use FAO to promote OXPHOS [207]. Nevertheless, it is elusive whether resting cDCs or plasmacytoid dendritic cells (pDCs) similarly fuel OXPHOS via FAO. Speaking generally, further explorations regarding the network of interactions between DCs and metabolism allow researchers to achieve a comprehensive understanding of immune metabolism in cancer.

Natural killer cell

The process by which natural killer (NK) cells achieve functional maturation and self-tolerance is known as NK cell education (also known as NK cell licensing), and changes in cellular metabolism are associated with this NK cell education process [208]. Resting mouse NK cells have been shown to have a low basal metabolic rate, maintaining low levels of glycolysis and OXPHOS [209, 210]. Notably, prolonged exposure of human NK cells to IL-15 in vitro results in a reduced metabolic rate [211]. In human NK cells, inhibition of amino acid uptake by SLC1A5 and SLC7A5 prevents IFNγ production as well as degranulation after cross-linking of the activating receptor NKG2D17 antibody [212]. CD36 researchers investigated changes in NK cell function in hyperlipidemic mice, which found that DCs with more lipids in the cytoplasm relied on ROS to increase the expression of PD-L1, TGF-β1, and NKG2D ligands and inhibit NK cell activity [213]. Pre-NK cells (CD11blowCD27hi) undergo a proliferative burst that is associated with the expression of the amino acid transporter SLC3A2 and transferrin receptor [209, 214]. Indeed, NK cells do not use glutamine as a fuel to drive OXPHOS, and inhibition of glutaminase does not inhibit OXPHOS or affect the function of NK cell effectors [210, 215]. Furthermore, whether NK cells use FAs as a fuel source has not been extensively studied. Interestingly, the accumulation of excess FAs in NK cells is thought to be detrimental to NK cell metabolism and function [216]. Not to be overlooked, it is generally accepted that NK cells have long-term functions and are characterized by immune memory [217,218,219]. An important process in the formation and self-renewal of memory NK cells is the restoration of mitochondrial metabolic function (achieved by removing damaged mitochondria through mitochondrial autophagy) [220, 221]. Furthermore, CD16 cross-linking on adaptive NK cells induces stronger mTORC1 activity compared to non-adaptive NK cells [222]. Overall, normal cellular metabolic drives are critical for NK cell development (including NK cell education) and immune function, but research in this area may lead to non-negligible therapeutic opportunities.

Neutrophil

As the most common cell type among leukocytes, neutrophils are considered to be the most abundant innate immune effector cells in the human immune system [160, 223, 224]. Tumor-associated neutrophils (TANs) have become an important part of the tumor microenvironment and play a double-edged role [165, 225, 226]. Under basal conditions, neutrophils predominantly undergo glycolysis with little mitochondria and oxygen exposure deleteriously affects neutrophil viability [227,228,229]. Defects in neutrophil glucose cycling (e.g., G6P transporter deletion) result in reduced glucose uptake and lower intracellular G6P, and also impair energy metabolism [230,231,232]. Arginase 1 (ARG1) blockade in combination with immune checkpoint inhibitors promotes CD8+ T cells in pancreatic ductal adenocarcinoma (PDAC) in vitro [166]. The induction of ER stress in neutrophils upregulates the expression of LOX1, a scavenger receptor involved in lipid metabolism, as well as potent immunosuppressive activity [233]. Neutrophil supply is tightly regulated by three mechanisms: phagocytosis, degranulation, and release of neutrophil extracellular traps (NETs) [160]. Mitochondrial ROS oxidize NET DNA, thereby enhancing its ability to activate the stimulator of interferon genes (STING) signaling and drive IFN production by pDCs [234, 235]. Based on the diversity and plasticity of neutrophil metabolism, it is reasonable to hypothesize that targeting TANs and NETs may become an integral and important component of immunotherapy.

Metabolically driven immunogenic cell death

Cancer immunoediting is the process by which immune cells constrain and promote tumor development through three phases: elimination, homeostasis, and escape [236, 237]. Through these processes, cancer immunogenicity declines as a result of the synergistic action of primary and adaptive immunosuppressive mechanisms. Immunogenic cell death (ICD) is a type of regulatory cell death that is sufficient to activate adaptive immunity in an immunocompetent host [238, 239]. Upon induction of ICD, dying tumor cells release or expose damage-associated molecular patterns (DAMPs). Immunogenic chemotherapy and radiotherapy both upregulate the expression of MHC class I and class II molecules on the surface of tumor cells, thereby enhancing their antigenicity [240, 241]. Of note, increasing numbers of ICD inducers have positively interacted with ICIs or other immunotherapies in cancer patients [242, 243]. Zhou et al. reported that ICD induction enhances anti-tumor immunity and inhibits tumor immune evasion through CD47 blockade, which may be expected to improve cancer chemoimmunotherapy [244]. Based on the ability to trigger cancer cell death and danger perception, ICD inducers can be categorized into two types, including type I (generating reactive oxygen species) and type II (inducing endoplasmic reticulum stress) inducers [245, 246]. Doxorubicin-induced ICD is caspase-dependent, and both doxorubicin and mitoxantrone induce tumor cells to expose CRT, secrete ATP, and release HMGB1 [247,248,249,250]. From the perspective of cancer immunotherapy, the exploration of the characterization of metabolism-associated ICD, the underlying cell biology, and the pathways by which immune effector cells sense ICD will be one of the important plates in future clinical strategies.

Immunoediting-driven metabolic adaptations in cancers

Cancer-immunity cycle and cancer-immunometabolism subcycle

Effective anti-tumor immune responses must initiate a series of step events and cycle back and forth, which are termed “cancer-immunity cycle” (CI cycle) [251, 252]. The left circle of Fig. 4 delineates the warrant steps involved in the CI cycle. Briefly, tumor antigens are presented by DCs and recognized by effector T cells to kill the tumor cells, and the killed tumor cells release more tumor-associated antigens to further promote the breadth and depth of the immune response [252]. In this cycle, the balance of the ratio of T effector cells to Treg cells is critical to the outcome. However, in cancer patients, the CI cycle often does not work optimally [252]. Namely, tumor antigens may not be detected or effectively activate DCs, DCs may recognize antigens as self-antigens and subsequently escape immune surveillance, and T cells may not properly home to the tumor bed for killing [253].

Fig. 4
figure 4

Disruptor of the virtuous cycle: the cancer-immunometabolism subcycle. In fact, antigen release occurs consistently in most patients with malignant tumors. Nevertheless, the CI cycle of Chen et al. suggests that it does not imply that the inevitable occurrence of cancer cell death events (the left cycle). Metabolic-related factors may be responsible for the disruption of the CI cycle. Considering the cancer-immunometabolism subcycle (the right cycle) proposed in this review, it is reasonable to assume that metabolism-related factors may contribute to the interruption of the CI cycle (the left cycle). When immune cells reach the tumor microenvironment through the vascular endothelium or basement membrane, nutrient deprivation as well as accumulation of local toxic substances accelerate the formation of tumor immunosuppressive microenvironment. Thereby, the infiltrating immune cells become dysfunctional, such as altered macrophage polarization, diminished killing effect of T cells and NK cells, and formation of NETs, and so on. As the result, the CI cycle is impaired and failed to stimulate a potent and sustainable immune response. CI cycle, cancer-immunity cycle; NK cells, natural killer cells; NETs, neutrophil extracellular traps

Nonetheless, metabolic disorders in cancer cells further create a vicious circle by creating a microenvironment that contains tumor metabolites conducive to cancer cell growth [254]. Scientists from Switzerland and other institutions have found that immune cell surveillance of cancer may itself induce metabolic adaptations in early-stage tumor cells, while also promoting their growth and giving them the ability to suppress the body’s lethal immune response [255]. In a workflow of influences on CD8+ T cell differentiation in cancer, Giles et al. argued that metabolism should be incorporated as a fourth signal to better execute the CI cycle [256]. In light of this, we propose the scenario of the cancer-immunometabolism subcycle (Fig. 4, the right circle). Filtrating the fundamental features of the CI cycle will help us to accurately characterize our understanding of the cancer immune response, and these insights will have profound implications for the establishment and application of immunometabolism.

Immunotherapy and precision medicine drug development

In oncology, immunotherapy and precision medicine drug development are in the ascendant. In this light, only about one-third of patients respond to immunotherapy, with the type of immunity playing a decisive role [251]. Table 2 briefly list the clinical prospects explored under several metabolic studies. VEGF blockade combined with tumor-derived glutamate has been found to induce systemic and intra-tumoral immunosuppression, and this effect can be prevented by Treg depletion, thereby enhancing anti-tumor efficacy [9]. Of interest, JHU083 was able to synergize with immunotherapy to enhance infiltration, proliferation, activation, and function of effector T cells in tumors [52]. Reducing GCPII expression through genetic alterations or pharmacological inhibition of glutamate carboxypeptidase II (GCPII) leads to reductions in glutamate concentration and tumor growth, which are enhanced by targeting GCPII in combination with glutaminase inhibition [257]. The combination of the drug with the lipoprotein transport pathway allows optimized lipophilic parent drug to be transported via the lymph, reducing ineffective exposure to the drug and subsequently enhancing efficacy [82, 258]. In addition, it may also target specific disease reservoirs in lymphatic vessels, providing advantages for advanced immunotherapeutic cancer strategies [259].

Table 2 Metabolic perspectives in immunotherapy and drug development

Given the Leu nutritional preference exhibited by leucine-tRNA-synthase-2-expressing B (LARS B) cells, Wang et al. proposed a leucine dieting regimen, which is considered to be a favorable option for colorectal cancer treatment [44]. In an allogeneic hematopoietic cell transplantation model, ceramide synthase 6-deficient T-cell proliferation and IFN-γ production were blocked, suggesting that targeting ceramide synthesis is expected to improve allogeneic hematopoietic cell transplantation therapy [264]. Additionally, the differentiation and survival of human monocyte-derived DCs were impaired by rapamycin [260, 261]. Metabolic inhibitor 2-deoxyglucose (2DG) limits glycolysis and OXPHOS and inhibits IFN-γ production and granzyme B expression in mouse and human NK cells [262, 263]. Neutrophil extracellular traps accumulate in the peripheral vasculature of tumor-bearing animals and impair organ function, and treatment with the autophagy-based inhibitor chloroquine blocks peripheral infiltration of neutrophils [265]. Currently, while next-generation checkpoint inhibitors may provide some benefit, it seems unlikely that they alone will overcome the hurdles specific to the CI cycle and immunotherapy. On the road to exploring immunotherapy, “metabolic checkpoints” also require more attention.

Pending challenges and clinical concerns

Based on the Warburg effect being observed in a variety of solid tumors, oncologists have been trying to reduce glucose utilization in tumors as a treatment for decades [110]. However, relevant attempts have not yet reached the stage of clinical application, such as attempts with hexokinase inhibitors. However, immunotherapy for tumors is on the rise. Herein, we present the following topics along with clinical issues that we hope will overcome in the future. (i) In the crosstalk of tumor metabolism and immune regulation, how do some key pathways, such as ROS signaling, TCR signaling, and co-stimulatory signaling, regulate each other, what are the important targets, are the differences between individuals significant, and are there clinical transformations (such as nutritional interventions) accessible? (ii) in fact, some carcinogenic signaling pathways, such as HIF-1α and RAS/PI3K/AKT, may generate drug resistance by enhancing aerobic glycolysis in cancer cells (the “Warburg effect”) [86]. This metabolic pattern that promotes cancer cell development and invasiveness also induces enhanced extracellular acidity, whereas most drugs are weakly basic molecules. On the one hand, citrate could act as a physiological inhibitor of PFK1 and PFK2, inhibiting glycolysis; On the other hand, citrate leads to acid-base imbalance in the extracellular environment, weakening the effective infiltration concentration of alkaline drug molecules in the local tumor microenvironment, and even triggering multi-drug resistance events. For example, sodium bicarbonate has been observed to increase responsiveness to immunotherapy in models of melanoma and pancreatic cancer [266]. Therefore, the effect of the disturbance of acid-base balance in the local microenvironment caused by metabolism on immunotherapy is still worth exploring. (iii) In a series of enzymatic reactions of immune metabolism, are there key enzymatic molecules that finely and efficiently regulate immune surveillance and killing effects, produce a dominant effect on the tumorigenesis and progression (exerting a similar role as “rate-limiting enzymes”)? (iv) How to determine the difference between metabolic studies based on tumor animal models (such as mice) and patients’ tumor microenvironment metabolism? (v) Metabolic enzymes maintain the metabolic balance of the whole body in a normal body. For instance, inhibitors of FASN have shown severe systemic side effects such as weight loss and anorexia in some clinical studies [267]. Therefore, how can strategies be developed to target dysregulated metabolism in cancer through nutritional interventions, and thus improve anti-tumor immunity? How far is a personalized treatment strategy for tumor metabolic inhibitors from clinical patients? (vi) Metabolic changes caused by aging have brought about a series of changes in pathophysiological processes, and the immune state of the body is inevitably affected. Which metabolic changes ultimately lead to an increased risk of cancer during this process? Can this condition be avoided or slowed down with age? (vii) Long-term accumulation of abnormal metabolism or sudden but intense metabolic changes, which has a greater impact on tumors, the former or the latter?

Concluding remarks

Undoubtedly, metabolic reprogramming is not unique to tumor cells, and immune cells share this feature. Immunology and oncology investigators are increasingly aware that different stages of immune cell activation coincide with different types of cellular metabolism. We hope to initiate or restart anti-tumor immune responses and maintain self-circulation while ensuring that unrestrained autoimmune inflammation is avoided. Herein, we have learned that there are checkpoints and inhibitors at every step of metabolism and immunity that impede the anti-tumor response from proceeding and expanding further, and that effective approaches are selective processing of patients at different limiting steps to adapt to tumor evolution. Immunophenotype variability and plasticity have profoundly inspired researchers to explore new directions for cancer treatment. Likewise, as is currently understood in the era of tumor immunology, it is reasonable to assume that metabolism is far less accurately and comprehensively delineated than it is perceived to be. The concept of immunometabolism could assist in the development of efficient treatment options, as well as the understanding of immune regulatory mechanisms focusing on metabolic perspectives will bring out a profound impact on the design of clinical therapies.

Availability of data and materials

No datasets were generated or analysed during the current study.

Abbreviations

ATP:

Adenosine 5’-triphosphate (ATP)

ROS:

Reactive oxygen species

Leu:

Leucine

Arg:

Arginine

TME:

Tumor microenvironment

Glu:

Glutamate

Gln:

Glutamine

CNS:

Central nervous system

Trp:

Tryptophan

Asn:

Asparagine

IDO:

Indoleamine 2,3-dioxygenase

Kyn:

Kynurenine

IFN-γ:

Interferon-gamma

MCs:

Mesenchymal cells

NK:

Natural killer

ER:

Endoplasmic reticulum

ACC1:

Acetyl-CoA carboxylase 1

AMPK:

AMP-activated protein kinase

CSCs:

Cancer stem cells

αKG:

α-ketoglutarate

cDC1:

Conventional type 1 dendritic cell

HK2:

Hexokinase 2

GC:

Gastric cancer

PD-L1:

Programmed cell death ligand 1

AKAP:

A-kinase anchoring protein

EMT:

Epithelial-mesenchymal transformation

PFK1:

Phosphofructokinase 1

F6P:

Fructose-6 phosphate

F1,6BP:

Fructose-1, 6-diphosphate

PFKL:

Phosphofructokinase-1 liver type

TGF-β:

Transforming growth factor-β

FA:

Fatty acid

FASN:

Fatty acid synthase

FAO:

Fatty acid oxidation

PI3Kα:

Phosphatidylinositol 3-kinase alpha

LAL:

Lysosomal acid lipase

MDSCs:

Myeloid-derived suppressor cells

PC:

Pyruvate carboxylase

Acly:

ATP citrate lyase

IDH:

Isocitrate dehydrogenase

d-2HG:

d-2-hydroxyglutarate

HAase:

Hyaluronidase

HA:

Hyaluronic acid

Mtor:

Mammalian target of rapamycin

mTORC1:

mammalian target of rapamycin complex 1

TCR:

T cell receptor

AML:

Acute myeloid leukemia

RCC:

Renal cell carcinoma

SCFAs:

Short-chain fatty acids

LXR:

Liver X receptor

SREBP:

Sterol regulatory binding protein

Treg cell:

T regulatory cell

S1P:

Sphingosine 1-phosphate

TRMs:

Tissue-resident macrophages

DCs:

Dendritic cells

PPARγ:

Peroxisome proliferator-activated receptor-γ

cDCs:

conventional DCs

pDCs:

plasmacytoid dendritic cells

NSCLC:

Non-small cell lung cancer

BMDCs:

Bone marrow-derived DCs

isoDCA:

3β-hydroxydeoxycholic acid

TANs:

Tumor-associated neutrophils

NETs:

Neutrophil extracellular traps

ARG1:

Arginase 1

PDAC:

Pancreatic ductal adenocarcinoma

pDCs:

plasmacytoid dendritic cells

STING:

Stimulator of interferon genes

ICD:

Immunogenic cell death

DAMPs:

Damage-associated molecular patterns

CI cycle:

Cancer-immunity cycle

GCPII:

Glutamate carboxypeptidase II

LARS B cells:

Leucine-tRNA-synthase-2-expressing B cells

2DG:

2-deoxyglucose

References

  1. Warburg O, Posener K, Negelein E. On the metabolism of carcinoma cells [J]. Biochem Z. 1924;152:309–44.

    CAS  Google Scholar 

  2. Koppenol WH, Bounds PL, Dang CV. Otto Warburg’s contributions to current concepts of cancer metabolism [J]. Nat Rev Cancer. 2011;11(5):325–37.

    Article  CAS  PubMed  Google Scholar 

  3. Dey P, Kimmelman AC, Depinho RA. Metabolic Codependencies in the Tumor Microenvironment [J]. Cancer Discov. 2021;11(5):1067–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Yang Z, Yan C, Ma J, et al. Lactylome analysis suggests lactylation-dependent mechanisms of metabolic adaptation in hepatocellular carcinoma [J]. Nat Metab. 2023;5(1):61–79.

    Article  CAS  PubMed  Google Scholar 

  5. Morrissey SM, Zhang F, Ding C, et al. Tumor-derived exosomes drive immunosuppressive macrophages in a pre-metastatic niche through glycolytic dominant metabolic reprogramming [J]. Cell Metab. 2021;33(10):2040–58 e10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Harris IS, Denicola GM. The complex interplay between antioxidants and ROS in cancer [J]. Trends Cell Biol. 2020;30(6):440–51.

    Article  CAS  PubMed  Google Scholar 

  7. Boroughs LK, Deberardinis RJ. Metabolic pathways promoting cancer cell survival and growth [J]. Nat Cell Biol. 2015;17(4):351–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Martinez-Reyes I, Chandel NS. Cancer metabolism: looking forward [J]. Nat Rev Cancer. 2021;21(10):669–80.

    Article  CAS  PubMed  Google Scholar 

  9. Long Y, Tao H, Karachi A, et al. Dysregulation of glutamate transport enhances Treg function that promotes VEGF blockade resistance in glioblastoma [J]. Cancer Res. 2020;80(3):499–509.

    Article  CAS  PubMed  Google Scholar 

  10. Nakaya M, Xiao Y, Zhou X, et al. Inflammatory T cell responses rely on amino acid transporter ASCT2 facilitation of glutamine uptake and mTORC1 kinase activation [J]. Immunity. 2014;40(5):692–705.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Lim SA, Wei J, Nguyen TM, et al. Lipid signalling enforces functional specialization of T(reg) cells in tumours [J]. Nature. 2021;591(7849):306–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Chen T, Xu ZG, Luo J, et al. NSUN2 is a glucose sensor suppressing cGAS/STING to maintain tumorigenesis and immunotherapy resistance [J]. Cell Metab. 2023;35(10):1782–98 e8.

    Article  CAS  PubMed  Google Scholar 

  13. Reina-Campos M, Scharping NE, Goldrath AW. CD8(+) T cell metabolism in infection and cancer [J]. Nat Rev Immunol. 2021;21(11):718–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Nemazee D. Mechanisms of central tolerance for B cells [J]. Nat Rev Immunol. 2017;17(5):281–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Ryan DG, O’Neill LAJ. Krebs cycle reborn in macrophage immunometabolism [J]. Annu Rev Immunol. 2020;38:289–313.

    Article  CAS  PubMed  Google Scholar 

  16. Campbell C, McKenney PT, Konstantinovsky D, et al. Bacterial metabolism of bile acids promotes generation of peripheral regulatory T cells [J]. Nature. 2020;581(7809):475–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Guerra L, Bonetti L, Brenner D. Metabolic modulation of immunity: a new concept in cancer immunotherapy [J]. Cell Rep. 2020;32(1):107848.

    Article  CAS  PubMed  Google Scholar 

  18. Jung J, Zeng H, Horng T. Metabolism as a guiding force for immunity [J]. Nat Cell Biol. 2019;21(1):85–93.

    Article  CAS  PubMed  Google Scholar 

  19. Chang CH, Qiu J, O’Sullivan D, et al. Metabolic competition in the tumor microenvironment is a driver of cancer progression [J]. Cell. 2015;162(6):1229–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Terry S, Engelsen AST, Buart S, et al. Hypoxia-driven intratumor heterogeneity and immune evasion [J]. Cancer Lett. 2020;492:1–10.

    Article  CAS  PubMed  Google Scholar 

  21. Huang B, Song BL, Xu C. Cholesterol metabolism in cancer: mechanisms and therapeutic opportunities [J]. Nat Metab. 2020;2(2):132–41.

    Article  PubMed  Google Scholar 

  22. Jin J, Zhao Q, Wei Z, et al. Glycolysis-cholesterol metabolic axis in immuno-oncology microenvironment: emerging role in immune cells and immunosuppressive signaling [J]. Cell Biosci. 2023;13(1):189.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Baek AE, Yu YA, He S, et al. The cholesterol metabolite 27 hydroxycholesterol facilitates breast cancer metastasis through its actions on immune cells [J]. Nat Commun. 2017;8(1):864.

    Article  PubMed  PubMed Central  Google Scholar 

  24. Gauci ML, Lanoy E, Champiat S, et al. Long-term survival in patients responding to Anti-PD-1/PD-L1 therapy and disease outcome upon treatment discontinuation [J]. Clin Cancer Res. 2019;25(3):946–56.

    Article  PubMed  Google Scholar 

  25. Shyer JA, Flavell RA, Bailis W. Metabolic signaling in T cells [J]. Cell Res. 2020;30(8):649–59.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Artyomov MN, van den Bossche J. Immunometabolism in the single-cell era [J]. Cell Metab. 2020;32(5):710–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Deretic V. Autophagy in inflammation, infection, and immunometabolism [J]. Immunity. 2021;54(3):437–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. O’Neill LA, Kishton RJ, Rathmell J. A guide to immunometabolism for immunologists [J]. Nat Rev Immunol. 2016;16(9):553–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Vrieling F, Stienstra R. Obesity and dysregulated innate immune responses: impact of micronutrient deficiencies [J]. Trends Immunol. 2023;44(3):217–30.

    Article  CAS  PubMed  Google Scholar 

  30. Xue C, Li G, Zheng Q, et al. Tryptophan metabolism in health and disease [J]. Cell Metab. 2023;35(8):1304–26.

    Article  CAS  PubMed  Google Scholar 

  31. Zaslona Z, O’Neill LAJ. Cytokine-like roles for metabolites in immunity [J]. Mol Cell. 2020;78(5):814–23.

    Article  CAS  PubMed  Google Scholar 

  32. Haas R, Cucchi D, Smith J, et al. Intermediates of metabolism: from bystanders to signalling molecules [J]. Trends Biochem Sci. 2016;41(5):460–71.

    Article  CAS  PubMed  Google Scholar 

  33. Nagata N, Takeuchi T, Masuoka H, et al. Human gut microbiota and its metabolites impact immune responses in COVID-19 and its complications [J]. Gastroenterology. 2023;164(2):272–88.

    Article  CAS  PubMed  Google Scholar 

  34. Ryan DG, Murphy MP, Frezza C, et al. Coupling Krebs cycle metabolites to signalling in immunity and cancer [J]. Nat Metab. 2019;1:16–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Zhang Z, Li X, Yang F, et al. DHHC9-mediated GLUT1 S-palmitoylation promotes glioblastoma glycolysis and tumorigenesis [J]. Nat Commun. 2021;12(1):5872.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Klein Geltink RI, Edwards-Hicks J, Apostolova P, et al. Metabolic conditioning of CD8(+) effector T cells for adoptive cell therapy [J]. Nat Metab. 2020;2(8):703–16.

    Article  CAS  PubMed  Google Scholar 

  37. He J, Shangguan X, Zhou W, et al. Glucose limitation activates AMPK coupled SENP1-Sirt3 signalling in mitochondria for T cell memory development [J]. Nat Commun. 2021;12(1):4371.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Alsheikh HAM, Metge BJ, Ha CM, et al. Normalizing glucose levels reconfigures the mammary tumor immune and metabolic microenvironment and decreases metastatic seeding [J]. Cancer Lett. 2021;517:24–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Afonso J, Santos LL, Longatto-Filho A, et al. Competitive glucose metabolism as a target to boost bladder cancer immunotherapy [J]. Nat Rev Urol. 2020;17(2):77–106.

    Article  CAS  PubMed  Google Scholar 

  40. Lieu EL, Nguyen T, Rhyne S, et al. Amino acids in cancer [J]. Exp Mol Med. 2020;52(1):15–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Ling ZN, Jiang YF, Ru JN, et al. Amino acid metabolism in health and disease [J]. Signal Transduct Target Ther. 2023;8(1):345.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Peng H, Wang Y, Luo W. Multifaceted role of branched-chain amino acid metabolism in cancer [J]. Oncogene. 2020;39(44):6747–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Faber J, Berkhout M, Fiedler U, et al. Rapid EPA and DHA incorporation and reduced PGE2 levels after one week intervention with a medical food in cancer patients receiving radiotherapy, a randomized trial [J]. Clin Nutr. 2013;32(3):338–45.

    Article  CAS  PubMed  Google Scholar 

  44. Wang Z, Lu Z, Lin S, et al. Leucine-tRNA-synthase-2-expressing B cells contribute to colorectal cancer immunoevasion [J]. Immunity. 2022;55(6):1067–81 e8.

    Article  CAS  PubMed  Google Scholar 

  45. Manfroi B, Fillatreau S. Regulatory B cells gain muscles with a leucine-rich diet [J]. Immunity. 2022;55(6):970–2.

    Article  CAS  PubMed  Google Scholar 

  46. Matos A, Carvalho M, Bicho M, et al. Arginine and arginases modulate metabolism, tumor microenvironment and prostate cancer progression [J]. Nutrients. 2021;13(12):4503.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Geiger R, Rieckmann JC, Wolf T, et al. L-Arginine modulates T cell metabolism and enhances survival and anti-tumor activity [J]. Cell. 2016;167(3):829–42 e13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Steggerda SM, Bennett MK, Chen J, et al. Inhibition of arginase by CB-1158 blocks myeloid cell-mediated immune suppression in the tumor microenvironment [J]. J Immunother Cancer. 2017;5(1):101.

    Article  PubMed  PubMed Central  Google Scholar 

  49. Miret JJ, Kirschmeier P, Koyama S, et al. Suppression of myeloid cell arginase activity leads to therapeutic response in a NSCLC mouse model by activating anti-tumor immunity [J]. J Immunother Cancer. 2019;7(1):32.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Magi S, Piccirillo S, Amoroso S. The dual face of glutamate: from a neurotoxin to a potential survival factor-metabolic implications in health and disease [J]. Cell Mol Life Sci. 2019;76(8):1473–88.

    Article  CAS  PubMed  Google Scholar 

  51. Cui L, Guo J, Cranfill SL, et al. Glutamate in primary afferents is required for itch transmission [J]. Neuron. 2022;110(5):809–23 e5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Leone RD, Zhao L, Englert JM, et al. Glutamine blockade induces divergent metabolic programs to overcome tumor immune evasion [J]. Science. 2019;366(6468):1013–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Edwards DN, Ngwa VM, Raybuck AL, et al. Selective glutamine metabolism inhibition in tumor cells improves antitumor T lymphocyte activity in triple-negative breast cancer [J]. J Clin Invest. 2021;131(4):e140100.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Platten M, Nollen EAA, Rohrig UF, et al. Tryptophan metabolism as a common therapeutic target in cancer, neurodegeneration and beyond [J]. Nat Rev Drug Discov. 2019;18(5):379–401.

    Article  CAS  PubMed  Google Scholar 

  55. Montgomery TL, Eckstrom K, Lile KH, et al. Lactobacillus reuteri tryptophan metabolism promotes host susceptibility to CNS autoimmunity [J]. Microbiome. 2022;10(1):198.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Chen C, Hou G, Zeng C, et al. Metabolomic profiling reveals amino acid and carnitine alterations as metabolic signatures in psoriasis [J]. Theranostics. 2021;11(2):754–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Qin R, Zhao C, Wang CJ, et al. Tryptophan potentiates CD8(+) T cells against cancer cells by TRIP12 tryptophanylation and surface PD-1 downregulation [J]. J Immunother Cancer. 2021;9(7):e002840.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Greene LI, Bruno TC, Christenson JL, et al. A Role for Tryptophan-2,3-dioxygenase in CD8 T-cell Suppression and Evidence of Tryptophan Catabolism in Breast Cancer Patient Plasma [J]. Mol Cancer Res. 2019;17(1):131–9.

    Article  CAS  PubMed  Google Scholar 

  59. Friedrich M, Sankowski R, Bunse L, et al. Tryptophan metabolism drives dynamic immunosuppressive myeloid states in IDH-mutant gliomas [J]. Nat Cancer. 2021;2(7):723–40.

    Article  CAS  PubMed  Google Scholar 

  60. Bender MJ, McPherson AC, Phelps CM, et al. Dietary tryptophan metabolite released by intratumoral Lactobacillus reuteri facilitates immune checkpoint inhibitor treatment [J]. Cell. 2023;186(9):1846–622 e6.

    Article  CAS  PubMed  Google Scholar 

  61. Fong W, Li Q, Ji F, et al. Lactobacillus gallinarum-derived metabolites boost anti-PD1 efficacy in colorectal cancer by inhibiting regulatory T cells through modulating IDO1/Kyn/AHR axis [J]. Gut. 2023;72(12):2272–85.

    Article  CAS  PubMed  Google Scholar 

  62. Siska PJ, Jiao J, Matos C, et al. Kynurenine induces T cell fat catabolism and has limited suppressive effects in vivo [J]. EBioMedicine. 2021;74:103734.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Xue C, Gu X, Zheng Q, et al. Effects of 3-HAA on HCC by Regulating the Heterogeneous Macrophages-A scRNA-Seq Analysis [J]. Adv Sci (Weinh). 2023;10(16):e2207074.

    Article  PubMed  Google Scholar 

  64. Bessede A, Peyraud F, le Moulec S, et al. Upregulation of indoleamine 2,3-dioxygenase 1 in tumor cells and tertiary lymphoid structures is a hallmark of inflamed non-small cell lung cancer [J]. Clin Cancer Res. 2023;29(23):4883–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Wu J, Li G, Li L, et al. Asparagine enhances LCK signalling to potentiate CD8(+) T-cell activation and anti-tumour responses [J]. Nat Cell Biol. 2021;23(1):75–86.

    Article  CAS  PubMed  Google Scholar 

  66. Gnanaprakasam JNR, Kushwaha B, Liu L, et al. Asparagine restriction enhances CD8(+) T cell metabolic fitness and antitumoral functionality through an NRF2-dependent stress response [J]. Nat Metab. 2023;5(8):1423–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Gong Z, Li Q, Shi J, et al. Lipid-laden lung mesenchymal cells foster breast cancer metastasis via metabolic reprogramming of tumor cells and natural killer cells [J]. Cell Metab. 2022;34(12):1960–76 e9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Ecker C, Guo L, Voicu S, et al. Differential reliance on lipid metabolism as a salvage pathway underlies functional differences of T cell subsets in poor nutrient environments [J]. Cell Rep. 2018;23(3):741–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. King RJ, Singh PK, Mehla K. The cholesterol pathway: impact on immunity and cancer [J]. Trends Immunol. 2022;43(1):78–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Yang W, Bai Y, Xiong Y, et al. Potentiating the antitumour response of CD8(+) T cells by modulating cholesterol metabolism [J]. Nature. 2016;531(7596):651–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Pistollato F, Forbes-Hernandez TY, Iglesias RC, et al. Effects of caloric restriction on immunosurveillance, microbiota and cancer cell phenotype: possible implications for cancer treatment [J]. Semin Cancer Biol. 2021;73:45–57.

    Article  CAS  PubMed  Google Scholar 

  72. Ma X, Bi E, Lu Y, et al. Cholesterol induces CD8(+) T cell exhaustion in the tumor microenvironment [J]. Cell Metab. 2019;30(1):143–56 e5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Koundouros N, Poulogiannis G. Reprogramming of fatty acid metabolism in cancer [J]. Br J Cancer. 2020;122(1):4–22.

    Article  CAS  PubMed  Google Scholar 

  74. Xu H, Chen Y, Gu M, et al. Fatty acid metabolism reprogramming in advanced prostate cancer [J]. Metabolites. 2021;11(11):765.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Berod L, Friedrich C, Nandan A, et al. De novo fatty acid synthesis controls the fate between regulatory T and T helper 17 cells [J]. Nat Med. 2014;20(11):1327–33.

    Article  CAS  PubMed  Google Scholar 

  76. Endo Y, Asou HK, Matsugae N, et al. Obesity drives Th17 cell differentiation by inducing the lipid metabolic kinase, ACC1 [J]. Cell Rep. 2015;12(6):1042–55.

    Article  CAS  PubMed  Google Scholar 

  77. Endo Y, Onodera A, Obata-Ninomiya K, et al. ACC1 determines memory potential of individual CD4(+) T cells by regulating de novo fatty acid biosynthesis [J]. Nat Metab. 2019;1(2):261–75.

    Article  CAS  PubMed  Google Scholar 

  78. Pearce EL, Walsh MC, Cejas PJ, et al. Enhancing CD8 T-cell memory by modulating fatty acid metabolism [J]. Nature. 2009;460(7251):103–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Da BorgesSilva H, Beura LK, Wang H, et al. The purinergic receptor P2RX7 directs metabolic fitness of long-lived memory CD8(+) T cells [J]. Nature. 2018;559(7713):264–8.

    Article  Google Scholar 

  80. Grajchen E, Loix M, Baeten P, et al. Fatty acid desaturation by stearoyl-CoA desaturase-1 controls regulatory T cell differentiation and autoimmunity [J]. Cell Mol Immunol. 2023;20(6):666–79.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Zhang C, Yue C, Herrmann A, et al. STAT3 activation-induced fatty acid oxidation in CD8(+) T effector cells is critical for obesity-promoted breast tumor growth [J]. Cell Metab. 2020;31(1):148–61 e5.

    Article  PubMed  Google Scholar 

  82. Yanez JA, Wang SW, Knemeyer IW, et al. Intestinal lymphatic transport for drug delivery [J]. Adv Drug Deliv Rev. 2011;63(10–11):923–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Lee JB, Zgair A, Malec J, et al. Lipophilic activated ester prodrug approach for drug delivery to the intestinal lymphatic system [J]. J Control Release. 2018;286:10–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Li Y, Li YC, Liu XT, et al. Blockage of citrate export prevents TCA cycle fragmentation via Irg1 inactivation [J]. Cell Rep. 2022;38(7):110391.

    Article  CAS  PubMed  Google Scholar 

  85. Chowdhury S, Kar A, Bhowmik D, et al. Intracellular acetyl CoA potentiates the therapeutic efficacy of antitumor CD8+ T cells [J]. Cancer Res. 2022;82(14):2640–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Icard P, Simula L, Wu Z, et al. Why may citrate sodium significantly increase the effectiveness of transarterial chemoembolization in hepatocellular carcinoma? [J]. Drug Resist Updat. 2021;59:100790.

    Article  CAS  PubMed  Google Scholar 

  87. Ding C, Yi X, Chen X, et al. Warburg effect-promoted exosomal circ_0072083 releasing up-regulates NANGO expression through multiple pathways and enhances temozolomide resistance in glioma [J]. J Exp Clin Cancer Res. 2021;40(1):164.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Catapano J, Luty M, Wrobel T, et al. Acquired drug resistance interferes with the susceptibility of prostate cancer cells to metabolic stress [J]. Cell Mol Biol Lett. 2022;27(1):100.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Morrow MR, Batchuluun B, Wu J, et al. Inhibition of ATP-citrate lyase improves NASH, liver fibrosis, and dyslipidemia [J]. Cell Metab. 2022;34(6):919–36 e8.

    Article  CAS  PubMed  Google Scholar 

  90. Umemoto T, Johansson A, Ahmad SAI, et al. ATP citrate lyase controls hematopoietic stem cell fate and supports bone marrow regeneration [J]. EMBO J. 2022;41(8):e109463.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Liu J, Peng Y, Shi L, et al. Skp2 dictates cell cycle-dependent metabolic oscillation between glycolysis and TCA cycle [J]. Cell Res. 2021;31(1):80–93.

    Article  CAS  PubMed  Google Scholar 

  92. Martinez-Reyes I, Diebold LP, Kong H, et al. TCA cycle and mitochondrial membrane potential are necessary for diverse biological functions [J]. Mol Cell. 2016;61(2):199–209.

    Article  CAS  PubMed  Google Scholar 

  93. Mills EL, Kelly B, Logan A, et al. Succinate dehydrogenase supports metabolic repurposing of mitochondria to drive inflammatory macrophages [J]. Cell. 2016;167(2):457–70 e13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Hubert S, Rissiek B, Klages K, et al. Extracellular NAD+ shapes the Foxp3+ regulatory T cell compartment through the ART2-P2X7 pathway [J]. J Exp Med. 2010;207(12):2561–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Chen X, Sunkel B, Wang M, et al. Succinate dehydrogenase/complex II is critical for metabolic and epigenetic regulation of T cell proliferation and inflammation [J]. Sci Immunol. 2022;7(70):eabm8161.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Liu PS, Wang H, Li X, et al. alpha-ketoglutarate orchestrates macrophage activation through metabolic and epigenetic reprogramming [J]. Nat Immunol. 2017;18(9):985–94.

    Article  CAS  PubMed  Google Scholar 

  97. Bottcher JP, Bonavita E, Chakravarty P, et al. NK cells stimulate recruitment of cdc1 into the tumor microenvironment promoting cancer immune control [J]. Cell. 2018;172(5):1022–37 e14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Bonavita E, Bromley CP, Jonsson G, et al. Antagonistic inflammatory phenotypes dictate tumor fate and response to immune checkpoint blockade [J]. Immunity. 2020;53(6):1215–29 e8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Watson MJ, Vignali PDA, Mullett SJ, et al. Metabolic support of tumour-infiltrating regulatory T cells by lactic acid [J]. Nature. 2021;591(7851):645–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Feng Q, Liu Z, Yu X, et al. Lactate increases stemness of CD8 + T cells to augment anti-tumor immunity [J]. Nat Commun. 2022;13(1):4981.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Johnston RJ, Su LJ, Pinckney J, et al. VISTA is an acidic pH-selective ligand for PSGL-1 [J]. Nature. 2019;574(7779):565–70.

    Article  CAS  PubMed  Google Scholar 

  102. Elia I, Rowe JH, Johnson S, et al. Tumor cells dictate anti-tumor immune responses by altering pyruvate utilization and succinate signaling in CD8(+) T cells [J]. Cell Metab. 2022;34(8):1137–50 e6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Pan C, Li B, Simon MC. Moonlighting functions of metabolic enzymes and metabolites in cancer [J]. Mol Cell. 2021;81(18):3760–74.

    Article  CAS  PubMed  Google Scholar 

  104. Hinrichsen F, Hamm J, Westermann M, et al. Microbial regulation of hexokinase 2 links mitochondrial metabolism and cell death in colitis [J]. Cell Metab. 2021;33(12):2355–66 e8.

    Article  CAS  PubMed  Google Scholar 

  105. Shangguan X, He J, Ma Z, et al. SUMOylation controls the binding of hexokinase 2 to mitochondria and protects against prostate cancer tumorigenesis [J]. Nat Commun. 2021;12(1):1812.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Jiao L, Zhang HL, Li DD, et al. Regulation of glycolytic metabolism by autophagy in liver cancer involves selective autophagic degradation of HK2 (hexokinase 2) [J]. Autophagy. 2018;14(4):671–84.

    Article  CAS  PubMed  Google Scholar 

  107. Wang J, Huang Q, Hu X, et al. Disrupting circadian rhythm via the PER1-HK2 axis reverses trastuzumab resistance in gastric cancer [J]. Cancer Res. 2022;82(8):1503–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Guo D, Tong Y, Jiang X, et al. Aerobic glycolysis promotes tumor immune evasion by hexokinase2-mediated phosphorylation of IkappaBalpha [J]. Cell Metab. 2022;34(9):1312–24 e6.

    Article  CAS  PubMed  Google Scholar 

  109. Blaha CS, Ramakrishnan G, Jeon SM, et al. A non-catalytic scaffolding activity of hexokinase 2 contributes to EMT and metastasis [J]. Nat Commun. 2022;13(1):899.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Xu S, Herschman HR. A tumor agnostic therapeutic strategy for hexokinase 1-Null/Hexokinase 2-positive cancers [J]. Cancer Res. 2019;79(23):5907–14.

    Article  CAS  PubMed  Google Scholar 

  111. Webb BA, Forouhar F, Szu FE, et al. Structures of human phosphofructokinase-1 and atomic basis of cancer-associated mutations [J]. Nature. 2015;523(7558):111–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Feng J, Li J, Wu L, et al. Emerging roles and the regulation of aerobic glycolysis in hepatocellular carcinoma [J]. J Exp Clin Cancer Res. 2020;39(1):126.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Webb BA, Chimenti M, Jacobson MP, et al. Dysregulated pH: a perfect storm for cancer progression [J]. Nat Rev Cancer. 2011;11(9):671–7.

    Article  CAS  PubMed  Google Scholar 

  114. Corbet C, Feron O. Tumour acidosis: from the passenger to the driver’s seat [J]. Nat Rev Cancer. 2017;17(10):577–93.

    Article  CAS  PubMed  Google Scholar 

  115. Taylor S, Spugnini EP, Assaraf YG, et al. Microenvironment acidity as a major determinant of tumor chemoresistance: proton pump inhibitors (PPIs) as a novel therapeutic approach [J]. Drug Resist Updat. 2015;23:69–78.

    Article  PubMed  Google Scholar 

  116. Amara N, Cooper MP, Voronkova MA, et al. Selective activation of PFKL suppresses the phagocytic oxidative burst [J]. Cell. 2021;184(17):4480–94 e15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Gauthier T, Yao C, Dowdy T, et al. TGF-beta uncouples glycolysis and inflammation in macrophages and controls survival during sepsis [J]. Sci Signal. 2023;16(797):eade0385.

    Article  CAS  PubMed  Google Scholar 

  118. Icard P, Coquerel A, Wu Z, et al. Understanding the central role of citrate in the metabolism of cancer cells and tumors: an update [J]. Int J Mol Sci. 2021;22(12):6587.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Ando H, Eshima K, Ishida T. Neutralization of acidic tumor microenvironment (TME) with daily oral dosing of sodium potassium citrate (K/Na Citrate) increases therapeutic effect of anti-cancer agent in pancreatic cancer xenograft mice model [J]. Biol Pharm Bull. 2021;44(2):266–70.

    Article  CAS  PubMed  Google Scholar 

  120. Menendez JA, Lupu R. Fatty acid synthase and the lipogenic phenotype in cancer pathogenesis [J]. Nat Rev Cancer. 2007;7(10):763–77.

    Article  CAS  PubMed  Google Scholar 

  121. Jiang L, Fang X, Wang H, et al. Ovarian cancer-intrinsic fatty acid synthase prevents anti-tumor immunity by disrupting tumor-infiltrating dendritic cells [J]. Front Immunol. 2018;9:2927.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Zhang M, Yu L, Sun Y, et al. Comprehensive analysis of FASN in tumor immune infiltration and prognostic value for immunotherapy and promoter DNA methylation [J]. Int J Mol Sci. 2022;23(24):15603.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. O’Farrell M, Duke G, Crowley R, et al. FASN inhibition targets multiple drivers of NASH by reducing steatosis, inflammation and fibrosis in preclinical models [J]. Sci Rep. 2022;12(1):15661.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Sun P, Zhang X, Wang RJ, et al. PI3Kalpha inhibitor CYH33 triggers antitumor immunity in murine breast cancer by activating CD8(+)T cells and promoting fatty acid metabolism [J]. J Immunother Cancer. 2021;9(8):e003093.

    Article  PubMed  PubMed Central  Google Scholar 

  125. Zhao T, Liu S, Ding X, et al. Lysosomal acid lipase, CSF1R, and PD-L1 determine functions of CD11c+ myeloid-derived suppressor cells [J]. JCI Insight. 2022;7(17):e156623.

    Article  PubMed  PubMed Central  Google Scholar 

  126. Gomaraschi M, Bonacina F, Norata GD. Lysosomal acid lipase: from cellular lipid handler to immunometabolic target [J]. Trends Pharmacol Sci. 2019;40(2):104–15.

    Article  CAS  PubMed  Google Scholar 

  127. Huang SC, Everts B, Ivanova Y, et al. Cell-intrinsic lysosomal lipolysis is essential for alternative activation of macrophages [J]. Nat Immunol. 2014;15(9):846–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Zhao T, Du H, Ding X, et al. Activation of mTOR pathway in myeloid-derived suppressor cells stimulates cancer cell proliferation and metastasis in lal(-/-) mice [J]. Oncogene. 2015;34(15):1938–48.

    Article  CAS  PubMed  Google Scholar 

  129. Ding X, Du H, Yoder MC, et al. Critical role of the mTOR pathway in development and function of myeloid-derived suppressor cells in lal-/- mice [J]. Am J Pathol. 2014;184(2):397–408.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Ding X, Zhao T, Lee CC, et al. Lysosomal acid lipase deficiency controls T- and B-Regulatory cell homeostasis in the lymph nodes of mice with human cancer xenotransplants [J]. Am J Pathol. 2021;191(2):353–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. O’Sullivan D, van der Windt GJ, Huang SC, et al. Memory CD8(+) T cells use cell-intrinsic lipolysis to support the metabolic programming necessary for development [J]. Immunity. 2014;41(1):75–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Dominguez M, Brune B, Namgaladze D. Exploring the role of ATP-Citrate lyase in the immune system [J]. Front Immunol. 2021;12:632526.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Zaidi N, Swinnen JV, Smans K. ATP-citrate lyase: a key player in cancer metabolism [J]. Cancer Res. 2012;72(15):3709–14.

    Article  CAS  PubMed  Google Scholar 

  134. Wellen KE, Hatzivassiliou G, Sachdeva UM, et al. ATP-citrate lyase links cellular metabolism to histone acetylation [J]. Science. 2009;324(5930):1076–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Tan M, Mosaoa R, Graham GT, et al. Inhibition of the mitochondrial citrate carrier, Slc25a1, reverts steatosis, glucose intolerance, and inflammation in preclinical models of NAFLD/NASH [J]. Cell Death Differ. 2020;27(7):2143–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Lauterbach MA, Hanke JE, Serefidou M, et al. Toll-like receptor signaling rewires macrophage metabolism and promotes histone acetylation via ATP-Citrate lyase [J]. Immunity. 2019;51(6):997–1011 e7.

    Article  CAS  PubMed  Google Scholar 

  137. Hochrein SM, Wu H, Eckstein M, et al. The glucose transporter GLUT3 controls T helper 17 cell responses through glycolytic-epigenetic reprogramming [J]. Cell Metab. 2022;34(4):516-32 e11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Tian M, Hao F, Jin X, et al. ACLY ubiquitination by CUL3-KLHL25 induces the reprogramming of fatty acid metabolism to facilitate iTreg differentiation [J]. Elife. 2021;10:e62394.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Yang H, Ye D, Guan KL, et al. IDH1 and IDH2 mutations in tumorigenesis: mechanistic insights and clinical perspectives [J]. Clin Cancer Res. 2012;18(20):5562–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Wu MJ, Shi L, Merritt J, et al. Biology of IDH mutant cholangiocarcinoma [J]. Hepatology. 2022;75(5):1322–37.

    Article  CAS  PubMed  Google Scholar 

  141. McClellan BL, Haase S, Nunez FJ, et al. Impact of epigenetic reprogramming on antitumor immune responses in glioma [J]. J Clin Invest. 2023;133(2):e163450.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Itsumi M, Inoue S, Elia AJ, et al. Idh1 protects murine hepatocytes from endotoxin-induced oxidative stress by regulating the intracellular NADP(+)/NADPH ratio [J]. Cell Death Differ. 2015;22(11):1837–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Notarangelo G, Spinelli JB, Perez EM, et al. Oncometabolite d-2HG alters T cell metabolism to impair CD8(+) T cell function [J]. Science. 2022;377(6614):1519–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Jha AK, Huang SC, Sergushichev A, et al. Network integration of parallel metabolic and transcriptional data reveals metabolic modules that regulate macrophage polarization [J]. Immunity. 2015;42(3):419–30.

    Article  CAS  PubMed  Google Scholar 

  145. Wu MJ, Shi L, Dubrot J, et al. Mutant IDH inhibits IFNgamma-TET2 signaling to promote immunoevasion and tumor maintenance in cholangiocarcinoma [J]. Cancer Discov. 2022;12(3):812–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Xiang X, Liu Z, Zhang C, et al. IDH mutation subgroup status associates with intratumor heterogeneity and the tumor microenvironment in intrahepatic cholangiocarcinoma [J]. Adv Sci (Weinh). 2021;8(17):e2101230.

    Article  PubMed  Google Scholar 

  147. Bunse L, Pusch S, Bunse T, et al. Suppression of antitumor T cell immunity by the oncometabolite (R)-2-hydroxyglutarate [J]. Nat Med. 2018;24(8):1192–203.

    Article  CAS  PubMed  Google Scholar 

  148. Yin W, Ping YF, Li F, et al. A map of the spatial distribution and tumour-associated macrophage states in glioblastoma and grade 4 IDH-mutant astrocytoma [J]. J Pathol. 2022;258(2):121–35.

    Article  CAS  PubMed  Google Scholar 

  149. Kohanbash G, Carrera DA, Shrivastav S, et al. Isocitrate dehydrogenase mutations suppress STAT1 and CD8+ T cell accumulation in gliomas [J]. J Clin Invest. 2017;127(4):1425–37.

    Article  PubMed  PubMed Central  Google Scholar 

  150. Schinzel RT, Higuchi-Sanabria R, Shalem O, et al. The Hyaluronidase, TMEM2, promotes ER homeostasis and longevity Independent of the UPR(ER) [J]. Cell. 2019;179(6):1306–18 e18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Hu W, Peng T, Huang Y, et al. Hyaluronidase-powered microneedles for significantly enhanced transdermal delivery efficiency [J]. J Control Release. 2023;353:380–90.

    Article  CAS  PubMed  Google Scholar 

  152. Ruschen H, Aravinth K, Bunce C, et al. Use of hyaluronidase as an adjunct to local anaesthetic eye blocks to reduce intraoperative pain in adults [J]. Cochrane Database Syst Rev. 2018;3(3):CD010368.

    PubMed  Google Scholar 

  153. Fatima K, Masood N, Ahmad Wani Z, et al. Neomenthol prevents the proliferation of skin cancer cells by restraining tubulin polymerization and hyaluronidase activity [J]. J Adv Res. 2021;34:93–107.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Fronza M, Caetano GF, Leite MN, et al. Hyaluronidase modulates inflammatory response and accelerates the cutaneous wound healing [J]. PLoS One. 2014;9(11):e112297.

    Article  PubMed  PubMed Central  Google Scholar 

  155. Jiang D, Liang J, Noble PW. Hyaluronan as an immune regulator in human diseases [J]. Physiol Rev. 2011;91(1):221–64.

    Article  CAS  PubMed  Google Scholar 

  156. Kolar SL, Kyme P, Tseng CW, et al. Group B streptococcus evades host immunity by degrading hyaluronan [J]. Cell Host Microbe. 2015;18(6):694–704.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Blair AB, Wang J, Davelaar J, et al. Dual stromal targeting sensitizes pancreatic adenocarcinoma for anti-programmed cell death protein 1 therapy [J]. Gastroenterology. 2022;163(5):1267–80 e7.

    Article  CAS  PubMed  Google Scholar 

  158. Liu Y, Xu D, Liu Y, et al. Remotely boosting hyaluronidase activity to normalize the hypoxic immunosuppressive tumor microenvironment for photothermal immunotherapy [J]. Biomaterials. 2022;284:121516.

    Article  CAS  PubMed  Google Scholar 

  159. Wang Y, Hou P, Li W, et al. The influence of different current-intensity transcranial alternating current stimulation on the eyes-open and eyes-closed resting-state electroencephalography [J]. Front Hum Neurosci. 2022;16:934382.

    Article  PubMed  PubMed Central  Google Scholar 

  160. Papayannopoulos V. Neutrophil extracellular traps in immunity and disease [J]. Nat Rev Immunol. 2018;18(2):134–47.

    Article  CAS  PubMed  Google Scholar 

  161. Ma X, Xiao L, Liu L, et al. CD36-mediated ferroptosis dampens intratumoral CD8(+) T cell effector function and impairs their antitumor ability [J]. Cell Metab. 2021;33(5):1001–12 e5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Peeters R, Cuenca-Escalona J, Zaal EA, et al. Fatty acid metabolism in aggressive B-cell lymphoma is inhibited by tetraspanin CD37 [J]. Nat Commun. 2022;13(1):5371.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Ma X, Bi E, Huang C, et al. Cholesterol negatively regulates IL-9-producing CD8(+) T cell differentiation and antitumor activity [J]. J Exp Med. 2018;215(6):1555–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Bauman DR, Bitmansour AD, McDonald JG, et al. 25-Hydroxycholesterol secreted by macrophages in response to Toll-like receptor activation suppresses immunoglobulin A production [J]. Proc Natl Acad Sci U S A. 2009;106(39):16764–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Jaillon S, Ponzetta A, di Mitri D, et al. Neutrophil diversity and plasticity in tumour progression and therapy [J]. Nat Rev Cancer. 2020;20(9):485–503.

    Article  CAS  PubMed  Google Scholar 

  166. Cane S, Barouni RM, Fabbi M, et al. Neutralization of NET-associated human ARG1 enhances cancer immunotherapy [J]. Sci Transl Med. 2023;15(687):eabq6221.

    Article  CAS  PubMed  Google Scholar 

  167. Cui G, Staron MM, Gray SM, et al. IL-7-induced glycerol transport and TAG synthesis promotes memory CD8+ T cell longevity [J]. Cell. 2015;161(4):750–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Kaech SM, Cui W. Transcriptional control of effector and memory CD8+ T cell differentiation [J]. Nat Rev Immunol. 2012;12(11):749–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Lim SA, Su W, Chapman NM, et al. Lipid metabolism in T cell signaling and function [J]. Nat Chem Biol. 2022;18(5):470–81.

    Article  CAS  PubMed  Google Scholar 

  170. Xu S, Chaudhary O, Rodriguez-Morales P, et al. Uptake of oxidized lipids by the scavenger receptor CD36 promotes lipid peroxidation and dysfunction in CD8(+) T cells in tumors [J]. Immunity. 2021;54(7):1561–77 e7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Chapman NM, Boothby MR, Chi H. Metabolic coordination of T cell quiescence and activation [J]. Nat Rev Immunol. 2020;20(1):55–70.

    Article  CAS  PubMed  Google Scholar 

  172. Yang K, Shrestha S, Zeng H, et al. T cell exit from quiescence and differentiation into Th2 cells depend on Raptor-mTORC1-mediated metabolic reprogramming [J]. Immunity. 2013;39(6):1043–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Nian Z, Zheng X, Dou Y, et al. Rapamycin pretreatment rescues the bone marrow AML cell elimination capacity of CAR-T cells [J]. Clin Cancer Res. 2021;27(21):6026–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Werter IM, Huijts CM, Lougheed SM, et al. Metronomic cyclophosphamide attenuates mTOR-mediated expansion of regulatory T cells, but does not impact clinical outcome in patients with metastatic renal cell cancer treated with everolimus [J]. Cancer Immunol Immunother. 2019;68(5):787–98.

    Article  CAS  PubMed  Google Scholar 

  175. Braun DA, Hou Y, Bakouny Z, et al. Interplay of somatic alterations and immune infiltration modulates response to PD-1 blockade in advanced clear cell renal cell carcinoma [J]. Nat Med. 2020;26(6):909–18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. van der Windt GJ, Everts B, Chang CH, et al. Mitochondrial respiratory capacity is a critical regulator of CD8+ T cell memory development [J]. Immunity. 2012;36(1):68–78.

    Article  PubMed  Google Scholar 

  177. Trompette A, Gollwitzer ES, Pattaroni C, et al. Dietary fiber confers protection against flu by shaping Ly6c(-) patrolling monocyte hematopoiesis and CD8(+) T cell metabolism [J]. Immunity. 2018;48(5):992–1005 e8.

    Article  CAS  PubMed  Google Scholar 

  178. Li J, Lu E, Yi T, et al. EBI2 augments Tfh cell fate by promoting interaction with IL-2-quenching dendritic cells [J]. Nature. 2016;533(7601):110–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Hang S, Paik D, Yao L, et al. Bile acid metabolites control T(H)17 and T(reg) cell differentiation [J]. Nature. 2019;576(7785):143–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Kidani Y, Elsaesser H, Hock MB, et al. Sterol regulatory element-binding proteins are essential for the metabolic programming of effector T cells and adaptive immunity [J]. Nat Immunol. 2013;14(5):489–99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  181. Su W, Chapman NM, Wei J, et al. Protein prenylation drives discrete signaling programs for the differentiation and maintenance of effector T(reg) cells [J]. Cell Metab. 2020;32(6):996–1011 e7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Boothby MR, Brookens SK, Raybuck AL, et al. Supplying the trip to antibody production-nutrients, signaling, and the programming of cellular metabolism in the mature B lineage [J]. Cell Mol Immunol. 2022;19(3):352–69.

    Article  CAS  PubMed  Google Scholar 

  183. Bommer GT, Macdougald OA. Regulation of lipid homeostasis by the bifunctional SREBF2-miR33a locus [J]. Cell Metab. 2011;13(3):241–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Luo W, Adamska JZ, Li C, et al. SREBP signaling is essential for effective B cell responses [J]. Nat Immunol. 2023;24(2):337–48.

    Article  CAS  PubMed  Google Scholar 

  185. Cheng J, Liu Y, Yan J, et al. Fumarate suppresses B-cell activation and function through direct inactivation of LYN [J]. Nat Chem Biol. 2022;18(9):954–62.

    Article  CAS  PubMed  Google Scholar 

  186. Soncini M, Corna G, Moresco M, et al. 24-Hydroxycholesterol participates in pancreatic neuroendocrine tumor development [J]. Proc Natl Acad Sci U S A. 2016;113(41):E6219–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Yi T, Wang X, Kelly LM, et al. Oxysterol gradient generation by lymphoid stromal cells guides activated B cell movement during humoral responses [J]. Immunity. 2012;37(3):535–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Blaho VA, Hla T. Regulation of mammalian physiology, development, and disease by the sphingosine 1-phosphate and lysophosphatidic acid receptors [J]. Chem Rev. 2011;111(10):6299–320.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Jozefczuk E, Guzik TJ, Siedlinski M. Significance of sphingosine-1-phosphate in cardiovascular physiology and pathology [J]. Pharmacol Res. 2020;156:104793.

    Article  CAS  PubMed  Google Scholar 

  190. Pereira JP, Xu Y, Cyster JG. A role for S1P and S1P1 in immature-B cell egress from mouse bone marrow [J]. PLoS One. 2010;5(2):e9277.

    Article  PubMed  PubMed Central  Google Scholar 

  191. Allende ML, Tuymetova G, Lee BG, et al. S1P1 receptor directs the release of immature B cells from bone marrow into blood [J]. J Exp Med. 2010;207(5):1113–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. He C, Wang S, Zhou C, et al. CD36 and LC3B initiated autophagy in B cells regulates the humoral immune response [J]. Autophagy. 2021;17(11):3577–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. Murray PJ, Allen JE, Biswas SK, et al. Macrophage activation and polarization: nomenclature and experimental guidelines [J]. Immunity. 2014;41(1):14–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Chen Y, Zhang J, Cui W, et al. CD36, a signaling receptor and fatty acid transporter that regulates immune cell metabolism and fate [J]. J Exp Med. 2022;219(6):e20211314.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Baardman J, Verberk SGS, van der Velden S, et al. Macrophage ATP citrate lyase deficiency stabilizes atherosclerotic plaques [J]. Nat Commun. 2020;11(1):6296.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Gomez Perdiguero E, Klapproth K, Schulz C, et al. Tissue-resident macrophages originate from yolk-sac-derived erythro-myeloid progenitors [J]. Nature. 2015;518(7540):547–51.

    Article  PubMed  Google Scholar 

  197. Hashimoto D, Chow A, Noizat C, et al. Tissue-resident macrophages self-maintain locally throughout adult life with minimal contribution from circulating monocytes [J]. Immunity. 2013;38(4):792–804.

    Article  CAS  PubMed  Google Scholar 

  198. Okabe Y, Medzhitov R. Tissue-specific signals control reversible program of localization and functional polarization of macrophages [J]. Cell. 2014;157(4):832–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Zhang X, Ji L, Li MO. Control of tumor-associated macrophage responses by nutrient acquisition and metabolism [J]. Immunity. 2023;56(1):14–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. Kane H, Lynch L. Innate immune control of adipose tissue homeostasis [J]. Trends Immunol. 2019;40(9):857–72.

    Article  CAS  PubMed  Google Scholar 

  201. Nakamura A, Ebina-Shibuya R, Itoh-Nakadai A, et al. Transcription repressor Bach2 is required for pulmonary surfactant homeostasis and alveolar macrophage function [J]. J Exp Med. 2013;210(11):2191–204.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. McCarthy C, Lee E, Bridges JP, et al. Statin as a novel pharmacotherapy of pulmonary alveolar proteinosis [J]. Nat Commun. 2018;9(1):3127.

    Article  PubMed  PubMed Central  Google Scholar 

  203. Ishikawa F, Niiro H, Iino T, et al. The developmental program of human dendritic cells is operated independently of conventional myeloid and lymphoid pathways [J]. Blood. 2007;110(10):3591–660.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. le Naour F, Hohenkirk L, Grolleau A, et al. Profiling changes in gene expression during differentiation and maturation of monocyte-derived dendritic cells using both oligonucleotide microarrays and proteomics [J]. J Biol Chem. 2001;276(21):17920–31.

    Article  PubMed  Google Scholar 

  205. Wang R, Dillon CP, Shi LZ, et al. The transcription factor Myc controls metabolic reprogramming upon T lymphocyte activation [J]. Immunity. 2011;35(6):871–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  206. Kc W, Satpathy AT, Rapaport AS, et al. L-Myc expression by dendritic cells is required for optimal T-cell priming [J]. Nature. 2014;507(7491):243–7.

    Article  CAS  PubMed  Google Scholar 

  207. Krawczyk CM, Holowka T, Sun J, et al. Toll-like receptor-induced changes in glycolytic metabolism regulate dendritic cell activation [J]. Blood. 2010;115(23):4742–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Schafer JR, Salzillo TC, Chakravarti N, et al. Education-dependent activation of glycolysis promotes the cytolytic potency of licensed human natural killer cells [J]. J Allergy Clin Immunol. 2019;143(1):346–58 e6.

    Article  CAS  PubMed  Google Scholar 

  209. Marcais A, Cherfils-Vicini J, Viant C, et al. The metabolic checkpoint kinase mTOR is essential for IL-15 signaling during the development and activation of NK cells [J]. Nat Immunol. 2014;15(8):749–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Keppel MP, Saucier N, Mah AY, et al. Activation-specific metabolic requirements for NK Cell IFN-gamma production [J]. J Immunol. 2015;194(4):1954–62.

    Article  CAS  PubMed  Google Scholar 

  211. Felices M, Lenvik AJ, McElmurry R, et al. Continuous treatment with IL-15 exhausts human NK cells via a metabolic defect [J]. JCI Insight. 2018;3(3):e96219.

    Article  PubMed  PubMed Central  Google Scholar 

  212. Jensen H, Potempa M, Gotthardt D, et al. Cutting edge: IL-2-Induced expression of the amino acid transporters SLC1A5 and CD98 is a prerequisite for NKG2D-mediated activation of human NK cells [J]. J Immunol. 2017;199(6):1967–72.

    Article  CAS  PubMed  Google Scholar 

  213. Hu X, Jia X, Xu C, et al. Downregulation of NK cell activities in Apolipoprotein C-III-induced hyperlipidemia resulting from lipid-induced metabolic reprogramming and crosstalk with lipid-laden dendritic cells [J]. Metabolism. 2021;120:154800.

    Article  CAS  PubMed  Google Scholar 

  214. Chiossone L, Chaix J, Fuseri N, et al. Maturation of mouse NK cells is a 4-stage developmental program [J]. Blood. 2009;113(22):5488–96.

    Article  CAS  PubMed  Google Scholar 

  215. Loftus RM, Assmann N, Kedia-Mehta N, et al. Amino acid-dependent cMyc expression is essential for NK cell metabolic and functional responses in mice [J]. Nat Commun. 2018;9(1):2341.

    Article  PubMed  PubMed Central  Google Scholar 

  216. Michelet X, Dyck L, Hogan A, et al. Metabolic reprogramming of natural killer cells in obesity limits antitumor responses [J]. Nat Immunol. 2018;19(12):1330–40.

    Article  CAS  PubMed  Google Scholar 

  217. Cerwenka A, Lanier LL. Natural killer cell memory in infection, inflammation and cancer [J]. Nat Rev Immunol. 2016;16(2):112–23.

    Article  CAS  PubMed  Google Scholar 

  218. O’Sullivan TE, Sun JC, Lanier LL. Natural killer cell memory [J]. Immunity. 2015;43(4):634–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  219. Paust S, von Andrian UH. Natural killer cell memory [J]. Nat Immunol. 2011;12(6):500–8.

    Article  CAS  PubMed  Google Scholar 

  220. O’Sullivan TE, Johnson LR, Kang HH, et al. BNIP3- and BNIP3L-Mediated mitophagy promotes the generation of natural killer cell memory [J]. Immunity. 2015;43(2):331–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Sun JC, Beilke JN, Lanier LL. Adaptive immune features of natural killer cells [J]. Nature. 2009;457(7229):557–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Liu LL, Landskron J, Ask EH, et al. Critical Role of CD2 Co-stimulation in adaptive natural killer cell responses revealed in NKG2C-Deficient humans [J]. Cell Rep. 2016;15(5):1088–99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  223. Filippi MD. Neutrophil transendothelial migration: updates and new perspectives [J]. Blood. 2019;133(20):2149–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  224. Hidalgo A, Chilvers ER, Summers C, et al. The neutrophil life cycle [J]. Trends Immunol. 2019;40(7):584–97.

    Article  CAS  PubMed  Google Scholar 

  225. Liew PX, Kubes P. The neutrophil’s role during health and disease [J]. Physiol Rev. 2019;99(2):1223–48.

    Article  CAS  PubMed  Google Scholar 

  226. Raccosta L, Fontana R, Maggioni D, et al. The oxysterol-CXCR2 axis plays a key role in the recruitment of tumor-promoting neutrophils [J]. J Exp Med. 2013;210(9):1711–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Monceaux V, Chiche-Lapierre C, Chaput C, et al. Anoxia and glucose supplementation preserve neutrophil viability and function [J]. Blood. 2016;128(7):993–1002.

    Article  CAS  PubMed  Google Scholar 

  228. Mecklenburgh KI, Walmsley SR, Cowburn AS, et al. Involvement of a ferroprotein sensor in hypoxia-mediated inhibition of neutrophil apoptosis [J]. Blood. 2002;100(8):3008–16.

    Article  CAS  PubMed  Google Scholar 

  229. Maianski NA, Geissler J, Srinivasula SM, et al. Functional characterization of mitochondria in neutrophils: a role restricted to apoptosis [J]. Cell Death Differ. 2004;11(2):143–53.

    Article  CAS  PubMed  Google Scholar 

  230. Veiga-Da-cunha M, Chevalier N, Stephenne X, et al. Failure to eliminate a phosphorylated glucose analog leads to neutropenia in patients with G6PT and G6PC3 deficiency [J]. Proc Natl Acad Sci U S A. 2019;116(4):1241–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  231. Jun HS, Weinstein DA, Lee YM, et al. Molecular mechanisms of neutrophil dysfunction in glycogen storage disease type Ib [J]. Blood. 2014;123(18):2843–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  232. Jun HS, Lee YM, Cheung YY, et al. Lack of glucose recycling between endoplasmic reticulum and cytoplasm underlies cellular dysfunction in glucose-6-phosphatase-beta-deficient neutrophils in a congenital neutropenia syndrome [J]. Blood. 2010;116(15):2783–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  233. Condamine T, Dominguez GA, Youn JI, et al. Lectin-type oxidized LDL receptor-1 distinguishes population of human polymorphonuclear myeloid-derived suppressor cells in cancer patients [J]. Sci Immunol. 2016;1(2):aaf8943.

    Article  PubMed  PubMed Central  Google Scholar 

  234. Gehrke N, Mertens C, Zillinger T, et al. Oxidative damage of DNA confers resistance to cytosolic nuclease TREX1 degradation and potentiates STING-dependent immune sensing [J]. Immunity. 2013;39(3):482–95.

    Article  CAS  PubMed  Google Scholar 

  235. Lood C, Blanco LP, Purmalek MM, et al. Neutrophil extracellular traps enriched in oxidized mitochondrial DNA are interferogenic and contribute to lupus-like disease [J]. Nat Med. 2016;22(2):146–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  236. Wang L, Qian J, Lu Y, et al. Immune evasion of mantle cell lymphoma: expression of B7–H1 leads to inhibited T-cell response to and killing of tumor cells [J]. Haematologica. 2013;98(9):1458–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  237. Braumuller H, Wieder T, Brenner E, et al. T-helper-1-cell cytokines drive cancer into senescence [J]. Nature. 2013;494(7437):361–5.

    Article  PubMed  Google Scholar 

  238. Galluzzi L, Vitale I, Warren S, et al. Consensus guidelines for the definition, detection and interpretation of immunogenic cell death [J]. J Immunother Cancer. 2020;8(1):e000337.

    Article  PubMed  PubMed Central  Google Scholar 

  239. Kroemer G, Galassi C, Zitvogel L, et al. Immunogenic cell stress and death [J]. Nat Immunol. 2022;23(4):487–500.

    Article  CAS  PubMed  Google Scholar 

  240. Zhou Y, Bastian IN, Long MD, et al. Activation of NF-kappaB and p300/CBP potentiates cancer chemoimmunotherapy through induction of MHC-I antigen presentation [J]. Proc Natl Acad Sci U S A. 2021;118(8):e2025840118.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  241. Galaine J, Turco C, Vauchy C, et al. CD4 T cells target colorectal cancer antigens upregulated by oxaliplatin [J]. Int J Cancer. 2019;145(11):3112–25.

    Article  CAS  PubMed  Google Scholar 

  242. Kepp O, Zitvogel L, Kroemer G. Clinical evidence that immunogenic cell death sensitizes to PD-1/PD-L1 blockade [J]. Oncoimmunology. 2019;8(10):e1637188.

    Article  PubMed  PubMed Central  Google Scholar 

  243. D’Amico L, Menzel U, Prummer M, et al. A novel anti-HER2 anthracycline-based antibody-drug conjugate induces adaptive anti-tumor immunity and potentiates PD-1 blockade in breast cancer [J]. J Immunother Cancer. 2019;7(1):16.

    Article  PubMed  PubMed Central  Google Scholar 

  244. Zhou F, Feng B, Yu H, et al. Tumor microenvironment-activatable prodrug vesicles for nanoenabled cancer chemoimmunotherapy combining immunogenic cell death induction and CD47 blockade [J]. Adv Mater. 2019;31(14):e1805888.

    Article  PubMed  Google Scholar 

  245. Krysko DV, Garg AD, Kaczmarek A, et al. Immunogenic cell death and DAMPs in cancer therapy [J]. Nat Rev Cancer. 2012;12(12):860–75.

    Article  CAS  PubMed  Google Scholar 

  246. Dudek AM, Garg AD, Krysko DV, et al. Inducers of immunogenic cancer cell death [J]. Cytokine Growth Factor Rev. 2013;24(4):319–33.

    Article  CAS  PubMed  Google Scholar 

  247. Apetoh L, Ghiringhelli F, Tesniere A, et al. The interaction between HMGB1 and TLR4 dictates the outcome of anticancer chemotherapy and radiotherapy [J]. Immunol Rev. 2007;220:47–59.

    Article  CAS  PubMed  Google Scholar 

  248. Michaud M, Martins I, Sukkurwala AQ, et al. Autophagy-dependent anticancer immune responses induced by chemotherapeutic agents in mice [J]. Science. 2011;334(6062):1573–7.

    Article  CAS  PubMed  Google Scholar 

  249. Obeid M, Tesniere A, Ghiringhelli F, et al. Calreticulin exposure dictates the immunogenicity of cancer cell death [J]. Nat Med. 2007;13(1):54–61.

    Article  CAS  PubMed  Google Scholar 

  250. Casares N, Pequignot MO, Tesniere A, et al. Caspase-dependent immunogenicity of doxorubicin-induced tumor cell death [J]. J Exp Med. 2005;202(12):1691–701.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  251. Mellman I, Chen DS, Powles T, et al. The cancer-immunity cycle: Indication, genotype, and immunotype [J]. Immunity. 2023;56(10):2188–205.

    Article  CAS  PubMed  Google Scholar 

  252. Chen DS, Mellman I. Oncology meets immunology: the cancer-immunity cycle [J]. Immunity. 2013;39(1):1–10.

    Article  PubMed  Google Scholar 

  253. Motz GT, Coukos G. Deciphering and reversing tumor immune suppression [J]. Immunity. 2013;39(1):61–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  254. You M, Xie Z, Zhang N, et al. Signaling pathways in cancer metabolism: mechanisms and therapeutic targets [J]. Signal Transduct Target Ther. 2023;8(1):196.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  255. Tsai CH, Chuang YM, Li X, et al. Immunoediting instructs tumor metabolic reprogramming to support immune evasion [J]. Cell Metab. 2023;35(1):118–33 e7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  256. Giles JR, Globig AM, Kaech SM, et al. CD8(+) T cells in the cancer-immunity cycle [J]. Immunity. 2023;56(10):2231–53.

    Article  CAS  PubMed  Google Scholar 

  257. Nguyen T, Kirsch BJ, Asaka R, et al. Uncovering the role of N-Acetyl-Aspartyl-Glutamate as a glutamate reservoir in cancer [J]. Cell Rep. 2019;27(2):491–501 e6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  258. Markovic M, Ben-Shabat S, Keinan S, et al. Lipidic prodrug approach for improved oral drug delivery and therapy [J]. Med Res Rev. 2019;39(2):579–607.

    Article  PubMed  Google Scholar 

  259. Elz AS, Trevaskis NL, Porter CJH, et al. Smart design approaches for orally administered lipophilic prodrugs to promote lymphatic transport [J]. J Control Release. 2022;341:676–701.

    Article  CAS  PubMed  Google Scholar 

  260. Haidinger M, Poglitsch M, Geyeregger R, et al. A versatile role of mammalian target of rapamycin in human dendritic cell function and differentiation [J]. J Immunol. 2010;185(7):3919–31.

    Article  CAS  PubMed  Google Scholar 

  261. Woltman AM, van der Kooij SW, Coffer PJ, et al. Rapamycin specifically interferes with GM-CSF signaling in human dendritic cells, leading to apoptosis via increased p27KIP1 expression [J]. Blood. 2003;101(4):1439–45.

    Article  CAS  PubMed  Google Scholar 

  262. Donnelly RP, Loftus RM, Keating SE, et al. mTORC1-dependent metabolic reprogramming is a prerequisite for NK cell effector function [J]. J Immunol. 2014;193(9):4477–84.

    Article  CAS  PubMed  Google Scholar 

  263. Keating SE, Zaiatz-Bittencourt V, Loftus RM, et al. Metabolic reprogramming supports IFN-gamma production by CD56bright NK cells [J]. J Immunol. 2016;196(6):2552–60.

    Article  CAS  PubMed  Google Scholar 

  264. Sofi MH, Heinrichs J, Dany M, et al. Ceramide synthesis regulates T cell activity and GVHD development [J]. JCI Insight. 2017;2(10):e91701.

    Article  PubMed  PubMed Central  Google Scholar 

  265. Cedervall J, Zhang Y, Huang H, et al. Neutrophil extracellular traps accumulate in peripheral blood vessels and compromise organ function in tumor-bearing animals [J]. Cancer Res. 2015;75(13):2653–62.

    Article  CAS  PubMed  Google Scholar 

  266. Pilon-Thomas S, Kodumudi KN, El-Kenawi AE, et al. Neutralization of tumor acidity improves antitumor responses to immunotherapy [J]. Cancer Res. 2016;76(6):1381–90.

    Article  CAS  PubMed  Google Scholar 

  267. Buckley D, Duke G, Heuer TS, et al. Fatty acid synthase - modern tumor cell biology insights into a classical oncology target [J]. Pharmacol Ther. 2017;177:23–31.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Part of the materials in the figures are drawn by Figdraw (www.figdraw.com) and this manuscript is shared with permission for copyright.

Funding

This work was supported by grants from National Natural Science Foundation of China (No. 82172948, 81972250, 82141129, and 82173281).

Author information

Authors and Affiliations

Authors

Contributions

Qin Dang, Borui Li and Bing Jin contributed equally.QD, XJY, XWX, and YQ provided direction and guidance throughout the preparation of this manuscript. QD, BRL, BJ, and ZY wrote and edited the manuscript. QD, CJZ, and SRJ reviewed and made significant revisions to the manuscript. XL, TW, YW, XP, QSH, and ZL collected and prepared the related papers. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Xianjun Yu, Yi Qin or Xiaowu Xu.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

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

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dang, Q., Li, B., Jin, B. et al. Cancer immunometabolism: advent, challenges, and perspective. Mol Cancer 23, 72 (2024). https://doi.org/10.1186/s12943-024-01981-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12943-024-01981-5

Keywords