Next Article in Journal
Amyotrophic Lateral Sclerosis as a Systemic Disease
Next Article in Special Issue
Intratumor Heterogeneity and Treatment Resistance of Solid Tumors with a Focus on Polyploid/Senescent Giant Cancer Cells (PGCCs)
Previous Article in Journal
PDGF Receptors and Signaling Are Required for 3D-Structure Formation and Differentiation of Human iPSC-Derived Hepatic Spheroids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulative Roles of Metabolic Plasticity Caused by Mitochondrial Oxidative Phosphorylation and Glycolysis on the Initiation and Progression of Tumorigenesis

1
Shenzhen Engineering Labortaory for Marine Algal Biotechnology, Longhua Innovation Institute for Biotechnology, College of Life Sciences and Oceanography, Lihu Campus of Shenzhen University, Shenzhen 518055, China
2
College of Physics and Optoelectronic Engineering, Canghai Campus of Shenzhen University, Shenzhen 518060, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(8), 7076; https://doi.org/10.3390/ijms24087076
Submission received: 3 February 2023 / Revised: 23 March 2023 / Accepted: 4 April 2023 / Published: 11 April 2023

Abstract

:
One important feature of tumour development is the regulatory role of metabolic plasticity in maintaining the balance of mitochondrial oxidative phosphorylation and glycolysis in cancer cells. In recent years, the transition and/or function of metabolic phenotypes between mitochondrial oxidative phosphorylation and glycolysis in tumour cells have been extensively studied. In this review, we aimed to elucidate the characteristics of metabolic plasticity (emphasizing their effects, such as immune escape, angiogenesis migration, invasiveness, heterogeneity, adhesion, and phenotypic properties of cancers, among others) on tumour progression, including the initiation and progression phases. Thus, this article provides an overall understanding of the influence of abnormal metabolic remodeling on malignant proliferation and pathophysiological changes in carcinoma.

1. Introduction

The energy metabolism of normal cells is maintained in a balanced state. In the presence of oxygen, partial metabolites are transmitted through respiratory chain enzyme complexes I–IV on the inner membrane of mitochondria, and adenosine triphosphates (ATPs) are finally synthesized by oxidative phosphorylation of adenosine triphosphate synthase (ATPase) [1,2,3], which only occupies a small part of the whole cell energy supply. More glycolysis-related enzymes are activated when cells are exposed to ischaemia and hypoxia to maintain the cellular energy supply [4]. However, in the 1920s, Otto Warburg’s pioneering study demonstrated that, unlike normal cells, which produce energy by oxidative phosphorylation of mitochondria, tumour cells mainly produce energy through glycolysis, even if the oxygen supply is sufficient [5]. This abnormal metabolic status in tumour cells is called the Warburg Effect, which is the most significant metabolic difference between normal and tumour cells. The Warburg Effect is responsible for various characteristics associated with the occurrence and development of tumour cells, such as their phenotype, material metabolism, molecular regulation, proliferation, invasion, immune escape, and angiogenesis stimulation [6,7,8,9,10].
Typically, in the presence of oxygen, glycolysis produces only two ATPs per glucose molecule, which is much lower than the ~38 ATPs produced per glucose molecule via oxidative phosphorylation. Thus, for energy supply, mitochondrial oxidative phosphorylation, rather than glycolysis (usually enhanced in the latter phase of energy production) should be used to meet the complex and vigorous biological and pathological needs of tumour cells. Generally, the contribution of glycolysis to total ATPs in tumour cells ranges from 1% to 64% [11]. However, with more research data, it has been revealed that tumour cells still retain relatively complete mitochondria and the ability of oxidative phosphorylation.
Tumour cells with glycolysis could be stimulated to use mitochondrial oxidative phosphorylation under specific conditions, and the capacity of mitochondrial oxidation that tumour cells retain is not just to meet energy demand [12,13,14]. These adjusted actions are mainly induced by their abnormal phenotype, functional transformation, or other physical and chemical factors caused by internal and external stimulators [15,16]. Glycolysis and mitochondrial oxidative phosphorylation constitute a diversified and sophisticated metabolic phenotype of tumour cells called metabolic plasticity. The metabolic plasticity of tumour cells allows them to have strong adaptability and elasticity to cope with intracellular and extracellular environments, which makes them highly plastic and pervasive in all aspects of tumourigenesis and development [17,18,19], including their initiation and progression. However, it remains unknown how tumour cells coordinate their gene regulation to balance their glycolysis and oxidative phosphorylation (OXPHOS) and subsequently affect their occurrence and development by regulating metabolic plasticity. These topic questions are also the hotspot and difficulty of current research and bring opportunities and challenges for clinical treatment.
The tumourigenesis process involves several stages: initiation of tumourigenesis, uncontrolled proliferation and unprogrammed cell death, development of tumour microenvironment (TME), and gaining the ability to invade and migrate [20]. The progression of tumour cells following the initial stage is caused by a series of gene mutations or genomic changes, as well as disorders of the cell cycle and frustration of immune surveillance function, which causes cells to divide continuously, escape apoptosis, immune phagocytosis, and inhibit growth [21,22,23,24]. This uncontrolled cell state gradually accumulates, eventually resulting in tumourigenesis. At the initial stages of tumourigenesis, many metabolic pathways may play important regulatory roles in tumourigenesis and development [25,26,27]. More importantly, to meet the needs of vigorous growth, cancer cells need to adjust, integrate, and change their metabolic pattern, which involves many oncogenes [28,29]. For example, the loss of NAD-dependent protein deacetylase sirtuin-6 (SIRT6) can produce a powerful metabolic rearrangement that can induce tumourigenesis [27,30]. The overexpression of glucose transporter type 3 (GLUT3) in non-malignant breast cancer cells can enhance the activation of epidermal growth factor receptor (EGFR), MAP kinase-activated protein kinase (MAPK), RAC-beta serine/threonine-protein kinase (Akt), and other signaling pathways as well as improve glucose metabolism and the expression of related oncogenes, thereby increasing the risk of cell transformation [31].
Tumour cells with immortality and unlimited proliferation ability gradually replace normal cells in the surrounding environment, absorb all possible nutrients, promote angiogenesis, and further develop cancer cells. When the environment cannot meet its growth needs, tumour cells change the expression pattern of endogenous regulatory molecules and the cell phenotype, secrete invasion factors, affect the TME externally, penetrate the blood circulation, and spread outwards. In this process, tumour cells adjust their adaptive strategies through metabolic plasticity regulation, such as adjusting the absorption mode of nutrients, which is accompanied by the transformation of glycolysis to OXPHOS [32,33]. Another example is to adjust the tumour microenvironment by secreting lactic acid to allow the infiltrating immune cells to undergo metabolic rearrangement and eventually escape [34,35,36,37].
Subsequently, metastatic tumour cells travel through the blood and lymphatic circulation, escape the surveillance of immune cells in various ways, infiltrate the blood vessel wall, and then reach the second plantable area for reciprocal circulation. In this process, tumour cells can also affect their ability to metastasize and invade in many ways by remodelling the metabolic processes of glycolysis and OXPHOS, such as adjusting epithelial–mesenchymal transition(EMT)-related proteins, changing the composition of the extracellular matrix, and antagonizing apoptosis [38,39,40,41].
Therefore, during the entire stage of tumour cell development, even though some factors and pathways can affect the future trend of tumours, they directly or indirectly regulate metabolism, and the regulation of metabolic plasticity will also affect the vigorous growth and other needs of tumour cells [42]. The regulation of plasticity between glycolysis and mitochondrial oxidative phosphorylation is one of the basic characteristics of tumour cells [29,42]. Correspondingly, this metabolic plasticity also affects the performance of anticancer drugs. Thus, this review aims to summarize and discuss the characteristics and regulatory processes of metabolic plasticity between glycolysis and mitochondrial oxidative phosphorylation during tumour initiation and progression to provide a new vision for the treatment and understanding of tumourigenesis and development.

2. Metabolic Plasticity in the Initiation of Tumourigenesis

The regulation of metabolic plasticity, dependent on glycolysis and OXPHOS, initiates tumour progression through a variety of complex and hard-to-explain mechanisms for its impact on tumour initiation. Researchers have proposed different theories regarding the initiation of tumour cells, such as oncogene theory [43], mutation theory [43], cell fusion theory [44], and two-hit hypothesis [45]. Therefore, we have discussed different aspects of the influence of metabolic plasticity on tumour initiation. While Potter and Pierce proposed that tumours might result from the inhibition of stem cell differentiation and maturation [46,47], many researchers believe that tumours may originate from a single clone of stem cells, which have undergone malignant transformation and show increased aggressiveness. Proliferating and differentiating stem cells have a high probability of gene mutation [48,49], genomic instability [50,51], and abnormal cell fusion [52,53], which lead to tumorigenesis [54,55,56,57].
Therefore, more researchers believe that tumourigenesis is initiated by the development of stem cells. That is, tumour stem cells, also known as tumour-initiating cells (TICs), are self-renewal cells with multiple heterogeneities. Tumour stem cells can divide asymmetrically to produce daughter cells with different phenotypes. This led to some doubts regarding whether the metabolic phenotype of tumour depends on a single tumour cell or composed of tumour cells with different metabolic phenotypes. To study the metabolic plasticity mediated by glycolysis and mitochondrial oxidative phosphorylation in the initial stage of the tumour, we focused on the effect of metabolic plasticity on the tumourigenesis in stem cells.

2.1. The Importance of Glycolysis and OXPHOS in the Transformation of Stem Cells into Tumour Stem Cells

Later, researchers began to explore the mechanism of metabolic regulation of stem cell-derived tumour stem cells. Many reports have shown that resting stem cells have typical metabolic characteristics and are more prone to glycolysis [58,59]. For instance, the transcription factor homeobox protein Meis1 (MEIS1) in haematopoietic stem cells (HSC) can regulate hypoxia inducible factor-1α (HIF-1α) expression [60], which further controls the expression of many glycolysis-related molecules. Differentiated daughter cells derived from HSCs also rely on glycolysis to meet their metabolic needs [61]. Some pyruvate dehydrogenase kinase (PDK) isoforms can restrict pyruvate entry into mitochondria to control the resting state of long-term repopulating haematopoietic stem cells (LT-HSCs) [59,62,63]. These studies have provided insights into the role of glycolysis in stem cells.
However, it is unclear whether the metabolic characteristics of glycolysis in stem cells are retained after they transform into cancer cells. As previously reported, the energy source of some glioma stem cells (GSCs) primarily depends on glycolysis [64,65,66]. In addition, the activation of some glycolytic enzymes, rather than oncogenes, can cause tumourigenesis and transformation, providing favourable evidence for the metabolic influence on tumorigenesis [67,68,69]. Notably, the biological and biochemical processes of some transcription factors, such as Myc proto-oncogene protein (c-Myc) and Nuclear factor NF-kappa-B (NF-κB), are mainly dependent on glycolytic reactions [9,70,71,72]. These transcription factors can activate epithelial stem cell(ESC)-like transcription programs [73] and induce dedifferentiation of epithelial non-stem cells to obtain characteristics of intestinal stem cells (ISC) [74], which could further promote the tumourigenesis.
Previous studies have also revealed a possible connection between glycolysis and tumourigenesis in normal somatic cells, except for reports on glycolysis in the cancer process of stem cells. For example, mutations in KRAS proto-oncogene and GTPase(KRAS) exist in most tumours, such as lung cancer, colon cancer, and pancreatic cancer [75,76,77], and introduce mutated KRAS (KRASG12D) into normal human breast cells, which can initiate breast cancer [78]. KRAS can also directly regulate the key enzymes of glycolysis, such as hexokinase (HK), phosphofructokinase-1 (PFK1), lactate dehydrogenase A (LDHA), and glucose transporter Type 1 (GLUT1), thus affecting the glycolysis and impacting cell canceration [77,79].
Therefore, can it be considered that the metabolic phenotype of the initial process of tumour cells is dominated by glycolysis? Alternatively, is glycolysis rather than OXPHOS involved in the initiation of tumour cells? Many researchers have used a more efficient and controllable method by artificially stimulating tumour initiation to study the metabolic plasticity at this stage and accurately explore its influence on tumour initiation. Using this strategy, researchers confirmed that the active cycling crypt base columnar (CBC) leucine-rich repeat-containing G protein-coupled receptor 5 positive (Lgr5+) intestinal stem cells are the main cause of colon epithelial cell carcinoma (CRC). On basis of this model, later studies found that, compared with the adjacent paneth cells, a high expression of Lgr5+ intestinal stem cells have high activity of mitochondrial oxidative phosphorylation [80,81,82].
Bonnay et al. reported that, upon an initiation of neural stem cell tumourigenesis, glutamine absorption was transformed into OXPHOS and increased NAD+ biosynthesis, which might be an important pathway for the initiation of neural stem cell tumourigenesis [83]. That is, OXPHOS produced by excessive mitochondrial fusion and increased oxidation reaction and NAD+ may initiate tumourigenesis [83,84]. This result is surprising because tumours are currently widely believed to be glycolytic. The importance of oxidative phosphorylation in different tumours has been reported in several studies [85,86]. For example, Transcription Factor A mitochondrial (TFAM) plays a key role in mitochondrial activity and gene transcription, and the deletion of TFAM leads to the loss of the self-renewal ability of stem cells and subsequent cell canceration [87]. Tumour stem cells from the brain, pancreas, lung, and ovary rely on oxidative phosphorylation [88,89,90,91,92,93].
However, tumourigenesis initiated by OXPHOS has created new problems. First, as mentioned above, researchers have found that molecules and events related to glycolysis are involved in tumour initiation. In addition, when stem cells develop into tumour stem cells, they not only have the ability of unlimited proliferation, but they also “evolve” a series of molecular mechanisms to escape death [94], which, under non-mechanical pressure, is vitally important. Although mitochondria are required to initiate cell death, they can also induce iron death and cell apoptosis [95,96]. Mitochondria are also the main source of reactive oxygen species (ROS) which causes cell damage [97]. Relevant reports have confirmed that the inhibition of the glycolytic pathway can induce apoptosis of GSCs [98,99], which is caused by ROS produced through OXPHOS [100]. Therefore, the maintenance of mitochondrial oxidative metabolism may not be conducive to tumour cells’ long-term growth and proliferation in complex and changeable internal and external environments [96]. Moreover, Bonnay’s study did not clarify how the increased oxidation reaction led to the immortality of tumour cells. They also did not completely exclude the high levels of tumour-expressed glycolytic enzymes. Furthermore, Lgr5+ ISCs have a lower level of pyruvate oxidation and a higher level of glycolysis [101], which suggests that glycolysis is highly related to tumour initiation. These results suggest that stem cell tumourigenesis is more complex than previously thought.

2.2. Implications and Significance of Research on Metabolic Plasticity of Tumour Stem Cells

It has long been considered that glycolysis and mitochondrial oxidative phosphorylation independently promote tumourigenesis [102]. However, it is not difficult to see from the above description that many difficulties and doubts exist in exploring which metabolic phenotype participates in tumourigenesis. Therefore, research on these issues has provided us with more thought and inspiration to further understand the occurrence and development of tumours.
Glycolysis and OXPHOS may exert different effects on the initial stages of tumourigenesis and execute different functions. At the earliest stage of tumourigenesis, mitochondrial dysfunction may be an additional factor causing increased gene instability [103] and plays an important role in cell transformation. For example, poly [ADP-ribose] polymerase(PARP)-mediated poly-ADP ribosylation and sirtuin-mediated deacetylation can have important effects on genome stability and cell fate and are key factors that cause cell canceration [84]. However, it cannot be concluded that OXPHOS causes cell transformation at the early stage of tumourigenesis because the study also showed that the inhibition of OXPHOS does not affect the accumulation of tumour stem cells [83]. During tumourigenesis, functionally activated glycolysis is important for the cell’s energy supply, biosynthesis, and escape from death. Many studies have shown that glycolysis cannot only make up for the lack of energy supply caused by mitochondrial dysfunction, but also avoid ROS-induced apoptosis caused by OXPHOS during the initial stage of tumour development t [100,104,105]. Glycolysis can provide more substrate sources for promoting further development of tumour cells [106,107,108]. Therefore, abnormal changes, or the overactivation of OXPHOS, may be an important transformation process in tumorigenesis initiation [109,110,111] as the transformed cancer cells have strong growth demand, and the energy supply of glycolysis is far from sufficient. In contrast, OXPHOS can create the necessary conditions for maintaining the sustained growth and immortality of tumours by promoting aspartate synthesis, subsequent nucleotide metabolism, and other biochemical reactions [83,85,86]. OXPHOS is involved in the immortalization and further transformation of TIC, and reducing the level of OXPHOS can affect genes associated with the growth of cancerous tumours [83]. Furthermore, glycolysis may be required for cell differentiation, basic metabolism, and the preparation for proliferation or adaptation to the external environment [100,101,112] (Figure 1).
Both glycolysis and OXPHOS may be involved in the process of stem cell development into tumour stem cells, equipping tumour stem cells with the potential for metabolic heterogeneity. Many factors affect the survival and development of tumours, such as hypoxia, nutrition, and microenvironment [113,114,115]. Therefore, metabolic heterogeneity during tumourigenesis may be an adaptation to the environment [116], which may be one of the important reasons for detecting different tumour metabolic types. For example, regardless of whether the glioma microenvironment is close to the blood vessels that are supplied with sufficient oxygen area or the hypoxic regions at the distal end of blood vessels, all these conditions can accommodate CSCs [117,118,119], which indicates the presence of at least two metabolic phenotypes in CSCs [102,113]. Isogenic murine-induced single stem cells are built into GSCs (Ink4a/Arf–null, H-RasV12-expressing murine GICs), while CSCs are induced artificially to achieve the initial tumourigenesis conditions to observe the metabolic phenotype at the initiation to promotion stage. The results showed that there were two metabolic phenotypes: one is under homeostatic conditions; CSCs dependent on OXPHOS could switch their metabolism to glycolysis in oxygen availability, which confirmed the metabolic plasticity of a single tumour cell [120]. Another is that CSCs dependent on glycolysis have a higher affinity for hypoxic niches, but CSCs dependent on OXPHOS have a better reoxygenation affinity that shows a preference for perivascular regions, which indicates that heterogeneous tumour cells with different metabolic phenotypes in the tumour stem cell population exist [120]. Therefore, the metabolic heterogeneity exhibited by this plastic hybrid metabolic phenotype is an important metabolic regulatory basis for cancer cells to adapt to different microenvironments and meet their vigorous growth and proliferation [121,122,123,124] (Figure 1).
Third, the results may be unreliable if the complex interrelationship between glycolysis and OXPHOS is not fully or systematically clarified. On the one hand, even if the marker molecules of glycolysis or OXPHOS are significantly increased, it does not mean that their metabolic level is increased. For instance, some studies have reported that lactate dehydrogenase (LDH) is upregulated after cell cancerization. Following this concept, the upregulation of LDH may be more inclined to the increasing level of glycolysis, and related research and treatment will also be inclined to it [125,126]. However, later studies indicated that high levels of LDH did not cause the accumulation of lactic acid and extracellular acids [127]. They also found that LDH produced pyruvate from lactic acid to fuel the Tricarboxylic acid (TCA) cycle, while lactic acid originated from tumour cells driven by glycolysis [128,129,130]. The same results were also found that lactate secreted by Paneth cells can provide a carbon source for Lgr5+ ISCs to meet mitochondrial energy demands [131]. In addition, some researchers use isotopes to label some nutrients (for example, C13-labelled glucose) in some cells that use glycolysis as metabolic energy. It has been found that the intermediate products of the TCA cycle mainly come from labelled glucose (13C-glucose), pyruvate carboxylase (PC), and pyruvate dehydrogenase (PDH), which help to enhance the activity of the TCA cycle [132,133]. This is noteworthy because these tumour cells have the capacity for glycolysis as an energy source, indicating that the material source of the TCA cycle may be directly or indirectly derived from glycolytic intermediates and demonstrating that glycolytic tumour cells may realize their anabolic demand for biosynthesis by using the TCA cycle [132] (Figure 2). In addition, to study the occurrence and development of tumours from the metabolism perspective, focusing only on one point may not achieve the expected drug intervention effect. For example, the glycolytic level of glioma stem cells isolated from surgery is lower than that of differentiated progeny [134]. However, the inhibition of glycolysis with dichloroacetate could promote mitochondrial oxidative phosphorylation, which might result in an increasing effect on cell viability and number [135].
From the above description, we know that there is complex metabolic heterogeneity and regulation of metabolic plasticity in tumour cells. Metabolic heterogeneity indicates the presence of various metabolic phenotypes, whereas metabolic plasticity indicates that metabolic phenotypes can be mutually transformed. Therefore, a full understanding of the key approaches in the process of tumour metabolic transformation by mining the key molecules involved in the process of mutual transformation of glycolysis and OXPHOS may provide a clearer explanation of the above issues and provide better research means and more objective theoretical and experimental basis for the mechanism and treatment of tumour progression. Relevant studies have also been conducted. For example, a hybrid metabolic phenotype has been observed in triple-negative breast cancer (TNBC) cells. The authors developed a mathematical model to unearth node proteins that regulate the metabolic phenotype and found that AMPK and HIF-1α are core regulatory loop molecules involved in tumour metabolism. By analyzing the expression characteristics of these two proteins, they found that tumour cells display three stable metabolic states: W (HIF-1high/pAMPKlow), O (HIF-1low/pAMPKhigh), and W/O (HIF-1high/pAMPKhigh), corresponding to the glycolytic phenotype, OXPHOS phenotype, and hybrid phenotype [136], which enables tumour cells to switch between OXPHOS and glycolytic phenotypes. By inhibiting the expression of AMPK and HIF-1, the ability of metabolic plasticity of cancer cells can be effectively eliminated. In this model, the team extensively studied the diversity of tumour metabolism and realized a more effective killing of tumours by mining node molecules that regulate metabolism and the use of targeted drugs, which may become a practical and reliable way to solve the clinical drug resistance of tumours.

3. Molecules Affecting Tumour Glycolysis and OXPHOS Transformation and Significance

3.1. Effect of Glycolytic Rate-Limiting Enzyme on Metabolic Plasticity

Glycolysis is mainly regulated by rate-limiting enzymes that catalyze three irreversible reactions: hexokinase (of the four HK isoenzymes, HK-2 has the most significant correlation with tumour cells) [137,138,139,140], phosphofructokinase (PFK), and pyruvate kinase (PK). Once glucose enters the cell, glycolysis is catalyzed by HK, and PK catalyzes the “export”. PFK is the most important rate-limiting enzyme of glycolysis, which can quickly react with ATP/ADP, citric acid, and other alternative fuels, such as fatty acids, ketones, and hormones, to regulate the whole process of glycolysis [141].

3.1.1. Hexokinase

Several in-depth reports on the effects of the three rate-limiting enzymes on metabolic plasticity are available. Early studies found that in some non-cancer cells that were dominated by glycolysis, the deletion of HK-2 would inhibit the glycolysis pathway, leading to 13C-glucose accumulation in the cytoplasm. Then, the number and size of mitochondria expanded, the TCA metabolic pathway was enhanced [142,143], and the activity of mitochondrial electron transport chain also increased [144,145].
However, this does not imply that the cell alleviates the metabolic pressure caused by the lack of HK-2 through the above metabolic pathway. HK-2 is not only the first rate-limiting enzyme of glycolysis, but it can also bind to the voltage-dependent anion channel (VDAC) on the outer membrane of the mitochondria. Mitochondrial VDAC is a channel for the binding and release of many apoptotic molecules, such as Cytochrome c and BCL2-associatedx(Bax) [146,147,148]. The combination of HK-2 and VDAC can, utilizing ATP produced by still-functioning mitochondria, realize the role of self-phosphorylation activation [149]. However, combining these two proteins can prevent some pro-apoptotic molecules from binding to VDAC, thus avoiding the initiation of mitochondrial apoptosis [150]. Therefore, the regulation of the metabolic plasticity process from glycolysis to OXPHOS, caused by the deletion of HK-2, is accompanied by the production of more ROS, the release of apoptosis-promoting molecules, and the accumulation of metabolites in the TCA cycle, which are also signs of mitochondrial dysfunction and apoptosis induction [151,152,153]. Based on this, some scientific researchers have developed precise combinations of drugs with HK-2 inhibitors, which have produced more effective killing effects on tumour cells. In general, the conversion of glycolysis to OXPHOS, caused by HK-2 inhibition, can be targeted by tumour drugs in two ways. One example is increasing tumours’ sensitivity to apoptosis during this process. Zhang et al. found that the selective inhibition of HK-2 expression by follicle-stimulating hormone (FSH) peptide-aligned PEG PEI copolymers loaded with HK2 shRNA could switch the metabolism from glycolysis to OXPHOS. In contrast, the expression of pro-apoptotic molecules increased. Therefore, treatment with chemotherapy drugs, such as cisplatin, can improve the killing effect of anticancer drugs on ovarian cancer [154]. Another example is changing the drug resistance of tumours. Dannielle et al. reported that the loss of HK2 could lead to an increase in OXPHOS, and the use of metformin in this process could simultaneously inhibit the activity of Complex I in OXPHOS, which can considerably reduce the drug resistance of tumour cells to sorafenib [155]. This observation highlights the impact of HK on metabolic remodelling and potential treatment options.

3.1.2. Phosphofructokinase

In contrast to HK, which is the most important rate-limiting enzyme of glycolysis, targeted drugs acting solely on PFK may have the same anti-cancer effect [156]. For instance, mitochondrial VDAC can bind not only with HK-2 but also with PFK [157,158]. The combination of VDAC and PFK has a significant inhibitory effect on PFK-mediated glycolysis [157], thus affecting the metabolic phenotype of tumour cells.
As previously reported, the metabolic phenotypes of GSCs and non-stem tumour cells (NSTCs) are very different. GSCs are more inclined to glycolysis than NSTCs, and their mechanism is realized through VDAC. NSTCs express higher levels of VDAC, which can be combined with PFKP, a dominant isoform of PFK in human glioblastoma (GBMs), making it bind to the outer membrane of mitochondria, thus inhibiting the expression of phosphofructokinase, platelet (PFKP) in the cytoplasm and inhibiting its glycolysis reaction [157,159]. Once VDAC is lost, PFKP is released, which activates glycolysis and causes the conversion of the metabolic phenotype [157]. Using this mechanism, clotrimazole, a PFK inhibitor, can alleviate the activation of glycolysis caused by VDAC disruption, thereby preventing the metabolic rearrangement process and phenotypic transformation of GSCs, as well as significantly inhibiting the growth and invasion of glioma cells [157]. This treatment mode has transformation significance because GSCs have strong drug resistance to current radio and chemotherapeutic regimens due to their strong cell plasticity regulation, which limits the treatment of many antineoplastic drugs.
In addition, avoiding apoptosis by regulating metabolism is a feature of tumour survival [160]. The Warburg effect is the cornerstone for tumours to realize the above process [161]. Therefore, by targeting the regulation of PFK expression to inhibit glycolysis and to improve the level of OXPHOS in a variety of tumour cells (such as pancreatic ductal adenocarcinoma (PDAC), non-small cell lung cancer (NSCLC) cells, and CRCs), improve the release of ROS and apoptosis-promoting molecules, and, finally, enhance the sensitivity of anti-cancer drugs, good anti-tumour effects have been achieved [158,162,163]. For example, DNASE1L3 (also named DNase1) is a DNA-degrading enzyme expressed at low levels in high-level glycolytic tumour cells [164]. A high expression of DNASE1L3 can induce tumour apoptosis [165], which is considered a prospective and personalized gene-based therapy for cancer [166,167,168]. The addition of DNASE1L3 can inhibit PFK1-mediated glycolysis and induce tumour apoptosis by activating the P53 pathway, thus inhibiting tumour growth, invasion, and migration [168]. Based on the above mechanism, human recombinant DNase1 (hrDNase1) can achieve high killing efficiency in tumour cells resistant to radiotherapy and chemotherapy [169]. Another example is that TNF receptor-associated Protein 1(TRAP1), a heat shock Protein 90 (HSP90) molecular chaperone that is upregulated in human CRCs and responsible for the downregulation of OXPHOS [163,170], can combine with PFK1, avoid the ubiquitination of PFK1, and enhance glycolysis activity, thus producing resistance to some drugs against glycolysis, such as cetuximab [163]. The inhibition of TRAP1 reduces its binding with PFK1, triggering the degradation of PFK1 through the ubiquitin-proteasome pathway. The downregulation of TRAP1 expression induces plasticity regulation, ultimately enhancing the sensitivity of CRCs to cetuximab [163].

3.1.3. Pyruvate Kinase

Pyruvate kinases have several isoforms. Compared with other pyruvate kinases, the M2 isoform of pyruvate kinase (PKM2), which is a specific PK of tumour cells, is significantly upregulated in tumours [171,172].
Compared to the above two rate-limiting enzymes, the effect of PKM2 on metabolic remodeling is more complex. The presence of PKM2 in tumour cells can inhibit OXPHOS by inhibiting the expression of monoubiquitinated histone H2B (H2Bub1) and enhance pyruvate phosphorylation in glycolysis to influence metabolic rearrangement. H2Bub1 can regulate a series of mitochondrial genes used for OXPHOS, thereby affecting the production of mitochondrial ATP and inhibiting glycolysis in human lung cancer cells (H1299 and A549 cell lines). PKM2 can combine with it to inhibit the expression level of H2Bub1 [173]. However, tumour cells can reduce mitochondrial fragmentation, promote mitochondrial fusion, protect mitochondrial function, and enhance OXPHOS by promoting PKM2 expression [174,175].
Therefore, PKM2 may have different beneficial effects on glycolysis and mitochondrial oxidation. The different functions of PKM2 may not depend on its specific protein activity patterns. PKM2 forms different polymers, including dimers and tetramers. Tetramers are the main functional and active units of PKM2 that play a role in glycolysis, providing energy for tumour growth and proliferation. Inactivated PKM2 is a dimer that can be oxidized and phosphorylated, resulting in the loss of catalytic function during glycolysis. However, it can enter the nucleus and combine with other transcription factors to exert transcriptional activity [176,177] and participate in the transcription and regulation of many molecules, affecting tumour proliferation, migration, and death [178,179]. For example, Dimer PKM2 can promote the release of exosomes through phosphorylating synaptosome-associated Protein 23 and affect the signal transduction between tumour cells and cells, and the microenvironment [180]. By inhibiting the encoding of E-cadherin and increasing the nuclear transcription of signal transducer and activator of Transcription 3 (STAT3), Dimer PKM2 can lead to a loss of EMT in tumour cells and promote tumour cell invasion and migration [181,182]. In addition, Dimer PKM2 can also mediate the disruption of NF-κB/miR-148a/152 and promote the expression of vascular endothelial growth factor (VEGF) as well as the activation of insulin like Growth Factor 1 receptor (IGF-IR), thereby promoting tumour angiogenesis [183,184,185,186].
Therefore, when PKM2 is used for drug intervention, more targeted measures are required to achieve better results. The oxidation and phosphorylation of PKM2 result in its nuclear localization and interaction with HIF-1α, which causes EMT-related molecules’ expression and increases tumour cell invasion [187]. Under these conditions, tumour cells are more likely to undergo OXPHOS reactions through lactic acid produced by cancer-associated fibroblasts (CAFs) dominated by glycolysis in the tumour microenvironment [187,188]. Using DASA, which can reactivate PKM2, could avoid its nuclear localization and enable tumour cells to restore glycolysis and reduce its function in EMT transcription. However, tumour cells can still maintain sufficient OXHPOS levels. When tumour cells are cultured in a non-sugar environment and treated with metformin, an inhibitor of mitochondrial complex I(COXI) activity, the growth and proliferation of tumour cells can be significantly inhibited [187]. Additionally, researchers have found that some tumour cells with high levels of both OXPHOS and glycolysis conduct multiple drug interventions based on metabolic phenotypes [175]. For example, the use of MLN4924, a small-molecule inhibitor of neoplasia modification, could induce the conversion of mitochondrial fission to fusion, thereby inhibiting the TCA cycle. Shikonin can inhibit the tetramer PKM2-mediated glycolysis reaction. This drug combination has achieved good results in both in vivo and in vitro experiments [175]. Due to its anti-cancer effect, MLN4924 is used as a clinical medication in Phase I/II [189]. Based on the mode of regulation of glycolysis-related molecules at the entry point, changing the metabolic phenotype of tumour cells and then using drug intervention according to the molecular regulation mechanism can be used to develop precision therapy.

3.2. Effects of Key Enzymes of Mitochondrial Oxidative Metabolism on Metabolic Plasticity

Mitochondrial oxidative phosphorylation is the ultimate destination for oxygen molecules during aerobic metabolism [190]. Energy substances are metabolized through the TCA cycle, which is the key link that causes a significant increase in OXPHOS levels [191,192,193]. Abnormal changes in key enzyme molecules, such as mutations, abnormal expression, and activity damage, lead to the accumulation of Krebs cycle intermediates [194,195], thus affecting the vigorous and diversified growth needs of tumour cells [196,197]. At the same time, the metabolic process has an important impact on the key molecules and events of metabolic transformation, such as HIF-1, AMPK, and ROS, to promote the metabolic transformation between glycolysis and OXPHOS [198,199,200,201]. Owing to space limitations, this article lists three key enzymes of the TCA cycle, namely citrate synthase (CS), isocitrate dehydrogenase (IDH), and α-ketoglutarate dehydrogenase complex(α-KGDC), to describe the influence of related enzymes on the metabolic plasticity process and the status of some targeted drug therapies.

3.2.1. Key Enzymes of the TCA Cycle

Citrate Synthase (CS)

CS can convert Acetyl Coenzyme A and oxaloacetic acid to citric acid. Citric acid is an important intermediate that affects metabolic transformation and can produce an important allosteric inhibition effect on PFK1/2 [202,203], thereby reducing glycolysis [204]. The insufficient energy supply caused by the inhibition of glycolysis due to the excessive increase in citric acid can be attributed to mitochondrial oxidative metabolism. In mitochondria, citric acid is a key regulator of mitochondrial oxidation enzymes, such as successful dehydrogenase (SDH), PDH, and Carnitine Palmitoyl Transfer I (CPT) [203,204,205,206,207]. However, as previously mentioned, the increase in the TCA cycle and the enhancement of OXPHOS caused by metabolic remodeling may also be features of mitochondrial dysfunction in tumour cells, and citric acid plays an important role in this process [151,152,153]. Citric acid can cause a decrease in the expression of anti-apoptotic molecules, such as BCL-2 and myeloid cell leukaemia 1(Mcl-1) [208,209,210], and an increase in the expression of apoptosis-initiating and executing molecules, such as caspase family-related proteins (caspase-2/3/8/9), which finally inhibit tumour cells [211,212].
Therefore, the inhibition of different types of tumour cells by regulating citrate synthase and its catalytecitric acid has been reported in many studies. For instance, oral citric acid has a good inhibitory effect on lung cancer [213,214], thyroid cancer [215,216], and pancreatic cancer [217]. Citric acid has a synergistic anti-tumour effect with some drugs, such as 3-bromo-pyruvate [212], cisplatin [210], and celecoxib [218], which can effectively improve the lethality of anti-cancer drugs. The effectiveness of anti-cancer drugs can also be greatly improved if metabolic plasticity is considered. For example, when T98G glioma cells were cultured in a medium containing galactose and an increased concentration of glucose (DMEM GAL medium), the citrate synthase activity was increased by 26%, and the corresponding citric acid was also increased. Correspondingly, TCA increased, and the level of glycolysis was weakened, reflecting the metabolic remodeling process from glycolysis to OXPHOS [205]. The authors believe mitochondria may become more sensitive to some mitochondria-targeted drugs under this condition. Therefore, the MI–D compound [4-phenyl-5-(4-nitro-cinnamoyl)-1,3,4-thiadiazolium-2-phenylamine], a potential candidate mitochondria-targeting drug [219], was used. They found that the MI–D compound can effectively enhance the activity of citrate synthase and citric acid and induce tumour metabolic remodelling, and, correspondingly, the tumour-killing effect of the drug is more obvious [205]. The above results explain the mechanism by which MI-D, an anticancer drug, can produce a more effective tumour-killing effect under the condition of inducing tumour metabolic remodelling. The above results reflect the effective antitumour effects of anti-tumour drugs under metabolic remodeling.

Isocitrate Dehydrogenase (IDH)

The effect of IDH on tumour cells mainly depends on the mutation state. Early reports stated that more than 70% of IDH mutations have been found in the middle- and high-grade (Grade II–III) astrocytomas, oligodendrogliomas, and secondary glioblastomas [220,221,222,223]. Later, IDH mutations were found in the early stages of some tumour cells, such as acute myeloid leukaemia and solid tumours [224,225,226,227]. Based on its important diagnostic value after mutation, it was defined as the classification standard for central nervous system tumours by the WHO in 2016 [228]. In the latest classification standard of central nervous system (CNS) tumours by the WHO in 2021, invasive glioma was clinically classified according to IDH1/2 mutations [229]. Therefore, to explore the effect of IDHs on tumour metabolic remodeling, we should focus on the effect of IDH mutations. Mammals mainly express IDH1, IDH2, and IDH3. IDH1 and IDH2 have more than 70% homology, and a recent report found that IDH1/2 is closely related to tumours. Wild-type IDHs (WT-IDHs) can catalyze the conversion of isocitrate to α-ketoglutarate (α-KG), CO2, and a certain amount of NADPH. The mutant IDH, especially mutation at the R132H site, can greatly change the enzyme activity, enhance its ability to bind to NAPDH, and reduce its affinity for isocitrate, which ultimately leads to its inability to catalyze isocitrate to α-KG and make the (R) enantiomer of 2-hydroxyglutarate(R-2HG) generate and release [230,231,232,233].
R-2HG is an oncometabolite [234,235,236,237] and is one of the best small molecule biomarkers for tumours [238,239]. R-2HG can cause abnormal expression of many molecules related to metabolic enzymes. For example, many studies have reported that R-2HG increases HIF-1α expression through its inhibitory effect on PDH, then enhances the HIF-1α-mediated glycolytic pathway [240,241]. However, it has also been reported that R-2HG can activate PDHs and degrade HIF-1α [242]. In addition, it has been reported that R-2HG can inhibit aerobic glycolysis by abolishing the FTO/m6A/YTHDF2-mediated post-transcriptional modification of PFKP and Lactate Dehydrogenase B (LDHB) [243]. In addition, R-2HG can also make mitochondria more susceptible to glucose starvation by inhibiting ATP production through its inhibitory effect on the oxidative phosphorylation of mitochondria [244].
However, the effect of IDH1/2 mutations on metabolic remodeling is not achieved only by inducing R-2HG. Many reports have shown that IDH1/2 mutation can increase the level of lactate, such as lactate levels in mutIDH1 (R132H) in U87 GBM [245], mutIDH1 (R132H), and mutIDH2 (R172K) in Grade II and III glioma [246]. Other reports have shown its inhibitory effect on lactate, such as mutIDH1 (R132H) in U87, NHA, and LN18 cells [247,248,249], and not changed in many other reports, such as mutIDH2 (R172K) in HOG cells [250]. Pyruvic acid has also been reported differently in relevant studies, such as unchanged studies on mutIDH1 (R132H) glioma and PDX mouse models, as well as in mutIDH1(R132H) expressing LN18 or HOG cell lines [249,250,251,252], and decreased studies, such as in mutIDH1 (R132H) PTBs [252,253]. In the TCA cycle, the mutation of IDH1/2 can also cause inconsistent changes in some metabolic intermediates, such as 2-OG, citrate, cis-aconite, isocitrate, and fumarate [245,249,250,251,252,253,254]. However, compared with WT–IDH cells that are dominated by glycolysis [255,256], IDH mutant tumour cells have a stronger TCA cycle [256,257,258], more OXPHOS [259], and lower glycolysis levels [256,260,261]. This may be related to using different carbon sources after the IDH mutation. Tumour cells with IDH mutation did not use glucose for oxidation [248,256,262,263]. Instead, they used lactic acid and glutamate as substrates for TCA metabolism [248,256,264] and metabolized pyruvate by increasing the expression of PC, which increased the production of oxaloacetate and succinate [258,265], thereby meeting the needs of the electronic transfer chain.
The different lactate levels may also be related to the uptake and release of lactic acid in IDH mutation tumour cells. Tumour cells with IDH mutation will express less MCT1/4, LDHA also decreases, and, thus, the production of lactic acid is reduced. However, pyruvate cannot be obtained from glycolysis due to the IDH mutation, which needs to be obtained from the tumour microenvironment or intracellular metabolism of lactic acid to provide a carbon source for TCA. Researchers have proposed that to study IDH1/2 mutant tumour cells, it is necessary to observe the dynamic intracellular and extracellular lactic acid content, but the current detection methods are not well-implemented [248,251,252].
Therefore, for IDH-mutated tumour cells, more intelligent treatment measures may be needed to achieve better drug intervention effects. For example, studies have shown that for acute myeloid leukaemia (AML) patients with IDH1 mutation, the use of drugs targeting the inhibition of mutant IDH1 alone may cause patients to be insensitive to drugs, mainly due to the inhibition of mutant IDH1, which could also cause the improvement of glycolysis [266,267]. Therefore, some researchers have identified several characteristic genes related to glycolysis by constructing IDH1-mutant GBM cells. Then, the risk scores of these genes were integrated and analyzed to correlate them with chemotherapy drugs, such as cells with mutant IDH1 having better sensitivity to doxorubicin and sorafenib [268]. A combination of glycolysis and mutant IDH inhibitors has been applied to treat other tumours to achieve better therapeutic efficacy [243,259].
However, the treatment mode cannot be realized simply by using multiple drugs, and certain conditions may need to be created before using a combined drug. For example, some tumour cells, such as IDH1-mutated CSCs, are mainly metabolized by OXPHOS. However, because they are in a long-term hypoxic state, drugs targeting OXPHOS are ineffective and retain glycolysis ability simultaneously. Therefore, the author proposes removing cells from the hypoxic area and then using the corresponding radiotherapy and chemotherapy methods, which can produce more ROS. This treatment has achieved good results [269,270]. In addition, other researchers have developed new drugs, such as BAY 87-2243 (B87), a special inhibitor of HIF-1α [271], which can also inhibit the activity of Complex I [272]. The oxygen consumption rate (OCR) level of cells decreases after B87 treatment, which reflects the inhibition of OXPHOS [273,274]. Although the drug can inhibit the proliferation of tumour cells, B87 is not very effective in inducing tumour cell death [274]. Later studies found that tumour cells treated with B87 had obvious anti-tumour effects under the environment of reduced glucose or combined with some glycolysis inhibitors, such as 2-deoxyglycose, indicating that the cancer cells treated with B87 were saccharophilous [20,275]. Inspired by this, researchers used synergistic molecules/drugs for B87 drugs and found that the α-KG precursor dimethyl-α-ketoglutarate (DMKG) can produce a strong tumour-killing effect together with B87 [274].
In addition to the above examples, compared with mutant IDH tumour cells, wild-type IDH has higher levels of glycolysis and TCA [255,276], which mainly depends on glucose and glutamine metabolism to generate pyruvate and α-KG [248,277,278]. Moreover, the survival rate of patients with wild-type IDH tumours is not as good as that of IDH mutant patients [279,280], which may be related to the different responses of these tumour patients to IDH inhibitors [267]. One of the more important mechanisms is that IDH can produce antioxidants and promote oxygen-sensitive signal transduction pathways [281,282,283]. However, for mutant IDH tumour cells, due to a lack of antioxidants and a high level of the TCA cycle, these tumour cells could produce more ROS [284,285], which leads to an increase in the sensitivity of radiotherapy and chemotherapy to mutant IDH patients, but not to the susceptibility of wild-type IDH tumour cells [270,286]. Therefore, according to the metabolic characteristics of IDH wild-type tumour cells, using drugs that induce ROS can produce better anti-cancer effects [286].

α-Ketoglutarate Dehydrogenase Complex(α-KGDC)

α-KGDC is a key regulatory point in the tricarboxylic acid cycle and is mainly used to catalyze α-KG, succinyl coenzyme A, and CO2, as well as convert NAD+ to NADH, which can be used later in the respiratory chain complex to produce ATP. Although many factors, such as pH, calcium ion concentration, and ROS [259], can regulate the activity of α-KGDC, current research on the direct effect of α-KGDC on metabolic plasticity is rare. The existing reports showed that the most important effect of α-KGDC is the mutual feedback loop regulation mode between its metabolic substrate α-KG and HIF-1. Under normal oxygen conditions, α-KG produced by α-KGDC can activate PHDs and hydroxylate HIF-1α, while hydroxylated HIF-1α can be degraded by the proteasome pathway [287,288,289,290]. However, in the absence of oxygen, the production of α-KG is limited, thus inhibiting the activity of PHDs. The degradation of HIF-1α decreases and its accumulation increases, which then targets the regulation of the E1 subunit of α-KGDC so that the E1 subunit is ubiquitinated under the action of ubiquitin-protein ligase SIAH2, resulting in its degradation [288,291]. The inhibition of α-KGDC leads to α-KG failure to persist in the mitochondrial TCA cycle. Instead, it enters the reductive pathway of the TCA cycle to produce acid, which ultimately promotes lipid synthesis. HIF-1α is an important promoter of glycolysis, which has a positive transcriptional regulatory effect on various glycolytic molecules, such as HK2, GLUT1, and PDK1 [292,293,294]. Therefore, the HIF-1α mediated regulation of α-KG function under hypoxia may have a potential anti-tumour effect [295,296].
In addition, the inhibition of α-KGDC activity leads to the accumulation of α-KG, which also affects the regulation of autophagy. Report presents that a decrease in mitochondrial oxidative phosphorylation is accompanied by a low level of α-KG, while a high level of α-KG can inhibit autophagy through its target regulation of rapamycin 1 signaling [297,298]. Autophagy can affect the process of metabolic plasticity in many aspects. For instance, the polyubiquitination of HK-2, catalyzed by E3 ligase TRAF6, can be recognized by autophagy Protein P62 before initiating the degradation enzyme of the autophagy pathway, thereby inhibiting glycolysis [299]. Mitochondrial autophagy can reduce glycolysis, increase OXPHOS, increase cell apoptosis, and inhibit ectopic metastatic growth of cancer in a P53-dependent manner [300]. Moreover, some studies have shown that the significance of autophagy induction lies in the fact that tumour cells can alleviate survival disorders caused by metabolic and therapeutic stress [301,302,303,304,305]. However, regardless of how autophagy affects tumour cells, in the selection of tumour-targeting drugs affected by α-KG, we can consider whether or how the activation process of autophagy is by α-KG. Then, we can use some small molecule drugs to target autophagy [306] and combine them with drugs to α-KG, such as JHU-083 [276], which may have a better curative effect.

3.3. Other Molecules Regulating Metabolism Plasticity Processes

In addition to the effects of key metabolic enzymes on metabolic plasticity, some molecules of non-rate-limiting enzymes also have an important impact on the metabolic plasticity process, thus providing potential value for tumour treatment (Figure 3). For example, FBP2, the isoenzyme of FBP (fructose-1, 6-bisphosphatase), is a molecule that can regulate the rate-limiting enzyme of glycolysis and inhibit the proliferation of soft tissue sarcomas. There are two mechanisms involved in its inhibition process. A high expression of FBP can inhibit HIF-1α, interfere with glycolysis, and hinder tumour cell proliferation, invasion, and migration [307]. While FBP2 can also be localized in the nucleus, with c-Myc, it exerts an inhibitory effect by binding to the promoter of mitochondrial transcription factor nuclear respiratory factors (NRFs) and TFAM, thereby limiting the expression of mitochondrial genes and reducing OXPHOS energy supply. This dual inhibition function provides a good direction for tumour treatment [307,308].
Another example is the influence of some non-key enzymes of the TCA cycle and mitochondrial proteins on metabolic plasticity. For instance, mutations in SDH in hereditary paragangliomas and fumarate hydratase (FH) in leiomyomatosis and kidney cancers lead to the accumulation of succinate and fumarate [309,310], thus reducing the degradation of HIF and promoting the expression of multiple glycolytic enzyme genes [311]. In renal cell carcinoma, the inhibition of HSP60, a chaperone protein that plays an important role in mitochondrial protein homeostasis, can strengthen the glycolysis reaction, reduce OXPHOS and the transition from glycolysis to OXPHOS, and enhance de novo nucleotide synthesis and lipid synthesis. Therefore, high HSP60 expression is related to a better prognosis of renal cell carcinoma, which can lead to the inhibition of cancer cell growth and proliferation [312].
In addition, regulating related molecules in the electron transport chain pathway is also one of the ways to change tumour metabolism. By directly regulating decaprenyl diphosphate synthase subunit 2 (PDSS2), a mitochondrial metabolic molecule that plays a crucial role in regulating the synthesis of coenzyme Q-binding protein 10 (CoQ10), the activity of CoQ10 and mitochondrial Complex I can be increased, and the TCA cycle activity can be enhanced, thus inducing the conversion of glycolysis to mitochondrial respiration in hepatoma cells. This effect can inhibit foci formation, colony formation in soft agar, and tumour formation in nude mice [313].
In addition, the stability of genes, such as tumour protein p53 (TP53), KRAS, kelch-like ECH-associated protein 1 (Keap1), serine/threonine kinase 11 (STK11), EGFR, neurofibromin 1 (NF1), and B-Raf proto-oncogene, serine/threonine kinase (BRAF) is related to metabolic changes [314,315,316], the most important of which is the effect of the p53 gene on metabolism. At present, the effects of P53 on metabolic remodeling have been studied in many aspects, such as TP53-induced glycolysis regulatory phosphatase (TIGAR), glycolysis, and glucose transporter; the synthesis of cytochrome c oxidase 2 (SCO2) and glutaminase 2 (GLS2); or protein interactions with metabolic enzymes, such as glucose-6-phosphate dehydrogenase (G-6-PD), peroxisome proliferative-activated receptor, gamma, coactivator 1 (PPARGC-1), and sterol regulatory element-binding protein (SREBP) [317,318]. Although for tumour cells, the effect of TP53 on metabolism plasticity has different aspects, many studies have shown that TP53 can inhibit tumours by inhibiting glycolysis and promoting OXPHOS [319]. Another study reported that TP53 could suppress tumorigenesis by inhibiting the conversion of OXPHOS to glycolytic processes during tumour initiation [318]. We believe that wild-type P53 (WTP53) is an anti-oncogene. The function of P53 in preventing the cell from canceration is related to its ability to inhibit OXPHOS. This regulation pattern of WTP53 could reduce harmful products, such as ROS, and obtain more stable genomic DNA, thus avoiding the initiation of tumorigenesis. It has a similar role on normal somatic cells that rely on OXPHOS [320,321,322,323]. However, due to the existence of a mutation in TP53 (MuTP53), the positive regulation of OXPHOS could be lost, causing the occurrence of tumours (Figure 4). The beneficial effect of WTP53 on cell survival has also been preserved in tumour cells, although more than 50% of human cancers have mutations in WTP53 [324]. In hepatocarcinoma (HCC), the mutation frequency of WTP53 is low [325]. By introducing WTP53, the mitochondrial pyruvate carrier (MPC) binding to pyruvate can be disrupted, which is realized by inducing BCL2 binding component 3 (Bbc3,also named PUMA) and inhibiting the uptake of pyruvate by mitochondria, thus promoting the survival of tumour cells and causing poor prognosis in HCC patients [326]. Therefore, these findings suggest that treatment or research measures for cancer should consider activating WTP53.

3.4. Summary

Studying the enzymes and related molecules directly involved in regulating glycolysis and mitochondrial oxidation may provide a more accurate perspective on regulating metabolic plasticity and developing effective therapeutic measures. The above section lists some promising reports that can be potentially used as drug therapies targeting metabolic plasticity (partial drugs are summarized in Table 1). Due to limited content, this article only lists the essential enzymes and introduces some important molecules (partial molecules are summarized in Table 2).

4. The Role of Metabolic Plasticity Regulation in the Tumour Microenvironment

Genes that regulate metabolic plasticity in tumour cells are the basis of metabolic remodeling in tumourigenesis. The tumour microenvironment is an important factor affecting metabolic plasticity, tumorigenesis, and development. In the advanced stage, the tumour volume at the primary site develops to a certain extent, and a microenvironment conducive to its survival has been formed. With further expansion of the tumour volume, the space and nutrition source of the primary site cannot meet the rapid growth of the tumour, and it begins to prepare for the next step, that is, diffusion; the process from promotion to progression is a stage in which tumour cells fight against immune cells and compete for nutrients in the surrounding environment to remodel and create the surrounding microenvironment (Figure 5).

4.1. Effects of Metabolic Plasticity Regulation on the Immune Microenvironment

Inducing the differentiation of naive CD8+ T cells is the primary anti-tumour mechanism of immune cells. Naive CD8+ T cells normally exist under quiescent conditions. Maintenance of this state is mainly mediated by two molecules: sphingosine 1-phosphate (S1P) and interleukin (IL)−7 [327,328]. S1P is important for OXPHOS [328,329,330], while IL-7 mainly promotes glucose uptake by GLUT-1, thus affecting the glycolytic process [331,332,333]. Glycolysis is primarily used for the activation of T cells, but OXPHOS is also important, and its deletion blocks the proliferation of T cells [334,335]. Once the anti-cancer signal is received, naive CD8+ T cells differentiate and become effector CD8+ T cells and memory CD8+ T cells. Compared with effector CD8+ T cells, memory CD8+ T cells have better mitochondrial activity [336]. Upon T cell receptor (TCR) and leukocyte surface differentiation antigen 28 (CD28) stimulation, acylglycerol kinase (AGK) can be combined with phosphatase and tensin homolog (PTEN), which leads to phosphorylation of the lipid phosphatase PTEN and loss of its activity, thereby activating the phosphatidylinositol 3-kinase (PI3K) mechanistic target of rapamycin kinase (mTOR)-signaling pathway, promoting its glycolytic metabolism, and differentiating into effector T cells [337], exerting its anti-tumour effect.
Overall, immune cells in the tumour microenvironment, and tumour cells, have similar metabolic modes. The two compete for resources to inhibit the other’s survival. At the initial stage, when resources are sufficient, T cells use glucose for glycolysis and mitochondrial metabolism to generate T cells with tumour-killing effects, which involves the activation of HIF-1α and/or Myc [10]. Cancer-associated antigens (CAAs) produced by the death of tumour cells can be taken up by dendritic cells (DCs) and transmitted to naive T cells in lymph nodes with tumour infiltration. After stimulation by CAAs, differentiated effector CD8+ T cells move and infiltrate the tumour area, recognize the tumour through the relevant antigens on the surface of tumour cells, and initiate the death of the tumour.
The dead tumour cells release CAAs and repeat the above process [338], called the cancer immunity cycle (CIC). Failure of the CIC process leads to the immune escape of tumour cells [20]. CIC realizes the above process through two points: immunogenicity of the tumour itself and the effector function of infiltrating T cells. Sufficient tumour mutations can increase immunogenicity. The uptake of glucose, glutamine, and other nutrients can increase the mutation effect [339,340]. In addition to the uptake of nutrients, which can be utilized by glycolysis, OXPHOS can also be used for its metabolism. At the same time, due to the specificity of mitochondrial function, the process of mitochondrial oxidation can produce ROS, causing more damage to tumour cells and inducing DNA mutations [341,342,343,344,345]. PI3K, Myc, RAS, and programmed cell death 1 ligand 1 (PD-L1) are the main pathway molecules involved in this mutation process [345,346,347]. In addition, aerobic glycolysis may have an inhibitory effect on tumour immunogenicity [348]. The transition of human OXPHOSS to glycolysis can reduce the occurrence of mutations caused by ROS (nuclear and mitochondrial) [349], lactic acid produced by glycolysis can reduce the immunogenicity of cells [37,350,351,352], and increased glycolysis can inhibit the synthesis of antigenic peptides by increasing cell cycle progression [353,354]. Therefore, the change in metabolic mode caused by various factors is an important inducer of increased tumour immunogenicity.
In addition, with the growth of a tumour and the intake of more glucose, the expression of high levels of Myc and HIF can increase the glycolysis flux of tumour cells, resulting in the excessive release of immunosuppressive molecules such as lactic acid and inhibitory cytokines in the TME, which has a significant impact on the activation of effector T and Treg cells. Lactic acid is an inhibitory molecule for all anti-tumour immune cells. Effector T cells in humans and mice lost their killing ability to tumour cells in the lactate microenvironment [35,355,356]. However, the high lactate environment can enable Treg cells to absorb lactate through monocarboxylate transporter 1 (MCT1) and enhance the expression of PD-1, thereby activating regulatory T (Treg) cells [34,357]. An increase in activated Treg cells can inhibit effector CD8+ T cells [358,359]. Therefore, when MCT1 on the surface of Treg cells is deleted, lactate importation can be disrupted, and anti-tumour ability can be improved [34].
The tumour microenvironment is a complex cellular environment. Besides lymphocytes, macrophages, CAFs, and natural killer (NK) cells are also important components [360,361]. Macrophages play a role in all stages of tumourigenesis and development [362]. A high level of lactic acid can make macrophages develop into M2-type macrophages, inhibit T cell activation and proliferation, and exert its immunosuppressive function by expressing arginase 1 (ARG1) protein [363,364,365]. Macrophages intaking glucose can develop into M1-type macrophages after receiving interferon gamma (IFN-γ) secreted by Th1 cells [364], and the anti-tumour ability of M1-macrophages can be enhanced. However, IFN-γ is unstable in an acidic microenvironment, thus inhibiting the differentiation of Th1 cells to Th2, which antagonizes the differentiation of M1-type macrophages and promotes tumorigenesis [366,367,368,369]. Macrophages affected by lactic acid can also secrete IL-6, increasing 3-phosphoinositide dependent protein kinase 1 (PDPK1)-dependent phosphoglycerate kinase (PGK1) phosphorylation by inducing PDPK1 in tumour cells, resulting in enhanced glycolysis and tumourigenesis [370]. Therefore, it is not difficult to see that the deprivation of surrounding nutrients and elimination of metabolites by the tumour affect the antitumour effect of macrophages from a long-term perspective.
Similar results were also found for NK cells. At the initial stage of KRAS-driven lung cancer, NK cells can exhibit strong cytotoxic effects and secrete cytotoxic factors to kill target cells directly [371]. This process is closely related to metabolic changes. That is, the activation of NK cells needs to switch their metabolism from mitochondrial oxidation to glycolysis [372]. However, for advanced tumour cells, the function of NK cells in tumour cells infiltrated in tumour cells, but not in blood circulation, is suppressed, reflecting the regulatory effect of the tumour microenvironment on NK cells. Mechanistically, FBP1 weakens the glycolytic levels of NK cells, thereby causing dysfunction [373,374]. While FBP1 is mediated by transforming growth factor beta (TGF-β) in the tumour microenvironment [375], CAFs are the origin of TGF-β, and CAFs are the most abundant cells in the tumour stroma that are recruited by platelet-derived growth factor secreted through the tumour [376]. The emergence of CAFs generally means that tumours occupy a dominant position in their environment through the activation of GLUT1 [188,377], thereby increasing glucose absorption and MCT4-mediated lactate release, finally remodeling tumour metabolism [187,188,378] and promoting tumour progression and invasion [379,380]. Therefore, it is not difficult to explain why NK cells can play a role in the initial stage but lose their role in the promotion and progression of tumour development (Figure 5).
The above interaction process between immunity and tumours makes it easy to see that immune cells play an inhibitory role in tumours and are effective in the early stage of tumour development. However, with the development of tumours, many immunosuppressive molecules are released to induce the differentiation of immune cells and inhibit their anticancer effects. Among these, the accumulation of lactic acid produced by metabolism plays an important role. Lactate is involved in glycolysis, an important component of the tumour microenvironment. Higher lactate levels can be found in advanced tumours and patients with poor prognoses [129,130]. Excessive lactic acid can inhibit the differentiation and maturation of T cells, DC and NK cells [381,382], and cytokine production [35,382]. However, lactate is not only produced as a by-product of glycolysis, nor is it an inhibitory effect on immune cells. It is an important tool for tumours to realize a metabolic adjustment to adapt to the environment and promote tumour progression. In the tumour microenvironment with sufficient nutrition, tumour cells release lactic acid by absorbing glucose, glutamine, etc., whereas in the tumour microenvironment with nutrition deprivation, lactic acid can enter the TCA cycle and be used as the metabolic source of tumour cells for energy supply. Moreover, lactic acid is more important than glucose under these conditions [129]. The accumulation of lactic acid makes the tumour microenvironment to be acidic.
However, in addition to the glycolysis of tumour cells, CAF, Treg, M2 type macrophages, and other immune cells that are “threatened” by tumours are also sources of lactic acid, resulting in an increase in H+ in the microenvironment. The increase in H+ in the tumour microenvironment during metabolic remodeling and tumour progression is a universal phenomenon, and lactate is an important source [383], which is very important for tumour survival.
The H+ is aimed at eliminating the persistent neutralizing stress (such as the neutralization of OH produced by the Fenton reaction and the overproduction of other end products), to maintain homeostasis between synthesis and decomposition (for example, for DNA synthesis and nucleic acid synthesis to facilitate cell division, for attacking immune cells through phospholipid synthesis, and for the synthesis and decomposition cycle through triglycerides) [383]. This was not a hypothesis and was confirmed in subsequent experiments. First, in the differential analysis of the expression profiles in the process of metabolic remodeling, it was found that the signaling pathways affected by metabolic plasticity were related to endoplasmic reticulum (ER) stress genes, and these genes were mainly induced by intracellular and extracellular stress [383]. Second, the acid-base imbalance pressure caused by different pH gradients was examined. Studies have shown that the pH in tumour cells is alkaline (pH > 7.4), while the extracellular pH is acidic (~6.5~−7), which shows a very different pH compared to normal cells (pH ≈ 7.2 intracellular, ≈7.4 in extracellular) [383]. The pH gradient has a dual effect on tumour cells. An intracellular alkaline environment can be conducive to glycolysis to produce more acids and neutralize the OH produced by the Fenton reaction to avoid the harmful effects of ROS on cells [383,384], and an extracellular acidic environment can inhibit immune response.
Therefore, the significance of tumour cells after metabolic rearrangement is preferentially dealing with stress (such as hypoxia, glucose deprivation, and acid increase) rather than proliferation. However, this increases the heterogeneity of tumour metabolism. Lactic acid causes metabolic symbiosis of the tumour. That is, tumour cell survival in blood vessels rich in oxygen and nutrition can establish a symbiotic metabolic relationship with tumour cells in the hypoxic region. The tumour cells in the oxygen-rich area will absorb the lactic acid produced by hypoxic tumour cells through MCTs [32,33], these tumour cells with rich oxidation and absorption of lactic acid are more likely to use glutamine for TCA cycle metabolism. However, the tumour cells in the hypoxic area are mainly glycolytic, and glutamine utilization is not preferred [36]. This heterogeneous metabolic mode produced by the absorption of lactic acid enables tumour cells to make full use of all microenvironment resources and does not need to compete for resources or interfere with each other. This regulation of metabolic plasticity can reasonably allocate environmental resources to facilitate tumour growth.

4.2. Effects of Metabolic Plasticity Regulation on Angiogenesis

Angiogenesis already exists in primary tumour growth, local tumour invasion, and distal metastasis [385], and lactic acid produced by glycolysis plays an important role in angiogenesis [386]. VEGF, fibroblast growth factor 2 (bFGF), and HIF-1 are involved in the process of angiogenesis [387,388,389], while lactic acid can inhibit prolyl hydroxylase-domain 2 (PHD2) and stabilize HIF-1α expression. In non-malignant epithelial cells, HIF-1α induced by lactate can synergistically upregulate bFGF and VEGFR2, with VEGF upregulated synergistically [33,390,391]. Therefore, the tumour needs to maintain neovascularization by introducing metabolism into glycolysis through regulating metabolic plasticity (Figure 5). Tumour cells try to activate glycolysis to facilitate angiogenesis, such as by inducing HIF-1 expression and increasing GLUT1/GLUT3 to enhance glucose absorption [392,393], inducing key enzymes of glycolysis, inhibiting the conversion of pyruvate to acetyl CoA into the TCA cycle [394,395], promoting LDHA transcription, and converting pyruvate into lactate [396,397].
In addition, HIF-1 can activate mitophagy and prevent glucose and fatty acids from entering oxidative metabolism [398,399]. Therefore, considering that the induction of HIF-1 is more significant under hypoxia, tumour cells in this state are more prone to angiogenesis. Thereafter, vascular bud formation is crucial for angiogenesis. It is guided and initiated by tip endothelial cells and then proliferates using stalk cells to extend the vascular wall. This process is accompanied by activating the VEGF/VEGFR2-signaling pathway in endothelial cells [400,401,402]. Glycolysis can promote endothelial cells to compete for “tip” sites and “stalk” sites [403], this process is accompanied by the overexpression of the glycolytic regulatory protein 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase 3 (PFKFB3) [403] (Figure 5). The knockdown of PFKFB3 can cause endothelial cells to lose these abilities, mainly by increasing cell adhesion and reducing filopodia formation. An in-depth mechanism revealed that PFKFB3 can drive glycolysis, thereby regulating VE-cadherin adhesion. ATP produced by PFKFB3-driven glycolysis can phosphorylate VE-cadherin before endocytosis [404]. Therefore, when PFKFB3 is deleted, glycolysis does not produce ATP, which blocks the endocytosis of VE-cadherin, causing extensive VE-cadherin on the cell surface and impeding cell movement [405]. From the above process, we can recognize the significance of glycolysis in tumour angiogenesis.

4.3. Effects of Metabolic Plasticity Regulation on Tumour Migration and Invasion

The regulatory mode of metabolic plasticity is crucial for the successful migration of tumour cells from proximal to distal regions. Thus, tumour cells can adjust the appropriate metabolic mode to adapt to the new microenvironment at the time of metastasis or after reaching the target organ [406]. In general, tumour cell metastasis to the distal end must proceed through the following stages: moving away from the primary site, invading the surrounding tissues, entering the circulation, planting to the distal end (adhesion), and completing migration [407,408].
Cell motility is the basis of tumour metastasis, and motile cells exhibit higher levels of glycolysis [409,410]. The enhancement of glycolytic flux (increased lactate release, enhanced glucose absorption, low OCR/extracellular acidification rate(ECAR), but high NAD+/NADH ratios) activates EMT-related molecules snail family transcriptional repressor 1(SNAI), twist-related protein (TWIST), and zinc finger E-box binding homeobox 1(ZEB1), and stimulates yes1-associated transcriptional regulator/Tafazzin (Yap/TAZ)-signaling to enhance cell movement [411]. Correspondingly, the plasticity of metabolic-to-glycolytic transition in tumour cell metastasis [8,406,412] is accompanied by the loss of mitochondrial function [413,414]. For instance, during the invasion and migration of breast cancer cells, the expression of delta-like non-canonical Notch ligand 2 (DLK2), a member of the epithelial growth factor (EGF) repeat superfamily protein, is regulated to enter the mitochondria and interact with the cox assembly factor 3 (COA3) to reduce its activity, thereby decreasing mitochondrial oxidative metabolism and increasing lactate release and glucose absorption, thus promoting the metabolic transition process from OXPHOS to glycolysis and realizing its metastasis [415].
Before leaving the primary site, cancerous epithelial cells must overcome anoikis, a form of apoptosis caused by leaving support from the extracellular matrix [40,41]. Notably, the induction of glycolysis can also enable tumour cells to shift to functional mitochondrial oxidation in an AMPK-dependent manner, thus leading to Snail-driven metastasis and subsequent anti-anoikis apoptosis [38]. This is a fascinating metabolic regulatory mechanism during cell migration. Tumour cells from the primary site require energy to survive when they transfer, and the glycolytic phenotype is the main metabolic mode in this process [416,417]. Although the energy supply to tumour cells is relatively small, its energy conversion efficiency is 100 times that of OXPHOS [42]. Currently, the transferred tumour cells need to exchange materials rapidly during migration, rather than OXPHOSS, which is a “slow-paced” energy supply mode. Contrastingly, the activation of AMPK2 induced by glycolysis enhances the expression of sodium-dependent neutral amino acid transporter type 2 (SLC1A5), a protein that transports glutamine into the TCA cycle for metabolism and increases mitochondrial oxidative stress. The resulting imbalance of redox balance is alleviated by the activation of Nrf2 by AMPK to maintain the energy balance and redox homeostasis of tumour cells and to avoid anoikis. Therefore, some researchers have proposed that glutamine metabolism occurs more easily in aggressive cancer cell lines (owing to stronger metastasis and higher survivability) [418,419,420], and that migrating and invading cells use more glutamine for metabolism [421,422].
Cell metastasis dominated by glutamine metabolism often occurs when the source of cell energy supply substrate (mainly glucose) is reduced, whereas tumour cells in remote hypoxia mainly rely on glucose for metabolism, and lactate is secreted through MCT4 and then taken up by tumour cells in oxidative metabolism through MCT1 [33]. However, tumour cells with the oxidative metabolic phenotype of lactate are more likely to undergo the TCA cycle by absorbing glutamine, which is inconsistent with the nutrient source (mainly glucose) taken up by cells undergoing glycolysis [36]. This metabolic symbiosis process (tumour cells prefer glycolysis to absorb glucose to produce lactic acid, and tumour cells prefer mitochondrial oxidative phosphorylation that absorbs lactic acid to metabolise glutamine) also fully reflects that tumour cells with metabolic heterogeneity can be more flexible in using environmental resources for survival. Notably, when glucose starvation occurs, tumour cells enhance the absorption of glutamine and, at the same time, upregulate matrix metallopeptidase 9 (MMP9) to enhance the metastasis of tumour cells and make them leave the harsh environment faster [423]. Therefore, the effect of glutamine on metastasis is not caused by its activation of OXPHOS, and the migration mechanism of glutamine to tumour cells is mainly realized by activating EMT [424,425]. EMT causes epithelial cells to lose polarization and cell–cell adhesion, resulting in a mesenchymal phenotype that can migrate and invade. Meaningfully, compared with the early progression stage in the 20 tumour types, when tumour cells developed, the upregulation of the EMT process was accompanied by the downregulation of mitochondrial-related genes [426]. Correspondingly, the molecules related to the glycolytic phenotype were significantly upregulated in interstitial cells [427], and key molecules that affect glycolysis, such as HIF-1α, PKM2, and HK2, can directly regulate the expression of EMT protein, such as E-cadherin, vimentin, and Twist1 [39,428]. Therefore, the increase in cell migration may be due to the interaction between glycolysis flux and EMT. However, we need to identify the factors that trigger enhanced mitochondrial metabolism (such as absorption of glutamine and activation of AMPK) also exist, whose purpose is to avoid excessive tumour cell death events (such as anoikis) during migration. The regulation of metabolic plasticity at this time is not only for cell energy supply. The by-product of glycolysis activates HIF-1α and promotes the upregulation of VEGF, bFGF, and their receptors. Thus, the adaptive expression of genes enables the tumour to adapt to the new environment faster when it migrates [429]. Therefore, during tumour cell transfer from the primary site to the secondary region, the plastic regulation of energy metabolism is not only a single stimulus but is a collaborative operation (Figure 6).
Therefore, glycolysis may be the main metabolic mode that enhances cell migration. In contrast, tumour cells with glycolytic phenotype can enhance mitochondrial metabolism (such as the absorption of glutamine and activation of AMPK); its real purpose is to avoid excessive tumour cell death events during migration (such as anti-anoikis by activation of AMPK). These glycolytic tumour cells are more aggressive (owing to stronger metastasis and higher survivability) [418,419,420] with more glutamine for metabolism [421,422]. Therefore, there may be at least two types of metastatic tumour cells, according to metabolic heterogeneity. One is glycolytic tumour cells with glucose for metabolism, and with reduced glucose, hypoxia, and/or high LA, these cells can develop into glycolytic tumour cells with glutamine for metabolism; these cells were more aggressive (Figure 6). Therefore, regulation of metabolic plasticity at this time is not only for cell energy supply but also as a by-product of glycolysis with multiple functions. The adaptive expression of genes enables the tumour to adapt to the new environment faster when it migrates.

5. Conclusions

Plasticity is the most prominent characteristic of tumourigenesis. It is also an advantageous biological that helps tumour cells adapt to environmental changes. Metabolic plasticity is an essential component of cellular plasticity, of which glycolysis and mitochondrial oxidative phosphorylation are two important components. Gene or genome instability caused by various factors initiates tumourigenesis and is accompanied by glycolysis and mitochondrial oxidative phosphorylation, enabling tumour cells to “arbitrarily” adjust their metabolic patterns. This change in metabolic mode also enables the daughter cells to survive better in “residential areas” with different nutrients, oxygen stresses, and immune cells by regulating metabolic plasticity according to the characteristics of the environment. Moreover, regulating metabolic plasticity partly prevents cell death and decreases viability caused by their respective metabolic defects. Therefore, regulating metabolic plasticity is a compensation for energy and a mechanism for adaptation and survival by optimizing a series of physiological, biochemical, and molecular regulatory processes. Therefore, to achieve an effective treatment of tumours, especially the occurrence and development of tumours through metabolism, multiple intervention strategies, which can block the metabolic changes generated by tumour cells, should be adopted. As tumourigenesis and its progression can occur and be regulated in multiple dimensions, more comprehensive metabolomic profiling and characteristic analyses on different stages and types of tumour cells are warranted in future studies.

Author Contributions

Conceptualization, N.N.; J.Z. and Y.W.; data collection, N.N. and J.Y.; writing —original draft preparation, N.N.; writing—review and editing, N.N.; Z.H. and Y.W.; supervision, Y.W. and J.Z.; project administration, Y.W.; funding acquisition, Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by Shenzhen Science and Technology Innovation Commission, the program name is Special program of sustainable development, subject title: Study on the techniques for comprehensive early diagnosis of liver carcer with clinical application (Grant No.: KCXFZ202002011006448, September 2020—September 2023).

Data Availability Statement

Data available in a publicly accessible repository.

Acknowledgments

The authors are especially grateful to Yu Zhang for his selfless help to the article.

Conflicts of Interest

The authors declare that they have no competing interest.

Abbreviations

ATPsAdenosine triphosphates
OXPHOSOxidative phosphorylation
TMETumour microenvironment
SIRT6NAD-dependent protein deacylase sirtuin-6
GLUT3Glucose transporter type 3
EGFREpidermal growth factor receptor
MAPKMAP kinase-activated protein kinase
LT-HSCsLong-term repopulating haematopoietic stem cells
c-MycMyc proto-oncogene protein
ESCEpithelial stem cells
KRASKirsten rat sarcoma viral oncogene
PFK1Phosphofructokinase-1
GLUT1Glucose transporter type 1
Lgr5Leucine rich repeat containing G protein-coupled receptor 5
TFAMTranscription factor A mitochondrial
PARPPoly [ADP-ribose] polymerase
TCATricarboxylic acid
PDHPyruvate dehydrogenase
Nrf-2Nuclear factor erythroid 2-related factor 2
TNBCTriple-negative breast cancer
PKPyruvate kinase
FSHFollicle-stimulating hormone
GBMsGlioblastoma
PDACPancreatic ductal adenocarcinoma
hrDNase1Human recombinant DNase1
HSP90Teat shock protein 90
H2Bub1Monoubiquitinated histone H2B
VEGFVascular endothelial growth factor
CAFsTumour associated fibroblasts
CSCitrate synthase
α-KGDCα-ketoglutarate dehydrogenase complex
CPTCarnitine palmitoyl transfer I
CNSCentral nervous system
R-2HG(R) enantiomer of 2-hydroxyglutarate
AMLAcute myeloid leukaemia
DMKGDimethyl-α-ketoglutarate
FBPFructose-1, 6-bisphosphatase
FHFumarate hydratase
CoQ10Coenzyme Q-binding protein 10
PINK1PTEN induced kinase 1
Keap1Kelch like ECH associated protein 1
NF1Neurofibromin 1
TIGARTP53 induced glycolysis regulaTory phosphatase
GLS2Glutaminase 2
PPARGC-1Peroxisome proliferative activated Receptor, gamma, coactivator 1
WTP53Wild-type P53
HCCHepatocarcinoma
Bbc3BCL2 binding component 3
PGMPhosphoglucomutase
S1PSphingosine 1-phosphate
TCRT cell receptor
PTENPhosphatase and tensin homolog
mTORMechanistic target of rapamycin kinase
CICCancer immunity cycle
MCTMonocarboxylate transporter
NKNatural killer
IFN-γInterferon gamma
PGK1Phosphoglycerate kinase
EREndoplasmic reticulum
PHD2Prolyl hydroxylase-domain 2
ECARExtracellular acidification rate
PDGFPlatelet-derived growth factor
TWISTtwist-related protein
YapYes1 associated transcriptional regulator
DLK2Delta like non-canonical Notch ligand 2
COA3Cox assembly factor 3
MMP9Matrix metallopeptidase 9
AktRAC-beta serine/threonine-protein kinase
EMTEpithelial-mesenchymal transition
TICsTumour-initiating cells
MEIS1Homeobox protein Meis1
HSCHaematopoietic stem cells
HIF-1αHypoxia inducible factor-1α
PDKPyruvate dehydrogenase kinase
GSCsGlioma stem cells
NF-κBNuclear factor NF-kappa-B
ISCIntestinal stem cells
HKHexokinase
LDHALactate dehydrogenase A
CBCCrypt base columnar
CRCColon epithelial cell carcinoma
ROSReactive oxygen species
LDHLactate dehydrogenase
PCPyruvate carboxylase
BCL-2B-cell lymphoma-2
AMPK5′-AMP-activated protein kinase
PFKPhosphofructokinase
VDACVoltage-dependent anion channel
NSTCsNon-stem tumour cells
PFKPPhosphofructokinase, platelet
NSCLCNon-small cell lung cancer
TRAP1TNF receptor associated protein 1
PKM2M2 isoform of pyruvate kinase
STAT3Signal transducer and activator of transcription 3
IGF-IRInsulin like growth factor 1 receptor
COXIComplex I
IDHIsocitrate dehydrogenase
SDHSuccessful dehydrogenase
Mcl-1Myeloid cell leukaemia1
α-KGα-ketoglutarate
LDHBLactate dehydrogenase B
OCROxygen consumption rate
mTORC1Mammalian target of rapamycin complex 1
NRFsNuclear respiratory factors
PDSS2Decaprenyl diphosphate synthase subunit 2
PEPPhosphoenolpyruvate
TP53Tumour protein p53
STK11Serine/threonine kinase 11
BRAFB-Raf proto-oncogene, serine/threonine kinase
SCO2Synthesis of cytochrome c oxidase 2
G-6-PDGlucose-6-phosphate dehydrogenase
SREBPSterol regulatory element binding protein
MuTP53Mutation in TP53
MPCMitochondrial pyruvate carrier
AIFApoptosis-induced factor
CAAsCancer-associated antigens
IL-7Interleukin-7
AGKAcylglycerol kinase
PI3KPhosphatidylinositol 3-kinase
DCsDendritic cells
PD-L1Programmed cell death 1 ligand 1
TregRegulatory T
ARG1Arginase 1
PDPK1Phosphoinositide dependent protein kinase 1
TGF-βTransforming growth factor beta
bFGFFibroblast growth factor 2
PFKFB3Protein 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase 3
LALactic acid
SNAISnail family transcriptional repressor
ZEB1Zinc finger E-box binding homeobox 1
TAZTafazzin
EGFEpithelial growth factor
SLC1A5Sodium-dependent neutral amino acid transporter type 2
CD28Leukocyte surface differentiation antigen 28

References

  1. Bausewein, T.; Nussberger, S.; Kuhlbrandt, W. Cryo-EM structure of Neurospora crassa respiratory complex IV. IUCr J. 2019, 6 Pt 4, 773–780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Mahad, D.; Ziabreva, I.; Lassmann, H.; Turnbull, D. Mitochondrial defects in acute multiple sclerosis lesions. Brain 2008, 131 Pt 7, 1722–1735. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, D.; Kim, J. Mitochondrial Retrograde Signalling and Metabolic Alterations in the Tumour Microenvironment. Cells 2019, 8, 275. [Google Scholar] [CrossRef] [Green Version]
  4. Fernie, A.R.; Carrari, F.; Sweetlove, L.J. Respiratory metabolism: Glycolysis, the TCA cycle and mitochondrial electron transport. Curr. Opin. Plant Biol. 2004, 7, 254–261. [Google Scholar] [CrossRef]
  5. Warburg, O.; Wind, F.; Negelein, E. The Metabolism of Tumors in the Body. J. Gen. Physiol. 1927, 8, 519–530. [Google Scholar] [CrossRef] [Green Version]
  6. Pouysségur, J.; Marchiq, I.; Parks, S.K.; Durivault, J.; Ždralević, M.; Vucetic, M. Warburg effect’ controls tumor growth, bacterial, viral infections and immunity–Genetic deconstruction and therapeutic perspectives. In Seminars in Cancer Biology; Academic Press: Cambridge, MA, USA, 2022. [Google Scholar]
  7. Vaupel, P.; Schmidberger, H.; Mayer, A. The Warburg effect: Essential part of metabolic reprogramming and central contributor to cancer progression. Int. J. Radiat. Biol. 2019, 95, 912–919. [Google Scholar] [CrossRef] [PubMed]
  8. Lu, J. The Warburg metabolism fuels tumor metastasis. Cancer Metastasis Rev. 2019, 38, 157–164. [Google Scholar] [CrossRef]
  9. Yang, J.; Liu, D.J.; Zheng, J.H.; He, R.Z.; Xu, D.P.; Yang, M.W.; Yao, H.F.; Fu, X.L.; Yang, J.Y.; Huo, Y.M.; et al. IRAK2-NF-kappaB signaling promotes glycolysis-dependent tumor growth in pancreatic cancer. Cell Oncol. 2022, 45, 367–379. [Google Scholar] [CrossRef]
  10. Reinfeld, B.I.; Rathmell, W.K.; Kim, T.K.; Rathmell, J.C. The therapeutic implications of immunosuppressive tumor aerobic glycolysis. Cell Mol. Immunol. 2022, 19, 46–58. [Google Scholar] [CrossRef]
  11. Zu, X.L.; Guppy, M. Cancer metabolism: Facts, fantasy, and fiction. Biochem. Biophys. Res. Commun. 2004, 313, 459–465. [Google Scholar] [CrossRef]
  12. Fantin, V.R.; St-Pierre, J.; Leder, P. Attenuation of LDH-A expression uncovers a link between glycolysis, mitochondrial physiology, and tumor maintenance. Cancer Cell 2006, 9, 425–434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Lee, M.; Yoon, J.H. Metabolic interplay between glycolysis and mitochondrial oxidation: The reverse Warburg effect and its therapeutic implication. World J. Biol. Chem. 2015, 6, 148–161. [Google Scholar] [CrossRef] [PubMed]
  14. Manley, S.J.; Liu, W.; Welch, D.R. The KISS1 metastasis suppressor appears to reverse the Warburg effect by shifting from glycolysis to mitochondrial beta-oxidation. J. Mol. Med. 2017, 95, 951–963. [Google Scholar] [CrossRef]
  15. Yuen, C.A.; Asuthkar, S.; Guda, M.R.; Tsung, A.J.; Velpula, K.K. Cancer stem cell molecular reprogramming of the Warburg effect in glioblastomas: A new target gleaned from an old concept. CNS Oncol. 2016, 5, 101–108. [Google Scholar] [CrossRef]
  16. Xu, J.; Richard, S. Cellular pathways influenced by protein arginine methylation: Implications for cancer. Mol. Cell 2021, 81, 4357–4368. [Google Scholar] [CrossRef]
  17. Chembazhi, U.V.; Bangru, S.; Hernaez, M.; Kalsotra, A. Cellular plasticity balances the metabolic and proliferation dynamics of a regenerating liver. Genome Res. 2021, 31, 576–591. [Google Scholar] [CrossRef] [PubMed]
  18. Yuan, S.; Norgard, R.; Stanger, B. Cellular Plasticity in Cancer. Cancer Discov. 2019, 9, 837–851. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Goldenring, J.R.; Mills, J.C. Cellular Plasticity, Reprogramming, and Regeneration: Metaplasia in the Stomach and Beyond. Gastroenterology 2022, 162, 415–430. [Google Scholar] [CrossRef]
  20. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  21. Greten, F.R.; Grivennikov, S.I. Inflammation and Cancer: Triggers, Mechanisms, and Consequences. Immunity 2019, 51, 27–41. [Google Scholar] [CrossRef]
  22. Liu, C.; Qiang, J.; Deng, Q.; Xia, J.; Deng, L.; Zhou, L.; Wang, D.; He, X.; Liu, Y.; Zhao, B.; et al. ALDH1A1 Activity in Tumor-Initiating Cells Remodels Myeloid-Derived Suppressor Cells to Promote Breast Cancer Progression. Cancer Res. 2021, 81, 5919–5934. [Google Scholar] [CrossRef] [PubMed]
  23. Nakatsuka, T.; Tateishi, K.; Kato, H.; Fujiwara, H.; Yamamoto, K.; Kudo, Y.; Nakagawa, H.; Tanaka, Y.; Ijichi, H.; Ikenoue, T.; et al. Inhibition of histone methyltransferase G9a attenuates liver cancer initiation by sensitizing DNA-damaged hepatocytes to p53-induced apoptosis. Cell Death Dis. 2021, 12, 99. [Google Scholar] [CrossRef]
  24. Inamura, K. Clinicopathological Characteristics and Mutations Driving Development of Early Lung Adenocarcinoma: Tumor Initiation and Progression. Int. J. Mol. Sci. 2018, 19, 1259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Zou, Z.W.; Ma, C.; Medoro, L.; Chen, L.; Wang, B.; Gupta, R.; Liu, T.; Yang, X.Z.; Chen, T.T.; Wang, R.Z.; et al. LncRNA ANRIL is up-regulated in nasopharyngeal carcinoma and promotes the cancer progression via increasing proliferation, reprograming cell glucose metabolism and inducing side-population stem-like cancer cells. Oncotarget 2016, 7, 61741–61754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Zhang, W.C.; Shyh-Chang, N.; Yang, H.; Rai, A.; Umashankar, S.; Ma, S.; Soh, B.S.; Sun, L.L.; Tai, B.C.; Nga, M.E.; et al. Glycine decarboxylase activity drives non-small cell lung cancer tumor-initiating cells and tumorigenesis. Cell 2012, 148, 259–272. [Google Scholar] [CrossRef] [Green Version]
  27. Sebastián, C.; Zwaans, B.M.; Silberman, D.M.; Gymrek, M.; Goren, A.; Zhong, L.; Ram, O.; Truelove, J.; Guimaraes, A.R.; Toiber, D.; et al. The histone deacetylase SIRT6 is a tumor suppressor that controls cancer metabolism. Cell 2012, 151, 1185–1199. [Google Scholar] [CrossRef] [Green Version]
  28. Dong, Y.; Tu, R.; Liu, H.; Qing, G. Regulation of cancer cell metabolism: Oncogenic MYC in the driver’s seat. Signal Transduct. Target. Ther. 2020, 5, 124. [Google Scholar] [CrossRef]
  29. Lunt, S.Y.; Vander, H.M. Aerobic glycolysis: Meeting the metabolic requirements of cell proliferation. Annu. Rev. Cell Dev. Biol. 2011, 27, 441–464. [Google Scholar] [CrossRef] [Green Version]
  30. Lyssiotis, C.A.; Cantley, L.C. SIRT6 puts cancer metabolism in the driver’s seat. Cell 2012, 151, 1155–1156. [Google Scholar] [CrossRef] [Green Version]
  31. Onodera, Y.; Nam, J.M.; Bissell, M.J. Increased sugar uptake promotes oncogenesis via EPAC/RAP1 and O-GlcNAc pathways. J. Clin. Investig. 2014, 124, 367–384. [Google Scholar] [CrossRef] [Green Version]
  32. Ippolito, L.; Morandi, A.; Giannoni, E.; Chiarugi, P. Lactate: A Metabolic Driver in the Tumour Landscape. Trends Biochem. Sci. 2019, 44, 153–166. [Google Scholar] [CrossRef] [PubMed]
  33. Sonveaux, P.; Végran, F.; Schroeder, T.; Wergin, M.C.; Verrax, J.; Rabbani, Z.N.; De Saedeleer, C.J.; Kennedy, K.M.; Diepart, C.; Jordan, B.F.; et al. Targeting lactate-fueled respiration selectively kills hypoxic tumor cells in mice. J. Clin. Investig. 2008, 118, 3930–3942. [Google Scholar] [CrossRef] [Green Version]
  34. Watson, M.J.; Vignali, P.D.A.; Mullett, S.J.; Overacre-Delgoffe, A.E.; Peralta, R.M.; Grebinoski, S.; Menk, A.V.; Rittenhouse, N.L.; DePeaux, K.; Whetstone, R.D.; et al. Metabolic support of tumour-infiltrating regulatory T cells by lactic acid. Nature 2021, 591, 645–651. [Google Scholar] [CrossRef] [PubMed]
  35. Brand, A.; Singer, K.; Koehl, G.E.; Kolitzus, M.; Schoenhammer, G.; Thiel, A.; Matos, C.; Bruss, C.; Klobuch, S.; Peter, K.; et al. LDHA-Associated Lactic Acid Production Blunts Tumor Immunosurveillance by T and NK Cells. Cell Metab. 2016, 24, 657–671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Pérez-Escuredo, J.; Dadhich, R.K.; Dhup, S.; Cacace, A.; Van Hée, V.F.; De Saedeleer, C.J.; Sboarina, M.; Rodriguez, F.; Fontenille, M.J.; Brisson, L.; et al. Lactate promotes glutamine uptake and metabolism in oxidative cancer cells. Cell Cycle 2016, 15, 72–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Brown, T.P.; Bhattacharjee, P.; Ramachandran, S.; Sivaprakasam, S.; Ristic, B.; Sikder, M.O.F.; Ganapathy, V. The lactate receptor GPR81 promotes breast cancer growth via a paracrine mechanism involving antigen-presenting cells in the tumor microenvironment. Oncogene 2020, 39, 3292–3304. [Google Scholar] [CrossRef]
  38. Zhao, H.; Yan, G.; Zheng, L.; Zhou, Y.; Sheng, H.; Wu, L.; Zhang, Q.; Lei, J.; Zhang, J.; Xin, R.; et al. STIM1 is a metabolic checkpoint regulating the invasion and metastasis of hepatocellular carcinoma. Theranostics 2020, 10, 6483–6499. [Google Scholar] [CrossRef]
  39. Dong, P.; Xiong, Y.; Konno, Y.; Ihira, K.; Kobayashi, N.; Yue, J.; Watari, H. Long non-coding RNA DLEU2 drives EMT and glycolysis in endometrial cancer through HK2 by competitively binding with miR-455 and by modulating the EZH2/miR-181a pathway. J. Exp. Clin. Cancer Res. 2021, 40, 216. [Google Scholar] [CrossRef]
  40. Park, H.A.; Brown, S.R.; Kim, Y. Cellular Mechanisms of Circulating Tumor Cells During Breast Cancer Metastasis. Int. J. Mol. Sci. 2020, 21, 5040. [Google Scholar] [CrossRef]
  41. Endo, H.; Owada, S.; Inagaki, Y.; Shida, Y.; Tatemichi, M. Metabolic reprogramming sustains cancer cell survival following extracellular matrix detachment. Redox Biol. 2020, 36, 101643. [Google Scholar] [CrossRef]
  42. Cairns, R.A.; Harris, I.S.; Mak, T.W. Regulation of cancer cell metabolism. Nat. Rev. Cancer 2011, 11, 85–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Todaro, G.J.; Huebner, R.J. NAS symposium: New evidence as the basis for increased efforts in cancer research. Proc. Natl. Acad. Sci. USA 1972, 69, 1009–1015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Pawelek, J.M. Cancer-cell fusion with migratory bone-marrow-derived cells as an explanation for metastasis: New therapeutic paradigms. Future Oncol. 2008, 4, 449–452. [Google Scholar] [CrossRef]
  45. Knudson, A.J. Mutation and cancer: Statistical study of retinoblastoma. Proc. Natl. Acad. Sci. USA 1971, 68, 820–823. [Google Scholar] [CrossRef] [Green Version]
  46. Potter, V.R. Phenotypic diversity in experimental hepatomas: The concept of partially blocked ontogeny. The 10th Walter Hubert Lecture. Br. J. Cancer 1978, 38, 1–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Pierce, G.B. Relationship between differentiation and carcinogenesis. J. Toxicol. Environ. Health 1977, 2, 1335–1342. [Google Scholar] [CrossRef]
  48. Zhao, Z.; Zuber, J.; Diaz-Flores, E.; Lintault, L.; Kogan, S.C.; Shannon, K.; Lowe, S.W. p53 loss promotes acute myeloid leukemia by enabling aberrant self-renewal. Genes Dev. 2010, 24, 1389–1402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Reya, T.; Morrison, S.J.; Clarke, M.F.; Weissman, I.L. Stem cells, cancer, and cancer stem cells. Nature 2001, 414, 105–111. [Google Scholar] [CrossRef] [Green Version]
  50. Pino, M.S.; Chung, D.C. The chromosomal instability pathway in colon cancer. Gastroenterology 2010, 138, 2059–2072. [Google Scholar] [CrossRef] [Green Version]
  51. Martin, U. Genome stability of programmed stem cell products. Adv. Drug. Deliv. Rev. 2017, 120, 108–117. [Google Scholar] [CrossRef]
  52. Yu, X.; Mengsteab, P.Y.; Narayanan, G.; Nair, L.S.; Laurencin, C.T. Enhancing the Surface Properties of a Bioengineered Anterior Cruciate Ligament Matrix for Use with Point-of-Care Stem Cell Therapy. Engineering 2021, 7, 153–161. [Google Scholar] [CrossRef] [PubMed]
  53. Halley-Stott, R.P.; Pasque, V.; Gurdon, J.B. Nuclear reprogramming. Development 2013, 140, 2468–2471. [Google Scholar] [CrossRef] [Green Version]
  54. Duelli, D.; Lazebnik, Y. Cell fusion: A hidden enemy? Cancer Cell 2003, 3, 445–448. [Google Scholar] [CrossRef] [Green Version]
  55. Pawelek, J.M. Tumour cell hybridization and metastasis revisited. Melanoma Res. 2000, 10, 507–514. [Google Scholar] [CrossRef]
  56. Rizvi, A.Z.; Swain, J.R.; Davies, P.S.; Bailey, A.S.; Decker, A.D.; Willenbring, H.; Grompe, M.; Fleming, W.H.; Wong, M.H. Bone marrow-derived cells fuse with normal and transformed intestinal stem cells. Proc. Natl. Acad. Sci. USA 2006, 103, 6321–6325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Aractingi, S.; Kanitakis, J.; Euvrard, S.; Le Danff, C.; Peguillet, I.; Khosrotehrani, K.; Lantz, O.; Carosella, E.D. Skin carcinoma arising from donor cells in a kidney transplant recipient. Cancer Res. 2005, 65, 1755–1760. [Google Scholar] [CrossRef] [Green Version]
  58. Suda, T.; Takubo, K.; Semenza, G.L. Metabolic regulation of hematopoietic stem cells in the hypoxic niche. Cell Stem Cell 2011, 9, 298–310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Simsek, T.; Kocabas, F.; Zheng, J.; Deberardinis, R.J.; Mahmoud, A.I.; Olson, E.N.; Schneider, J.W.; Zhang, C.C.; Sadek, H.A. The distinct metabolic profile of hematopoietic stem cells reflects their location in a hypoxic niche. Cell Stem Cell 2010, 7, 380–390. [Google Scholar] [CrossRef] [Green Version]
  60. Kocabas, F.; Zheng, J.; Thet, S.; Copeland, N.G.; Jenkins, N.A.; DeBerardinis, R.J.; Zhang, C.; Sadek, H.A. Meis1 regulates the metabolic phenotype and oxidant defense of hematopoietic stem cells. Blood 2012, 120, 4963–4972. [Google Scholar] [CrossRef] [Green Version]
  61. Shyh-Chang, N.; Daley, G.Q.; Cantley, L.C. Stem cell metabolism in tissue development and aging. Development 2013, 140, 2535–2547. [Google Scholar] [CrossRef] [Green Version]
  62. Takubo, K.; Nagamatsu, G.; Kobayashi, C.I.; Nakamura-Ishizu, A.; Kobayashi, H.; Ikeda, E.; Goda, N.; Rahimi, Y.; Johnson, R.S.; Soga, T.; et al. Regulation of glycolysis by Pdk functions as a metabolic checkpoint for cell cycle quiescence in hematopoietic stem cells. Cell Stem Cell 2013, 12, 49–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Klimmeck, D.; Hansson, J.; Raffel, S.; Vakhrushev, S.Y.; Trumpp, A.; Krijgsveld, J. Proteomic cornerstones of hematopoietic stem cell differentiation: Distinct signatures of multipotent progenitors and myeloid committed cells. Mol. Cell. Proteomics 2012, 11, 286–302. [Google Scholar] [CrossRef] [Green Version]
  64. Morfouace, M.; Lalier, L.; Bahut, M.; Bonnamain, V.; Naveilhan, P.; Guette, C.; Oliver, L.; Gueguen, N.; Reynier, P.; Vallette, F.M. Comparison of spheroids formed by rat glioma stem cells and neural stem cells reveals differences in glucose metabolism and promising therapeutic applications. J. Biol. Chem. 2012, 287, 33664–33674. [Google Scholar] [CrossRef] [Green Version]
  65. Pistollato, F.; Abbadi, S.; Rampazzo, E.; Viola, G.; Della Puppa, A.; Cavallini, L.; Frasson, C.; Persano, L.; Panchision, D.M.; Basso, G. Hypoxia and succinate antagonize 2-deoxyglucose effects on glioblastoma. Biochem. Pharmacol. 2010, 80, 1517–1527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Zhou, Y.; Zhou, Y.; Shingu, T.; Feng, L.; Chen, Z.; Ogasawara, M.; Keating, M.J.; Kondo, S.; Huang, P. Metabolic alterations in highly tumorigenic glioblastoma cells: Preference for hypoxia and high dependency on glycolysis. J. Biol. Chem. 2011, 286, 32843–32853. [Google Scholar] [CrossRef] [Green Version]
  67. Kondoh, H.; Lleonart, M.E.; Gil, J.; Wang, J.; Degan, P.; Peters, G.; Martinez, D.; Carnero, A.; Beach, D. Glycolytic enzymes can modulate cellular life span. Cancer Res. 2005, 65, 177–185. [Google Scholar] [CrossRef] [PubMed]
  68. Zahra, K.; Dey, T.; Ashish; Mishra, S.P.; Pandey, U. Pyruvate Kinase M2 and Cancer: The Role of PKM2 in Promoting Tumorigenesis. Front Oncol. 2020, 10, 159. [Google Scholar] [CrossRef] [Green Version]
  69. Chen, Y.; Yang, H.; Chen, S.; Lu, Z.; Li, B.; Jiang, T.; Xuan, M.; Ye, R.; Liang, H.; Liu, X.; et al. SIRT1 regulated hexokinase-2 promoting glycolysis is involved in hydroquinone-enhanced malignant progression in human lymphoblastoid TK6 cells. Ecotoxicol. Environ. Saf. 2022, 241, 113757. [Google Scholar] [CrossRef]
  70. Wang, X.Y.; Wei, Y.; Hu, B.; Liao, Y.; Wang, X.; Wan, W.H.; Huang, C.X.; Mahabati, M.; Liu, Z.Y.; Qu, J.R.; et al. c-Myc-driven glycolysis polarizes functional regulatory B cells that trigger pathogenic inflammatory responses. Signal Transduct. Target. Ther. 2022, 7, 105. [Google Scholar] [CrossRef]
  71. Qu, X.; Sun, J.; Zhang, Y.; Li, J.; Hu, J.; Li, K.; Gao, L.; Shen, L. c-Myc-driven glycolysis via TXNIP suppression is dependent on glutaminase-MondoA axis in prostate cancer. Biochem. Biophys. Res. Commun. 2018, 504, 415–421. [Google Scholar] [CrossRef]
  72. Ye, H.; Zhou, Q.; Zheng, S.; Li, G.; Lin, Q.; Wei, L.; Fu, Z.; Zhang, B.; Liu, Y.; Li, Z.; et al. Tumor-associated macrophages promote progression and the Warburg effect via CCL18/NF-kB/VCAM-1 pathway in pancreatic ductal adenocarcinoma. Cell Death Dis. 2018, 9, 453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Wong, D.J.; Liu, H.; Ridky, T.W.; Cassarino, D.; Segal, E.; Chang, H.Y. Module map of stem cell genes guides creation of epithelial cancer stem cells. Cell Stem Cell 2008, 2, 333–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Schwitalla, S.; Fingerle, A.A.; Cammareri, P.; Nebelsiek, T.; Göktuna, S.I.; Ziegler, P.K.; Canli, O.; Heijmans, J.; Huels, D.J.; Moreaux, G.; et al. Intestinal tumorigenesis initiated by dedifferentiation and acquisition of stem-cell-like properties. Cell 2013, 152, 25–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Formica, V.; Sera, F.; Cremolini, C.; Riondino, S.; Morelli, C.; Arkenau, H.T.; Roselli, M. KRAS and BRAF Mutations in Stage II and III Colon Cancer: A Systematic Review and Meta-Analysis. J. Natl. Cancer Inst. 2022, 114, 517–527. [Google Scholar] [CrossRef] [PubMed]
  76. Judd, J.; Abdel Karim, N.; Khan, H.; Naqash, A.R.; Baca, Y.; Xiu, J.; VanderWalde, A.M.; Mamdani, H.; Raez, L.E.; Nagasaka, M.; et al. Characterization of KRAS Mutation Subtypes in Non-small Cell Lung Cancer. Mol. Cancer Ther. 2021, 20, 2577–2584. [Google Scholar] [CrossRef]
  77. Sasaki, H.; Shitara, M.; Yokota, K.; Hikosaka, Y.; Moriyama, S.; Yano, M.; Fujii, Y. Overexpression of GLUT1 correlates with Kras mutations in lung carcinomas. Mol. Med. Rep. 2012, 5, 599–602. [Google Scholar] [CrossRef]
  78. Lefort, S.; Tan, S.; Balani, S.; Rafn, B.; Pellacani, D.; Hirst, M.; Sorensen, P.H.; Eaves, C.J. Initiation of human mammary cell tumorigenesis by mutant KRAS requires YAP inactivation. Oncogene 2020, 39, 1957–1968. [Google Scholar] [CrossRef] [Green Version]
  79. Bryant, K.L.; Mancias, J.D.; Kimmelman, A.C.; Der, C.J. KRAS: Feeding pancreatic cancer proliferation. Trends Biochem. Sci. 2014, 39, 91–100. [Google Scholar] [CrossRef] [Green Version]
  80. Barker, N.; Ridgway, R.A.; van Es, J.H.; van de Wetering, M.; Begthel, H.; van den Born, M.; Danenberg, E.; Clarke, A.R.; Sansom, O.J.; Clevers, H. Crypt stem cells as the cells-of-origin of intestinal cancer. Nature 2009, 457, 608–611. [Google Scholar] [CrossRef]
  81. Kemper, K.; Prasetyanti, P.R.; De Lau, W.; Rodermond, H.; Clevers, H.; Medema, J.P. Monoclonal antibodies against Lgr5 identify human colorectal cancer stem cells. Stem Cells 2012, 30, 2378–2386. [Google Scholar] [CrossRef]
  82. Merlos-Suárez, A.; Barriga, F.M.; Jung, P.; Iglesias, M.; Céspedes, M.V.; Rossell, D.; Sevillano, M.; Hernando-Momblona, X.; da Silva-Diz, V.; Muñoz, P.; et al. The intestinal stem cell signature identifies colorectal cancer stem cells and predicts disease relapse. Cell Stem Cell 2011, 8, 511–524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Bonnay, F.; Veloso, A.; Steinmann, V.; Köcher, T.; Abdusselamoglu, M.D.; Bajaj, S.; Rivelles, E.; Landskron, L.; Esterbauer, H.; Zinzen, R.P.; et al. Oxidative Metabolism Drives Immortalization of Neural Stem Cells during Tumorigenesis. Cell 2020, 182, 1490–1507.e19. [Google Scholar] [CrossRef] [PubMed]
  84. Hottiger, M.O. Nuclear ADP-Ribosylation and Its Role in Chromatin Plasticity, Cell Differentiation, and Epigenetics. Annu. Rev. Biochem. 2015, 84, 227–263. [Google Scholar] [CrossRef]
  85. Birsoy, K.; Wang, T.; Chen, W.W.; Freinkman, E.; Abu-Remaileh, M.; Sabatini, D.M. An Essential Role of the Mitochondrial Electron Transport Chain in Cell Proliferation Is to Enable Aspartate Synthesis. Cell 2015, 162, 540–551. [Google Scholar] [CrossRef] [Green Version]
  86. Sullivan, L.B.; Gui, D.Y.; Hosios, A.M.; Bush, L.N.; Freinkman, E.; Vander Heiden, M.G. Supporting Aspartate Biosynthesis Is an Essential Function of Respiration in Proliferating Cells. Cell 2015, 162, 552–563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Srivillibhuthur, M.; Warder, B.N.; Toke, N.H.; Shah, P.P.; Feng, Q.; Gao, N.; Bonder, E.M.; Verzi, M.P. TFAM is required for maturation of the fetal and adult intestinal epithelium. Dev. Biol. 2018, 439, 92–101. [Google Scholar] [CrossRef] [PubMed]
  88. Janiszewska, M.; Suvà, M.L.; Riggi, N.; Houtkooper, R.H.; Auwerx, J.; Clément-Schatlo, V.; Radovanovic, I.; Rheinbay, E.; Provero, P.; Stamenkovic, I. Imp2 controls oxidative phosphorylation and is crucial for preserving glioblastoma cancer stem cells. Genes Dev. 2012, 26, 1926–1944. [Google Scholar] [CrossRef] [Green Version]
  89. Kuntz, E.M.; Baquero, P.; Michie, A.M.; Dunn, K.; Tardito, S.; Holyoake, T.L.; Helgason, G.V.; Gottlieb, E. Targeting mitochondrial oxidative phosphorylation eradicates therapy-resistant chronic myeloid leukemia stem cells. Nat. Med. 2017, 23, 1234–1240. [Google Scholar] [CrossRef] [Green Version]
  90. Franco, J.; Balaji, U.; Freinkman, E.; Witkiewicz, A.K.; Knudsen, E.S. Metabolic Reprogramming of Pancreatic Cancer Mediated by CDK4/6 Inhibition Elicits Unique Vulnerabilities. Cell Rep. 2020, 32, 107793. [Google Scholar] [CrossRef]
  91. Haq, R.; Shoag, J.; Andreu-Perez, P.; Yokoyama, S.; Edelman, H.; Rowe, G.C.; Frederick, D.T.; Hurley, A.D.; Nellore, A.; Kung, A.L.; et al. Oncogenic BRAF regulates oxidative metabolism via PGC1alpha and MITF. Cancer Cell 2013, 23, 302–315. [Google Scholar] [CrossRef] [Green Version]
  92. Rao, S.; Mondragón, L.; Pranjic, B.; Hanada, T.; Stoll, G.; Köcher, T.; Zhang, P.; Jais, A.; Lercher, A.; Bergthaler, A.; et al. AIF-regulated oxidative phosphorylation supports lung cancer development. Cell Res. 2019, 29, 579–591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Pastò, A.; Bellio, C.; Pilotto, G.; Ciminale, V.; Silic-Benussi, M.; Guzzo, G.; Rasola, A.; Frasson, C.; Nardo, G.; Zulato, E.; et al. Cancer stem cells from epithelial ovarian cancer patients privilege oxidative phosphorylation, and resist glucose deprivation. Oncotarget 2014, 5, 4305–4319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Puisieux, A.; Pommier, R.M.; Morel, A.P.; Lavial, F. Cellular Pliancy and the Multistep Process of Tumorigenesis. Cancer Cell 2018, 33, 164–172. [Google Scholar] [CrossRef] [Green Version]
  95. Liu, M.; Fan, Y.; Li, D.; Han, B.; Meng, Y.; Chen, F.; Liu, T.; Song, Z.; Han, Y.; Huang, L.; et al. Ferroptosis inducer erastin sensitizes NSCLC cells to celastrol through activation of the ROS-mitochondrial fission-mitophagy axis. Mol. Oncol. 2021, 15, 2084–2105. [Google Scholar] [CrossRef]
  96. Lopez, J.; Tait, S.W. Mitochondrial apoptosis: Killing cancer using the enemy within. Br. J. Cancer 2015, 112, 57–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Tian, S.; Saravanan, K.; Mothana, R.A.; Ramachandran, G.; Rajivgandhi, G.; Manoharan, N. Anti-cancer activity of biosynthesized silver nanoparticles using Avicennia marina against A549 lung cancer cells through ROS/mitochondrial damages. Saudi J. Biol. Sci. 2020, 27, 3018–3024. [Google Scholar] [CrossRef]
  98. Pacchiana, R.; Mullappilly, N.; Pinto, A.; Bova, S.; Forciniti, S.; Cullia, G.; Dalla Pozza, E.; Bottani, E.; Decimo, I.; Dando, I.; et al. 3-Bromo-Isoxazoline Derivatives Inhibit GAPDH Enzyme in PDAC Cells Triggering Autophagy and Apoptotic Cell Death. Cancers 2022, 14, 3153. [Google Scholar] [CrossRef]
  99. Yi, W.; Clark, P.M.; Mason, D.E.; Keenan, M.C.; Hill, C.; Goddard, W.A., 3rd; Peters, E.C.; Driggers, E.M.; Hsieh-Wilson, L.C. Phosphofructokinase 1 glycosylation regulates cell growth and metabolism. Science 2012, 337, 75–80. [Google Scholar] [CrossRef] [Green Version]
  100. Hwang, I.; Tang, D.; Paik, J. Oxidative stress sensing and response in neural stem cell fate. Free Radic Biol. Med. 2021, 169, 74–83. [Google Scholar] [CrossRef]
  101. Schell, J.C.; Wisidagama, D.R.; Bensard, C.; Zhao, H.; Wei, P.; Tanner, J.; Flores, A.; Mohlman, J.; Sorensen, L.K.; Earl, C.S.; et al. Control of intestinal stem cell function and proliferation by mitochondrial pyruvate metabolism. Nat. Cell Biol. 2017, 19, 1027–1036. [Google Scholar] [CrossRef]
  102. Saga, I.; Shibao, S.; Okubo, J.; Osuka, S.; Kobayashi, Y.; Yamada, S.; Fujita, S.; Urakami, K.; Kusuhara, M.; Yoshida, K.; et al. Integrated analysis identifies different metabolic signatures for tumor-initiating cells in a murine glioblastoma model. Neuro Oncol. 2014, 16, 1048–1056. [Google Scholar] [CrossRef] [Green Version]
  103. Ichim, G.; Lopez, J.; Ahmed, S.U.; Muthalagu, N.; Giampazolias, E.; Delgado, M.E.; Haller, M.; Riley, J.S.; Mason, S.M.; Athineos, D.; et al. Limited mitochondrial permeabilization causes DNA damage and genomic instability in the absence of cell death. Mol. Cell 2015, 57, 860–872. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Bensard, C.L.; Wisidagama, D.R.; Olson, K.A.; Berg, J.A.; Krah, N.M.; Schell, J.C.; Nowinski, S.M.; Fogarty, S.; Bott, A.J.; Wei, P.; et al. Regulation of Tumor Initiation by the Mitochondrial Pyruvate Carrier. Cell Metab. 2020, 31, 284–300.e7. [Google Scholar] [CrossRef] [PubMed]
  105. Reichardt, I.; Bonnay, F.; Steinmann, V.; Loedige, I.; Burkard, T.R.; Meister, G.; Knoblich, J.A. The tumor suppressor Brat controls neuronal stem cell lineages by inhibiting Deadpan and Zelda. EMBO Rep. 2018, 19, 102–117. [Google Scholar] [CrossRef] [PubMed]
  106. Gatenby, R.A.; Gillies, R.J. Why do cancers have high aerobic glycolysis? Nat. Rev. Cancer 2004, 4, 891–899. [Google Scholar] [CrossRef] [PubMed]
  107. Khan, M.S.; Rehman, M.T.; Ismael, M.A.; AlAjmi, M.F.; Alruwaished, G.I.; Alokail, M.S.; Khan, M.R. Bioflavonoid (Hesperidin) Restrains Protein Oxidation and Advanced Glycation End Product Formation by Targeting AGEs and Glycolytic Enzymes. Cell Biochem. Biophys. 2021, 79, 833–844. [Google Scholar] [CrossRef]
  108. Brooks, G.A. Cell-cell and intracellular lactate shuttles. J. Physiol. 2009, 587 Pt 23, 5591–5600. [Google Scholar] [CrossRef]
  109. Kotredes, K.P.; Razmpour, R.; Lutton, E.; Alfonso-Prieto, M.; Ramirez, S.H.; Gamero, A.M. Characterization of cancer-associated IDH2 mutations that differ in tumorigenicity, chemosensitivity and 2-hydroxyglutarate production. Oncotarget 2019, 10, 2675–2692. [Google Scholar] [CrossRef] [Green Version]
  110. Gross, S.; Cairns, R.A.; Minden, M.D.; Driggers, E.M.; Bittinger, M.A.; Jang, H.G.; Sasaki, M.; Jin, S.; Schenkein, D.P.; Su, S.M.; et al. Cancer-associated metabolite 2-hydroxyglutarate accumulates in acute myelogenous leukemia with isocitrate dehydrogenase 1 and 2 mutations. J. Exp. Med. 2010, 207, 339–344. [Google Scholar] [CrossRef]
  111. Nakagawa, M.; Nakatani, F.; Matsunaga, H.; Seki, T.; Endo, M.; Ogawara, Y.; Machida, Y.; Katsumoto, T.; Yamagata, K.; Hattori, A.; et al. Selective inhibition of mutant IDH1 by DS-1001b ameliorates aberrant histone modifications and impairs tumor activity in chondrosarcoma. Oncogene 2019, 38, 6835–6849. [Google Scholar] [CrossRef] [Green Version]
  112. Peng, F.; Wang, J.H.; Fan, W.J.; Meng, Y.T.; Li, M.M.; Li, T.T.; Cui, B.; Wang, H.F.; Zhao, Y.; An, F.; et al. Glycolysis gatekeeper PDK1 reprograms breast cancer stem cells under hypoxia. Oncogene 2018, 37, 1062–1074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Peiris-Pagès, M.; Martinez-Outschoorn, U.E.; Pestell, R.G.; Sotgia, F.; Lisanti, M.P. Cancer stem cell metabolism. Breast Cancer Res. 2016, 18, 55. [Google Scholar] [CrossRef]
  114. Kim, C.F.; Dirks, P.B. Cancer and stem cell biology: How tightly intertwined? Cell Stem Cell 2008, 3, 147–150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Sun, L.; Suo, C.; Li, S.T.; Zhang, H.; Gao, P. Metabolic reprogramming for cancer cells and their microenvironment: Beyond the Warburg Effect. Biochim. Biophys. Acta Rev. Cancer 2018, 1870, 51–66. [Google Scholar] [CrossRef] [PubMed]
  116. DeBerardinis, R.J.; Chandel, N.S. Fundamentals of cancer metabolism. Sci. Adv. 2016, 2, e1600200. [Google Scholar] [CrossRef] [Green Version]
  117. Bao, S.; Wu, Q.; Sathornsumetee, S.; Hao, Y.; Li, Z.; Hjelmeland, A.B.; Shi, Q.; McLendon, R.E.; Bigner, D.D.; Rich, J.N. Stem cell-like glioma cells promote tumor angiogenesis through vascular endothelial growth factor. Cancer Res. 2006, 66, 7843–7848. [Google Scholar] [CrossRef] [Green Version]
  118. Calabrese, C.; Poppleton, H.; Kocak, M.; Hogg, T.L.; Fuller, C.; Hamner, B.; Oh, E.Y.; Gaber, M.W.; Finklestein, D.; Allen, M.; et al. A perivascular niche for brain tumor stem cells. Cancer Cell 2007, 11, 69–82. [Google Scholar] [CrossRef] [Green Version]
  119. Li, Z.; Bao, S.; Wu, Q.; Wang, H.; Eyler, C.; Sathornsumetee, S.; Shi, Q.; Cao, Y.; Lathia, J.; McLendon, R.E.; et al. Hypoxia-inducible factors regulate tumorigenic capacity of glioma stem cells. Cancer Cell 2009, 15, 501–513. [Google Scholar] [CrossRef] [Green Version]
  120. Shibao, S.; Minami, N.; Koike, N.; Fukui, N.; Yoshida, K.; Saya, H.; Sampetrean, O. Metabolic heterogeneity and plasticity of glioma stem cells in a mouse glioblastoma model. Neuro Oncol. 2018, 20, 343–354. [Google Scholar] [CrossRef] [Green Version]
  121. Jia, D.; Park, J.H.; Jung, K.H.; Levine, H.; Kaipparettu, B.A. Elucidating the Metabolic Plasticity of Cancer: Mitochondrial Reprogramming and Hybrid Metabolic States. Cells 2018, 7, 21. [Google Scholar] [CrossRef] [Green Version]
  122. Ishikawa, K.; Takenaga, K.; Akimoto, M.; Koshikawa, N.; Yamaguchi, A.; Imanishi, H.; Nakada, K.; Honma, Y.; Hayashi, J. ROS-generating mitochondrial DNA mutations can regulate tumor cell metastasis. Science 2008, 320, 661–664. [Google Scholar] [CrossRef] [Green Version]
  123. Piskounova, E.; Agathocleous, M.; Murphy, M.M.; Hu, Z.; Huddlestun, S.E.; Zhao, Z.; Leitch, A.M.; Johnson, T.M.; DeBerardinis, R.J.; Morrison, S.J. Oxidative stress inhibits distant metastasis by human melanoma cells. Nature 2015, 527, 186–191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Jia, D.; Lu, M.; Jung, K.H.; Park, J.H.; Yu, L.; Onuchic, J.N.; Kaipparettu, B.A.; Levine, H. Elucidating cancer metabolic plasticity by coupling gene regulation with metabolic pathways. Proc. Natl. Acad. Sci. USA 2019, 116, 3909–3918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Sharma, D.; Singh, M.; Rani, R. Role of LDH in tumor glycolysis: Regulation of LDHA by small molecules for cancer therapeutics. Semin. Cancer Biol. 2022, 87, 184–195. [Google Scholar] [CrossRef]
  126. Yeung, C.; Gibson, A.E.; Issaq, S.H.; Oshima, N.; Baumgart, J.T.; Edessa, L.D.; Rai, G.; Urban, D.J.; Johnson, M.S.; Benavides, G.A. Targeting Glycolysis through Inhibition of Lactate Dehydrogenase Impairs Tumor Growth in Preclinical Models of Ewing Sarcoma. Cancer Res. 2019, 79, 5060–5073. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Flores, A.; Sandoval-Gonzalez, S.; Takahashi, R.; Krall, A.; Sathe, L.; Wei, L.; Radu, C.; Joly, J.H.; Graham, N.A.; Christofk, H.R.; et al. Increased lactate dehydrogenase activity is dispensable in squamous carcinoma cells of origin. Nat. Commun. 2019, 10, 91. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Brisson, L.; Bański, P.; Sboarina, M.; Dethier, C.; Danhier, P.; Fontenille, M.J.; Van Hée, V.F.; Vazeille, T.; Tardy, M.; Falces, J.; et al. Lactate Dehydrogenase B Controls Lysosome Activity and Autophagy in Cancer. Cancer Cell 2016, 30, 418–431. [Google Scholar] [CrossRef] [Green Version]
  129. Faubert, B.; Li, K.Y.; Cai, L.; Hensley, C.T.; Kim, J.; Zacharias, L.G.; Yang, C.; Do, Q.N.; Doucette, S.; Burguete, D.; et al. Lactate Metabolism in Human Lung Tumors. Cell 2017, 171, 358–371.e9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Hui, S.; Ghergurovich, J.M.; Morscher, R.J.; Jang, C.; Teng, X.; Lu, W.; Esparza, L.A.; Reya, T.; Zhan, L.; Yanxiang Guo, J.; et al. Glucose feeds the TCA cycle via circulating lactate. Nature 2017, 551, 115–118. [Google Scholar] [CrossRef] [Green Version]
  131. Rodríguez-Colman, M.J.; Schewe, M.; Meerlo, M.; Stigter, E.; Gerrits, J.; Pras-Raves, M.; Sacchetti, A.; Hornsveld, M.; Oost, K.C.; Snippert, H.J.; et al. Interplay between metabolic identities in the intestinal crypt supports stem cell function. Nature 2017, 543, 424–427. [Google Scholar] [CrossRef]
  132. Hensley, C.T.; Faubert, B.; Yuan, Q.; Lev-Cohain, N.; Jin, E.; Kim, J.; Jiang, L.; Ko, B.; Skelton, R.; Loudat, L.; et al. Metabolic Heterogeneity in Human Lung Tumors. Cell 2016, 164, 681–694. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Sellers, K.; Fox, M.P.; Bousamra, M., 2nd; Slone, S.P.; Higashi, R.M.; Miller, D.M.; Wang, Y.; Yan, J.; Yuneva, M.O.; Deshpande, R.; et al. Pyruvate carboxylase is critical for non-small-cell lung cancer proliferation. J. Clin. Investig. 2015, 125, 687–698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Ahmed, A.U.; Thaci, B.; Tobias, A.L.; Auffinger, B.; Zhang, L.; Cheng, Y.; Kim, C.K.; Yunis, C.; Han, Y.; Alexiades, N.G.; et al. Metabolic state of glioma stem cells and nontumorigenic cells. Proc. Natl. Acad. Sci. USA 2011, 108, 16062–16067. [Google Scholar]
  135. Michelakis, E.D.; Sutendra, G.; Dromparis, P.; Webster, L.; Haromy, A.; Niven, E.; Maguire, C.; Gammer, T.L.; Mackey, J.R.; Fulton, D.; et al. Metabolic modulation of glioblastoma with dichloroacetate. Sci. Transl. Med. 2010, 2, 31ra34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Yu, L.; Lu, M.; Jia, D.; Ma, J.; Ben-Jacob, E.; Levine, H.; Kaipparettu, B.A.; Onuchic, J.N. Modeling the Genetic Regulation of Cancer Metabolism: Interplay between Glycolysis and Oxidative Phosphorylation. Cancer Res. 2017, 77, 1564–1574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Mathupala, S.P.; Ko, Y.H.; Pedersen, P.L. Hexokinase II: Cancer’s double-edged sword acting as both facilitator and gatekeeper of malignancy when bound to mitochondria. Oncogene 2006, 25, 4777–4786. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Anderson, M.; Marayati, R.; Moffitt, R.; Yeh, J.J. Hexokinase 2 promotes tumor growth and metastasis by regulating lactate production in pancreatic cancer. Oncotarget 2017, 8, 56081–56094. [Google Scholar] [CrossRef] [Green Version]
  139. Patra, K.C.; Wang, Q.; Bhaskar, P.T.; Miller, L.; Wang, Z.; Wheaton, W.; Chandel, N.; Laakso, M.; Muller, W.J.; Allen, E.L.; et al. Hexokinase 2 is required for tumor initiation and maintenance and its systemic deletion is therapeutic in mouse models of cancer. Cancer Cell 2013, 24, 213–228. [Google Scholar] [CrossRef] [Green Version]
  140. Xu, S.; Herschman, H.R. A Tumor Agnostic Therapeutic Strategy for Hexokinase 1-Null/Hexokinase 2-Positive Cancers. Cancer Res. 2019, 79, 5907–5914. [Google Scholar] [CrossRef] [Green Version]
  141. Fukushi, A.; Kim, H.D.; Chang, Y.C.; Kim, C.H. Revisited Metabolic Control and Reprogramming Cancers by Means of the Warburg Effect in Tumor Cells. Int. J. Mol. Sci. 2022, 23, 10037. [Google Scholar] [CrossRef]
  142. Zhang, R.; Shen, W.; Du, J.; Gillies, M.C. Selective knockdown of hexokinase 2 in rods leads to age-related photoreceptor degeneration and retinal metabolic remodeling. Cell Death Dis. 2020, 11, 885. [Google Scholar] [CrossRef]
  143. Petit, L.; Ma, S.; Cipi, J.; Cheng, S.Y.; Zieger, M.; Hay, N.; Punzo, C. Aerobic Glycolysis Is Essential for Normal Rod Function and Controls Secondary Cone Death in Retinitis Pigmentosa. Cell Rep. 2018, 23, 2629–2642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Petit, L.; Ma, S.; Cipi, J.; Cheng, S.Y.; Zieger, M.; Hay, N.; Punzo, C. Monoamine Oxidase A is a Major Mediator of Mitochondrial Homeostasis and Glycolysis in Gastric Cancer Progression. Cancer Manag. Res. 2020, 12, 8023–8035. [Google Scholar]
  145. Wolf, A.; Agnihotri, S.; Micallef, J.; Mukherjee, J.; Sabha, N.; Cairns, R.; Hawkins, C.; Guha, A. Hexokinase 2 is a key mediator of aerobic glycolysis and promotes tumor growth in human glioblastoma multiforme. J. Exp. Med. 2011, 208, 313–326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Shoshan-Barmatz, V.; Israelson, A.; Brdiczka, D.; Sheu, S.S. The voltage-dependent anion channel (VDAC): Function in intracellular signalling, cell life and cell death. Curr. Pharm. Des. 2006, 12, 2249–22470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Shoshan-Barmatz, V.; Zakar, M.; Rosenthal, K.; Abu-Hamad, S. Key regions of VDAC1 functioning in apoptosis induction and regulation by hexokinase. Biochim. Biophys. Acta 2009, 1787, 421–430. [Google Scholar] [CrossRef] [Green Version]
  148. Shoshan-Barmatz, V.; Maldonado, E.N.; Krelin, Y. VDAC1 at the crossroads of cell metabolism, apoptosis and cell stress. Cell Stress 2017, 1, 11–36. [Google Scholar] [CrossRef] [Green Version]
  149. Cesar, M.C.; Wilson, J.E. Further studies on the coupling of mitochondrially bound hexokinase to intramitochondrially compartmented ATP, generated by oxidative phosphorylation. Arch Biochem. Biophys. 1998, 350, 109–117. [Google Scholar] [CrossRef]
  150. Pastorino, J.G.; Hoek, J.B. Regulation of hexokinase binding to VDAC. J. Bioenerg. Biomembr. 2008, 40, 171–182. [Google Scholar] [CrossRef] [Green Version]
  151. Sas, K.M.; Kayampilly, P.; Byun, J.; Nair, V.; Hinder, L.M.; Hur, J.; Zhang, H.; Lin, C.; Qi, N.R.; Michailidis, G.; et al. Tissue-specific metabolic reprogramming drives nutrient flux in diabetic complications. JCI Insight 2016, 1, e86976. [Google Scholar] [CrossRef]
  152. Yam, M.; Engel, A.L.; Wang, Y.; Zhu, S.; Hauer, A.; Zhang, R.; Lohner, D.; Huang, J.; Dinterman, M.; Zhao, C.; et al. Proline mediates metabolic communication between retinal pigment epithelial cells and the retina. J. Biol. Chem. 2019, 294, 10278–10289. [Google Scholar] [CrossRef] [PubMed]
  153. Tretter, L.; Patocs, A.; Chinopoulos, C. Succinate, an intermediate in metabolism, signal transduction, ROS, hypoxia, and tumorigenesis. Biochim. Biophys. Acta 2016, 1857, 1086–1101. [Google Scholar] [CrossRef] [PubMed]
  154. Zhang, M.; Liu, Q.; Zhang, M.; Cao, C.; Liu, X.; Zhang, M.; Li, G.; Xu, C.; Zhang, X. Enhanced antitumor effects of follicle-stimulating hormone receptor-mediated hexokinase-2 depletion on ovarian cancer mediated by a shift in glucose metabolism. J. Nanobiotechnology 2020, 18, 161. [Google Scholar] [CrossRef] [PubMed]
  155. DeWaal, D.; Nogueira, V.; Terry, A.R.; Patra, K.C.; Jeon, S.M.; Guzman, G.; Au, J.; Long, C.P.; Antoniewicz, M.R.; Hay, N.; et al. Author Correction: Hexokinase-2 depletion inhibits glycolysis and induces oxidative phosphorylation in hepatocellular carcinoma and sensitizes to metformin. Nat. Commun. 2018, 9, 2539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Teixeira, C.S.S.; Sousa, S.F. Current Status of the Use of Multifunctional Enzymes as Anti-Cancer Drug Targets. Pharmaceutics. 2021, 14, 10. [Google Scholar] [CrossRef]
  157. Zhou, K.; Yao, Y.L.; He, Z.C.; Chen, C.; Zhang, X.N.; Yang, K.D.; Liu, Y.Q.; Liu, Q.; Fu, W.J.; Chen, Y.P.; et al. VDAC2 interacts with PFKP to regulate glucose metabolism and phenotypic reprogramming of glioma stem cells. Cell Death Dis. 2018, 9, 988. [Google Scholar] [CrossRef] [Green Version]
  158. Chen, J.; Zou, L.; Lu, G.; Grinchuk, O.; Fang, L.; Ong, D.S.T.; Taneja, R.; Ong, C.N.; Shen, H.M. PFKP alleviates glucose starvation-induced metabolic stress in lung cancer cells via AMPK-ACC2 dependent fatty acid oxidation. Cell Discov. 2022, 8, 52. [Google Scholar] [CrossRef]
  159. Lee, J.H.; Liu, R.; Li, J.; Zhang, C.; Wang, Y.; Cai, Q.; Qian, X.; Xia, Y.; Zheng, Y.; Piao, Y.; et al. Stabilization of phosphofructokinase 1 platelet isoform by AKT promotes tumorigenesis. Nat. Commun. 2017, 8, 949. [Google Scholar] [CrossRef] [Green Version]
  160. Matsuura, K.; Canfield, K.; Feng, W.; Kurokawa, M. Metabolic Regulation of Apoptosis in Cancer. Int. Rev. Cell Mol. Biol. 2016, 327, 43–87. [Google Scholar]
  161. Icard, P.; Shulman, S.; Farhat, D.; Steyaert, J.M.; Alifano, M.; Lincet, H. How the Warburg effect supports aggressiveness and drug resistance of cancer cells? Drug Resist. Updates 2018, 38, 1–11. [Google Scholar] [CrossRef]
  162. Yang, Y.; Ishak Gabra, M.B.; Hanse, E.A.; Lowman, X.H.; Tran, T.Q.; Li, H.; Milman, N.; Liu, J.; Reid, M.A.; Locasale, J.W.; et al. MiR-135 suppresses glycolysis and promotes pancreatic cancer cell adaptation to metabolic stress by targeting phosphofructokinase-1. Nat. Commun. 2019, 10, 809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Maddalena, F.; Condelli, V.; Matassa, D.S.; Pacelli, C.; Scrima, R.; Lettini, G.; Li Bergolis, V.; Pietrafesa, M.; Crispo, F.; Piscazzi, A.; et al. TRAP1 enhances Warburg metabolism through modulation of PFK1 expression/activity and favors resistance to EGFR inhibitors in human colorectal carcinomas. Mol. Oncol. 2020, 14, 3030–3047. [Google Scholar] [CrossRef] [PubMed]
  164. Snyder, M.W.; Kircher, M.; Hill, A.J.; Daza, R.M.; Shendure, J. Cell-free DNA Comprises an In Vivo Nucleosome Footprint that Informs Its Tissues-Of-Origin. Cell 2016, 164, 57–68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Keyel, P.A. Dnases in health and disease. Dev. Biol. 2017, 429, 1–11. [Google Scholar] [CrossRef] [PubMed]
  166. Wang, S.; Ma, H.; Li, X.; Mo, X.; Zhang, H.; Yang, L.; Deng, Y.; Yan, Y.; Yang, G.; Liu, X.; et al. DNASE1L3 as an indicator of favorable survival in hepatocellular carcinoma patients following resection. Aging 2020, 12, 1171–1185. [Google Scholar] [CrossRef] [PubMed]
  167. Xu, B.; Lv, W.; Li, X.; Zhang, L.; Lin, J. Prognostic genes of hepatocellular carcinoma based on gene coexpression network analysis. J. Cell Biochem. 2019, 120, 11616–11623. [Google Scholar] [CrossRef] [PubMed]
  168. Xiao, Y.; Yang, K.; Liu, P.; Ma, D.; Lei, P.; Liu, Q. Deoxyribonuclease 1-like 3 Inhibits Hepatocellular Carcinoma Progression by Inducing Apoptosis and Reprogramming Glucose Metabolism. Int. J. Biol. Sci. 2022, 18, 82–95. [Google Scholar] [CrossRef]
  169. Rosner, K. DNase1: A new personalized therapy for cancer? Expert Rev. Anticancer Ther. 2011, 11, 981–984. [Google Scholar] [CrossRef]
  170. Joshi, A.; Dai, L.; Liu, Y.; Lee, J.; Ghahhari, N.M.; Segala, G.; Beebe, K.; Jenkins, L.M.; Lyons, G.C.; Bernasconi, L.; et al. The mitochondrial HSP90 paralog TRAP1 forms an OXPHOS-regulated tetramer and is involved in mitochondrial metabolic homeostasis. BMC Biol. 2020, 18, 10. [Google Scholar] [CrossRef] [Green Version]
  171. Li, L.; Wang, Y.; Wang, Q.; Qu, J.; Wei, X.; Xu, J.; Wang, Y.; Suo, F.; Zhang, Y. High developmental pluripotency-associated 4 expression promotes cell proliferation and glycolysis, and predicts poor prognosis in non-small-cell lung cancer. Mol. Med. Rep. 2019, 20, 445–454. [Google Scholar] [CrossRef] [Green Version]
  172. Mazurek, S.; Boschek, C.B.; Hugo, F.; Eigenbrodt, E. Pyruvate kinase type M2 and its role in tumor growth and spreading. Semin. Cancer Biol. 2005, 15, 300–308. [Google Scholar] [CrossRef] [PubMed]
  173. Jing, Y.Y.; Cai, F.F.; Zhang, L.; Han, J.; Yang, L.; Tang, F.; Li, Y.B.; Chang, J.F.; Sun, F.; Yang, X.M.; et al. Epigenetic regulation of the Warburg effect by H2B monoubiquitination. Cell Death Differ. 2020, 27, 1660–1676. [Google Scholar] [CrossRef] [PubMed]
  174. Li, T.; Han, J.; Jia, L.; Hu, X.; Chen, L.; Wang, Y. PKM2 coordinates glycolysis with mitochondrial fusion and oxidative phosphorylation. Protein Cell 2019, 10, 583–594. [Google Scholar] [CrossRef] [Green Version]
  175. Zhou, Q.; Li, H.; Li, Y.; Tan, M.; Fan, S.; Cao, C.; Meng, F.; Zhu, L.; Zhao, L.; Guan, M.X.; et al. Inhibiting neddylation modification alters mitochondrial morphology and reprograms energy metabolism in cancer cells. JCI Insight 2019, 4, e121582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Gao, X.; Wang, H.; Yang, J.J.; Liu, X.; Liu, Z.R. Pyruvate kinase M2 regulates gene transcription by acting as a protein kinase. Mol. Cell 2012, 45, 98–609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Zhu, S.; Guo, Y.; Zhang, X.; Liu, H.; Yin, M.; Chen, X.; Peng, C. Pyruvate kinase M2 (PKM2) in cancer and cancer therapeutics. Cancer Lett. 2021, 503, 240–248. [Google Scholar] [CrossRef]
  178. Gui, D.Y.; Lewis, C.A.; Vander Heiden, M.G. Allosteric regulation of PKM2 allows cellular adaptation to different physiological states. Sci. Signal 2013, 6, e7. [Google Scholar] [CrossRef]
  179. Feng, C.; Gao, Y.; Wang, C.; Yu, X.; Zhang, W.; Guan, H.; Shan, Z.; Teng, W. Aberrant overexpression of pyruvate kinase M2 is associated with aggressive tumor features and the BRAF mutation in papillary thyroid cancer. J. Clin. Endocrinol. Metab. 2013, 98, E1524–E1533. [Google Scholar] [CrossRef] [Green Version]
  180. McAllister, S.S.; Weinberg, R.A. The tumour-induced systemic environment as a critical regulator of cancer progression and metastasis. Nat. Cell Biol. 2014, 16, 717–727. [Google Scholar] [CrossRef]
  181. Hamabe, A.; Konno, M.; Tanuma, N.; Shima, H.; Tsunekuni, K.; Kawamoto, K.; Nishida, N.; Koseki, J.; Mimori, K.; Gotoh, N.; et al. Role of pyruvate kinase M2 in transcriptional regulation leading to epithelial-mesenchymal transition. Proc. Natl. Acad. Sci. USA 2014, 111, 15526–15531. [Google Scholar] [CrossRef] [Green Version]
  182. Amin, S.; Yang, P.; Li, Z. Pyruvate kinase M2: A multifarious enzyme in non-canonical localization to promote cancer progression. Biochim. Biophys. Acta Rev. Cancer 2019, 1871, 331–341. [Google Scholar] [CrossRef] [PubMed]
  183. Xu, Q.; Liu, L.Z.; Yin, Y.; He, J.; Li, Q.; Qian, X.; You, Y.; Lu, Z.; Peiper, S.C.; Shu, Y.; et al. Regulatory circuit of PKM2/NF-kappaB/miR-148a/152-modulated tumor angiogenesis and cancer progression. Oncogene 2015, 34, 5482–5493. [Google Scholar] [CrossRef] [PubMed]
  184. Li, Z.; Yang, P.; Li, Z. The multifaceted regulation and functions of PKM2 in tumor progression. Biochim. Biophys. Acta 2014, 1846, 285–296. [Google Scholar] [CrossRef] [PubMed]
  185. Li, L.; Zhang, Y.; Qiao, J.; Yang, J.J.; Liu, Z.R. Pyruvate kinase M2 in blood circulation facilitates tumor growth by promoting angiogenesis. J. Biol. Chem. 2014, 289, 25812–25821. [Google Scholar] [CrossRef] [Green Version]
  186. Azoitei, N.; Becher, A.; Steinestel, K.; Rouhi, A.; Diepold, K.; Genze, F.; Simmet, T.; Seufferlein, T. PKM2 promotes tumor angiogenesis by regulating HIF-1alpha through NF-kappaB activation. Mol. Cancer 2016, 15, 3. [Google Scholar] [CrossRef] [Green Version]
  187. Giannoni, E.; Taddei, M.L.; Morandi, A.; Comito, G.; Calvani, M.; Bianchini, F.; Richichi, B.; Raugei, G.; Wong, N.; Tang, D.; et al. Targeting stromal-induced pyruvate kinase M2 nuclear translocation impairs oxphos and prostate cancer metastatic spread. Oncotarget 2015, 6, 24061–24074. [Google Scholar] [CrossRef] [Green Version]
  188. Fiaschi, T.; Marini, A.; Giannoni, E.; Taddei, M.L.; Gandellini, P.; De Donatis, A.; Lanciotti, M.; Serni, S.; Cirri, P.; Chiarugi, P. Reciprocal metabolic reprogramming through lactate shuttle coordinately influences tumor-stroma interplay. Cancer Res. 2012, 72, 5130–5140. [Google Scholar] [CrossRef] [Green Version]
  189. Zhou, L.; Zhang, W.; Sun, Y.; Jia, L. Protein neddylation and its alterations in human cancers for targeted therapy. Cell Signal 2018, 44, 92–102. [Google Scholar] [CrossRef] [PubMed]
  190. Papa, S.; Martino, P.L.; Capitanio, G.; Gaballo, A.; De Rasmo, D.; Signorile, A.; Petruzzella, V. The oxidative phosphorylation system in mammalian mitochondria. Adv. Exp. Med. Biol. 2012, 942, 3–37. [Google Scholar]
  191. Curry, J.M.; Tuluc, M.; Whitaker-Menezes, D.; Ames, J.A.; Anantharaman, A.; Butera, A.; Leiby, B.; Cognetti, D.M.; Sotgia, F.; Lisanti, M.P.; et al. Cancer metabolism, stemness and tumor recurrence: MCT1 and MCT4 are functional biomarkers of metabolic symbiosis in head and neck cancer. Cell Cycle 2013, 12, 1371–1384. [Google Scholar] [CrossRef] [Green Version]
  192. De Luca, A.; Fiorillo, M.; Peiris-Pagès, M.; Ozsvari, B.; Smith, D.L.; Sanchez-Alvarez, R.; Martinez-Outschoorn, U.E.; Cappello, A.R.; Pezzi, V.; Lisanti, M.P.; et al. Mitochondrial biogenesis is required for the anchorage-independent survival and propagation of stem-like cancer cells. Oncotarget 2015, 6, 4777–47795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Mikkilineni, L.; Whitaker-Menezes, D.; Domingo-Vidal, M.; Sprandio, J.; Avena, P.; Cotzia, P.; Dulau-Florea, A.; Gong, J.; Uppal, G.; Zhan, T.; et al. Hodgkin lymphoma: A complex metabolic ecosystem with glycolytic reprogramming of the tumor microenvironment. Semin. Oncol. 2017, 44, 218–225. [Google Scholar] [CrossRef] [PubMed]
  194. Khatami, F.; Aghamir, S.; Tavangar, S.M. Oncometabolites: A new insight for oncology. Mol. Genet. Genomic. Med. 2019, 7, e873. [Google Scholar] [CrossRef] [PubMed]
  195. Fontana, F.; Limonta, P. The multifaceted roles of mitochondria at the crossroads of cell life and death in cancer. Free Radic. Biol. Med. 2021, 176, 203–221. [Google Scholar] [CrossRef] [PubMed]
  196. Whitaker-Menezes, D.; Martinez-Outschoorn, U.E.; Flomenberg, N.; Birbe, R.C.; Witkiewicz, A.K.; Howell, A.; Pavlides, S.; Tsirigos, A.; Ertel, A.; Pestell, R.G.; et al. Hyperactivation of oxidative mitochondrial metabolism in epithelial cancer cells in situ: Visualizing the therapeutic effects of metformin in tumor tissue. Cell Cycle 2011, 10, 4047–4064. [Google Scholar] [CrossRef] [Green Version]
  197. Choi, B.W.; Jeong, Y.J.; Park, S.H.; Oh, H.K.; Kang, S. Reverse Warburg Effect-Related Mitochondrial Activity and (18)F-FDG Uptake in Invasive Ductal Carcinoma. Nucl. Med. Mol. Imaging 2019, 53, 396–405. [Google Scholar] [CrossRef]
  198. Moldogazieva, N.T.; Mokhosoev, I.M.; Terentiev, A.A. Metabolic Heterogeneity of Cancer Cells: An Interplay between HIF-1, GLUTs, and AMPK. Cancers 2020, 12, 862. [Google Scholar] [CrossRef] [Green Version]
  199. Sajnani, K.; Islam, F.; Smith, R.A.; Gopalan, V.; Lam, A.K. Genetic alterations in Krebs cycle and its impact on cancer pathogenesis. Biochimie 2017, 135, 164–172. [Google Scholar] [CrossRef]
  200. Herzig, S.; Shaw, R.J. AMPK: Guardian of metabolism and mitochondrial homeostasis. Nat. Rev. Mol. Cell Biol. 2018, 19, 121–135. [Google Scholar] [CrossRef] [Green Version]
  201. Moldogazieva, N.T.; Mokhosoev, I.M.; Feldman, N.B.; Lutsenko, S.V. ROS and RNS signalling: Adaptive redox switches through oxidative/nitrosative protein modifications. Free Radic. Res. 2018, 52, 507–543. [Google Scholar] [CrossRef]
  202. Chesney, J. 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase and tumor cell glycolysis. Curr. Opin. Clin. Nutr. Metab. Care 2006, 9, 535–539. [Google Scholar] [CrossRef] [PubMed]
  203. Iacobazzi, V.; Infantino, V. Citrate—New functions for an old metabolite. Biol. Chem. 2014, 395, 387–399. [Google Scholar] [CrossRef]
  204. Icard, P.; Coquerel, A.; Wu, Z.; Gligorov, J.; Fuks, D.; Fournel, L.; Lincet, H.; Simula, L. Understanding the Central Role of Citrate in the Metabolism of Cancer Cells and Tumors: An Update. Int. J. Mol. Sci. 2021, 22, 6587. [Google Scholar] [CrossRef]
  205. Corrêa-Ferreira, M.L.; do Rocio Andrade Pires, A.; Barbosa, I.R.; Echevarria, A.; Pedrassoli, G.H.; Winnischofer, S.M.B.; Noleto, G.R.; Cadena, S.M.S.C. The mesoionic compound MI-D changes energy metabolism and induces apoptosis in T98G glioma cells. Mol. Cell Biochem. 2022, 477, 2033–2045. [Google Scholar] [CrossRef]
  206. Liang, L.; Zhang, G.; Cheng, C.; Li, H.; Jin, T.; Su, C.; Xiao, Y.; Bradley, J.; Peberdy, M.A.; Ornato, J.P.; et al. High-resolution respirometry for evaluation of mitochondrial function on brain and heart homogenates in a rat model of cardiac arrest and cardiopulmonary resuscitation. Biomed. Pharmacother. 2021, 142, 111935. [Google Scholar] [CrossRef] [PubMed]
  207. Djafarzadeh, S.; Jakob, S.M. High-resolution Respirometry to Assess Mitochondrial Function in Permeabilized and Intact Cells. J. Vis. Exp. 2017, 120, e54985. [Google Scholar]
  208. Lu, Y.; Zhang, X.; Zhang, H.; Lan, J.; Huang, G.; Varin, E.; Lincet, H.; Poulain, L.; Icard, P. Citrate induces apoptotic cell death: A promising way to treat gastric carcinoma? Anticancer Res. 2011, 31, 797–805. [Google Scholar] [PubMed]
  209. Yousefi, S.; Owens, J.W.; Cesario, T.C. Citrate shows specific, dose-dependent lympholytic activity in neoplastic cell lines. Leuk. Lymphoma 2004, 45, 1657–1665. [Google Scholar] [CrossRef]
  210. Zhang, X.; Varin, E.; Allouche, S.; Lu, Y.; Poulain, L.; Icard, P. Effect of citrate on malignant pleural mesothelioma cells: A synergistic effect with cisplatin. Anticancer Res. 2009, 29, 1249–1254. [Google Scholar]
  211. Kruspig, B.; Nilchian, A.; Orrenius, S.; Zhivotovsky, B.; Gogvadze, V. Citrate kills tumor cells through activation of apical caspases. Cell Mol. Life Sci. 2012, 69, 4229–4237. [Google Scholar] [CrossRef] [Green Version]
  212. Nishimura, T.; Machida, S.; Kondo, M.; Terasaki, H.; Yokoyama, D.; Kurosaka, D. Enhancement of ON-bipolar cell responses of cone electroretinograms in rabbits with the Pro347Leu rhodopsin mutation. Investig. Ophthalmol. Vis. Sci. 2011, 52, 7610–7617. [Google Scholar] [CrossRef] [PubMed]
  213. Ren, J.G.; Seth, P.; Ye, H.; Guo, K.; Hanai, J.I.; Husain, Z.; Sukhatme, V.P. Citrate Suppresses Tumor Growth in Multiple Models through Inhibition of Glycolysis, the Tricarboxylic Acid Cycle and the IGF-1R Pathway. Sci. Rep. 2017, 7, 4537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Wang, T.A.; Zhang, X.D.; Guo, X.Y.; Xian, S.L.; Lu, Y.F. 3-bromopyruvate and sodium citrate target glycolysis, suppress survivin, and induce mitochondrial-mediated apoptosis in gastric cancer cells and inhibit gastric orthotopic transplantation tumor growth. Oncol. Rep. 2016, 35, 1287–1296. [Google Scholar] [CrossRef] [Green Version]
  215. Halabe, B.A. Hypothesis proved...citric acid (citrate) does improve cancer: A case of a patient suffering from medullary thyroid cancer. Med. Hypotheses 2009, 73, 271. [Google Scholar] [CrossRef] [PubMed]
  216. Bucay, A.H. Clinical report: A patient with primary peritoneal mesothelioma that has improved after taking citric acid orally. Clin. Res. Hepatol. Gastroenterol. 2011, 35, 241. [Google Scholar] [CrossRef]
  217. Icard, P.; Fournel, L.; Alifano, M.; Lincet, H. Extracellular Citrate and Cancer Metabolism-Letter. Cancer Res. 2018, 78, 5176. [Google Scholar] [CrossRef] [Green Version]
  218. Vahidi, R.; Safi, S.; Farsinejad, A.; Panahi, N. Citrate and celecoxib induce apoptosis and decrease necrosis in synergistic manner in canine mammary tumor cells. Cell Mol. Biol. 2015, 61, 22–28. [Google Scholar]
  219. Cadena, S.M.; Carnieri, E.G.; Echevarria, A.; de Oliveira, M.B. Interference of MI-D, a new mesoionic compound, on artificial and native membranes. Cell Biochem. Funct. 2002, 20, 31–37. [Google Scholar] [CrossRef] [PubMed]
  220. Parsons, D.W.; Jones, S.; Zhang, X.; Lin, J.C.; Leary, R.J.; Angenendt, P.; Mankoo, P.; Carter, H.; Siu, I.M.; Gallia, G.L.; et al. An integrated genomic analysis of human glioblastoma multiforme. Science 2008, 321, 1807–1812. [Google Scholar] [CrossRef] [Green Version]
  221. Yan, H.; Parsons, D.W.; Jin, G.; McLendon, R.; Rasheed, B.A.; Yuan, W.; Kos, I.; Batinic-Haberle, I.; Jones, S.; Riggins, G.J.; et al. IDH1 and IDH2 mutations in gliomas. N. Engl. J. Med. 2009, 360, 765–773. [Google Scholar] [CrossRef]
  222. Kloosterhof, N.K.; Bralten, L.B.; Dubbink, H.J.; French, P.J.; van den Bent, M.J. Isocitrate dehydrogenase-1 mutations: A fundamentally new understanding of diffuse glioma? Lancet Oncol. 2011, 12, 83–91. [Google Scholar] [CrossRef] [PubMed]
  223. Yan, H.; Bigner, D.D.; Velculescu, V.; Parsons, D.W. Mutant metabolic enzymes are at the origin of gliomas. Cancer Res. 2009, 69, 9157–9159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Popovici-Muller, J.; Lemieux, R.M.; Artin, E.; Saunders, J.O.; Salituro, F.G.; Travins, J.; Cianchetta, G.; Cai, Z.; Zhou, D.; Cui, D.; et al. Discovery of AG-120 (Ivosidenib): A First-in-Class Mutant IDH1 Inhibitor for the Treatment of IDH1 Mutant Cancers. ACS Med. Chem. Lett. 2018, 9, 300–305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Chaturvedi, A.; Araujo Cruz, M.M.; Jyotsana, N.; Sharma, A.; Yun, H.; Görlich, K.; Wichmann, M.; Schwarzer, A.; Preller, M.; Thol, F.; et al. Mutant IDH1 promotes leukemogenesis in vivo and can be specifically targeted in human AML. Blood 2013, 122, 2877–2887. [Google Scholar] [CrossRef] [PubMed]
  226. Tesileanu, C.M.S.; Vallentgoed, W.R.; Sanson, M.; Taal, W.; Clement, P.M.; Wick, W.; Brandes, A.A.; Baurain, J.F.; Chinot, O.L.; Wheeler, H.; et al. Non-IDH1-R132H IDH1/2 mutations are associated with increased DNA methylation and improved survival in astrocytomas, compared to IDH1-R132H mutations. Acta Neuropathol. 2021, 141, 945–957. [Google Scholar] [CrossRef]
  227. Suzuki, H.; Aoki, K.; Chiba, K.; Sato, Y.; Shiozawa, Y.; Shiraishi, Y.; Shimamura, T.; Niida, A.; Motomura, K.; Ohka, F.; et al. Mutational landscape and clonal architecture in grade II and III gliomas. Nat. Genet. 2015, 47, 458–468. [Google Scholar] [CrossRef]
  228. Louis, D.N.; Perry, A.; Reifenberger, G.; von Deimling, A.; Figarella-Branger, D.; Cavenee, W.K.; Ohgaki, H.; Wiestler, O.D.; Kleihues, P.; Ellison, D.W. The 2016 World Health Organization Classification of Tumors of the Central Nervous System: A summary. Acta Neuropathol. 2016, 131, 803–820. [Google Scholar] [CrossRef] [Green Version]
  229. Louis, D.N.; Perry, A.; Wesseling, P.; Brat, D.J.; Cree, I.A.; Figarella-Branger, D.; Hawkins, C.; Ng, H.K.; Pfister, S.M.; Reifenberger, G.; et al. The 2021 WHO Classification of Tumors of the Central Nervous System: A summary. Neuro Oncol. 2021, 23, 1231–1251. [Google Scholar] [CrossRef]
  230. Rendina, A.R.; Pietrak, B.; Smallwood, A.; Zhao, H.; Qi, H.; Quinn, C.; Adams, N.D.; Concha, N.; Duraiswami, C.; Thrall, S.H.; et al. Mutant IDH1 enhances the production of 2-hydroxyglutarate due to its kinetic mechanism. Biochemistry 2013, 52, 4563–4577. [Google Scholar] [CrossRef]
  231. Dang, L.; White, D.W.; Gross, S.; Bennett, B.D.; Bittinger, M.A.; Driggers, E.M.; Fantin, V.R.; Jang, H.G.; Jin, S.; Keenan, M.C.; et al. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 2009, 462, 739–744. [Google Scholar] [CrossRef] [Green Version]
  232. Showalter, M.R.; Hatakeyama, J.; Cajka, T.; VanderVorst, K.; Carraway, K.L.; Fiehn, O. Replication Study: The common feature of leukemia-associated IDH1 and IDH2 mutations is a neomorphic enzyme activity converting alpha-ketoglutarate to 2-hydroxyglutarate. Elife 2017, 6, 225–234. [Google Scholar] [CrossRef] [PubMed]
  233. Jin, G.; Reitman, Z.J.; Duncan, C.G.; Spasojevic, I.; Gooden, D.M.; Rasheed, B.A.; Yang, R.; Lopez, G.Y.; He, Y.; McLendon, R.E.; et al. Disruption of wild-type IDH1 suppresses D-2-hydroxyglutarate production in IDH1-mutated gliomas. Cancer Res. 2013, 73, 496–501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Cuyàs, E.; Fernández-Arroyo, S.; Corominas-Faja, B.; Rodríguez-Gallego, E.; Bosch-Barrera, J.; Martin-Castillo, B.; De Llorens, R.; Joven, J.; Menendez, J.A. Oncometabolic mutation IDH1 R132H confers a metformin-hypersensitive phenotype. Oncotarget 2015, 6, 12279–12296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Xu, H.; Sun, Y.; You, B.; Huang, C.P.; Ye, D.; Chang, C. Androgen receptor reverses the oncometabolite R-2-hydroxyglutarate-induced prostate cancer cell invasion via suppressing the circRNA-51217/miRNA-646/TGFbeta1/p-Smad2/3 signaling. Cancer Lett. 2020, 472, 151–164. [Google Scholar] [CrossRef]
  236. Chen, J.Y.; Lai, Y.S.; Tsai, H.J.; Kuo, C.C.; Yen, B.L.; Yeh, S.P.; Sun, H.S.; Hung, W.C. The oncometabolite R-2-hydroxyglutarate activates NF-kappaB-dependent tumor-promoting stromal niche for acute myeloid leukemia cells. Sci. Rep. 2016, 6, 32428. [Google Scholar] [CrossRef] [Green Version]
  237. Larsen, C.J. (R)-2-hydroxyglutarate or (R)-2HG is an oncometabolite. Bull. Cancer 2013, 100, 655. [Google Scholar] [CrossRef]
  238. Adam, J.; Yang, M.; Soga, T.; Pollard, P.J. Rare insights into cancer biology. Oncogene 2014, 33, 2547–2556. [Google Scholar] [CrossRef] [Green Version]
  239. Menendez, J.A.; Alarcon, T. Metabostemness: A new cancer hallmark. Front. Oncol. 2014, 4, 262. [Google Scholar] [CrossRef] [Green Version]
  240. Zhao, S.; Lin, Y.; Xu, W.; Jiang, W.; Zha, Z.; Wang, P.; Yu, W.; Li, Z.; Gong, L.; Peng, Y.; et al. Glioma-derived mutations in IDH1 dominantly inhibit IDH1 catalytic activity and induce HIF-1alpha. Science 2009, 324, 261–265. [Google Scholar] [CrossRef] [Green Version]
  241. Semukunzi, H.; Roy, D.; Li, H.; Khan, G.J.; Lyu, X.; Yuan, S.; Lin, S. IDH mutations associated impact on related cancer epidemiology and subsequent effect toward HIF-1alpha. Biomed. Pharmacother. 2017, 89, 805–811. [Google Scholar] [CrossRef]
  242. Koivunen, P.; Lee, S.; Duncan, C.G.; Lopez, G.; Lu, G.; Ramkissoon, S.; Losman, J.A.; Joensuu, P.; Bergmann, U.; Gross, S.; et al. Transformation by the (R)-enantiomer of 2-hydroxyglutarate linked to EGLN activation. Nature 2012, 483, 484–488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Qing, Y.; Dong, L.; Gao, L.; Li, C.; Li, Y.; Han, L.; Prince, E.; Tan, B.; Deng, X.; Wetzel, C.; et al. R-2-hydroxyglutarate attenuates aerobic glycolysis in leukemia by targeting the FTO/m (6)A/PFKP/LDHB axis. Mol. Cell 2021, 81, 922–939.e9. [Google Scholar] [CrossRef]
  244. Fu, X.; Chin, R.M.; Vergnes, L.; Hwang, H.; Deng, G.; Xing, Y.; Pai, M.Y.; Li, S.; Ta, L.; Fazlollahi, F.; et al. 2-Hydroxyglutarate Inhibits ATP Synthase and mTOR Signaling. Cell Metab. 2015, 22, 508–515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Wen, H.; Cho, H.R.; Yun, T.; Kim, H.; Park, C.K.; Lee, S.H.; Choi, S.H.; Park, S. Metabolomic comparison between cells over-expressing isocitrate dehydrogenase 1 and 2 mutants and the effects of an inhibitor on the metabolism. J. Neurochem. 2015, 132, 183–193. [Google Scholar] [CrossRef]
  246. Emir, U.E.; Larkin, S.J.; de Pennington, N.; Voets, N.; Plaha, P.; Stacey, R.; Al-Qahtani, K.; Mccullagh, J.; Schofield, C.J.; Clare, S.; et al. Noninvasive Quantification of 2-Hydroxyglutarate in Human Gliomas with IDH1 and IDH2 Mutations. Cancer Res. 2016, 76, 43–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Izquierdo-Garcia, J.L.; Viswanath, P.; Eriksson, P.; Chaumeil, M.M.; Pieper, R.O.; Phillips, J.J.; Ronen, S.M. Metabolic reprogramming in mutant IDH1 glioma cells. PLoS ONE 2015, 10, e0118781. [Google Scholar] [CrossRef] [Green Version]
  248. Lenting, K.; Khurshed, M.; Peeters, T.H.; van den Heuvel, C.N.A.M.; van Lith, S.A.M.; de Bitter, T.; Hendriks, W.; Span, P.N.; Molenaar, R.J.; Botman, D.; et al. Isocitrate dehydrogenase 1-mutated human gliomas depend on lactate and glutamate to alleviate metabolic stress. FASEB J. 2019, 33, 557–571. [Google Scholar] [CrossRef]
  249. Walsby-Tickle, J.; Gannon, J.; Hvinden, I.; Bardella, C.; Abboud, M.I.; Nazeer, A.; Hauton, D.; Pires, E.; Cadoux-Hudson, T.; Schofield, C.J.; et al. Anion-exchange chromatography mass spectrometry provides extensive coverage of primary metabolic pathways revealing altered metabolism in IDH1 mutant cells. Commun. Biol. 2020, 3, 247. [Google Scholar] [CrossRef]
  250. Reitman, Z.J.; Jin, G.; Karoly, E.D.; Spasojevic, I.; Yang, J.; Kinzler, K.W.; He, Y.; Bigner, D.D.; Vogelstein, B.; Yan, H. Profiling the effects of isocitrate dehydrogenase 1 and 2 mutations on the cellular metabolome. Proc. Natl. Acad. Sci. USA 2011, 108, 3270–3275. [Google Scholar] [CrossRef] [Green Version]
  251. Ohka, F.; Ito, M.; Ranjit, M.; Senga, T.; Motomura, A.; Motomura, K.; Saito, K.; Kato, K.; Kato, Y.; Wakabayashi, T.; et al. Quantitative metabolome analysis profiles activation of glutaminolysis in glioma with IDH1 mutation. Tumour. Biol. 2014, 35, 5911–5920. [Google Scholar] [CrossRef]
  252. Fack, F.; Tardito, S.; Hochart, G.; Oudin, A.; Zheng, L.; Fritah, S.; Golebiewska, A.; Nazarov, P.V.; Bernard, A.; Hau, A.C.; et al. Altered metabolic landscape in IDH-mutant gliomas affects phospholipid, energy, and oxidative stress pathways. EMBO Mol. Med. 2017, 9, 1681–1695. [Google Scholar] [CrossRef] [PubMed]
  253. Zhou, L.; Wang, Z.; Hu, C.; Zhang, C.; Kovatcheva-Datchary, P.; Yu, D.; Liu, S.; Ren, F.; Wang, X.; Li, Y.; et al. Integrated Metabolomics and Lipidomics Analyses Reveal Metabolic Reprogramming in Human Glioma with IDH1 Mutation. J. Proteome Res. 2019, 18, 960–969. [Google Scholar] [CrossRef] [PubMed]
  254. Hujber, Z.; Horváth, G.; Petővári, G.; Krencz, I.; Dankó, T.; Mészáros, K.; Rajnai, H.; Szoboszlai, N.; Leenders, W.P.J.; Jeney, A.; et al. GABA, glutamine, glutamate oxidation and succinic semialdehyde dehydrogenase expression in human gliomas. J. Exp. Clin. Cancer Res. 2018, 37, 271. [Google Scholar] [CrossRef] [Green Version]
  255. He, X.; Wu, N.; Li, R.; Zhang, H.; Zhao, Y.; Nie, Y.; Wu, J. IDH2, a novel target of OGT, facilitates glucose uptake and cellular bioenergy production via NF-kappaB signaling to promote colorectal cancer progression. Cell Oncol. 2022, 46, 145–164. [Google Scholar] [CrossRef] [PubMed]
  256. Khurshed, M.; Molenaar, R.J.; Lenting, K.; Leenders, W.P.; van Noorden, C.J.F. In silico gene expression analysis reveals glycolysis and acetate anaplerosis in IDH1 wild-type glioma and lactate and glutamate anaplerosis in IDH1-mutated glioma. Oncotarget 2017, 8, 49165–49177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Grassian, A.R.; Parker, S.J.; Davidson, S.M.; Divakaruni, A.S.; Green, C.R.; Zhang, X.; Slocum, K.L.; Pu, M.; Lin, F.; Vickers, C.; et al. IDH1 mutations alter citric acid cycle metabolism and increase dependence on oxidative mitochondrial metabolism. Cancer Res. 2014, 74, 3317–3331. [Google Scholar] [CrossRef] [Green Version]
  258. Strickland, M.; Stoll, E.A. Metabolic Reprogramming in Glioma. Front. Cell Dev. Biol. 2017, 5, 43. [Google Scholar] [CrossRef] [Green Version]
  259. Vatrinet, R.; Leone, G.; De Luise, M.; Girolimetti, G.; Vidone, M.; Gasparre, G.; Porcelli, A.M. The alpha-ketoglutarate dehydrogenase complex in cancer metabolic plasticity. Cancer Metab. 2017, 5, 3. [Google Scholar] [CrossRef] [Green Version]
  260. Dekker, L.J.M.; Verheul, C.; Wensveen, N.; Leenders, W.; Lamfers, M.L.M.; Leenstra, S.; Luider, T.M. Effects of the IDH1 R132H Mutation on the Energy Metabolism: A Comparison between Tissue and Corresponding Primary Glioma Cell Cultures. ACS Omega 2022, 7, 3568–3578. [Google Scholar] [CrossRef]
  261. Park, J.S.; Burckhardt, C.J.; Lazcano, R.; Solis, L.M.; Isogai, T.; Li, L.; Chen, C.S.; Gao, B.; Minna, J.D.; Bachoo, R.; et al. Mechanical regulation of glycolysis via cytoskeleton architecture. Nature 2020, 578, 621–626. [Google Scholar] [CrossRef]
  262. Izquierdo-Garcia, J.L.; Viswanath, P.; Eriksson, P.; Cai, L.; Radoul, M.; Chaumeil, M.M.; Blough, M.; Luchman, H.A.; Weiss, S.; Cairncross, J.G.; et al. IDH1 Mutation Induces Reprogramming of Pyruvate Metabolism. Cancer Res. 2015, 75, 2999–3009. [Google Scholar] [CrossRef] [Green Version]
  263. Garrett, M.; Sperry, J.; Braas, D.; Yan, W.; Le, T.M.; Mottahedeh, J.; Ludwig, K.; Eskin, A.; Qin, Y.; Levy, R.; et al. Metabolic characterization of isocitrate dehydrogenase (IDH) mutant and IDH wildtype gliomaspheres uncovers cell type-specific vulnerabilities. Cancer Metab. 2018, 6, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Dekker, L.J.M.; Wu, S.; Jurriëns, C.; Mustafa, D.A.N.; Grevers, F.; Burgers, P.C.; Sillevis Smitt, P.A.E.; Kros, J.M.; Luider, T.M. Metabolic changes related to the IDH1 mutation in gliomas preserve TCA-cycle activity: An investigation at the protein level. FASEB J. 2020, 34, 3646–3657. [Google Scholar] [CrossRef] [Green Version]
  265. Izquierdo-Garcia, J.L.; Cai, L.M.; Chaumeil, M.M.; Eriksson, P.; Robinson, A.E.; Pieper, R.O.; Phillips, J.J.; Ronen, S.M. Glioma cells with the IDH1 mutation modulate metabolic fractional flux through pyruvate carboxylase. PLoS ONE 2014, 9, e108289. [Google Scholar] [CrossRef]
  266. Roboz, G.J.; DiNardo, C.D.; Stein, E.M.; de Botton, S.; Mims, A.S.; Prince, G.T.; Altman, J.K.; Arellano, M.L.; Donnellan, W.; Erba, H.P.; et al. Ivosidenib induces deep durable remissions in patients with newly diagnosed IDH1-mutant acute myeloid leukemia. Blood 2020, 135, 463–471. [Google Scholar] [CrossRef] [Green Version]
  267. Stein, E.M.; DiNardo, C.D.; Pollyea, D.A.; Fathi, A.T.; Roboz, G.J.; Altman, J.K.; Stone, R.M.; DeAngelo, D.J.; Levine, R.L.; Flinn, I.W.; et al. Enasidenib in mutant IDH2 relapsed or refractory acute myeloid leukemia. Blood 2017, 130, 722–731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Cai, X.; Chen, Z.; Huang, C.; Shen, J.; Zeng, W.; Feng, S.; Liu, Y.; Li, S.; Chen, M. Development of a novel glycolysis-related genes signature for isocitrate dehydrogenase 1-associated glioblastoma multiforme. Front. Immunol. 2022, 13, 950917. [Google Scholar] [CrossRef]
  269. van Noorden, C.J.F.; Hira, V.V.V.; van Dijck, A.J.; Novak, M.; Breznik, B.; Molenaar, R.J. Energy Metabolism in IDH1 Wild-Type and IDH1-Mutated Glioblastoma Stem Cells: A Novel Target for Therapy? Cells 2021, 10, 705. [Google Scholar] [CrossRef] [PubMed]
  270. Kessler, J.; Hohmann, T.; Güttler, A.; Petrenko, M.; Ostheimer, C.; Hohmann, U.; Bache, M.; Dehghani, F.; Vordermark, D. Radiosensitization and a Less Aggressive Phenotype of Human Malignant Glioma Cells Expressing Isocitrate Dehydrogenase 1 (IDH1) Mutant Protein: Dissecting the Mechanisms. Cancers 2019, 11, 889. [Google Scholar] [CrossRef] [Green Version]
  271. Masson, N.; Ratcliffe, P.J. Hypoxia signaling pathways in cancer metabolism: The importance of co-selecting interconnected physiological pathways. Cancer Metab. 2014, 2, 3. [Google Scholar] [CrossRef] [Green Version]
  272. Ellinghaus, P.; Heisler, I.; Unterschemmann, K.; Haerter, M.; Beck, H.; Greschat, S.; Ehrmann, A.; Summer, H.; Flamme, I.; Oehme, F.; et al. BAY 87-2243, a highly potent and selective inhibitor of hypoxia-induced gene activation has antitumor activities by inhibition of mitochondrial complex I. Cancer Med. 2013, 2, 611–624. [Google Scholar] [CrossRef] [PubMed]
  273. Sica, V.; Bravo-San Pedro, J.M.; Pietrocola, F.; Izzo, V.; Maiuri, M.C.; Kroemer, G.; Galluzzi, L. Assessment of Glycolytic Flux and Mitochondrial Respiration in the Course of Autophagic Responses. Methods Enzymol. 2017, 588, 155–170. [Google Scholar] [PubMed]
  274. Sica, V.; Bravo-San Pedro, J.M.; Izzo, V.; Pol, J.; Pierredon, S.; Enot, D.; Durand, S.; Bossut, N.; Chery, A.; Souquere, S.; et al. Lethal Poisoning of Cancer Cells by Respiratory Chain Inhibition plus Dimethyl alpha-Ketoglutarate. Cell Rep. 2019, 27, 820–834.e9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Sica, V.; Bravo-San, P.J.; Kroemer, G. A strategy for poisoning cancer cell metabolism: Inhibition of oxidative phosphorylation coupled to anaplerotic saturation. Int. Rev. Cell Mol. Biol. 2019, 347, 27–37. [Google Scholar]
  276. Chung, C.; Sweha, S.R.; Pratt, D.; Tamrazi, B.; Panwalkar, P.; Banda, A.; Bayliss, J.; Hawes, D.; Yang, F.; Lee, H.J.; et al. Integrated Metabolic and Epigenomic Reprograming by H3K27M Mutations in Diffuse Intrinsic Pontine Gliomas. Cancer Cell 2020, 38, 334–349.e9. [Google Scholar] [CrossRef]
  277. Sperry, J.; Condro, M.C.; Guo, L.; Braas, D.; Vanderveer-Harris, N.; Kim, K.K.O.; Pope, W.B.; Divakaruni, A.S.; Lai, A.; Christofk, H.; et al. Glioblastoma Utilizes Fatty Acids and Ketone Bodies for Growth Allowing Progression during Ketogenic Diet Therapy. iScience 2020, 23, 101453. [Google Scholar] [CrossRef]
  278. Kant, S.; Kesarwani, P.; Prabhu, A.; Graham, S.F.; Buelow, K.L.; Nakano, I.; Chinnaiyan, P. Enhanced fatty acid oxidation provides glioblastoma cells metabolic plasticity to accommodate to its dynamic nutrient microenvironment. Cell Death Dis. 2020, 11, 253. [Google Scholar] [CrossRef] [Green Version]
  279. Brat, D.J.; Verhaak, R.G.; Aldape, K.D.; Yung, W.K.; Salama, S.R.; Cooper, L.A.; Rheinbay, E.; Miller, C.R.; Vitucci, M.; Morozova, O.; et al. Comprehensive, Integrative Genomic Analysis of Diffuse Lower-Grade Gliomas. N. Engl. J. Med. 2015, 372, 2481–2498. [Google Scholar]
  280. Patel, J.P.; Gönen, M.; Figueroa, M.E.; Fernandez, H.; Sun, Z.; Racevskis, J.; Van Vlierberghe, P.; Dolgalev, I.; Thomas, S.; Aminova, O.; et al. Prognostic relevance of integrated genetic profiling in acute myeloid leukemia. N. Engl. J. Med. 2012, 366, 1079–1089. [Google Scholar] [CrossRef] [Green Version]
  281. Cuperlovic-Culf, M.; Ferguson, D.; Culf, A.; Morin, P., Jr.; Touaibia, M. 1H NMR metabolomics analysis of glioblastoma subtypes: Correlation between metabolomics and gene expression characteristics. J. Biol. Chem. 2012, 287, 20164–20175. [Google Scholar] [CrossRef] [Green Version]
  282. Lewis, C.A.; Parker, S.J.; Fiske, B.P.; McCloskey, D.; Gui, D.Y.; Green, C.R.; Vokes, N.I.; Feist, A.M.; Vander Heiden, M.G.; Metallo, C.M. Tracing compartmentalized NADPH metabolism in the cytosol and mitochondria of mammalian cells. Mol. Cell 2014, 55, 253–263. [Google Scholar] [CrossRef] [Green Version]
  283. Kim, S.H.; Kim, H.; Lee, J.H.; Park, J.W. Oxalomalate suppresses metastatic melanoma through IDH-targeted stress response to ROS. Free Radic. Res. 2019, 53, 418–429. [Google Scholar] [CrossRef]
  284. Erol, A. Type 2 diabetes and cancer as redox diseases? Lancet 2014, 384, 853–854. [Google Scholar] [CrossRef] [PubMed]
  285. Erol, A. Systemic DNA damage response and metabolic syndrome as a premalignant state. Curr. Mol. Med. 2010, 10, 321–334. [Google Scholar] [CrossRef] [PubMed]
  286. Wei, Z.; Ye, S.; Feng, H.; Zeng, C.; Dong, X.; Zeng, X.; Zeng, L.; Lin, X.; Liu, Q.; Yao, J. Silybin suppresses ovarian cancer cell proliferation by inhibiting isocitrate dehydrogenase 1 activity. Cancer Sci. 2022, 113, 3032–3043. [Google Scholar] [CrossRef] [PubMed]
  287. Kappler, M.; Pabst, U.; Weinholdt, C.; Taubert, H.; Rot, S.; Kaune, T.; Kotrba, J.; Porsch, M.; Güttler, A.; Bache, M.; et al. Causes and Consequences of a Glutamine Induced Normoxic HIF1 Activity for the Tumor Metabolism. Int. J. Mol. Sci. 2019, 20, 4742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  288. Sun, R.C.; Denko, N.C. Hypoxic regulation of glutamine metabolism through HIF1 and SIAH2 supports lipid synthesis that is necessary for tumor growth. Cell Metab. 2014, 19, 285–292. [Google Scholar] [CrossRef] [Green Version]
  289. Burr, S.P.; Costa, A.S.; Grice, G.L.; Timms, R.T.; Lobb, I.T.; Freisinger, P.; Dodd, R.B.; Dougan, G.; Lehner, P.J.; Frezza, C.; et al. Mitochondrial Protein Lipoylation and the 2-Oxoglutarate Dehydrogenase Complex Controls HIF1alpha Stability in Aerobic Conditions. Cell Metab. 2016, 24, 740–752. [Google Scholar] [CrossRef] [Green Version]
  290. Ziello, J.E.; Jovin, I.S.; Huang, Y. Hypoxia-Inducible Factor (HIF)-1 regulatory pathway and its potential for therapeutic intervention in malignancy and ischemia. Yale J. Biol. Med. 2007, 80, 51–60. [Google Scholar]
  291. Hansen, G.E.; Gibson, G.E. The alpha-Ketoglutarate Dehydrogenase Complex as a Hub of Plasticity in Neurodegeneration and Regeneration. Int. J. Mol. Sci. 2022, 23, 12403. [Google Scholar] [CrossRef]
  292. You, Q.; Wang, J.; Yu, Y.; Li, F.; Meng, L.; Chen, M.; Yang, Q.; Xu, Z.; Sun, J.; Zhuo, W.; et al. The histone deacetylase SIRT6 promotes glycolysis through the HIF-1alpha/HK2 signaling axis and induces erlotinib resistance in non-small cell lung cancer. Apoptosis 2022, 27, 883–898. [Google Scholar] [CrossRef] [PubMed]
  293. Zhou, L.; Zhao, Y.; Pan, L.C.; Wang, J.; Shi, X.J.; Du, G.S.; He, Q. Sirolimus increases the anti-cancer effect of Huai Er by regulating hypoxia inducible factor-1alpha-mediated glycolysis in hepatocellular carcinoma. World J. Gastroenterol. 2022, 28, 4600–4619. [Google Scholar] [CrossRef] [PubMed]
  294. Semba, H.; Takeda, N.; Isagawa, T.; Sugiura, Y.; Honda, K.; Wake, M.; Miyazawa, H.; Yamaguchi, Y.; Miura, M.; Jenkins, D.M.; et al. HIF-1alpha-PDK1 axis-induced active glycolysis plays an essential role in macrophage migratory capacity. Nat. Commun. 2016, 7, 11635. [Google Scholar] [CrossRef] [Green Version]
  295. Williams, N.C.; Ryan, D.G.; Costa, A.S.H.; Mills, E.L.; Jedrychowski, M.P.; Cloonan, S.M.; Frezza, C.; O’Neill, L.A. Signaling metabolite L-2-hydroxyglutarate activates the transcription factor HIF-1alpha in lipopolysaccharide-activated macrophages. J. Biol. Chem. 2022, 298, 101501. [Google Scholar] [CrossRef] [PubMed]
  296. Intlekofer, A.M.; Dematteo, R.G.; Venneti, S.; Finley, L.W.; Lu, C.; Judkins, A.R.; Rustenburg, A.S.; Grinaway, P.B.; Chodera, J.D.; Cross, J.R.; et al. Hypoxia Induces Production of L-2-Hydroxyglutarate. Cell Metab. 2015, 22, 304–311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  297. Durán, R.V.; Oppliger, W.; Robitaille, A.M.; Heiserich, L.; Skendaj, R.; Gottlieb, E.; Hall, M.N. Glutaminolysis activates Rag-mTORC1 signaling. Mol. Cell 2012, 47, 349–358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  298. Durán, R.V.; MacKenzie, E.D.; Boulahbel, H.; Frezza, C.; Heiserich, L.; Tardito, S.; Bussolati, O.; Rocha, S.; Hall, M.N.; Gottlieb, E. HIF-independent role of prolyl hydroxylases in the cellular response to amino acids. Oncogene 2013, 32, 4549–4556. [Google Scholar] [CrossRef] [Green Version]
  299. Jiao, L.; Zhang, H.L.; Li, D.D.; Yang, K.L.; Tang, J.; Li, X.; Ji, J.; Yu, Y.; Wu, R.Y.; Ravichandran, S.; et al. Regulation of glycolytic metabolism by autophagy in liver cancer involves selective autophagic degradation of HK2 (hexokinase 2). Autophagy 2018, 14, 671–684. [Google Scholar] [CrossRef]
  300. Yin, K.; Lee, J.; Liu, Z.; Kim, H.; Martin, D.R.; Wu, D.; Liu, M.; Xue, X. Mitophagy protein PINK1 suppresses colon tumor growth by metabolic reprogramming via p53 activation and reducing acetyl-CoA production. Cell Death Differ. 2021, 28, 2421–2435. [Google Scholar] [CrossRef]
  301. Lue, H.W.; Podolak, J.; Kolahi, K.; Cheng, L.; Rao, S.; Garg, D.; Xue, C.H.; Rantala, J.K.; Tyner, J.W.; Thornburg, K.L.; et al. Metabolic reprogramming ensures cancer cell survival despite oncogenic signaling blockade. Genes Dev. 2017, 31, 2067–2084. [Google Scholar] [CrossRef]
  302. Amaravadi, R.K.; Lippincott-Schwartz, J.; Yin, X.M.; Weiss, W.A.; Takebe, N.; Timmer, W.; DiPaola, R.S.; Lotze, M.T.; White, E. Principles and current strategies for targeting autophagy for cancer treatment. Clin. Cancer Res. 2011, 17, 654–666. [Google Scholar] [CrossRef] [Green Version]
  303. Sehgal, A.R.; Konig, H.; Johnson, D.E.; Tang, D.; Amaravadi, R.K.; Boyiadzis, M.; Lotze, M.T. You eat what you are: Autophagy inhibition as a therapeutic strategy in leukemia. Leukemia 2015, 29, 517–525. [Google Scholar] [CrossRef]
  304. Liu, L.; Li, L.; Li, M.; Luo, Z. Autophagy-Dependent Ferroptosis as a Therapeutic Target in Cancer. ChemMedChem 2021, 16, 942–2950. [Google Scholar] [CrossRef] [PubMed]
  305. Rebecca, V.W.; Amaravadi, R.K. Emerging strategies to effectively target autophagy in cancer. Oncogene 2016, 35, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Zhang, J.; Wang, G.; Zhou, Y.; Chen, Y.; Ouyang, L.; Liu, B. Mechanisms of autophagy and relevant small-molecule compounds for targeted cancer therapy. Cell Mol. Life Sci. 2018, 75, 1803–1826. [Google Scholar] [CrossRef] [PubMed]
  307. Li, L.; Yang, L.; Fan, Z.; Xue, W.; Shen, Z.; Yuan, Y.; Sun, X.; Wang, D.; Lian, J.; Wang, L.; et al. Hypoxia-induced GBE1 expression promotes tumor progression through metabolic reprogramming in lung adenocarcinoma. Signal Transduct. Target. Ther. 2020, 5, 54. [Google Scholar] [CrossRef] [PubMed]
  308. Huangyang, P.; Li, F.; Lee, P.; Nissim, I.; Weljie, A.M.; Mancuso, A.; Li, B.; Keith, B.; Yoon, S.S.; Simon, M.C. Fructose-1,6-Bisphosphatase 2 Inhibits Sarcoma Progression by Restraining Mitochondrial Biogenesis. Cell Metab. 2020, 31, 174–188.e7. [Google Scholar] [CrossRef] [Green Version]
  309. Tomlinson, I.P.; Alam, N.A.; Rowan, A.J.; Barclay, E.; Jaeger, E.E.; Kelsell, D.; Leigh, I.; Gorman, P.; Lamlum, H.; Rahman, S.; et al. Germline mutations in FH predispose to dominantly inherited uterine fibroids, skin leiomyomata and papillary renal cell cancer. Nat. Genet. 2002, 30, 406–410. [Google Scholar]
  310. Baysal, B.E.; Ferrell, R.E.; Willett-Brozick, J.E.; Lawrence, E.C.; Myssiorek, D.; Bosch, A.; van der Mey, A.; Taschner, P.E.; Rubinstein, W.S.; Myers, E.N.; et al. Mutations in SDHD, a mitochondrial complex II gene, in hereditary paraganglioma. Science 2000, 287, 848–851. [Google Scholar] [CrossRef]
  311. King, A.; Selak, M.A.; Gottlieb, E. Succinate dehydrogenase and fumarate hydratase: Linking mitochondrial dysfunction and cancer. Oncogene 2006, 25, 4675–4782. [Google Scholar] [CrossRef] [Green Version]
  312. Wang, X.H.; Jiang, Z.H.; Yang, H.M.; Zhang, Y.; Xu, L.H. Hypoxia-induced FOXO4/LDHA axis modulates gastric cancer cell glycolysis and progression. Clin. Transl. Med. 2021, 11, e279. [Google Scholar] [CrossRef]
  313. Li, Y.; Lin, S.; Li, L.; Tang, Z.; Hu, Y.; Ban, X.; Zeng, T.; Zhou, Y.; Zhu, Y.; Gao, S.; et al. PDSS2 Deficiency Induces Hepatocarcinogenesis by Decreasing Mitochondrial Respiration and Reprogramming Glucose Metabolism. Cancer Res. 2018, 78, 4471–4481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  314. Kim, J.; DeBerardinis, R.J. Mechanisms and Implications of Metabolic Heterogeneity in Cancer. Cell Metab. 2019, 30, 434–446. [Google Scholar] [CrossRef] [PubMed]
  315. Cancer Genome Atlas Research Network, Comprehensive molecular profiling of lung adenocarcinoma. Nature 2014, 511, 543–550. [CrossRef] [PubMed] [Green Version]
  316. Li, C.; Tian, C.; Liu, Y.; Liang, J.; Zeng, Y.; Yang, Q.; Liu, Y.; Wu, D.; Wu, J.; Wang, J.; et al. Comprehensive Profiling Reveals Distinct Microenvironment and Metabolism Characterization of Lung Adenocarcinoma. Front. Genet. 2021, 12, 619821. [Google Scholar] [CrossRef]
  317. Humpton, T.J.; Vousden, K.H. Regulation of Cellular Metabolism and Hypoxia by p53. Cold Spring Harb. Perspect Med. 2016, 6, a026146. [Google Scholar] [CrossRef] [Green Version]
  318. Kruiswijk, F.; Labuschagne, C.F.; Vousden, K.H. p53 in survival, death and metabolic health: A lifeguard with a licence to kill. Nat. Rev. Mol. Cell Biol. 2015, 16, 393–405. [Google Scholar] [CrossRef]
  319. Shen, L.; Sun, X.; Fu, Z.; Yang, G.; Li, J.; Yao, L. The fundamental role of the p53 pathway in tumor metabolism and its implication in tumor therapy. Clin. Cancer Res. 2012, 18, 1561–1567. [Google Scholar] [CrossRef] [Green Version]
  320. García-Cao, I.; García-Cao, M.; Martín-Caballero, J.; Criado, L.M.; Klatt, P.; Flores, J.M.; Weill, J.C.; Blasco, M.A.; Serrano, M. “Super p53” mice exhibit enhanced DNA damage response, are tumor resistant and age normally. EMBO J. 2002, 21, 6225–6235. [Google Scholar] [CrossRef] [Green Version]
  321. Matheu, A.; Maraver, A.; Klatt, P.; Flores, I.; Garcia-Cao, I.; Borras, C.; Flores, J.M.; Viña, J.; Blasco, M.A.; Serrano, M. Delayed ageing through damage protection by the Arf/p53 pathway. Nature 2007, 448, 375–379. [Google Scholar] [CrossRef]
  322. Carrasco-Garcia, E.; Moreno, M.; Moreno-Cugnon, L.; Matheu, A. Increased Arf/p53 activity in stem cells, aging and cancer. Aging Cell 2017, 16, 219–225. [Google Scholar] [CrossRef] [PubMed]
  323. Mendrysa, S.M.; O’Leary, K.A.; McElwee, M.K.; Michalowski, J.; Eisenman, R.N.; Powell, D.A.; Perry, M.E. Tumor suppression and normal aging in mice with constitutively high p53 activity. Genes Dev. 2006, 20, 16–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Bykov, V.J.N.; Eriksson, S.E.; Bianchi, J.; Wiman, K.G. Targeting mutant p53 for efficient cancer therapy. Nat. Rev. Cancer 2018, 18, 89–102. [Google Scholar] [CrossRef]
  325. Soussi, T.; Wiman, K.G. Shaping genetic alterations in human cancer: The p53 mutation paradigm. Cancer Cell 2007, 12, 303–312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  326. Kim, J.; Yu, L.; Chen, W.; Xu, Y.; Wu, M.; Todorova, D.; Tang, Q.; Feng, B.; Jiang, L.; He, J.; et al. Wild-Type p53 Promotes Cancer Metabolic Switch by Inducing PUMA-Dependent Suppression of Oxidative Phosphorylation. Cancer Cell 2019, 35, 191–203.e8. [Google Scholar] [CrossRef] [Green Version]
  327. Goronzy, J.J.; Fang, F.; Cavanagh, M.M.; Qi, Q.; Weyand, C.M. Naive T cell maintenance and function in human aging. J. Immunol. 2015, 194, 4073–4080. [Google Scholar] [CrossRef] [Green Version]
  328. Mendoza, A.; Fang, V.; Chen, C.; Serasinghe, M.; Verma, A.; Muller, J.; Chaluvadi, V.S.; Dustin, M.L.; Hla, T.; Elemento, O.; et al. Lymphatic endothelial S1P promotes mitochondrial function and survival in naive T cells. Nature 2017, 546, 158–161. [Google Scholar] [CrossRef] [Green Version]
  329. Fox, C.J.; Hammerman, P.S.; Thompson, C.B. Fuel feeds function: Energy metabolism and the T-cell response. Nat. Rev. Immunol. 2005, 5, 844–852. [Google Scholar] [CrossRef]
  330. Wofford, J.A.; Wieman, H.L.; Jacobs, S.R.; Zhao, Y.; Rathmell, J.C. IL-7 promotes Glut1 trafficking and glucose uptake via STAT5-mediated activation of Akt to support T-cell survival. Blood 2008, 111, 2101–2111. [Google Scholar] [CrossRef] [Green Version]
  331. Wang, R.; Dillon, C.P.; Shi, L.Z.; Milasta, S.; Carter, R.; Finkelstein, D.; McCormick, L.L.; Fitzgerald, P.; Chi, H.; Munger, J.; et al. The transcription factor Myc controls metabolic reprogramming upon T lymphocyte activation. Immunity 2011, 35, 871–882. [Google Scholar] [CrossRef] [Green Version]
  332. Sinclair, L.V.; Rolf, J.; Emslie, E.; Shi, Y.B.; Taylor, P.M.; Cantrell, D.A. Control of amino-acid transport by antigen receptors coordinates the metabolic reprogramming essential for T cell differentiation. Nat. Immunol. 2013, 14, 500–508. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  333. Heinzel, S.; Binh Giang, T.; Kan, A.; Marchingo, J.M.; Lye, B.K.; Corcoran, L.M.; Hodgkin, P.D. A Myc-dependent division timer complements a cell-death timer to regulate T cell and B cell responses. Nat. Immunol. 2017, 18, 96–103. [Google Scholar] [CrossRef] [PubMed]
  334. Chang, C.H.; Curtis, J.D.; Maggi, L.B., Jr.; Faubert, B.; Villarino, A.V.; O’Sullivan, D.; Huang, S.C.; van der Windt, G.J.; Blagih, J.; Qiu, J.; et al. Posttranscriptional control of T cell effector function by aerobic glycolysis. Cell 2013, 153, 1239–1251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  335. Tarasenko, T.N.; Pacheco, S.E.; Koenig, M.K.; Gomez-Rodriguez, J.; Kapnick, S.M.; Diaz, F.; Zerfas, P.M.; Barca, E.; Sudderth, J.; DeBerardinis, R.J.; et al. Cytochrome c Oxidase Activity Is a Metabolic Checkpoint that Regulates Cell Fate Decisions During T Cell Activation and Differentiation. Cell Metab. 2017, 25, 1254–1268.e7. [Google Scholar] [CrossRef] [Green Version]
  336. Geltink, R.; Kyle, R.L.; Pearce, E.L. Unraveling the Complex Interplay Between T Cell Metabolism and Function. Annu. Rev. Immunol. 2018, 36, 461–488. [Google Scholar] [CrossRef] [PubMed]
  337. Hu, Z.; Qu, G.; Yu, X.; Jiang, H.; Teng, X.L.; Ding, L.; Hu, Q.; Guo, X.; Zhou, Y.; Wang, F.; et al. Acylglycerol Kinase Maintains Metabolic State and Immune Responses of CD8(+) T Cells. Cell Metab. 2019, 30, 290–302.e5. [Google Scholar] [CrossRef]
  338. Chen, D.S.; Mellman, I. Oncology meets immunology: The cancer-immunity cycle. Immunity 2013, 39, 1–10. [Google Scholar] [CrossRef] [Green Version]
  339. Reinfeld, B.I.; Madden, M.Z.; Wolf, M.M.; Chytil, A.; Bader, J.E.; Patterson, A.R.; Sugiura, A.; Cohen, A.S.; Ali, A.; Do, B.T.; et al. Cell-programmed nutrient partitioning in the tumour microenvironment. Nature 2021, 593, 282–288. [Google Scholar] [CrossRef]
  340. Peng, X.; Chen, Z.; Farshidfar, F.; Xu, X.; Lorenzi, P.L.; Wang, Y.; Cheng, F.; Tan, L.; Mojumdar, K.; Du, D.; et al. Molecular Characterization and Clinical Relevance of Metabolic Expression Subtypes in Human Cancers. Cell Rep. 2018, 23, 255–269.e4. [Google Scholar] [CrossRef] [Green Version]
  341. Wellen, K.E.; Thompson, C.B. Cellular metabolic stress: Considering how cells respond to nutrient excess. Mol. Cell 2010, 40, 323–332. [Google Scholar] [CrossRef] [Green Version]
  342. Wang, T.; Liu, L.; Chen, X.; Shen, Y.; Lian, G.; Shah, N.; Davidoff, A.M.; Yang, J.; Wang, R. MYCN drives glutaminolysis in neuroblastoma and confers sensitivity to an ROS augmenting agent. Cell Death Dis. 2018, 9, 220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  343. Lee, J.V.; Carrer, A.; Shah, S.; Snyder, N.W.; Wei, S.; Venneti, S.; Worth, A.J.; Yuan, Z.F.; Lim, H.W.; Liu, S.; et al. Akt-dependent metabolic reprogramming regulates tumor cell histone acetylation. Cell Metab. 2014, 20, 306–319. [Google Scholar] [CrossRef] [Green Version]
  344. Weinberg, F.; Hamanaka, R.; Wheaton, W.W.; Weinberg, S.; Joseph, J.; Lopez, M.; Kalyanaraman, B.; Mutlu, G.M.; Budinger, G.R.; Chandel, N.S. Mitochondrial metabolism and ROS generation are essential for Kras-mediated tumorigenicity. Proc. Natl. Acad. Sci. USA 2010, 107, 8788–8793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  345. Ogrunc, M.; Di Micco, R.; Liontos, M.; Bombardelli, L.; Mione, M.; Fumagalli, M.; Gorgoulis, V.G.; d’Adda di Fagagna, F. Oncogene-induced reactive oxygen species fuel hyperproliferation and DNA damage response activation. Cell Death Differ. 2014, 21, 998–1012. [Google Scholar] [CrossRef] [Green Version]
  346. Bailey, M.H.; Tokheim, C.; Porta-Pardo, E.; Sengupta, S.; Bertrand, D.; Weerasinghe, A.; Colaprico, A.; Wendl, M.C.; Kim, J.; Reardon, B.; et al. Comprehensive Characterization of Cancer Driver Genes and Mutations. Cell 2018, 173, 371–385.e18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  347. Vafa, O.; Wade, M.; Kern, S.; Beeche, M.; Pandita, T.K.; Hampton, G.M.; Wahl, G.M. c-Myc can induce DNA damage, increase reactive oxygen species, and mitigate p53 function: A mechanism for oncogene-induced genetic instability. Mol. Cell 2002, 9, 1031–1044. [Google Scholar] [CrossRef]
  348. Somarribas, P.L.; Vardhana, S.A. Metabolic regulation of the cancer-immunity cycle. Trends Immunol. 2021, 42, 975–993. [Google Scholar] [CrossRef]
  349. Tigano, M.; Vargas, D.C.; Tremblay-Belzile, S.; Fu, Y.; Sfeir, A. Nuclear sensing of breaks in mitochondrial DNA enhances immune surveillance. Nature 2021, 591, 477–481. [Google Scholar] [CrossRef]
  350. Zhang, W.; Wang, G.; Xu, Z.G.; Tu, H.; Hu, F.; Dai, J.; Chang, Y.; Chen, Y.; Lu, Y.; Zeng, H.; et al. Lactate Is a Natural Suppressor of RLR Signaling by Targeting MAVS. Cell 2019, 178, 176–189.e15. [Google Scholar] [CrossRef]
  351. Caronni, N.; Simoncello, F.; Stafetta, F.; Guarnaccia, C.; Ruiz-Moreno, J.S.; Opitz, B.; Galli, T.; Proux-Gillardeaux, V.; Benvenuti, F. Downregulation of Membrane Trafficking Proteins and Lactate Conditioning Determine Loss of Dendritic Cell Function in Lung Cancer. Cancer Res. 2018, 78, 1685–1699. [Google Scholar] [CrossRef] [Green Version]
  352. Nasi, A.; Fekete, T.; Krishnamurthy, A.; Snowden, S.; Rajnavölgyi, E.; Catrina, A.I.; Wheelock, C.E.; Vivar, N.; Rethi, B. Dendritic cell reprogramming by endogenously produced lactic acid. J. Immunol. 2013, 191, 3090–3099. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  353. Goel, S.; DeCristo, M.J.; Watt, A.C.; BrinJones, H.; Sceneay, J.; Li, B.B.; Khan, N.; Ubellacker, J.M.; Xie, S.; Metzger-Filho, O.; et al. CDK4/6 inhibition triggers anti-tumour immunity. Nature 2017, 548, 471–475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Schaer, D.A.; Beckmann, R.P.; Dempsey, J.A.; Huber, L.; Forest, A.; Amaladas, N.; Li, Y.; Wang, Y.C.; Rasmussen, E.R.; Chin, D.; et al. The CDK4/6 Inhibitor Abemaciclib Induces a T Cell Inflamed Tumor Microenvironment and Enhances the Efficacy of PD-L1 Checkpoint Blockade. Cell Rep. 2018, 22, 2978–2994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  355. Fischer, K.; Hoffmann, P.; Voelkl, S.; Meidenbauer, N.; Ammer, J.; Edinger, M.; Gottfried, E.; Schwarz, S.; Rothe, G.; Hoves, S.; et al. Inhibitory effect of tumor cell-derived lactic acid on human T cells. Blood 2007, 109, 3812–3819. [Google Scholar] [CrossRef] [Green Version]
  356. Mendler, A.N.; Hu, B.; Prinz, P.U.; Kreutz, M.; Gottfried, E.; Noessner, E. Tumor lactic acidosis suppresses CTL function by inhibition of p38 and JNK/c-Jun activation. Int. J. Cancer 2012, 131, 633–640. [Google Scholar] [CrossRef]
  357. Angelin, A.; Gil-de-Gómez, L.; Dahiya, S.; Jiao, J.; Guo, L.; Levine, M.H.; Wang, Z.; Quinn, W.J., 3rd; Kopinski, P.K.; Wang, L.; et al. Foxp3 Reprograms T Cell Metabolism to Function in Low-Glucose, High-Lactate Environments. Cell Metab. 2017, 25, 1282–1293.e7. [Google Scholar] [CrossRef] [Green Version]
  358. Wherry, E.J.; Kurachi, M. Molecular and cellular insights into T cell exhaustion. Nat. Rev. Immunol. 2015, 15, 486–499. [Google Scholar] [CrossRef]
  359. Williams, M.A.; Bevan, M.J. Effector and memory CTL differentiation. Annu. Rev. Immunol. 2007, 25, 171–192. [Google Scholar] [CrossRef]
  360. Joyce, J.A.; Fearon, D.T. T cell exclusion, immune privilege, and the tumor microenvironment. Science 2015, 348, 74–80. [Google Scholar] [CrossRef] [Green Version]
  361. Spill, F.; Reynolds, D.S.; Kamm, R.D.; Zaman, M.H. Impact of the physical microenvironment on tumor progression and metastasis. Curr. Opin. Biotechnol. 2016, 40, 41–48. [Google Scholar] [CrossRef] [Green Version]
  362. Qian, B.Z.; Pollard, J.W. Macrophage diversity enhances tumor progression and metastasis. Cell 2010, 141, 39–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  363. Colegio, O.R.; Chu, N.Q.; Szabo, A.L.; Chu, T.; Rhebergen, A.M.; Jairam, V.; Cyrus, N.; Brokowski, C.E.; Eisenbarth, S.C.; Phillips, G.M.; et al. Functional polarization of tumour-associated macrophages by tumour-derived lactic acid. Nature 2014, 513, 559–563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  364. Andrejeva, G.; Rathmell, J.C. Similarities and Distinctions of Cancer and Immune Metabolism in Inflammation and Tumors. Cell Metab. 2017, 26, 49–70. [Google Scholar] [CrossRef] [Green Version]
  365. Ohashi, T.; Akazawa, T.; Aoki, M.; Kuze, B.; Mizuta, K.; Ito, Y.; Inoue, N. Dichloroacetate improves immune dysfunction caused by tumor-secreted lactic acid and increases antitumor immunoreactivity. Int. J. Cancer 2013, 133, 1107–1118. [Google Scholar] [CrossRef] [PubMed]
  366. Kareva, I.; Hahnfeldt, P. The emerging “hallmarks” of metabolic reprogramming and immune evasion: Distinct or linked? Cancer Res. 2013, 73, 2737–2742. [Google Scholar] [CrossRef] [Green Version]
  367. Ruffell, B.; DeNardo, D.G.; Affara, N.I.; Coussens, L.M. Lymphocytes in cancer development: Polarization towards pro-tumor immunity. Cytokine Growth Factor Rev. 2010, 21, 3–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  368. de Visser, K.E.; Eichten, A.; Coussens, L.M. Paradoxical roles of the immune system during cancer development. Nat. Rev. Cancer 2006, 6, 24–37. [Google Scholar] [CrossRef]
  369. Kawaguchi, T.; Mednis, A.; Golde, D.W.; David, J.R.; Remold, H.R. Immune (gamma) interferon produced by a human T-lymphoblast cell line. Nature 1981, 292, 842–844. [Google Scholar]
  370. Zhang, Y.; Yu, G.; Chu, H.; Wang, X.; Xiong, L.; Cai, G.; Liu, R.; Gao, H.; Tao, B.; Li, W.; et al. Macrophage-Associated PGK1 Phosphorylation Promotes Aerobic Glycolysis and Tumorigenesis. Mol. Cell 2018, 71, 201–215.e7. [Google Scholar] [CrossRef] [Green Version]
  371. Morvan, M.G.; Lanier, L.L. NK cells and cancer: You can teach innate cells new tricks. Nat. Rev. Cancer 2016, 16, 7–19. [Google Scholar] [CrossRef]
  372. Assmann, N.; O’Brien, K.L.; Donnelly, R.P.; Dyck, L.; Zaiatz-Bittencourt, V.; Loftus, R.M.; Heinrich, P.; Oefner, P.J.; Lynch, L.; Gardiner, C.M.; et al. Srebp-controlled glucose metabolism is essential for NK cell functional responses. Nat. Immunol. 2017, 18, 1197–1206. [Google Scholar] [CrossRef]
  373. Isaacson, B.; Mandelboim, O. Sweet Killers: NK Cells Need Glycolysis to Kill Tumors. Cell Metab. 2018, 28, 183–184. [Google Scholar] [CrossRef] [PubMed]
  374. Cong, J.; Wang, X.; Zheng, X.; Wang, D.; Fu, B.; Sun, R.; Tian, Z.; Wei, H. Dysfunction of Natural Killer Cells by FBP1-Induced Inhibition of Glycolysis during Lung Cancer Progression. Cell Metab. 2018, 28, 243–255.e5. [Google Scholar] [CrossRef] [Green Version]
  375. Flavell, R.A.; Sanjabi, S.; Wrzesinski, S.H.; Licona-Limón, P. The polarization of immune cells in the tumour environment by TGFbeta. Nat. Rev. Immunol. 2010, 10, 554–567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  376. Cadamuro, M.; Nardo, G.; Indraccolo, S.; Dall’olmo, L.; Sambado, L.; Moserle, L.; Franceschet, I.; Colledan, M.; Massani, M.; Stecca, T.; et al. Platelet-derived growth factor-D and Rho GTPases regulate recruitment of cancer-associated fibroblasts in cholangiocarcinoma. Hepatology 2013, 58, 1042–1053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  377. Pavlides, S.; Whitaker-Menezes, D.; Castello-Cros, R.; Flomenberg, N.; Witkiewicz, A.K.; Frank, P.G.; Casimiro, M.C.; Wang, C.; Fortina, P.; Addya, S.; et al. The reverse Warburg effect: Aerobic glycolysis in cancer associated fibroblasts and the tumor stroma. Cell Cycle 2009, 8, 3984–4001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  378. Jena, B.C.; Das, C.K.; Banerjee, I.; Bharadwaj, D.; Majumder, R.; Das, S.; Biswas, A.; Kundu, M.; Roy, P.K.; Kundu, C.N.; et al. TGF-β1 induced autophagy in cancer associated fibroblasts during hypoxia contributes EMT and glycolysis via MCT4 upregulation. Exp. Cell Res. 2022, 417, 113195. [Google Scholar] [CrossRef]
  379. Wagner, E.F. Cancer: Fibroblasts for all seasons. Nature 2016, 530, 42–43. [Google Scholar] [CrossRef] [Green Version]
  380. Pallangyo, C.K.; Ziegler, P.K.; Greten, F.R. IKKbeta acts as a tumor suppressor in cancer-associated fibroblasts during intestinal tumorigenesis. J. Exp. Med. 2015, 212, 2253–2266. [Google Scholar] [CrossRef] [Green Version]
  381. Utley, A.; Chavel, C.; Lightman, S.; Holling, G.A.; Cooper, J.; Peng, P.; Liu, W.; Barwick, B.G.; Gavile, C.M.; Maguire, O.; et al. CD28 Regulates Metabolic Fitness for Long-Lived Plasma Cell Survival. Cell Rep. 2020, 31, 107815. [Google Scholar] [CrossRef]
  382. Gottfried, E.; Kunz-Schughart, L.A.; Ebner, S.; Mueller-Klieser, W.; Hoves, S.; Andreesen, R.; Mackensen, A.; Kreutz, M. Tumor-derived lactic acid modulates dendritic cell activation and antigen expression. Blood 2006, 107, 2013–2021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  383. Sun, H.; Zhou, Y.; Skaro, M.F.; Wu, Y.; Qu, Z.; Mao, F.; Zhao, S.; Xu, Y. Metabolic Reprogramming in Cancer Is Induced to Increase Proton Production. Cancer Res. 2020, 80, 1143–1155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  384. Persi, E.; Duran-Frigola, M.; Damaghi, M.; Roush, W.R.; Aloy, P.; Cleveland, J.L.; Gillies, R.J.; Ruppin, E. Systems analysis of intracellular pH vulnerabilities for cancer therapy. Nat. Commun. 2018, 9, 2997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  385. Cantelmo, A.R.; Pircher, A.; Kalucka, J.; Carmeliet, P. Vessel pruning or healing: Endothelial metabolism as a novel target? Expert Opin. Ther. Targets 2017, 21, 239–247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  386. Goetze, K.; Walenta, S.; Ksiazkiewicz, M.; Kunz-Schughart, L.A.; Mueller-Klieser, W. Lactate enhances motility of tumor cells and inhibits monocyte migration and cytokine release. Int. J. Oncol. 2011, 39, 453–463. [Google Scholar] [CrossRef] [Green Version]
  387. Ribatti, D.; Nico, B.; Crivellato, E.; Roccaro, A.M.; Vacca, A. The history of the angiogenic switch concept. Leukemia 2007, 21, 44–52. [Google Scholar] [CrossRef] [Green Version]
  388. Chappell, J.C.; Wiley, D.M.; Bautch, V.L. Regulation of blood vessel sprouting. Semin. Cell Dev. Biol. 2011, 22, 1005–1011. [Google Scholar] [CrossRef] [Green Version]
  389. Carmeliet, P. Angiogenesis in life, disease and medicine. Nature 2005, 438, 932–936. [Google Scholar] [CrossRef]
  390. Lu, H.; Dalgard, C.L.; Mohyeldin, A.; McFate, T.; Tait, A.S.; Verma, A. Reversible inactivation of HIF-1 prolyl hydroxylases allows cell metabolism to control basal HIF-1. J. Biol. Chem. 2005, 280, 41928–41939. [Google Scholar] [CrossRef] [Green Version]
  391. Sonveaux, P.; Copetti, T.; De Saedeleer, C.J.; Végran, F.; Verrax, J.; Kennedy, K.M.; Moon, E.J.; Dhup, S.; Danhier, P.; Frérart, F.; et al. Targeting the lactate transporter MCT1 in endothelial cells inhibits lactate-induced HIF-1 activation and tumor angiogenesis. PLoS ONE 2012, 7, e33418. [Google Scholar] [CrossRef] [Green Version]
  392. Yang, Z.; Huang, Y.; Zhu, L.; Yang, K.; Liang, K.; Tan, J.; Yu, B. SIRT6 promotes angiogenesis and hemorrhage of carotid plaque via regulating HIF-1alpha and reactive oxygen species. Cell Death Dis. 2021, 12, 77. [Google Scholar] [CrossRef]
  393. Huang, Y.H.; Kuo, C.H.; Peng, I.C.; Chang, Y.S.; Tseng, S.H.; Conway, E.M.; Wu, H.L. Recombinant thrombomodulin domain 1 rescues pathological angiogenesis by inhibition of HIF-1alpha-VEGF pathway. Cell Mol. Life Sci. 2021, 78, 7681–7692. [Google Scholar] [CrossRef]
  394. Kim, J.W.; Tchernyshyov, I.; Semenza, G.L.; Dang, C.V. HIF-1-mediated expression of pyruvate dehydrogenase kinase: A metabolic switch required for cellular adaptation to hypoxia. Cell Metab. 2006, 3, 177–185. [Google Scholar] [CrossRef] [Green Version]
  395. Papandreou, I.; Cairns, R.A.; Fontana, L.; Lim, A.L.; Denko, N.C. HIF-1 mediates adaptation to hypoxia by actively downregulating mitochondrial oxygen consumption. Cell Metab. 2006, 3, 187–197. [Google Scholar] [CrossRef] [Green Version]
  396. Iyer, N.V.; Kotch, L.E.; Agani, F.; Leung, S.W.; Laughner, E.; Wenger, R.H.; Gassmann, M.; Gearhart, J.D.; Lawler, A.M.; Yu, A.Y.; et al. Cellular and developmental control of O2 homeostasis by hypoxia-inducible factor 1 alpha. Genes Dev. 1998, 12, 149–162. [Google Scholar] [CrossRef] [Green Version]
  397. Semenza, G.L.; Jiang, B.H.; Leung, S.W.; Passantino, R.; Concordet, J.P.; Maire, P.; Giallongo, A. Hypoxia response elements in the aldolase A, enolase 1, and lactate dehydrogenase A gene promoters contain essential binding sites for hypoxia-inducible factor 1. J. Biol. Chem. 1996, 271, 32529–32537. [Google Scholar] [CrossRef] [Green Version]
  398. Zhang, H.; Bosch-Marce, M.; Shimoda, L.A.; Tan, Y.S.; Baek, J.H.; Wesley, J.B.; Gonzalez, F.J.; Semenza, G.L. Mitochondrial autophagy is an HIF-1-dependent adaptive metabolic response to hypoxia. J. Biol. Chem. 2008, 283, 10892–10903. [Google Scholar] [CrossRef] [Green Version]
  399. Bellot, G.; Garcia-Medina, R.; Gounon, P.; Chiche, J.; Roux, D.; Pouysségur, J.; Mazure, N.M. Hypoxia-induced autophagy is mediated through hypoxia-inducible factor induction of BNIP3 and BNIP3L via their BH3 domains. Mol. Cell Biol. 2009, 29, 570–581. [Google Scholar] [CrossRef] [Green Version]
  400. Phng, L.K.; Gerhardt, G. Angiogenesis: A team effort coordinated by notch. Dev. Cell 2009, 16, 196–208. [Google Scholar] [CrossRef]
  401. Melincovici, C.S.; Boşca, A.B.; Şuşman, S.; Mărginean, M.; Mihu, C.; Istrate, M.; Moldovan, I.M.; Roman, A.L.; Mihu, C.M. Vascular endothelial growth factor (VEGF)–Key factor in normal and pathological angiogenesis. Rom. J. Morphol. Embryol. 2018, 59, 455–467. [Google Scholar]
  402. Zhou, H.; Binmadi, N.O.; Yang, Y.H.; Proia, P.; Basile, J.R. Retraction Note to: Semaphorin 4D cooperates with VEGF to promote angiogenesis and tumor progression. Angiogenesis 2020, 23, 267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  403. De Bock, K.; Georgiadou, M.; Schoors, S.; Kuchnio, A.; Wong, B.W.; Cantelmo, A.R.; Quaegebeur, A.; Ghesquière, B.; Cauwenberghs, S.; Eelen, G.; et al. Role of PFKFB3-driven glycolysis in vessel sprouting. Cell 2013, 154, 651–663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  404. Gavard, J.; Gutkind, J.S. VEGF controls endothelial-cell permeability by promoting the beta-arrestin-dependent endocytosis of VE-cadherin. Nat. Cell Biol. 2006, 8, 1223–12234. [Google Scholar] [CrossRef] [PubMed]
  405. Cruys, B.; Wong, B.W.; Kuchnio, A.; Verdegem, D.; Cantelmo, A.R.; Conradi, L.C.; Vandekeere, S.; Bouché, A.; Cornelissen, I.; Vinckier, S.; et al. Glycolytic regulation of cell rearrangement in angiogenesis. Nat. Commun. 2016, 7, 2240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  406. Lehuédé, C.; Dupuy, F.; Rabinovitch, R.; Jones, R.G.; Siegel, P.M. Metabolic Plasticity as a Determinant of Tumor Growth and Metastasis. Cancer Res. 2016, 76, 5201–5528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  407. Jin, X.; Mu, P. Targeting Breast Cancer Metastasis. Breast Cancer 2015, 9 (Suppl. S1), 23–34. [Google Scholar] [CrossRef] [Green Version]
  408. Valastyan, S.; Weinberg, R.A. Tumor metastasis: Molecular insights and evolving paradigms. Cell 2011, 147, 275–292. [Google Scholar] [CrossRef] [Green Version]
  409. Kaushik, D.K.; Bhattacharya, A.; Mirzaei, R.; Rawji, K.S.; Ahn, Y.; Rho, J.M.; Yong, V.W. Enhanced glycolytic metabolism supports transmigration of brain-infiltrating macrophages in multiple sclerosis. J. Clin. Investig. 2019, 129, 3277–3292. [Google Scholar] [CrossRef] [Green Version]
  410. Kishore, M.; Cheung, K.C.P.; Fu, H.; Bonacina, F.; Wang, G.; Coe, D.; Ward, E.J.; Colamatteo, A.; Jangani, M.; Baragetti, A.; et al. Regulatory T Cell Migration Is Dependent on Glucokinase-Mediated Glycolysis. Immunity 2018, 48, 831–832. [Google Scholar] [CrossRef]
  411. Bhattacharya, D.; Azambuja, A.P.; Simoes-Costa, M. Metabolic Reprogramming Promotes Neural Crest Migration via Yap/Tead Signaling. Dev. Cell 2020, 53, 199–211.e6. [Google Scholar] [CrossRef]
  412. Yang, J.; Ren, B.; Yang, G.; Wang, H.; Chen, G.; You, L.; Zhang, T.; Zhao, Y. A mitochondrial switch promotes tumor metastasis. Cell Rep. 2014, 8, 754–766. [Google Scholar]
  413. Hsu, C.C.; Tseng, L.M.; Lee, H.C. Role of mitochondrial dysfunction in cancer progression. Exp. Biol. Med. 2016, 241, 1281–1295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  414. Wang, S.F.; Chen, S.; Tseng, L.M.; Lee, H.C. Role of the mitochondrial stress response in human cancer progression. Exp. Biol. Med. 2020, 245, 861–878. [Google Scholar] [CrossRef]
  415. Meng, F.; Wu, L.; Dong, L.; Mitchell, A.V.; James Block, C.; Liu, J.; Zhang, H.; Lu, Q.; Song, W.M.; Zhang, B.; et al. EGFL9 promotes breast cancer metastasis by inducing cMET activation and metabolic reprogramming. Nat. Commun. 2019, 10, 5033. [Google Scholar] [CrossRef] [Green Version]
  416. Payen, V.L.; Porporato, P.E.; Baselet, B.; Sonveaux, P. Metabolic changes associated with tumor metastasis, part 1: Tumor pH, glycolysis and the pentose phosphate pathway. Cell Mol. Life Sci. 2016, 73, 1333–1348. [Google Scholar] [CrossRef] [PubMed]
  417. Dupuy, F.; Tabariès, S.; Andrzejewski, S.; Dong, Z.; Blagih, J.; Annis, M.G.; Omeroglu, A.; Gao, D.; Leung, S.; Amir, E.; et al. PDK1-Dependent Metabolic Reprogramming Dictates Metastatic Potential in Breast Cancer. Cell Metab. 2015, 22, 577–589. [Google Scholar] [CrossRef] [Green Version]
  418. Le Grand, M.; Mukha, A.; Püschel, J.; Valli, E.; Kamili, A.; Vittorio, O.; Dubrovska, A.; Kavallaris, M. Interplay between MycN and c-Myc regulates radioresistance and cancer stem cell phenotype in neuroblastoma upon glutamine deprivation. Theranostics 2020, 10, 6411–6429. [Google Scholar] [CrossRef] [PubMed]
  419. Li, B.; Cao, Y.; Meng, G.; Qian, L.; Xu, T.; Yan, C.; Luo, O.; Wang, S.; Wei, J.; Ding, Y.; et al. Targeting glutaminase 1 attenuates stemness properties in hepatocellular carcinoma by increasing reactive oxygen species and suppressing Wnt/beta-catenin pathway. EBioMedicine 2019, 39, 239–254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  420. Kawan, M.A.; Kyrou, I.; Ramanjaneya, M.; Williams, K.; Jeyaneethi, J.; Randeva, H.S.; Karteris, E. Involvement of the glutamine RFamide peptide and its cognate receptor GPR103 in prostate cancer. Oncol. Rep. 2019, 41, 1140–1150. [Google Scholar]
  421. Mestre-Farrera, A.; Bruch-Oms, M.; Peña, R.; Rodríguez-Morató, J.; Alba-Castellón, L.; Comerma, L.; Quintela-Fandino, M.; Duñach, M.; Baulida, J.; Pozo, Ó.J.; et al. Glutamine-Directed Migration of Cancer-Activated Fibroblasts Facilitates Epithelial Tumor Invasion. Cancer Res. 2021, 81, 438–451. [Google Scholar] [CrossRef]
  422. Kodama, M.; Oshikawa, K.; Shimizu, H.; Yoshioka, S.; Takahashi, M.; Izumi, Y.; Bamba, T.; Tateishi, C.; Tomonaga, T.; Matsumoto, M.; et al. A shift in glutamine nitrogen metabolism contributes to the malignant progression of cancer. Nat. Commun. 2020, 11, 1320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  423. Endo, H.; Owada, S.; Inagaki, Y.; Shida, Y.; Tatemichi, M. Glucose starvation induces LKB1-AMPK-mediated MMP-9 expression in cancer cells. Sci. Rep. 2018, 8, 10122. [Google Scholar] [CrossRef] [Green Version]
  424. Xiang, L.; Mou, J.; Shao, B.; Wei, Y.; Liang, H.; Takano, N.; Semenza, G.L.; Xie, G. Glutaminase 1 expression in colorectal cancer cells is induced by hypoxia and required for tumor growth, invasion, and metastatic colonization. Cell Death Dis. 2019, 10, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  425. Lee, S.Y.; Jeon, H.M.; Ju, M.K.; Jeong, E.K.; Kim, C.H.; Park, H.G.; Han, S.I.; Kang, H.S. Dlx-2 and glutaminase upregulate epithelial-mesenchymal transition and glycolytic switch. Oncotarget 2016, 7, 7925–7939. [Google Scholar] [CrossRef] [Green Version]
  426. Gaude, E.; Frezza, C. Tissue-specific and convergent metabolic transformation of cancer correlates with metastatic potential and patient survival. Nat. Commun. 2016, 7, 13041. [Google Scholar] [CrossRef] [Green Version]
  427. Zhu, S.; Wei, L.; Lin, G.; Tong, Y.; Zhang, J.; Jiang, X.; He, Q.; Lu, X.; Zhu, D.D.; Chen, Y.Q. Metabolic Alterations Induced by Kudingcha Lead to Cancer Cell Apoptosis and Metastasis Inhibition. Nutr. Cancer 2020, 72, 696–707. [Google Scholar] [CrossRef]
  428. Liu, H.; Takagaki, Y.; Kumagai, A.; Kanasaki, K.; Koya, D. The PKM2 activator TEPP-46 suppresses kidney fibrosis via inhibition of the EMT program and aberrant glycolysis associated with suppression of HIF-1alpha accumulation. J. Diabetes Investig. 2021, 12, 697–709. [Google Scholar] [CrossRef] [PubMed]
  429. Draoui, N.; de Zeeuw, P.; Carmeliet, P. Angiogenesis revisited from a metabolic perspective: Role and therapeutic implications of endothelial cell metabolism. Open Biol. 2017, 7, 170219. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Influence of metabolic plasticity regulation on tumour initiation. Compared with other periods, stem cells at the resting stage (resting stem cells) are more susceptible to external stimuli during division and differentiation, resulting in increased gene instability, tumorigenesis initiation, and appearance of TICs. At the same time, mitochondrial OXPHOS may exist in the early stage, that is, for the immortalization and further transformation of TIC by promoting aspartate synthesis and subsequent nucleotide metabolism and other biochemical reactions. Glycolysis has essential roles during tumour initiation and affects the entire TIC stage, which is necessary for tumour cell initiation. These processes generate at least two types of metabolic heterogeneity; the metabolic plasticity originates from a single tumour cell and is caused by the heterogeneity of the tumour cell population, which helps tumour cells cope with complex internal and external microenvironments.
Figure 1. Influence of metabolic plasticity regulation on tumour initiation. Compared with other periods, stem cells at the resting stage (resting stem cells) are more susceptible to external stimuli during division and differentiation, resulting in increased gene instability, tumorigenesis initiation, and appearance of TICs. At the same time, mitochondrial OXPHOS may exist in the early stage, that is, for the immortalization and further transformation of TIC by promoting aspartate synthesis and subsequent nucleotide metabolism and other biochemical reactions. Glycolysis has essential roles during tumour initiation and affects the entire TIC stage, which is necessary for tumour cell initiation. These processes generate at least two types of metabolic heterogeneity; the metabolic plasticity originates from a single tumour cell and is caused by the heterogeneity of the tumour cell population, which helps tumour cells cope with complex internal and external microenvironments.
Ijms 24 07076 g001
Figure 2. Tumour cells with glycolytic phenotype have high mitochondrial oxidative metabolism and can switch to OXPHOS. Glycolytic tumour cells absorb glucose (C13 labelled) and produce pyruvate. Pyruvate can be catalysed by PC and PDH for producing carbon-derived metabolites (such as citrate and oxaloacetate). Tumour cells that perform this process are “hypofunctive on dehydrogenase”. The TCA cycle in this process can be used for biosynthesis (such as synthesizing fatty acid, pyrimidines, and pyruvate). Switching from glycolysis to OXPHOS can cause a reverse Warburg Effect and change the phenotype of tumour metabolism. The hyperfunction successful dehydrogenase activated by OXPHOS function can generate more oxidative stress, causing cell death, which is not beneficial for the long-term survival of tumour cells. Tumour cells can then counter the endogenous damage caused by increasing the anti-apoptotic proteins such as B-cell lymphoma-2 (BCL-2), nuclear factor erythroid 2-related factor 2(Nrf-2), and 5′-AMP-activated protein kinase (AMPK). Tumour cells that perform this process are “hyperfunctive on dehydrogenase”.
Figure 2. Tumour cells with glycolytic phenotype have high mitochondrial oxidative metabolism and can switch to OXPHOS. Glycolytic tumour cells absorb glucose (C13 labelled) and produce pyruvate. Pyruvate can be catalysed by PC and PDH for producing carbon-derived metabolites (such as citrate and oxaloacetate). Tumour cells that perform this process are “hypofunctive on dehydrogenase”. The TCA cycle in this process can be used for biosynthesis (such as synthesizing fatty acid, pyrimidines, and pyruvate). Switching from glycolysis to OXPHOS can cause a reverse Warburg Effect and change the phenotype of tumour metabolism. The hyperfunction successful dehydrogenase activated by OXPHOS function can generate more oxidative stress, causing cell death, which is not beneficial for the long-term survival of tumour cells. Tumour cells can then counter the endogenous damage caused by increasing the anti-apoptotic proteins such as B-cell lymphoma-2 (BCL-2), nuclear factor erythroid 2-related factor 2(Nrf-2), and 5′-AMP-activated protein kinase (AMPK). Tumour cells that perform this process are “hyperfunctive on dehydrogenase”.
Ijms 24 07076 g002
Figure 3. Partial metabolic plasticity regulation in tumour cells. Many molecules can cause tumour cells to switch between glycolysis and OXPHOS phenotype. (A) For tumour cells with glycolysis metabolic phenotype, many factors can directly or indirectly regulate the glycolysis process, such as HIF-1α, which can regulate HK-2, PDK1, GLUT1; HSP60, which can strengthen glycolysis reaction, reduce OXPHOS, and reduce the transition from glycolysis to OXPHOS; KRAS, which can regulate the critical enzymes of HK, PFK1, LDHA, and GLUT1, thus affecting the glycolysis process; inactivated PKM2 (Dimer PKM2), which manifest that metabolites accumulate in tumour cells mainly by glycolysis; and FBP2, which can bind to and activate PKM2, making cells for executing glycolysis reactions. In addition, by affecting the mitochondria-related molecules, which can also shift the metabolism to glycolysis phenotype, such as FBP2, it can present the inhibitory effect on mitochondrial transcription factor NRFs and TFAM. Some PDK isoforms can restrict pyruvate entry into mitochondria, and the mutation of SDH can promote the expression of multiple glycolytic enzyme genes. High glycolysis flux levels can help basic metabolism, lipid metabolism, biosynthesis, and denovo synthesis of nucleic acid. In addition, glycolysis flux does not produce oxidative stress molecules to harm tumour cells. Therefore, the glycolysis metabolic phenotype is beneficial for the long-term survival of tumour cells. (B) Tumour cells with OXPHOS metabolic phenotype also have many factors that can directly or indirectly regulate the OXPHOS process. For example, activating PKM2 (Tetrameric PKM2) can catalyze phosphoenolpyruvate (PEP) to pyruvate and promote its entry into the TCA cycle. An overexpression of PTEN-induced kinase 1 (PINK1) can promote mitophagy, reduce glycolysis, and increase OXPHOS. Autophagy protein P62 can recognize polyubiquitination of HK-2, catalysed by E3 ligase TRAF6, and then initiate the degradation of HK-2, thereby inhibiting the glycolysis levels. In addition, some effector molecules can directly promote OXPHOS, such as by directly regulating PDSS2, increasing the activity of CoQ10 and mitochondrial Complex I. H2Bub1 regulates a series of mitochondrial genes that can be used for OXPHOS and inhibit the glycolysis level. AMPK is a core molecule regulating mitochondrial energy metabolism, which can enhance mitochondrial metabolism in multiple ways. TFAM plays a critical role in mitochondrial activity and gene transcription. High levels of OXPHOS can help with proliferation and tumour foci formation. However, OXPHOS is not good for the long-term survival of tumour cells because excessive OXPHOS may lead to more oxidative stress molecules that can damage tumour cells and cause the death of tumour cells.
Figure 3. Partial metabolic plasticity regulation in tumour cells. Many molecules can cause tumour cells to switch between glycolysis and OXPHOS phenotype. (A) For tumour cells with glycolysis metabolic phenotype, many factors can directly or indirectly regulate the glycolysis process, such as HIF-1α, which can regulate HK-2, PDK1, GLUT1; HSP60, which can strengthen glycolysis reaction, reduce OXPHOS, and reduce the transition from glycolysis to OXPHOS; KRAS, which can regulate the critical enzymes of HK, PFK1, LDHA, and GLUT1, thus affecting the glycolysis process; inactivated PKM2 (Dimer PKM2), which manifest that metabolites accumulate in tumour cells mainly by glycolysis; and FBP2, which can bind to and activate PKM2, making cells for executing glycolysis reactions. In addition, by affecting the mitochondria-related molecules, which can also shift the metabolism to glycolysis phenotype, such as FBP2, it can present the inhibitory effect on mitochondrial transcription factor NRFs and TFAM. Some PDK isoforms can restrict pyruvate entry into mitochondria, and the mutation of SDH can promote the expression of multiple glycolytic enzyme genes. High glycolysis flux levels can help basic metabolism, lipid metabolism, biosynthesis, and denovo synthesis of nucleic acid. In addition, glycolysis flux does not produce oxidative stress molecules to harm tumour cells. Therefore, the glycolysis metabolic phenotype is beneficial for the long-term survival of tumour cells. (B) Tumour cells with OXPHOS metabolic phenotype also have many factors that can directly or indirectly regulate the OXPHOS process. For example, activating PKM2 (Tetrameric PKM2) can catalyze phosphoenolpyruvate (PEP) to pyruvate and promote its entry into the TCA cycle. An overexpression of PTEN-induced kinase 1 (PINK1) can promote mitophagy, reduce glycolysis, and increase OXPHOS. Autophagy protein P62 can recognize polyubiquitination of HK-2, catalysed by E3 ligase TRAF6, and then initiate the degradation of HK-2, thereby inhibiting the glycolysis levels. In addition, some effector molecules can directly promote OXPHOS, such as by directly regulating PDSS2, increasing the activity of CoQ10 and mitochondrial Complex I. H2Bub1 regulates a series of mitochondrial genes that can be used for OXPHOS and inhibit the glycolysis level. AMPK is a core molecule regulating mitochondrial energy metabolism, which can enhance mitochondrial metabolism in multiple ways. TFAM plays a critical role in mitochondrial activity and gene transcription. High levels of OXPHOS can help with proliferation and tumour foci formation. However, OXPHOS is not good for the long-term survival of tumour cells because excessive OXPHOS may lead to more oxidative stress molecules that can damage tumour cells and cause the death of tumour cells.
Ijms 24 07076 g003
Figure 4. Effects of WTP53 and mutation of MuTP53 on tumour metabolism regulation. P53 can affect the metabolic plasticity process in multiple ways, thus affecting cancer pathogenesis and development. For the positive response to mitochondrial metabolism, WTP53 can promote the transcriptional level of GLS2 to catalyse the conversion of glutamine to glutamate, which is the source material of the TCA cycle. In addition, WTP53 can maintain the expression of apoptosis-induced factor (AIF), while SCO2, AIF, and SCO2 are required for mitochondrial cytochrome c oxidase assembly and maintain the electron transport chain. In contrast, for the negative response to glycolysis, WTP53 can directly regulate the first step of glycolysis by downregulating the transcriptional level of GLUT1/4 or suppressing NF-κB that can activate GLUT3. Moreover, the TIGAR can be encoded by WTP53 before hydrolysing and reducing the expression level of PFK1, which is the rate-limiting enzyme of glycolysis. However, in liver cancer cells, WTP53 can disrupt the binding of MPC to pyruvate by inducing PUMA and inhibit the uptake of pyruvate from mitochondria, thus restraining OXPHOS. These data indicate that P53 may promote tumour progression but prevent cell carcinogenesis. In contrast with WTP53, MuTP53 may have opposite effects on tumour metabolism regulation. For glycolysis, MuTP53 can promote multiple glycolytic enzymes expression through different molecular pathways, such as HK-2, PFK-1, phosphoglucomutase (PGM), and PKM2. In contrast, for the mitochondrial metabolism, MuTP53 can inhibit the activity of α-KGDC and COXIV, then reduce the level of OXPHOS. In short, WTP53 or MuTP53 will have a complete effect on the type of energy metabolism.
Figure 4. Effects of WTP53 and mutation of MuTP53 on tumour metabolism regulation. P53 can affect the metabolic plasticity process in multiple ways, thus affecting cancer pathogenesis and development. For the positive response to mitochondrial metabolism, WTP53 can promote the transcriptional level of GLS2 to catalyse the conversion of glutamine to glutamate, which is the source material of the TCA cycle. In addition, WTP53 can maintain the expression of apoptosis-induced factor (AIF), while SCO2, AIF, and SCO2 are required for mitochondrial cytochrome c oxidase assembly and maintain the electron transport chain. In contrast, for the negative response to glycolysis, WTP53 can directly regulate the first step of glycolysis by downregulating the transcriptional level of GLUT1/4 or suppressing NF-κB that can activate GLUT3. Moreover, the TIGAR can be encoded by WTP53 before hydrolysing and reducing the expression level of PFK1, which is the rate-limiting enzyme of glycolysis. However, in liver cancer cells, WTP53 can disrupt the binding of MPC to pyruvate by inducing PUMA and inhibit the uptake of pyruvate from mitochondria, thus restraining OXPHOS. These data indicate that P53 may promote tumour progression but prevent cell carcinogenesis. In contrast with WTP53, MuTP53 may have opposite effects on tumour metabolism regulation. For glycolysis, MuTP53 can promote multiple glycolytic enzymes expression through different molecular pathways, such as HK-2, PFK-1, phosphoglucomutase (PGM), and PKM2. In contrast, for the mitochondrial metabolism, MuTP53 can inhibit the activity of α-KGDC and COXIV, then reduce the level of OXPHOS. In short, WTP53 or MuTP53 will have a complete effect on the type of energy metabolism.
Ijms 24 07076 g004
Figure 5. Effects of metabolic plasticity regulation on tumour microenvironment. The stages of tumour development are also per stage of adapting and transforming for the microenvironment. (A) At the early stage of tumour development, CAAs produced by the death of tumour cells can be taken up by DCs, and CAAs are transmitted to the naive CD8+ T cells and differentiate into effector CD8+ T cells, then recognize the tumour cells and initiate the death of the tumour. This repetitive process is called CIC. The uptake of glucose, glutamine, and other nutrients can be utilized by glycolysis and OXPHOS. At the same time, OXPHOS can produce ROS, causing more damage to the tumour cells and then inducing DNA mutations, which increase its immunogenicity. Moreover, S1P and IL-7 can influence the differentiation of naive CD8+ T cells by S1P-driving OXPHOS and IL-7-driving glycolysis. Differentiated memory CD8+ T cells have better mitochondrial activity, and upon TCR and CD28 stimulation, their glycolytic metabolism can be promoted by activating the PI3K-mTOR-signaling pathway and differentiating into effector T cells. In addition, IFN-γ secreted by Th1 can activate macrophages to develop into M1-type macrophages but is unstable in an acidic microenvironment. NK cells can kill target cells directly, which is closely related to metabolic changes such as the activation of NK cells needed to switch its metabolism from mitochondrial oxidation to glycolysis. (B) In the progression of tumour development, the microenvironment favourable for the tumour was formed. The transition of OXPHOS to glycolysis or lactic acid can reduce cells’ immunogenicity, thus inhibiting the CIC. Excessive lactic acid (LA) produced by glycolysis is the inhibitory molecule of all anti-tumour immune cells, which can be absorbed by Treg cells, then suppressing effector CD8+ T cells. An acidic microenvironment caused by LA can make IFN-γ unstable. Thus, it inhibits the differentiation of Th1 cells and turns to Th2, which antagonizes the differentiation of M1-type macrophages. Except for tumour cells, CAFs are the main source of lactic acid, which is recruited by platelet-derived growth factor (PDGF) secreted through the tumour cells. Moreover, CAFs are the origin of TGF- β, which can induce FBP1 expression. This weakened the glycolytic level of NK cells, thereby inactivating NK cells. Angiogenesis is an important characteristic of tumour cells, and glycolysis contributes to this process. Lactic acid and hypoxia can promote this process by stabilizing HIF-1, VEGF, bFGF, VEGFR2, etc. The formation of vascular buds is crucial in the process of angiogenesis, which is guided and initiated by tip endothelial cells and stalk cells. Glycolysis can promote endothelial cells to compete for “tip” sites and “stalk” sites, and this process is accompanied by the overexpression of the glycolytic regulatory PFKFB3.
Figure 5. Effects of metabolic plasticity regulation on tumour microenvironment. The stages of tumour development are also per stage of adapting and transforming for the microenvironment. (A) At the early stage of tumour development, CAAs produced by the death of tumour cells can be taken up by DCs, and CAAs are transmitted to the naive CD8+ T cells and differentiate into effector CD8+ T cells, then recognize the tumour cells and initiate the death of the tumour. This repetitive process is called CIC. The uptake of glucose, glutamine, and other nutrients can be utilized by glycolysis and OXPHOS. At the same time, OXPHOS can produce ROS, causing more damage to the tumour cells and then inducing DNA mutations, which increase its immunogenicity. Moreover, S1P and IL-7 can influence the differentiation of naive CD8+ T cells by S1P-driving OXPHOS and IL-7-driving glycolysis. Differentiated memory CD8+ T cells have better mitochondrial activity, and upon TCR and CD28 stimulation, their glycolytic metabolism can be promoted by activating the PI3K-mTOR-signaling pathway and differentiating into effector T cells. In addition, IFN-γ secreted by Th1 can activate macrophages to develop into M1-type macrophages but is unstable in an acidic microenvironment. NK cells can kill target cells directly, which is closely related to metabolic changes such as the activation of NK cells needed to switch its metabolism from mitochondrial oxidation to glycolysis. (B) In the progression of tumour development, the microenvironment favourable for the tumour was formed. The transition of OXPHOS to glycolysis or lactic acid can reduce cells’ immunogenicity, thus inhibiting the CIC. Excessive lactic acid (LA) produced by glycolysis is the inhibitory molecule of all anti-tumour immune cells, which can be absorbed by Treg cells, then suppressing effector CD8+ T cells. An acidic microenvironment caused by LA can make IFN-γ unstable. Thus, it inhibits the differentiation of Th1 cells and turns to Th2, which antagonizes the differentiation of M1-type macrophages. Except for tumour cells, CAFs are the main source of lactic acid, which is recruited by platelet-derived growth factor (PDGF) secreted through the tumour cells. Moreover, CAFs are the origin of TGF- β, which can induce FBP1 expression. This weakened the glycolytic level of NK cells, thereby inactivating NK cells. Angiogenesis is an important characteristic of tumour cells, and glycolysis contributes to this process. Lactic acid and hypoxia can promote this process by stabilizing HIF-1, VEGF, bFGF, VEGFR2, etc. The formation of vascular buds is crucial in the process of angiogenesis, which is guided and initiated by tip endothelial cells and stalk cells. Glycolysis can promote endothelial cells to compete for “tip” sites and “stalk” sites, and this process is accompanied by the overexpression of the glycolytic regulatory PFKFB3.
Ijms 24 07076 g005
Figure 6. Effects of metabolic regulation on tumour metastasis. Glycolysis and Mitochondrial oxidative metabolism jointly contribute to the invasion and migration of tumour cells. Under the condition of sufficient glucose, normoxia, and low LA, glycolytic tumour cells can enable tumour cells to shift to functional mitochondrial oxidation in a manner of activation dependent on AMPK, thus leading to anti-anoikis apoptosis. In addition, glycolytic flux can activate EMT-related molecules SNAI1, SNAI2, TWIST, and ZEB1, and stimulate Yap/YAZ-signaling pathway to drive cell movement. With the starvation of glucose, hypoxia, and increased lactate acid, glycolytic tumour cells can absorb LA and use glutamine for mitochondrial oxidative metabolism. AMPK activated by glycolysis also can enable tumour cells to shift to functional mitochondrial oxidation. The glycolytic tumour cells with glutamine for metabolism showed more aggression, and together with glycolytic tumour cells with glucose for metabolism, both can activate EMT to drive cell metastasis. The dotted arrows represent that there is a multi-step regulatory pathway between the upstream and downstream of molecular events, which is a brief description rather than one-step regulation.
Figure 6. Effects of metabolic regulation on tumour metastasis. Glycolysis and Mitochondrial oxidative metabolism jointly contribute to the invasion and migration of tumour cells. Under the condition of sufficient glucose, normoxia, and low LA, glycolytic tumour cells can enable tumour cells to shift to functional mitochondrial oxidation in a manner of activation dependent on AMPK, thus leading to anti-anoikis apoptosis. In addition, glycolytic flux can activate EMT-related molecules SNAI1, SNAI2, TWIST, and ZEB1, and stimulate Yap/YAZ-signaling pathway to drive cell movement. With the starvation of glucose, hypoxia, and increased lactate acid, glycolytic tumour cells can absorb LA and use glutamine for mitochondrial oxidative metabolism. AMPK activated by glycolysis also can enable tumour cells to shift to functional mitochondrial oxidation. The glycolytic tumour cells with glutamine for metabolism showed more aggression, and together with glycolytic tumour cells with glucose for metabolism, both can activate EMT to drive cell metastasis. The dotted arrows represent that there is a multi-step regulatory pathway between the upstream and downstream of molecular events, which is a brief description rather than one-step regulation.
Ijms 24 07076 g006
Table 1. Drug therapy for key enzymes that can cause metabolic remodeling.
Table 1. Drug therapy for key enzymes that can cause metabolic remodeling.
DrugMedication ConditionsMechanism
CisplatinHK2 shRNADeletion or decrease of HK-2 could produce more ROS, pro-apoptotic molecules, and mitochondrial dysfunction, reduce the drug resistance of tumour cells [154,155,156].
Metformin + sorafenibHK2 deletion
ClotrimazoleVDAC disruptionThe use of PFK inhibitor clotrimazole can reduce glycolysis under the condition of VDAC disruption, prevent the metabolic rearrangement process, and inhibit the growth and invasion of tumour cells [157].
HrDNase1DNASE1L3 low expressionHrDNase1 could increase the level of DNASE1L3, then activate the P53-apoptosis pathway, thus inhibiting tumour growth, invasion, and migration [169].
CetuximabTRAP1 inhibitionThe inhibition of TRAP1 could reduce its binding with PFK1, then trigger the degradation of PFK1, enhancing the sensitivity of CRCs to cetuximab [163].
DASA + metforminnon-sugar environmentDASA inhibits the invasion of tumour cells caused by PKM2 nuclear localisation, and metformin inhibits COXI activity of tumour cells, reduces the growth and promotion of tumour cells [187].
MLN4924 + shikonin + metforminTumour cells with high levels of both OXPHOS and glycolysisMLN4924 inhibits TCA level, shikonin inhibits PKM2 mediated glycolysis, and metformin was used for drug intervention [175,189].
MI-DDMEM GAL mediumCitric acid could decrease the level of anti-apoptotic molecule expression and increase the level of apoptosis, initiating and executing molecule expression. Then, tumour cells become more sensitive to anti-tumour drugs [210,212,218].
Citric acid + 3-bromo-pyruvate, cisplatin, or celecoxib
Doxorubicin + SorafenibInsensitive to drugs for mutant IDH1The risk scores of glycolysis-related genes were correlated with those two drugs and had better sensitivity [268].
Normoxia + radiotherapy and chemotherapytumour cells have the potential of OXPHOS and glycolysisRemove the cells from the hypoxic area, and then use the corresponding radiotherapy and chemotherapy methods, which can produce more ROS [274,279].
BAY 87-2243(B87) + DMKGB87 alone is not very effective in inducing tumour cell deathDMKG can produce a strong tumour-killing effect together with B87 [274].
JHU-083 Inhibition of α-KG [276].
Table 2. Partial molecules that influence Metabolic plasticity process.
Table 2. Partial molecules that influence Metabolic plasticity process.
Metabolic TypeFactors that Affecting Metabolic PlasticityTumour Cell TypesSignificance
From Glycolysis to OXPHOSResting stem cells are more susceptible to external stimuli during division and differentiation, resulting in increased gene instability, tumorigenesis initiation, and the appearance of tumour-initiating cells (TICs)The early stage of TICFor the immortalization and further transformation of TIC by promoting aspartate synthesis and subsequent nucleotide metabolism and other biochemical reactions [83,84,85,86].
Be inclined to GlycolysisThe complete stage of TICAvoid apoptosis, and provide more substrate sources and preparation for adaption to the external environment [58,59].
OXPHOSWith sufficient oxygenGlioma stem cells (CSCs)To adapt to different microenvironments and meet their vigorous growth and proliferation [117,118,119,120].
Be inclined to GlycolysisIn the hypoxic regions
Be inclined to OXPHOSHIF-1low/pAMPKhighLiver
hepatocellular carcinoma (HCC), lung adenocarcinoma (LUAD), breast invasive carcinoma, stomach
adenocarcinoma, acute myeloid leukaemia and pancreatic adenocarcinoma (PAC)
Hybrid phenotype contributes to metabolic plasticity, allowing
tumour cells to adapt to various microenvironments [136].
Be inclined to GlycolysisHIF-1high/pAMPKlow
Hybrid phenotypeHIF-1high/pAMPKhigh
From Glycolysis to OXPHOSDeletion of HK-2Retinal cells, Gastric Cancer, glioblastoma multiformeMore ROS and apoptosis-promoting molecules produce [142,143,144,145].
From Glycolysis to OXPHOSInhibitory of PFK expressionPancreatic ductal adenocarcinoma (PDAC), non-small cell lung cancer (NSCLC) cells, and colorectal cancer (CRCs)Provide energy for the vigorous metabolism of tumour cells [157,158,159,162,163].
Be inclined to OXPHOSDimer PKM2-Entering the nucleus to participate in the transcription and regulation of many molecules, affecting the process of tumour proliferation, migration, angiogenesis, and death [180,181,182].
Be inclined to GlycolysisTetramer PKM2Providing energy for tumour growth and proliferation [176,177].
Be inclined to OXPHOSCitrate synthase (CS)-Increase the burden of mitochondrial, causing apoptosis [208,209,210,211,212,213,214].
Be inclined to OXPHOSMutation of Isocitrate Dehydrogenase (IDH)-Make the (R) enantiomer of 2-hydroxyglutarate (R)-2HG) generate and release abnormal metabolism of lactate and pyruvic acid [248,249,250,251,252,253,254].
Be inclined to OXPHOSHigh expression of FBP-Inhibit HIF-1α, interfere with glycolysis, and hinder tumour cell proliferation, invasion, and migration [307].
Be inclined to GlycolysisFBP2 located in the nucleus with c-Myc-limiting the expression of mitochondrial genes and reducing OXPHOS energy supply [308].
Be inclined to GlycolysisMutations in succinate dehydrogenase (SDH) and fumarate hydratase (FH)Hereditary paragangliomas, leiomyomatosis and kidney cancersReducing the degradation of HIF and promoting the expression of multiple glycolytic enzyme genes [309,310,311].
From OXPHOS to Glycolysisinhibition of HSP60renal cell carcinomaEnhance de novo nucleotide synthesis and lipid synthesis [312].
From Glycolysis to OXPHOSDirectly regulating PDSS2hepatoma cellsInhibit foci formation, colony formation in soft agar, and tumour formation in nude mice [313].
OXPHOSWTP53-Reduce harmful products, such as ROS, and obtain more stable genomic DNA, thus avoiding the initiation of tumorigenesis [318,319,320,321,322,323].
From OXPHOS to GlycolysisMuTP53 Promotion of tumour proliferation and growth [319].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Niu, N.; Ye, J.; Hu, Z.; Zhang, J.; Wang, Y. Regulative Roles of Metabolic Plasticity Caused by Mitochondrial Oxidative Phosphorylation and Glycolysis on the Initiation and Progression of Tumorigenesis. Int. J. Mol. Sci. 2023, 24, 7076. https://doi.org/10.3390/ijms24087076

AMA Style

Niu N, Ye J, Hu Z, Zhang J, Wang Y. Regulative Roles of Metabolic Plasticity Caused by Mitochondrial Oxidative Phosphorylation and Glycolysis on the Initiation and Progression of Tumorigenesis. International Journal of Molecular Sciences. 2023; 24(8):7076. https://doi.org/10.3390/ijms24087076

Chicago/Turabian Style

Niu, Nan, Jinfeng Ye, Zhangli Hu, Junbin Zhang, and Yun Wang. 2023. "Regulative Roles of Metabolic Plasticity Caused by Mitochondrial Oxidative Phosphorylation and Glycolysis on the Initiation and Progression of Tumorigenesis" International Journal of Molecular Sciences 24, no. 8: 7076. https://doi.org/10.3390/ijms24087076

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

Article Metrics

Back to TopTop