Next Article in Journal
Systemic Delivery Strategies for Oncolytic Viruses: Advancing Targeted and Efficient Tumor Therapy
Previous Article in Journal
Novel Roles for Urokinase- and Tissue-Type Plasminogen Activators in the Pathogenesis of Mood Disorders
Previous Article in Special Issue
Distinct Hepatic Metabolic Reprogramming in Acute and Chronic Sleep Deprivation and the Protective Effects of the Chalcone Analogue TAK
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Spatiotemporal Heterogeneity of Tumor Glucose Metabolism Reprogramming: From Single-Cell Mechanisms to Precision Interventions

Cancer Epigenetics Laboratory, Department of Clinical Oncology, State Key Laboratory of Translational Oncology, Sir YK Pao Center for Cancer, The Chinese University of Hong Kong, Shatin, Hong Kong
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(14), 6901; https://doi.org/10.3390/ijms26146901
Submission received: 20 May 2025 / Revised: 7 July 2025 / Accepted: 15 July 2025 / Published: 18 July 2025

Abstract

Glucose metabolism reprogramming as a defining hallmark of cancer has become a pivotal frontier in oncology research. Recent technological advances in single-cell sequencing, spatial omics, and metabolic imaging have transformed the field from static bulk analyses to dynamic investigations of spatiotemporal heterogeneity at a single-cell resolution. This review systematically summarizes the current knowledge on tumor glucose metabolism dynamics, discussing spatial heterogeneity and temporal evolution patterns, metabolic subpopulation interactions revealed by single-cell metabolomics, the glucose metabolism–epigenetics–immunology regulatory axis, and therapeutic strategies targeting metabolic vulnerabilities. Recent technological advances in single-cell sequencing and spatial omics have transformed our understanding of tumor glucose metabolism by providing high-resolution insights into metabolic heterogeneity and regulatory mechanisms, contrasting with classical bulk analyses. Spatiotemporal heterogeneity critically influences therapeutic outcomes by enabling tumor cells to adapt metabolically under selective pressures (e.g., hypoxia, nutrient deprivation), fostering treatment resistance and relapse. Deciphering these dynamics is essential for developing spatiotemporally targeted strategies that address intratumoral diversity and microenvironmental fluctuations. By integrating cutting-edge advances, this review deepens our understanding of tumor metabolic complexity and provides a conceptual framework for developing spatiotemporally precise interventions.

1. Introduction: Paradigm Shift in Tumor Glucose Metabolism Research

From Warburg’s aerobic glycolysis discovery to current single-cell metabolic mapping, cancer metabolism research now reveals spatiotemporal heterogeneity and therapeutic vulnerabilities. The field has progressed from classical enzymology and bulk metabolic analyses to single-cell transcriptomics, enabling a systems-level metabolic–immune–epigenetic network. This progress has facilitated the clinical development of targeted metabolic drugs, with numerous agents now in ongoing oncology trials.
This review summarizes the current advances in tumor glucose metabolism research, presenting three new insights: a single-cell metabolic topology mapping glycolytic network complexity; the glucose metabolism–epigenetics–immune axis as a multilayer regulatory circuit; and temporal targeting for metabolic oscillations. By integrating multi-omics data with clinical insights, we connect mechanistic understanding to therapeutic development, advancing next-generation metabolic interventions.

2. Multidimensional Characteristics of Spatiotemporal Heterogeneity in Tumor Glucose Metabolic Reprogramming

Tumor glucose metabolism reprogramming exhibits spatiotemporal heterogeneity characterized by microenvironmental pressures and cellular signaling. Understanding these regulatory mechanisms enables targeted metabolic therapies. This section analyzes glucose metabolic heterogeneity across tumor evolution and its role in treatment resistance and immune modulation. Figure 1 shows the spatiotemporal heterogeneity of tumor glucose metabolism.
Single-cell transcriptomics provides an unprecedented resolution, revealing intricate metabolic zonation within tumors—far beyond the limitations of bulk measurements in classical studies. By dissecting the cell-to-cell variability, this approach highlights how microenvironmental gradients (e.g., hypoxia, nutrient availability) shape distinct metabolic phenotypes, offering new mechanistic insights into tumor metabolic adaptation.

2.1. Spatial Heterogeneity

2.1.1. Intratumoral Metabolic Zonation

Tumors establish spatially structured metabolic networks shaped by oxygen and nutrient gradients. Oxygen-rich rims utilize oxidative phosphorylation (OXPHOS), along with accumulated tricarboxylic acid (TCA) cycle intermediates (citrate, malate, succinate) and glutaminase 1 (GLS1)-dependent glutamine catabolism, to meet both the biosynthetic needs and invasive capacity [2,3]. Conversely, hypoxic cores exhibit hypoxia-inducible factor 1 alpha (HIF-1α)-dependent metabolic reprogramming by glycolytic dominance (elevated lactate/pyruvate) [2] and the suppression of oxidative metabolism via lactate dehydrogenase A (LDHA)-driven pyruvate diversion [4,5]. In high-grade osteosarcomas, hypoxic cores selectively upregulate hypoxia-inducible factor 2 alpha (HIF-2α) (not HIF-1α) to activate proline dehydrogenase (PRODH)-dependent proline metabolism, maintaining the redox balance under nutrient stress [6]. Pancreatic neuroendocrine tumors (PanNETs) lack metabolic zonation due to dense vascularization, showing homogeneous glycolytic phenotypes throughout via aberrant mammalian target of rapamycin (mTOR)-vascular endothelial growth factor (VEGF) crosstalk [7]. Spatial transcriptomics in glioblastoma revealed a progressive decline in hypoxia-responsive genes (e.g., vascular endothelial growth factor A (VEGFA)) and glycolytic genes (e.g., glucose transporter 1 (GLUT1)) with increasing distance from necrotic regions [8], demonstrating how metabolic adaptation correlates with microenvironmental topography. This compartmentalization creates therapeutic challenges, as regions may exhibit differential sensitivity to metabolic interventions (see Table 1 for cross-tumor comparisons).

2.1.2. Metabolic Symbiosis Networks

Metabolic symbiosis refers to cooperative metabolic interactions between distinct tumor cell populations or between tumor cells and stromal components, where metabolic byproducts or substrates are exchanged to fulfill complementary bioenergetic or biosynthetic demands. Tumors exhibit oxygen-dependent metabolic zonation, creating symbiotic networks between compartments. Hypoxic cells export glycolysis-derived lactate via the low-affinity monocarboxylate transporter 4 (MCT4), while oxygen-rich cells uptake it via the high-affinity monocarboxylate transporter 1 (MCT1) for mitochondrial OXPHOS (the reverse Warburg effect) [9,10]. Beyond its bioenergetic role, this “reverse Warburg effect” creates a pro-tumorigenic niche through lactate-driven VEGF-mediated angiogenesis and acidification. Single-cell studies in esophageal squamous cell carcinoma (ESCC) have uncovered metabolite exchange with cytotoxic T cells and pro-tumor M2 macrophages, but the metabolic exclusion of B cells [11], highlighting immunometabolic crosstalk in tumor progression.
Tumor ecosystems additionally employ stromal metabolic support, where cancer-associated fibroblasts (CAFs) and tumor-associated macrophages provide alternative metabolic substrates (e.g., glutamine, ketone bodies) but competitively consume glucose and arginine to impair immune function [11]. Cancer stem cells (CSCs) display OXPHOS-dominant, low-glycolysis phenotypes to maintain their stem-like properties, further contributing to tumor metabolic heterogeneity. Single-cell metabolomics has identified therapy-resistant fatty acid oxidation (FAO)-dependent clones coexisting with glycolytic proliferative populations in breast and esophageal cancers [11]. These findings highlight metabolic symbiosis as a multifaceted driver of tumor progression and therapeutic resistance.
Table 1. Comparison of spatial metabolic characteristics of different types of tumors.
Table 1. Comparison of spatial metabolic characteristics of different types of tumors.
Tumor TypeMetabolic Characteristics of the Core RegionMetabolic Characteristics of the Marginal ZoneKey Molecules of
Metabolic Interactions
Clinical Significance
Glioblastoma [12]Enhanced glycolysis and hypoxia-induced HIF-1αOXPHOS is active and aggressive.LDHA, MCT4Hypoxic regions are resistant to radiotherapy.
Glioblastoma [13]Significant glycolytic phenotype, high expression of hypoxia-related genes, significant chromosomal copy number variations (CNAs).The metabolic state is more similar to that of normal tissues and more dependent on oxidative phosphorylation. Infiltration of T cells and myeloid cells was observed, but the degree of immunosuppression was low.HIF-1α, VEGFAMultimodal treatment: core zone (glycolysis inhibitor + radiotherapy) combined with marginal zone (OXPHOS inhibitor + immunotherapy).
OSCC [14]Energy is obtained mainly through aerobic glycolysis, which produces large amounts of lactic acid.Immune cells and stromal cells in the TME take up lactate to produce energy.HIF-1α, CXCL12, TGF-βCombined targeting of lactic acid metabolism (such as MCT inhibitors) and immune checkpoint inhibitors may enhance the therapeutic effect.
OSCC [15]Significant glycolytic activity was enhanced, with localized enrichment of glycolytic metabolites.The retention of higher glucose levels may serve as a “reservoir” of glucose to support tumor metabolic requirements.Hexokinase,
pyruvate kinase
Targeting key glycolytic enzymes (such as hexokinase and lactate dehydrogenase) or the inhibition of glycolytic pathways (such as the use of 2-deoxyglucose) may inhibit the tumor energy supply, especially for OSCC subtypes that are resistant to conventional chemoradiotherapy.
Cervical squamous cell carcinoma [16]OXPHOS activity was low and hypoxia-related metabolic pathways dominated.The OXPHOS pathway was significantly activated.PRKCE, ITGA2, PKMA combination strategy is appropriate.
Gastric cancer [17]Glycolytic activity was highest and lactate accumulation was accompanied by the upregulation of tricarboxylic acid cycle intermediates, such as succinate.Some regions may retain OXPHOS capacity, but are affected by tumor–stroma interactions and have complex metabolic phenotypes (e.g., immune cell-infiltrating regions may inhibit glycolysis).PKM2, PFKL, ENO1Early screening and staging based on glycolysis imaging; combined strategies targeting glycolysis (e.g., PFKFB3 inhibitors) with immunometabolic regulation (e.g., GLS inhibitors).
Breast cancer [18]The glucose content is high and more prone to glycolytic metabolism.With a preference for mitochondrial metabolism.PI3K, GLUT1Combined inhibition of PI3K and the bromodomain can overcome drug resistance and reduce metabolic heterogeneity.
Osteosarcoma [6,19]Nucleotide/amino acid pathway amplification (alanine, aspartate, proline); glycolytic dominance; mitochondrial dysfunction.FAO-enhanced invasion; ACLY-mediated acetyl-CoA production.HIF-2α, PRODH, CPT1APRODH inhibitors sensitize to hypoxia-targeted therapy.
Leiomyosarcoma [20,21] Highest glycolytic activity; lactate accumulation.Retained OXPHOS capacity; modulated by tumor–stroma interactions.RAS/PI3K pathway, FBP2High glycolysis correlates with poor prognosis.
PanNETs [7]Homogeneous glycolysis (mTOR-VEGF axis dominance); under hypoxic conditions, invasive PanNET achieved metabolic adaptation by upregulating glycolysis and downregulating oxidative phosphorylation. Lactate shuttling to stromal fibroblasts.mTOR, VEGF, MCT4, HDAC1/2mTOR inhibitors reduce glycolytic flux but increase metastasis risk.
PDAC [22]Enhanced glycolysis, hypoxia-induced HIF-1α stabilization, lactate accumulation, upregulation of glucose transporters (e.g., GLUT1), and key glycolytic enzymes (e.g., HK2, PFK1, LDHA).OXPHOS activity is present but often compromised due to tumor–stroma interactions, lactate uptake via MCT1.HIF-1α, GLUT1, HK2, PFK1, LDHA, MCT1, MCT4, CD147Hypoxic regions are resistant to radiotherapy; metabolic heterogeneity contributes to treatment resistance and relapse; targeting metabolic vulnerabilities (e.g., glycolysis inhibitors, MCT inhibitors) may enhance therapeutic efficacy.
Note: HIF-1α, hypoxia-inducible factor 1 alpha; OXPHOS, oxidative phosphorylation; LDHA, lactate dehydrogenase A; MCT4, monocarboxylate transporter 4; VEGFA, vascular endothelial growth factor A; OSCC, oral squamous cell carcinoma; TME, tumor microenvironment; CXCL12, C-X-C motif chemokine ligand 12; TGF-β, transforming growth factor beta; PRKCE, protein kinase C epsilon; ITGA2, integrin subunit alpha 2; PKM, pyruvate kinase M; PFKL, phosphofructokinase, liver type; ENO1, enolase 1; PFKFB3, 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 3; PI3K, phosphoinositide 3-kinase; GLUT1, glucose transporter 1; FAO, fatty acid oxidation; ACLY, ATP citrate lyase; HIF-2α, hypoxia-inducible factor 2 alpha; PRODH, proline dehydrogenase; CPT1A, carnitine palmitoyltransferase 1A; FBP2, fructose-1,6-bisphosphatase 2; PanNET, pancreatic neuroendocrine tumor; mTOR, mechanistic target of rapamycin (or mammalian target of rapamycin); VEGF, vascular endothelial growth factor; HDAC1/2, histone deacetylase 1/2; PDAC, pancreatic ductal adenocarcinoma; HK2, hexokinase 2; PFK1, phosphofructokinase 1.

2.1.3. Organ-Specific Metabolic Adaptation of Metastatic Lesions

Metastatic tumors exhibit remarkable metabolic plasticity by adapting to organ-specific microenvironmental conditions through distinct reprogramming strategies. Liver metastases exploit high lactate concentrations by upregulating MCT1 and lactate dehydrogenase B (LDHB) to convert lactate into pyruvate for the TCA cycle anaplerosis [23]. Oxygen-rich lung metastases preferentially utilize OXPHOS and glutaminolysis to overcome glucose limitation. Cerebral metastases circumvent blood–brain barrier restrictions by metabolizing astrocyte-derived ketone bodies [24]. Interestingly, organ-specific metabolic adaptations show striking contrasts between pancreatic and colorectal metastases. Pancreatic metastases exhibit mitochondrial dysfunction through mitochondrial DNA (mtDNA) mutations (e.g., Complex I/III defects) that reduce OXPHOS capacity, forcing glycolytic dependence, although a subset of CSCs maintain elevated OXPHOS activity via peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α) to preserve stemness [25,26]. Conversely, colorectal metastases display metabolic compartmentalization: liver lesions develop glycolytic edges (lactate dehydrogenase A (LDHA)/hexokinase 2 (HK2)) for invasion but OXPHOS/glutamine synthetase (GLUL)-dependent cores for hypoxic survival via HIF-1α/pyruvate dehydrogenase kinase 4 (PDK4) [27], while lung metastases upregulate carnitine palmitoyltransferase 1A (CPT1A)-driven FAO to exploit alveolar niches [28]. This comparison underscores how the primary tumor origin dictates distinct metastatic metabolic programs.
Single-cell metabolomics reveals that breast cancer bone metastases activate osteolytic destruction via parathyroid hormone-related protein (PTHrP)-induced transforming growth factor beta (TGF-β)/calcium signaling, concurrently increasing glycolytic demands [8]. Such adaptations are achieved through two mechanisms: microenvironmental selection and exosome-mediated preconditioning, facilitated by metabolic enzymes (e.g., HK2, LDHA) that remodel distant niches prior to colonization [29]. Thus, metastases develop metabolic diversity through spatial patterns like intratumoral zonation and organ adaptation, and temporal changes including therapy resistance and progression.

2.2. Temporal Heterogeneity: Metabolic Dynamic Evolution and Therapeutic Adaptation

Extending our analysis of spatial dynamics, we next investigate the temporal evolution of tumor glucose metabolism across disease progression and therapeutic contexts, focusing on three key processes: metabolic phase transitions, therapy-induced adaptive remodeling, and circadian regulation mechanisms. Figure 2 illustrates the metabolic flux alterations during tumor progression and under therapy.

2.2.1. Metabolic Phase Transitions During Tumor Progression

Tumor metabolism undergoes dynamic temporal evolution during progression. Early tumors adopt aerobic glycolysis (the Warburg effect) to gain selective advantages: rapid adenosine triphosphate (ATP) production for energy demands [30], reduced oxidative stress during hypoxia via nicotinamide adenine dinucleotide (NAD+) recycling [29,31], and microenvironment remodeling that promotes angiogenesis [32]. This metabolic shift promotes tumor growth and suppresses mitochondrial function, facilitating cancer progression. Fatty acids exhibit stage-specific roles, supporting invasion, circulating cell survival, and metastasis. Hypoxia induces glutamine dependence via c-Myc/HIF-2α-mediated upregulation of GLS1 and solute carrier family 1 member 5 (SLC1A5), thereby facilitating mitochondrial oxidation [3,33]. Metastases reactivate mitochondrial metabolism through peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC1α)-driven TCA cycle enhancement and glutamine reprogramming [34]. This metabolic rewiring not only meets progression-specific energy and redox demands but also creates therapeutic vulnerabilities at each progression stage, revealing the remarkable adaptability of cancer cells during dissemination.

2.2.2. Metabolic Adaptive Remodeling Under Therapeutic Pressure

Metabolic adaptive remodeling under therapeutic pressure is a major clinical challenge, as targeted therapies can rapidly reprogram tumor glucose metabolism, driving compensatory resistance mechanisms (Table 2). Cancer cells evade therapy primarily by enhancing glycolysis and rewiring mitochondrial function. In gefitinib-resistant epidermal growth factor receptor (EGFR)-mutant non-small cell lung cancer (NSCLC), lactate is overproduced by the upregulation of GLUT1, HK2, and LDHA [35]. Vemurafenib-resistant v-Raf murine sarcoma viral oncogene homolog B1 (BRAF)-mutant melanomas activate mitogen-activated protein kinase (MEK)/extracellular signal-regulated kinase (ERK) signaling to increase GLUT1/3 and HK2 expression, enhancing glucose uptake and glycolytic flux [36]. Trastuzumab-resistant human epidermal growth factor receptor 2 (HER2+) breast cancers drive glycolysis through 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 3 (PFKFB3)-mediated fructose-2,6-bisphosphate synthesis [37]. Mitochondrial reprogramming represents another resistance strategy. Osimertinib-resistant NSCLC cells enhance OXPHOS activity [38], and BRAF inhibitor-resistant melanomas activate PGC-1α-mediated mitochondrial biogenesis, further increasing the cristae density and elevating oxidative capacity [39].
Metabolic adaptations also drive immunotherapy resistance. Under programmed cell death protein 1 (PD-1) inhibition, tumors select for clones with enhanced OXPHOS that frequently co-express T cell immunoglobulin and mucin domain 3 (TIM-3), linking metabolic reprogramming directly to immune evasion mechanisms [40]. These findings demonstrate that metabolic plasticity serves as a fundamental mechanism of therapeutic resistance across diverse anticancer treatment modalities.

2.2.3. Circadian Regulation of Tumor Metabolism

Circadian metabolic rhythmicity offers new insights into tumor heterogeneity, with the core clock components (Clock Circadian Regulator (CLOCK), brain and muscle ARNT-like 1 (BMAL1/ARNTL), Period Circadian Regulator (PER)) acting as master regulators of oncogenic metabolism. Dysregulation of core clock genes not only correlates with disease progression and poor prognosis across multiple cancer types [41] but also directly initiates tumor formation by rewriting metabolic processes [41,42]. For example, the Zinc Finger Protein 704 (ZNF704)-SIN3 Transcription Regulator Family Member A (SIN3A) transcriptional repressor complex destabilizes period circadian regulator 2 (PER2) transcription, destabilizing circadian rhythms and driving breast cancer progression [43]. At the metabolic interface, core clock genes directly regulate metabolic enzymes: BMAL1 or CLOCK knockdown disrupts both mitochondrial respiration and glycolysis [44].
Circadian systems in cancer dynamically interact with microenvironmental stressors. Under hypoxic conditions, BMAL1 stabilizes HIF-1α to facilitate metabolic adaptation, while fluctuating oxygen levels synchronize circadian rhythms via HIF-1α-mediated regulation of clock genes [45]. This reciprocal relationship between circadian and hypoxic signaling establishes a self-reinforcing loop that promotes genomic instability and metabolic plasticity. Thus, circadian disruption emerges as a critical contributor to spatiotemporal tumor heterogeneity, highlighting the potential for time-based metabolic therapies. With this temporal framework established, we next will integrate the spatial and temporal dimensions to elucidate the molecular regulatory networks governing metabolic heterogeneity.

3. Regulatory Networks of Spatiotemporal Heterogeneity in Tumor Glucose Metabolism Reprogramming

3.1. Spatial Heterogeneity: Topographic Metabolic Landscapes and Microenvironmental Crosstalk

3.1.1. Oxygen Gradient-Driven Metabolic Zonation Mechanism and Epigenetic Interplay

Tumors develop spatially heterogeneous oxygen gradients that establish distinct metabolic zones, with HIF-1α orchestrating the cellular response to hypoxia. Under hypoxia, stabilized HIF-1α dimerizes with HIF-1β to transactivate genes governing glycolysis (GLUT1/3, HK2, LDHA), angiogenesis (VEGF), and pH regulation (Carbonic Anhydrase IX (CAIX), MCT4), simultaneously suppressing oxidative metabolism through pyruvate dehydrogenase kinase 1 (PDK1)-mediated inhibition of pyruvate dehydrogenase [46,47,48]. This metabolic reprogramming is further reinforced by crosstalk with oncogenic signaling pathways. The PI3K/Akt/mTOR pathway, for instance, promotes both HIF-1α production and stabilization [49,50].
Metabolic zonation promotes the accumulation of oncometabolites, which coordinate signaling and epigenetic remodeling (Table 3). Lactate promotes epithelial-mesenchymal transition (EMT) via histone lactylation and histone deacetylase (HDAC) inhibition [51,52]. Succinate stabilizes HIF-1α via competitive prolyl hydroxylase (PHD) inhibition [53,54,55], establishing a feedforward hypoxic response. Moreover, succinate and fumarate inhibit alpha-ketoglutarate (α-KG)-dependent dioxygenases (ten–eleven translocation methylcytosine dioxygenases (TETs), histone lysine demethylases (KDMs)), inducing DNA and histone hypermethylation [54]. In addition, 2-hydroxyglutarate disrupts fat mass and obesity-associated protein (FTO)-mediated RNA demethylation [56]. This intricate coupling of oxygen sensing, metabolic reconfiguration, and epigenetic rearrangement establishes a self-reinforcing cycle that sustains tumor heterogeneity, accelerates progression, and confers therapeutic resistance.

3.1.2. Metabolic Competition in the Immunometabolic Microenvironment

Cancer cells escape immune surveillance through metabolic reprogramming. Two key mechanisms drive this process: (1) the depletion of essential nutrients from the tumor microenvironment; and (2) the production of metabolites that suppress immune cell function. Through aerobic glycolysis (the Warburg effect), cancer cells aggressively consume glucose, starving adjacent T cells. This nutrient deprivation disrupts mechanistic target of rapamycin complex 1 (mTORC1) signaling, impairing T-cell activation and proliferation, and interferon gamma (IFN-γ) production [59,60,61].
Lactate accumulation contributes to immune evasion through several mechanisms: (1) impairing dendritic cell function by reducing MHC-II expression and IL-12 secretion via G-protein-coupled receptor 81 (GPR81) [62,63]; (2) expanding regulatory T cells (Tregs) through metabolic–epigenetic reprogramming [64,65]; (3) promoting OXPHOS-dependent Tregs with high PD-1 expression via GPR81/PI3K/AKT/mTOR signaling [64,66], particularly in PD-1-resistant tumors [66,67].
Beyond lactate, other key oncometabolites contribute significantly to immune-suppression within the glucose-deprived tumor microenvironment (TME) and intersect with cancer glucose metabolism. Tumors in the glucose-limited environments enhance glutamine uptake via glutaminase (GLS), converting it to glutamate for TCA cycle replenishment [68]. This glutamine depletion directly impairs T-cell activation and cytotoxicity [69]. Ketone bodies like β-hydroxybutyrate (β-HB) serve as alternative energy sources in glucose-limited conditions. Although certain immune cells (e.g., memory T cells) use β-HB for metabolic homeostasis, its accumulation in the TME may promote immune tolerance independently of HDAC inhibition [70,71]. The hypoxic and nutrient-poor TMEs also accelerate ATP breakdown, elevating extracellular adenosine levels. Through adenosine Receptor A2a (A2aR) binding, adenosine suppresses T-cell and NK cell activity, further promoting immune evasion [72]. Thus, glutamine metabolism supplying the TCA cycle and glycolytic ATP generating adenosine create metabolically linked nodes within tumor glucose metabolism. This interconnected network sustains an immunosuppressive TME and facilitates tumor immune evasion.
Immune evasion is also supported by glycolytic enzymes through non-metabolic mechanisms. The C-terminal domain of HK2 activates NF-κB to upregulate PD-L1 in glioblastoma and breast cancer [73,74,75]. Glycolysis–PI3K/AKT/mTOR crosstalk inhibits CD8+ T cells, expands Tregs, and promotes M2 macrophage polarization [76,77,78]. Lactate signaling through MCT4-lactate-PI3Kγ/Akt blocks M1 macrophage polarization via interferon regulatory factor 5 (IRF5) suppression [79]. Thus, these pathways establish an immunosuppressive microenvironment through both metabolic competition and direct immune cell regulation, revealing novel therapeutic targets.

3.1.3. Metabolic Plasticity Drives Tumor Metastasis

  • Metabolic Adaptations During Invasion and Dissemination
Tumor cells exploit metabolic flexibility to adapt metastasis, from initial invasion through distant colonization. As cells detach from primary sites, hypoxia and nutrient deprivation trigger HIF-1α/2α signaling, enhancing glycolysis and reactive oxygen species (ROS) accumulation, further inducing EMT [80]. Pyruvate kinase M2 (PKM2) reinforces this process by altering nuclear metabolite pools to modulate EMT gene expression [81]. Acidification from lactate and CO2 activates MMPs to remodel basement membranes, while glutamine-derived ammonia provides another mechanism driving invasion [82].
  • Survival Mechanisms of Circulating Tumor Cells (CTCs)
To survive in circulation, circulating tumor cells (CTCs) evade anoikis by upregulating pentose phosphate pathway (PPP)-driven nicotinamide adenine dinucleotide phosphate (NADPH) production to counteract ROS [83,84]. Exosomes mediate metabolic crosstalk: PKM2-containing exosomes from lung cancer cells activate stromal glycolysis, promoting bone destruction [85]. Exosomal miR-122 depletes extracellular glucose to fuel metastasis [86]. Furthermore, tumor exosomes modulate immune metabolism by inhibiting T-cell activation via glycolytic enzyme transfer, whereas glycolysis inhibition conversely restores antitumor immune responses [87]. These adaptations enable CTC survival during dissemination and prime metastatic niches.
  • Organ-Specific Metabolic Strategies for Colonization
Metabolic reprogramming drives organ-specific patterns of metastatic colonization. Brain metastases utilize acetate (via acyl-CoA synthetase short-chain family member 2 (ACSS2)) and branched-chain amino acids [88,89] in glucose-scarce environments [90]. Lung metastases enhance mitochondrial function through PGC-1α/peroxiredoxin 2(PRDX2) [91,92], favoring pyruvate over glutamine [93]. Liver metastases employ glycolysis and phosphocreatine cycling (PDK1/creatine kinase B-type (CKB)) for rapid ATP production [94,95]. Bone metastases actively remodel the niche through lactate and serine secretion, and drive osteoclastogenesis [88,96,97]. These organ-specific strategies highlight the role of metabolic reprogramming during the spread of metastases.

3.2. Temporal Dynamic Regulation: From Genetic Mutations to Epigenetic Memory

Following our analysis of spatial–microenvironmental interactions, we explore temporal–genetic mechanisms, analyzing how mutations and epigenetic modifications drive metabolic adaptations throughout tumor evolution and therapy, with a particular focus on clonal selection, treatment-induced remodeling, and molecular-level circadian regulation.

3.2.1. Metabolic Clonal Selection During Tumor Evolution

During early tumor development, oncogenic drivers like KRAS mutations increase the expression of glycolytic enzymes (HK2, PKM2) to induce the Warburg effect [98]. This metabolic shift favors glycolysis over oxidative phosphorylation, enabling rapid ATP production and generation of biosynthetic precursors (lipids, nucleotides), even in oxygen-rich conditions. This metabolic rewiring further promotes clonal expansion by some mechanisms: (1) glycolytic intermediates (e.g., PPP-generated NADPH) decrease oxidative damage and DNA damage [99]; and (2) p53 loss increases PPP flux through TP53-induced glycolysis and apoptosis regulator (TIGAR) downregulation and glucose-6-phosphate dehydrogenase (G6PD) depression, to enhance NADPH/glutathione (GSH) production [100]. The newly acquired metabolic phenotype confers both biosynthetic efficiency for proliferation and antioxidant protection against microenvironmental stress (hypoxia, inflammation), enabling early clonal dominance and malignant transformation.
In late-stage metastasis, EMT induces profound metabolic changes to facilitate cancer cell dissemination. Transcription factors such as Snail and Twist drive a glycolytic shift by suppressing mitochondrial genes (e.g., Cytochrome C Oxidase Subunit 4I1) to reduce OXPHOS and ROS but enhancing glycolysis for migration energy needs [101]. This metabolic rewiring promotes invasiveness through LDHA-mediated lactate production (activating matrix metalloproteinases (MMPs) for extracellular matrix (ECM) degradation) and glycolytic intermediate-supported cytoskeletal remodeling [102]. Emerging evidence, however, reveals substantial complexity in the OXPHOS–metastasis relationship, as many cancers maintain functional OXPHOS alongside glycolysis despite EMT-mediated suppression, including pancreatic ductal adenocarcinoma (PDAC), high-OXPHOS melanoma, and endometrial carcinomas [103,104,105]. This metabolic flexibility helps tumors adapt to microenvironmental changes during metastasis. It has been identified that OXPHOS inhibition enhances metastatic potential in specific biological contexts due to the mutational inactivation of tumor suppressors, such as the p53 mutation affecting mitochondrial function [106,107], EMT-related repression of cytochrome c oxidase via Snail [108], and deletion of KiSS-1 metastasis-suppressor (KISS1)-mediated PGC1α regulation [109]. Conversely, OXPHOS remains essential for metastatic colonization in other contexts: melanoma metastases require OXPHOS/PGC1α for lung/brain colonization [110] and leukemia stem cells depend critically on OXPHOS for persistence [111,112]. This functional duality explains the contradictory findings: OXPHOS suppression may facilitate early dissemination via glycolytic activation but impair colonization in OXPHOS-dependent tumors.
To overcome detachment stress, CTCs activate FAO through the peroxisome proliferator-activated receptor gamma (PPARγ)/CPT1A pathway. The roles of FAO in metabolic adaptation include: (1) generate ATP for energy needs during circulation [113,114]; (2) produce acetyl-CoA for histone acetylation of pro-survival genes (e.g., B-cell lymphoma 2 (BCL-2)) [115,116]; and (3) maintain mitochondrial membrane potential to suppress anoikis [28]. This FAO-dependent adaptation, combined with the initial glycolysis-driven EMT program that enables invasion, creates a coordinated metabolic cascade that drives successful metastatic progression.

3.2.2. Treatment-Induced Metabolic and Epigenetic Remodeling Mechanisms

Cancer cells evade therapy resistance by rewriting metabolic and epigenetic programs. These adaptations establish regulatory networks that manage stress and transmit drug-resistant phenotypes to daughter cells via epigenetic memory, ultimately leading to treatment failure and tumor recurrence.
Chemotherapy causes oxidative stress and DNA damage, forcing cancer cells to survive via metabolic adaptation. Drug-resistant cells combat oxidative stress by activating the nuclear factor erythroid 2-related factor 2 (NRF2)/G6PD axis to boost PPP-derived NADPH for ROS detoxification [117,118,119,120,121]. Platinum-based chemotherapy induces metabolic reprogramming by suppressing serine synthesis. Phosphoglycerate dehydrogenase (PHGDH) downregulation maintains the cellular NAD+ levels required for PARP-dependent DNA repair processes. In contrast, 5-FU-resistant cells adapt by upregulating Serine Hydroxymethyltransferase 2 (SHMT2) to enhance serine metabolism and maintain nucleotide production [122,123]. These adaptive changes underscore metabolic plasticity as a primary mechanism of resistance.
Ionizing radiation drives radioresistance through an integrated network of metabolic adaptations. HIF-1α upregulates GLUT1, which enhances glucose uptake, and p53 suppression enhances hypoxic survival [124,125]. ATM Serine/Threonine Kinase (ATM)/NRF2 activates PPP via G6PD for NADPH-mediated ROS clearance and DNA repair [126]. Radiation-induced lactate efflux via MCT1 acidifies the microenvironment, impairing CD8+ T-cell function and activating HIF-1α/myeloid-derived suppressor cells (MDSC)-mediated immunosuppression [126]. Concurrently, signal transducer and activator of transcription 5 (STAT5) induces glutamine synthetase to fuel nucleotide synthesis [127]. These synergistic adaptations establish diagnostic metabolic signatures in recurrent tumors [128].
Cancer therapies induce epigenetic–metabolic crosstalk through multiple mechanisms. DNA-damaging drugs modulate the expression of metabolic enzymes by activating DNA methyltransferase 3 alpha (DNMT3A). For example, isocitrate dehydrogenase 1 (IDH1) silencing by promoter methylation promotes glutamine metabolism [129]. LDHA or phosphoenolpyruvate carboxykinase 1 (PCK1) hypomethylation enhances glycolytic/gluconeogenic flux [130]. Enhancer of zeste homolog 2 (EZH2)-mediated H3K27me3 regulates chromatin accessibility [131], but its inhibition (e.g., GSK126) disrupts mitochondrial function and triggers glycolytic compensation [131]. This epigenetic plasticity facilitates metabolic adaptation to treatment stress.

3.2.3. Molecular Basis of Metabolic Circadian Rhythm

Tumor cells disrupt circadian metabolic regulation through multiple mechanisms, including clock gene dysregulation, temporal heterogeneity of metabolic enzyme expression, and time-dependent crosstalk between microenvironmental signals, ultimately contributing to tumor proliferation, metastasis, and resistance. Figure 3 shows the molecular crosstalk between circadian rhythm disruption and tumor metabolic reprogramming.
The deregulation of core clock genes drives tumor progression. BMAL1 deficiency activates cancer-associated fibroblasts, promoting invasion and immune escape [132,133]. PER2 recruits HIF-1α to suppress mitochondrial function, forcing glycolytic metabolism [134]. MYC-induced nuclear receptor subfamily 1 group D member 1 (REV-ERBα) represses BMAL1, upregulating HK2 and GLUT1 [42,135,136]. BMAL1/PER2 knockout accelerates lung tumor growth with glycolytic upregulation [137]. Post translationally, CLOCK phosphorylation at Ser106 promotes nucleotide biosynthesis and tumor proliferation [138]. Period circadian regulator 1 (PER1) mutations enhance glycolytic flux and tumor progression via the receptor for activated C kinase 1 (RACK1)/PI3K/AKT pathway [42,139]. These changes rewire the cancer metabolism, creating a permissive environment for progression.
Tumor metabolites like lactate and cytokines disrupt circadian regulation through feedback loops. The deacetylation of BMAL1 by lactate/interleukin-1 beta (IL-1β) amplifies LDHA expression and tumor growth [140]. Hypoxia stabilizes HIF-1α–BMAL1 interactions to suppress mitochondria and enhance glycolysis [42,141]. Systemically, breast cancers remodel liver metabolism via signal transducer and activator of transcription 3 (STAT3)-suppressor of cytokine signaling 3 (SOCS3)-AMPK to create metastatic niches [42,142]. Lung cancers induce CLOCK/BMAL1-mediated hepatic inflammation to evade immunity [42,143]. Such cross-organ circadian reprogramming highlights the integrated metabolic–immune crosstalk in cancer initiation and progression.

4. Targeted Intervention Strategies for Spatiotemporal Heterogeneity

Spatiotemporal heterogeneity in tumors arises from metabolic gradients (hypoxic cores to oxygenated rims), stromal interactions, and clonal evolution, creating distinct therapeutic microenvironments. In this section, we evaluate compartment-specific targeting strategies to address the spatial stratification of treatment-resistance mechanisms.

4.1. Compartment-Specific Metabolic Targeting

Hypoxic tumor cores exhibit glycolytic dependency mediated by HIF-1α stabilization, targetable through HIF-1α inhibitors [144] (adamantane derivatives [145], MO-2097 [146], NLG207 [147], NB-5-MT [148]), combined with glycolytic enzyme blockade (HK2 inhibitor PF-04691502 [149], LDHA inhibitor FX-11 [150,151]). The oxygen-rich perivascular niche demonstrates GLUT1-mediated glucose addiction, sensitive to GLUT1 inhibitors (BAY-876 [152,153], WZB117 [154]), MCT1 inhibitors (AZD3965 [155], AR-C155858 [156]), and the dual MCT1/MCT4 inhibitor (syrosingopine [157]), further impairing lactate efflux.
Metastatic sites develop organ-specific adaptations: brain metastases utilize ACSS2-dependent lipid metabolism inhibited by AD-5584 [158]; lung metastases are vulnerable to PGC-1α inhibitors (SR18292) [159,160] and PRDX2 blockade (Conoidin A) [161]; liver metastases depend on PDK1-regulated glycolysis targeted by the DCA/CPT combination [162]; and bone metastases engage osteoclast metabolic symbiosis disrupted by the PHGDH inhibitor NCT-503 [96,163], MCT1 inhibitors, or cathepsin K blockade (odanacatib) [97,164].

4.2. Temporal Modulation of Metabolic Evolution

Therapeutic strategies must address the dynamic metabolic shifts during tumor progression. Table 4 shows inhibitors targeting glycolysis and the latest drug developments. Early carcinogenesis involves c-Myc-driven glycolytic reprogramming inhibited by c-Myc inhibitors (OmoMYC [165], EN4 [166]). G6PD inhibitors (DHEA [167,168], 6-AN [169]) counteract oxidative stress resistance in p53-deficient cancers by blocking NADPH/GSH production. Emerging compounds (THDOC, G6PDi1 [170]) show improved specificity but require tumor-selective delivery strategies to mitigate normal cell toxicity. Metastatic progression requires EMT–glycolysis pathway disruption through agents like GNE-140 and KIS37 [171,172].
Therapeutic targeting of metabolic reprogramming overcomes drug resistance in cancer, including glycolytic inhibition (2-DG [173], oxamate [174]) and OXPHOS disruption (metformin [175], CB-839 [176]). Circadian-targeted therapies (REV-ERB agonists [177,178], casein kinase 2 (CK2) inhibitors [138,179]) disrupt tumor metabolic rhythms and enhance chemoradiation efficacy in multiple cancers.
Table 4. Targeting glycolysis inhibitors and the latest drug developments.
Table 4. Targeting glycolysis inhibitors and the latest drug developments.
CategoryTargetMechanism of ActionRepresentative DrugsDevelopment Status
Glycolytic Key Enzyme Inhibitors
GLUT InhibitorsGLUT1, GLUT3Block glucose uptake; suppress glycolysis initiation.2-Deoxyglucose (2-DG), WZB117 [154]Preclinical
Hexokinase (HK) InhibitorsHK2, HK3Inhibit glycolysis initiation; reduce ATP/lactate production.HK2 siRNA [180], 3-Bromopyruvate [181]Preclinical
Pyruvate Kinase M2 (PKM2) InhibitorsPKM2Block nuclear PKM2; inhibit HIF-1α/TGF-β-driven EMT.Compound 3K [182]Preclinical
Lactate Dehydrogenase A (LDHA) InhibitorsLDHASuppress lactate production; reverse acidic TME and apoptosis resistance.FX11 [183], GNE-140 [184]FX11: Preclinical; GNE-140: Phase I (Breast/Pancreatic Cancer)
Pyruvate Dehydrogenase Kinase (PDK) InhibitorsPDK1, PDK4Restore OXPHOS; inhibit glycolysis-dependent EMT.DCA [185], KIS37 (Cryptotanshinone) [186]DCA: Phase II (NCT01111097); KIS37: Preclinical
Other Glycolysis-Related Inhibitors
Vitamin C (High-dose IV)Indirect (GAPDH)Depletes NAD+; disrupts microtubule dynamics; inhibits migration.Ascorbate + Gemcitabine [187]Phase I/II (NCT02905578, Pancreatic Cancer)
Aldolase A (ALDOA) InhibitorsALDOABlock glycolysis intermediates; suppress cytoskeletal remodeling.Raltegravir [188]Preclinical
Enolase 1 (ENO1) InhibitorsENO1Inhibit glycolysis; regulate PI3K/AKT signaling.ENO1 DNA vaccine [189,190]Preclinical
Latest Developments
2-DG + MetforminGlycolysis/AMPKSynergistically inhibits EMT and stemness in glioblastoma.2-DG + Metformin [191]Preclinical
DCA (Dichloroacetate)PDK1–EGFR axisReverses cisplatin resistance in ovarian cancer.DCA + Cisplatin [192]Phase II (NCT01111097)
Repurposed Drugs
MetforminAMPK/mTORIndirectly inhibits glycolysis and EMT.Metformin + Chemotherapy [193]Phase II (NCT01864096, Breast Cancer/Glioblastoma)
SimvastatinHMG-CoA reductaseDownregulates CXCR4, vimentin; inhibits TGF-β-induced EMT.Simvastatin [194]Phase II (NCT00944463, Pancreatic Cancer; NCT02161822, Advanced Rectal Cancer)
Immune-Metabolic Synergistic Approaches
MCT4 InhibitorsLactate exportBlock MCT4-mediated lactate efflux to restore T-cell function.AZD3965 [195,196], Diclofenac [197,198], Syrosingopine [157,195]Clinical (AZD3965: Phase I/II)
GLUT1 InhibitorsGlucose uptakeInhibit GLUT1-mediated glucose transport to counteract TME starvation.BAY-876 [152], Phloretin [199], Glutor [200]Preclinical
LDHA InhibitorsLactate productionSuppress LDHA activity to reduce immunosuppressive lactate accumulation.Oxamate [174], Compound 7 [201,202]Preclinical (Compound 7: Phase I)
GPR81 AntagonistsLactate signalingBlock GPR81-mediated immunosuppressive signaling in dendritic cells.3-OBA [203]Preclinical
SREBP2 InhibitorsLipid metabolismInhibit SREBP2 to disrupt lactate signaling in antigen-presenting cells.Fatostatin [63], Botulin [204]Preclinical
PI3K InhibitorsMetabolic–immune crosstalkDual inhibition of PI3K-AKT-mTOR pathway and lactate signaling.Copanlisib [205], Duvelisib [205]Clinical (Copanlisib: FDA-approved for lymphoma)
mTOR BlockersImmune metabolismSuppress mTOR-driven metabolic reprogramming in Tregs.Everolimus [206], CC-223 [207]Clinical (Everolimus: Approved for multiple cancers)
Note: GLUT, glucose transporter; HK2, hexokinase 2; HK3, hexokinase 3; HIF-1α, hypoxia-inducible factor 1 alpha; TGF-β, transforming growth factor beta; TME, tumor microenvironment; OXPHOS, oxidative phosphorylation; EMT, epithelial-mesenchymal transition; GAPDH, glyceraldehyde-3-phosphate dehydrogenase; NAD+, nicotinamide adenine dinucleotide; AMPK, AMP-activated protein kinase; mTOR, mechanistic target of rapamycin; HMG-CoA, 3-hydroxy-3-methylglutaryl-coenzyme A; CXCR4, C-X-C motif chemokine receptor 4; MCT4, monocarboxylate transporter 4; LDHA, lactate dehydrogenase A; GPR81, G protein-coupled receptor 81; SREBP2, sterol regulatory element-binding protein 2.

4.3. Immune–Metabolic Synergistic Approaches

The Warburg effect and lactate secretion create an immunosuppressive tumor microenvironment that is targetable through specific inhibitors: MCT4 inhibitors (AZD3965 [195,196], diclofenac [197,198], syrosingopine [157,195]) block lactate export to restore T-cell function; GLUT1 inhibitors (BAY-876 [152], phloretin [199], Glutor [200]) counteract glucose depletion; LDHA inhibitors (oxamate [174], Compound 7 [201,202]) suppress lactate production; GPR81 antagonism (3-OBA) [203] and sterol regulatory element-binding protein 2 (SREBP2) inhibition (fatostatin [63], botulin [204]) disrupt lactate signaling in dendritic cells; PI3K inhibitors (copanlisib [205], duvelisib [205]) and mTOR blockers (everolimus [206], CC-223 [207]) dually target tumor metabolism and immune suppression. These agents show enhanced efficacy when combined with PD-1/PD-L1 inhibitors, offering a multipronged approach to overcome metabolic immunosuppression.

4.4. Innovative Approaches for Targeting Metabolic–Immune Crosstalk

Recent breakthroughs in multimodal imaging, organoid models, nanodelivery systems, and artificial intelligence (AI) have enabled novel strategies for the precision modulation of the metabolic–immune microenvironment. Integrated multimodal imaging (PET/MRI/MALDI-TOF MS, NanoSIMS) elucidates the spatial mapping of metabolic heterogeneity from tissue to subcellular levels [208,209,210,211]. Fluorescent biosensors (FLII12Pglu, SoNar) achieve real-time single-cell metabolic profiling [18,212,213]. Organ-on-a-chip (OOC) systems are microfluidic platforms that recapitulate key physiological and pathological features of human organs, integrating living cells within engineered extracellular matrices and dynamic fluid flow to mimic tissue–tissue interactions and organ-level functions [214]. OOC systems interpret tumor microenvironments with vascular networks and immune interactions, enabling personalized drug testing [215,216]. Patient-derived tumor chips accurately predict clinical treatment responses, improving outcomes for cancers such as glioblastoma and colorectal carcinoma [217,218].
Two core nanosystems have been developed to target tumor microenvironment-specific biomarkers: hypoxia- and enzyme-responsive nanoparticles. Hypoxia-responsive nanocarriers with nitroimidazole moieties and pH-sensitive polymers achieve targeted drug release in acidic tumor microenvironments [219,220]. Enzyme-responsive nanoparticles (GOx-MnO2 [221], HAase-AuNR@HA-DC [222], MMP-liposomes [223]) enable precise drug release by responding to hyaluronidase, glucose oxidase, or matrix metalloproteinases in the microenvironment.
Machine learning analyzes tumor metabolism from PET scans and mass spectrometry data to predict treatment response. The REKINDLE model [224] uses reinforcement learning to simulate tumor metabolic adaptation to inhibitors. Machine learning applied to 13C-glucose tracing data predicts optimal drug timing in glioblastoma by disrupting the OXPHOS–glycolysis balance [225]. Supervised machine learning algorithms integrates multi-omics data for personalized regimens [226,227] and temporal metabolic modeling guides therapy adjustments in pancreatic cancer [228]. These technologies collectively address spatial and temporal heterogeneity, enabling targeted interventions that account for tumor evolution and microenvironment interactions.

5. Challenges and Future Perspectives

Although significant advances have been made, fundamental questions remain regarding the spatiotemporal regulation of tumor metabolism. Key unresolved issues include: (1) the molecular basis of metabolic heterogeneity across different cancer types; (2) the adaptive mechanisms underlying resistance to metabolism-targeted therapies. A key challenge is developing spatiotemporally precise metabolic modulation, which is a prerequisite for effective targeting.

5.1. Current Challenges and Technical Limitations

Current metabolic imaging technologies—from clinical FDG-PET to research-grade hyperpolarized 13C-MRI (1–2 mm resolution) and mass spectrometry imaging (50–100 μm)—face inherent trade-offs between spatial resolution and throughput. These limitations constrain the analysis of metabolic heterogeneity at cellular scales. Existing model systems, including patient-derived organoids lacking stromal components, and humanized mice with species-specific metabolic discrepancies, further complicate translational research. The integration of multi-omics data remains challenging, with current methods matching only 50–60% of cells across adjacent tumor sections. Advancing the field requires coordinated development of next-generation imaging, improved computational models, and more physiologically relevant experimental systems.

5.2. Clinical Translation Barriers

Metabolic biomarkers have two key limitations: poor specificity of circulating metabolites and unresolved spatial metabolic heterogeneity in tissue biopsies. Exosomal liquid biopsies and PET–genomic integration may improve monitoring. Drug delivery challenges still persist, such as nanocarriers facing stromal barriers. Metabolic-drug combinations risk, genome-scale modeling, and intervention atlases could optimize strategies. Moreover, current metabolic-drug combinations lack rationale, potentially causing antagonism, for example, glycolysis attenuates the efficacy of checkpoint immunotherapy. Computational modeling (genome-scale metabolic models) and systematic atlases of metabolic-drug interactions are needed to identify synergistic sequences and guide clinical translation.

5.3. Future Directions

Spatial multi-omics (multiplexed error-robust fluorescence in situ hybridization, Visium, imaging mass spectrometry) reveals new targets at a single-cell resolution. Standardized protocols and lower costs will soon enable metabolic atlases for precision therapies. Targeting microbiome-derived metabolites (e.g., short-chain fatty acids) and intratumoral microbes (Fusobacterium) via probiotics or phage editing may overcome immunosuppression and improve therapy responses. Expanding AI-based digital twins with continuous metabolic monitoring could achieve >75% prediction accuracy, enabling automated therapy optimization across cancer types. The “metabolic clock” framework will combine spatial tumor targeting, circadian timing, and systemic coordination using molecular profiling and adaptive algorithms for precise metabolic therapy.

6. Conclusions

Recent advances in single-cell and spatial omics have transformed our understanding of tumor glucose metabolism, revealing dynamic spatiotemporal regulation and metabolic plasticity. These findings establish direct links between glycolytic flux, epigenetic reprogramming, and immune evasion. Clinically, the “metabolic clock” concept provides a framework for timed therapeutic interventions. The remaining challenges include developing tumor-specific inhibitors and understanding metabolic crosstalk in resistant niches. Future work must bridge these mechanistic insights into clinical applications through rigorous translational studies.

Author Contributions

Conceptualization, L.L. and X.C.; writing—original draft preparation, X.C. and L.L.; writing—review and editing, L.L. and Q.T.; funding acquisition, L.L., X.C., and Q.T. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the NSFC/RGC Joint Research Scheme (#N_CUHK419/24), the RGC (#14111924, #14102923), and the Faculty Postdoctoral Fellowship Scheme, Chinese University of Hong Kong.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Jiang, S.; Li, H.; Zhang, L.; Mu, W.; Zhang, Y.; Chen, T.; Wu, J.; Tang, H.; Zheng, S.; Liu, Y.; et al. Generic Diagramming Platform (GDP): A comprehensive database of high-quality biomedical graphics. Nucleic Acids Res. 2025, 53, D1670–D1676. [Google Scholar] [CrossRef] [PubMed]
  2. Flint, L.E.; Hamm, G.; Ready, J.D.; Ling, S.; Duckett, C.J.; Cross, N.A.; Cole, L.M.; Smith, D.P.; Goodwin, R.J.A.; Clench, M.R. Characterization of an Aggregated Three-Dimensional Cell Culture Model by Multimodal Mass Spectrometry Imaging. Anal. Chem. 2020, 92, 12538–12547. [Google Scholar] [CrossRef] [PubMed]
  3. Demicco, M.; Liu, X.Z.; Leithner, K.; Fendt, S.M. Metabolic heterogeneity in cancer. Nat. Metab. 2024, 6, 18–38. [Google Scholar] [CrossRef] [PubMed]
  4. 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]
  5. Lee, P.; Chandel, N.S.; Simon, M.C. Cellular adaptation to hypoxia through hypoxia inducible factors and beyond. Nat. Rev. Mol. Cell Biol. 2020, 21, 268–283. [Google Scholar] [CrossRef] [PubMed]
  6. Miallot, R.; Galland, F.; Millet, V.; Blay, J.Y.; Naquet, P. Metabolic landscapes in sarcomas. J. Hematol. Oncol. 2021, 14, 114. [Google Scholar] [CrossRef] [PubMed]
  7. Tanaka, A.; Ogawa, M.; Zhou, Y.; Otani, Y.; Hendrickson, R.C.; Miele, M.M.; Li, Z.; Klimstra, D.S.; Wang, J.Y.; Roehrl, M.H. Proteogenomic characterization of pancreatic neuroendocrine tumors uncovers hypoxia and immune signatures in clinically aggressive subtypes. iScience 2024, 27, 110544. [Google Scholar] [CrossRef] [PubMed]
  8. Kueckelhaus, J.; Frerich, S.; Kada-Benotmane, J.; Koupourtidou, C.; Ninkovic, J.; Dichgans, M.; Beck, J.; Schnell, O.; Heiland, D.H. Inferring histology-associated gene expression gradients in spatial transcriptomic studies. Nat. Commun. 2024, 15, 7280. [Google Scholar] [CrossRef] [PubMed]
  9. Guillaumond, F.; Leca, J.; Olivares, O.; Lavaut, M.N.; Vidal, N.; Berthezene, P.; Dusetti, N.J.; Loncle, C.; Calvo, E.; Turrini, O.; et al. Strengthened glycolysis under hypoxia supports tumor symbiosis and hexosamine biosynthesis in pancreatic adenocarcinoma. Proc. Natl. Acad. Sci. USA 2013, 110, 3919–3924. [Google Scholar] [CrossRef] [PubMed]
  10. Chen, J.; Huang, Z.; Chen, Y.; Tian, H.; Chai, P.; Shen, Y.; Yao, Y.; Xu, S.; Ge, S.; Jia, R. Lactate and lactylation in cancer. Signal Transduct. Target. Ther. 2025, 10, 38. [Google Scholar] [CrossRef] [PubMed]
  11. Singh, L.; Nair, L.; Kumar, D.; Arora, M.K.; Bajaj, S.; Gadewar, M.; Mishra, S.S.; Rath, S.K.; Dubey, A.K.; Kaithwas, G.; et al. Hypoxia induced lactate acidosis modulates tumor microenvironment and lipid reprogramming to sustain the cancer cell survival. Front. Oncol. 2023, 13, 1034205. [Google Scholar] [CrossRef] [PubMed]
  12. Caniglia, J.L.; Jalasutram, A.; Asuthkar, S.; Sahagun, J.; Park, S.; Ravindra, A.; Tsung, A.J.; Guda, M.R.; Velpula, K.K. Beyond glucose: Alternative sources of energy in glioblastoma. Theranostics 2021, 11, 2048–2057. [Google Scholar] [CrossRef] [PubMed]
  13. Ravi, V.M.; Will, P.; Kueckelhaus, J.; Sun, N.; Joseph, K.; Salie, H.; Vollmer, L.; Kuliesiute, U.; von Ehr, J.; Benotmane, J.K.; et al. Spatially resolved multi-omics deciphers bidirectional tumor-host interdependence in glioblastoma. Cancer Cell 2022, 40, 639–655e613. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, Z.; Zhang, Z.; Zhang, Y.; Zhou, W.; Zhang, X.; Peng, C.; Ji, T.; Zou, X.; Zhang, Z.; Ren, Z. Spatial transcriptomics reveals that metabolic characteristics define the tumor immunosuppression microenvironment via iCAF transformation in oral squamous cell carcinoma. Int. J. Oral. Sci. 2024, 16, 9. [Google Scholar] [CrossRef] [PubMed]
  15. Zhi, Y.; Wang, Q.; Zi, M.; Zhang, S.; Ge, J.; Liu, K.; Lu, L.; Fan, C.; Yan, Q.; Shi, L.; et al. Spatial Transcriptomic and Metabolomic Landscapes of Oral Submucous Fibrosis-Derived Oral Squamous Cell Carcinoma and its Tumor Microenvironment. Adv. Sci. 2024, 11, e2306515. [Google Scholar] [CrossRef] [PubMed]
  16. Zhou, L.; Liu, J.; Yao, P.; Liu, X.; Chen, F.; Chen, Y.; Zhou, L.; Shen, C.; Zhou, Y.; Du, X.; et al. Spatial transcriptomics reveals unique metabolic profile and key oncogenic regulators of cervical squamous cell carcinoma. J. Transl. Med. 2024, 22, 1163. [Google Scholar] [CrossRef] [PubMed]
  17. Sun, C.; Wang, A.; Zhou, Y.; Chen, P.; Wang, X.; Huang, J.; Gao, J.; Wang, X.; Shu, L.; Lu, J.; et al. Spatially resolved multi-omics highlights cell-specific metabolic remodeling and interactions in gastric cancer. Nat. Commun. 2023, 14, 2692. [Google Scholar] [CrossRef] [PubMed]
  18. Kondo, H.; Ratcliffe, C.D.H.; Hooper, S.; Ellis, J.; MacRae, J.I.; Hennequart, M.; Dunsby, C.W.; Anderson, K.I.; Sahai, E. Single-cell resolved imaging reveals intra-tumor heterogeneity in glycolysis, transitions between metabolic states, and their regulatory mechanisms. Cell Rep. 2021, 34, 108750. [Google Scholar] [CrossRef] [PubMed]
  19. Jackson, M.; Serada, N.; Sheehan, M.; Srinivasan, S.; Mason, N.; Guha, M.; Avadhani, N. Mitochondrial genome and functional defects in osteosarcoma are associated with their aggressive phenotype. PLoS ONE 2018, 13, e0209489. [Google Scholar] [CrossRef] [PubMed]
  20. Babichev, Y.; Kabaroff, L.; Datti, A.; Uehling, D.; Isaac, M.; Al-Awar, R.; Prakesch, M.; Sun, R.X.; Boutros, P.C.; Venier, R.; et al. PI3K/AKT/mTOR inhibition in combination with doxorubicin is an effective therapy for leiomyosarcoma. J. Transl. Med. 2016, 14, 67. [Google Scholar] [CrossRef] [PubMed]
  21. Mao, L.; Dauchy, R.T.; Blask, D.E.; Dauchy, E.M.; Slakey, L.M.; Brimer, S.; Yuan, L.; Xiang, S.; Hauch, A.; Smith, K.; et al. Melatonin suppression of aerobic glycolysis (Warburg effect), survival signalling and metastasis in human leiomyosarcoma. J. Pineal Res. 2016, 60, 167–177. [Google Scholar] [CrossRef] [PubMed]
  22. Yan, L.; Raj, P.; Yao, W.; Ying, H. Glucose Metabolism in Pancreatic Cancer. Cancers 2019, 11, 1460. [Google Scholar] [CrossRef] [PubMed]
  23. Adams, A.; Scheckel, B.; Habsaoui, A.; Haque, M.; Kuhr, K.; Monsef, I.; Bohlius, J.; Skoetz, N. Intravenous iron versus oral iron versus no iron with or without erythropoiesis-stimulating agents (ESA) for cancer patients with anaemia: A systematic review and network meta-analysis. Cochrane Database Syst. Rev. 2022, 6, CD012633. [Google Scholar] [CrossRef] [PubMed]
  24. Prayoonhong, W.; Sonsingh, W.; Permsuwan, U. Clinical outcomes and economic evaluation of patient-centered care system versus routine-service system for patients with type 2 diabetes in Thailand. Heliyon 2024, 10, e25093. [Google Scholar] [CrossRef] [PubMed]
  25. Shan, T.; Chen, S.; Chen, X.; Lin, W.R.; Li, W.; Ma, J.; Wu, T.; Cui, X.; Ji, H.; Li, Y.; et al. Cancer-associated fibroblasts enhance pancreatic cancer cell invasion by remodeling the metabolic conversion mechanism. Oncol. Rep. 2017, 37, 1971–1979. [Google Scholar] [CrossRef] [PubMed]
  26. Guo, Q. Changes in mitochondrial function during EMT induced by TGFbeta-1 in pancreatic cancer. Oncol. Lett. 2017, 13, 1575–1580. [Google Scholar] [CrossRef] [PubMed]
  27. Shasha, T.; Gruijs, M.; van Egmond, M. Mechanisms of colorectal liver metastasis development. Cell Mol. Life Sci. 2022, 79, 607. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, Y.N.; Zeng, Z.L.; Lu, J.; Wang, Y.; Liu, Z.X.; He, M.M.; Zhao, Q.; Wang, Z.X.; Li, T.; Lu, Y.X.; et al. CPT1A-mediated fatty acid oxidation promotes colorectal cancer cell metastasis by inhibiting anoikis. Oncogene 2018, 37, 6025–6040. [Google Scholar] [CrossRef] [PubMed]
  29. Lee, W.D.; Weilandt, D.R.; Liang, L.; MacArthur, M.R.; Jaiswal, N.; Ong, O.; Mann, C.G.; Chu, Q.; Hunter, C.J.; Ryseck, R.P.; et al. Lactate homeostasis is maintained through regulation of glycolysis and lipolysis. Cell Metab. 2025, 37, 758–771.E8. [Google Scholar] [CrossRef] [PubMed]
  30. Zhou, D.; Duan, Z.; Li, Z.; Ge, F.; Wei, R.; Kong, L. The significance of glycolysis in tumor progression and its relationship with the tumor microenvironment. Front. Pharmacol. 2022, 13, 1091779. [Google Scholar] [CrossRef] [PubMed]
  31. Jing, X.; Yang, F.; Shao, C.; Wei, K.; Xie, M.; Shen, H.; Shu, Y. Role of hypoxia in cancer therapy by regulating the tumor microenvironment. Mol. Cancer 2019, 18, 157. [Google Scholar] [CrossRef] [PubMed]
  32. TeSlaa, T.; Bartman, C.R.; Jankowski, C.S.R.; Zhang, Z.; Xu, X.; Xing, X.; Wang, L.; Lu, W.; Hui, S.; Rabinowitz, J.D. The Source of Glycolytic Intermediates in Mammalian Tissues. Cell Metab. 2021, 33, 367–378.E5. [Google Scholar] [CrossRef] [PubMed]
  33. Yang, J.; Chen, F.; Lang, L.; Yang, F.; Fu, Z.; Martinez, J.; Cho, A.; Saba, N.F.; Teng, Y. Therapeutic Targeting of the GLS1-c-Myc Positive Feedback Loop Suppresses Glutaminolysis and Inhibits Progression of Head and Neck Cancer. Cancer Res. 2024, 84, 3223–3234. [Google Scholar] [CrossRef] [PubMed]
  34. Salhi, A.; Jordan, A.C.; Bochaca, I.I.; Izsak, A.; Darvishian, F.; Houvras, Y.; Giles, K.M.; Osman, I. Oxidative Phosphorylation Promotes Primary Melanoma Invasion. Am. J. Pathol. 2020, 190, 1108–1117. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, J.; Liu, X.; Huang, Y.; Li, P.; Yang, M.; Zeng, S.; Chen, D.; Wang, Q.; Liu, H.; Luo, K.; et al. Targeting nicotinamide N-methyltransferase overcomes resistance to EGFR-TKI in non-small cell lung cancer cells. Cell Death Discov. 2022, 8, 170. [Google Scholar] [CrossRef] [PubMed]
  36. Parmenter, T.J.; Kleinschmidt, M.; Kinross, K.M.; Bond, S.T.; Li, J.; Kaadige, M.R.; Rao, A.; Sheppard, K.E.; Hugo, W.; Pupo, G.M.; et al. Response of BRAF-mutant melanoma to BRAF inhibition is mediated by a network of transcriptional regulators of glycolysis. Cancer Discov. 2014, 4, 423–433. [Google Scholar] [CrossRef] [PubMed]
  37. Das, C.; Adhikari, S.; Bhattacharya, A.; Chakraborty, S.; Mondal, P.; Yadav, S.S.; Adhikary, S.; Hunt, C.R.; Yadav, K.K.; Pandita, S.; et al. Epigenetic-Metabolic Interplay in the DNA Damage Response and Therapeutic Resistance of Breast Cancer. Cancer Res. 2023, 83, 657–666. [Google Scholar] [CrossRef] [PubMed]
  38. Jin, P.; Jiang, J.; Zhou, L.; Huang, Z.; Nice, E.C.; Huang, C.; Fu, L. Mitochondrial adaptation in cancer drug resistance: Prevalence, mechanisms, and management. J. Hematol. Oncol. 2022, 15, 97. [Google Scholar] [CrossRef] [PubMed]
  39. Kumar, P.R.; Moore, J.A.; Bowles, K.M.; Rushworth, S.A.; Moncrieff, M.D. Mitochondrial oxidative phosphorylation in cutaneous melanoma. Br. J. Cancer 2021, 124, 115–123. [Google Scholar] [CrossRef] [PubMed]
  40. Kumar, A.; Chamoto, K. Immune metabolism in PD-1 blockade-based cancer immunotherapy. Int. Immunol. 2021, 33, 17–26. [Google Scholar] [CrossRef] [PubMed]
  41. Kinouchi, K.; Sassone-Corsi, P. Metabolic rivalry: Circadian homeostasis and tumorigenesis. Nat. Rev. Cancer 2020, 20, 645–661. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, Z.; Ma, L.; Meng, Y.; Fang, J.; Xu, D.; Lu, Z. The interplay of the circadian clock and metabolic tumorigenesis. Trends Cell Biol. 2024, 34, 742–755. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, C.; Wu, J.; Liu, X.; Wang, Y.; Liu, B.; Chen, X.; Wu, X.; Yan, D.; Han, L.; Liu, S.; et al. Circadian Rhythm Is Disrupted by ZNF704 in Breast Carcinogenesis. Cancer Res. 2020, 80, 4114–4128. [Google Scholar] [CrossRef] [PubMed]
  44. Dong, Z.; Zhang, G.; Qu, M.; Gimple, R.C.; Wu, Q.; Qiu, Z.; Prager, B.C.; Wang, X.; Kim, L.J.Y.; Morton, A.R.; et al. Targeting Glioblastoma Stem Cells through Disruption of the Circadian Clock. Cancer Discov. 2019, 9, 1556–1573. [Google Scholar] [CrossRef] [PubMed]
  45. Su, K.; Zeng, D.; Zhang, W.; Peng, F.; Cui, B.; Liu, Q. Integrating cancer medicine into metabolic rhythms. Trends Endocrinol. Metab. 2025. [Google Scholar] [CrossRef] [PubMed]
  46. Yfantis, A.; Mylonis, I.; Chachami, G.; Nikolaidis, M.; Amoutzias, G.D.; Paraskeva, E.; Simos, G. Transcriptional Response to Hypoxia: The Role of HIF-1-Associated Co-Regulators. Cells 2023, 12, 798. [Google Scholar] [CrossRef] [PubMed]
  47. Shi, Y.; Gilkes, D.M. HIF-1 and HIF-2 in cancer: Structure, regulation, and therapeutic prospects. Cell Mol. Life Sci. 2025, 82, 44. [Google Scholar] [CrossRef] [PubMed]
  48. Infantino, V.; Santarsiero, A.; Convertini, P.; Todisco, S.; Iacobazzi, V. Cancer Cell Metabolism in Hypoxia: Role of HIF-1 as Key Regulator and Therapeutic Target. Int. J. Mol. Sci. 2021, 22, 5703. [Google Scholar] [CrossRef] [PubMed]
  49. Tian, Y.; Zhao, L.; Gui, Z.; Liu, S.; Liu, C.; Yu, T.; Zhang, L. PI3K/AKT signaling activates HIF1alpha to modulate the biological effects of invasive breast cancer with microcalcification. NPJ Breast Cancer 2023, 9, 93. [Google Scholar] [CrossRef] [PubMed]
  50. Kilic-Eren, M.; Boylu, T.; Tabor, V. Targeting PI3K/Akt represses Hypoxia inducible factor-1alpha activation and sensitizes Rhabdomyosarcoma and Ewing’s sarcoma cells for apoptosis. Cancer Cell Int. 2013, 13, 36. [Google Scholar] [CrossRef] [PubMed]
  51. Latham, T.; Mackay, L.; Sproul, D.; Karim, M.; Culley, J.; Harrison, D.J.; Hayward, L.; Langridge-Smith, P.; Gilbert, N.; Ramsahoye, B.H. Lactate, a product of glycolytic metabolism, inhibits histone deacetylase activity and promotes changes in gene expression. Nucleic Acids Res. 2012, 40, 4794–4803. [Google Scholar] [CrossRef] [PubMed]
  52. Huimin, W.; Xin, W.; Shan, Y.; Junwang, Z.; Jing, W.; Yuan, W.; Qingtong, L.; Xiaohui, L.; Jia, Y.; Lili, Y. Lactate promotes the epithelial-mesenchymal transition of liver cancer cells via TWIST1 lactylation. Exp. Cell Res. 2025, 447, 114474. [Google Scholar] [CrossRef] [PubMed]
  53. Atallah, R.; Olschewski, A.; Heinemann, A. Succinate at the Crossroad of Metabolism and Angiogenesis: Roles of SDH, HIF1alpha and SUCNR1. Biomedicines 2022, 10, 3089. [Google Scholar] [CrossRef] [PubMed]
  54. Yong, C.; Stewart, G.D.; Frezza, C. Oncometabolites in renal cancer. Nat. Rev. Nephrol. 2020, 16, 156–172. [Google Scholar] [CrossRef] [PubMed]
  55. Fernie, G. Consent and the individual detained in custody. Med. Law. 2005, 24, 515–523. [Google Scholar] [PubMed]
  56. Fitzsimmons, C.M.; Mandler, M.D.; Lunger, J.C.; Chan, D.; Maligireddy, S.S.; Schmiechen, A.C.; Gamage, S.T.; Link, C.; Jenkins, L.M.; Chan, K.; et al. Rewiring of RNA methylation by the oncometabolite fumarate in renal cell carcinoma. NAR Cancer 2024, 6, zcae004. [Google Scholar] [CrossRef] [PubMed]
  57. Pianka, S.T.; Li, T.; Prins, T.J.; Eldred, B.S.C.; Kevan, B.M.; Liang, H.; Zapanta Rinonos, S.; Kornblum, H.I.; Nathanson, D.A.; Pellegrini, M.; et al. D-2-HG Inhibits IDH1mut Glioma Growth via FTO Inhibition and Resultant m6A Hypermethylation. Cancer Res. Commun. 2024, 4, 876–894. [Google Scholar] [CrossRef] [PubMed]
  58. Carbonneau, M.; Gagné, L.M.; Lalonde, M.E.; Germain, M.A.; Motorina, A.; Guiot, M.C.; Secco, B.; Vincent, E.E.; Tumber, A.; Hulea, L.; et al. The oncometabolite 2-hydroxyglutarate activates the mTOR signalling pathway. Nat. Commun. 2016, 7, 12700. [Google Scholar] [CrossRef] [PubMed]
  59. Villa, M.; O’Sullivan, D.; Pearce, E.L. Glucose makes Treg lose their temper. Cancer Cell 2021, 39, 460–462. [Google Scholar] [CrossRef] [PubMed]
  60. Zhang, S.; Zhang, X.; Yang, H.; Liang, T.; Bai, X. Hurdle or thruster: Glucose metabolism of T cells in anti-tumour immunity. Biochim. Biophys. Acta Rev. Cancer 2024, 1879, 189022. [Google Scholar] [CrossRef] [PubMed]
  61. Chang, C.H.; Qiu, J.; O’Sullivan, D.; Buck, M.D.; Noguchi, T.; Curtis, J.D.; Chen, Q.; Gindin, M.; Gubin, M.M.; van der Windt, G.J.; et al. Metabolic Competition in the Tumor Microenvironment Is a Driver of Cancer Progression. Cell 2015, 162, 1229–1241. [Google Scholar] [CrossRef] [PubMed]
  62. Llibre, A.; Kucuk, S.; Gope, A.; Certo, M.; Mauro, C. Lactate: A key regulator of the immune response. Immunity 2025, 58, 535–554. [Google Scholar] [CrossRef] [PubMed]
  63. Plebanek, M.P.; Xue, Y.; Nguyen, Y.V.; DeVito, N.C.; Wang, X.; Holtzhausen, A.; Beasley, G.M.; Theivanthiran, B.; Hanks, B.A. A lactate-SREBP2 signaling axis drives tolerogenic dendritic cell maturation and promotes cancer progression. Sci. Immunol. 2024, 9, eadi4191. [Google Scholar] [CrossRef] [PubMed]
  64. Togashi, Y.; Shitara, K.; Nishikawa, H. Regulatory T cells in cancer immunosuppression—Implications for anticancer therapy. Nat. Rev. Clin. Oncol. 2019, 16, 356–371. [Google Scholar] [CrossRef] [PubMed]
  65. Kim, J.; Li, J.; Wei, J.; Lim, S.A. Regulatory T Cell Metabolism: A Promising Therapeutic Target for Cancer Treatment? Immune Netw. 2025, 25, e13. [Google Scholar] [CrossRef] [PubMed]
  66. Kondo, M.; Kumagai, S.; Nishikawa, H. Metabolic advantages of regulatory T cells dictated by cancer cells. Int. Immunol. 2024, 36, 75–86. [Google Scholar] [CrossRef] [PubMed]
  67. Kumagai, S.; Koyama, S.; Itahashi, K.; Tanegashima, T.; Lin, Y.T.; Togashi, Y.; Kamada, T.; Irie, T.; Okumura, G.; Kono, H.; et al. Lactic acid promotes PD-1 expression in regulatory T cells in highly glycolytic tumor microenvironments. Cancer Cell 2022, 40, 201–218.E9. [Google Scholar] [CrossRef] [PubMed]
  68. DeBerardinis, R.J.; Cheng, T. Q’s next: The diverse functions of glutamine in metabolism, cell biology and cancer. Oncogene 2010, 29, 313–324. [Google Scholar] [CrossRef] [PubMed]
  69. 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]
  70. Newman, J.C.; Verdin, E. Ketone bodies as signaling metabolites. Trends Endocrinol. Metab. 2014, 25, 42–52. [Google Scholar] [CrossRef] [PubMed]
  71. Youm, Y.H.; Nguyen, K.Y.; Grant, R.W.; Goldberg, E.L.; Bodogai, M.; Kim, D.; D’Agostino, D.; Planavsky, N.; Lupfer, C.; Kanneganti, T.D.; et al. The ketone metabolite beta-hydroxybutyrate blocks NLRP3 inflammasome-mediated inflammatory disease. Nat. Med. 2015, 21, 263–269. [Google Scholar] [CrossRef] [PubMed]
  72. Antonioli, L.; Yegutkin, G.G.; Pacher, P.; Blandizzi, C.; Hasko, G. Anti-CD73 in cancer immunotherapy: Awakening new opportunities. Trends Cancer 2016, 2, 95–109. [Google Scholar] [CrossRef] [PubMed]
  73. Jiang, C.; Tang, M.; Su, Y.; Xie, J.; Shang, Q.; Guo, M.; An, X.; Lin, L.; Wang, R.; Huang, Q.; et al. Nanomedicine-driven tumor glucose metabolic reprogramming for enhanced cancer immunotherapy. Acta Pharm. Sin. B 2025, 15, 2845–2866. [Google Scholar] [CrossRef] [PubMed]
  74. Guo, D.; Tong, Y.; Jiang, X.; Meng, Y.; Jiang, H.; Du, L.; Wu, Q.; Li, S.; Luo, S.; Li, M.; et al. Aerobic glycolysis promotes tumor immune evasion by hexokinase2-mediated phosphorylation of IkappaBalpha. Cell Metab. 2022, 34, 1312–1324.E6. [Google Scholar] [CrossRef] [PubMed]
  75. Lin, J.; Fang, W.; Xiang, Z.; Wang, Q.; Cheng, H.; Chen, S.; Fang, J.; Liu, J.; Wang, Q.; Lu, Z.; et al. Glycolytic enzyme HK2 promotes PD-L1 expression and breast cancer cell immune evasion. Front. Immunol. 2023, 14, 1189953. [Google Scholar] [CrossRef] [PubMed]
  76. Lee, J.H.; Liu, R.; Li, J.; Wang, Y.; Tan, L.; Li, X.J.; Qian, X.; Zhang, C.; Xia, Y.; Xu, D.; et al. EGFR-Phosphorylated Platelet Isoform of Phosphofructokinase 1 Promotes PI3K Activation. Mol. Cell 2018, 70, 197–210.E7. [Google Scholar] [CrossRef] [PubMed]
  77. Sun, P.; Zhang, X.; Wang, R.J.; Ma, Q.Y.; Xu, L.; Wang, Y.; Liao, H.P.; Wang, H.L.; Hu, L.D.; Kong, X.; et al. PI3Kalpha inhibitor CYH33 triggers antitumor immunity in murine breast cancer by activating CD8+T cells and promoting fatty acid metabolism. J. Immunother. Cancer 2021, 9, e003093. [Google Scholar] [CrossRef] [PubMed]
  78. Peng, X.; Huang, X.; Lulu, T.B.; Jia, W.; Zhang, S.; Cohen, L.; Huang, S.; Fan, J.; Chen, X.; Liu, S.; et al. A novel pan-PI3K inhibitor KTC1101 synergizes with anti-PD-1 therapy by targeting tumor suppression and immune activation. Mol. Cancer 2024, 23, 54. [Google Scholar] [CrossRef] [PubMed]
  79. Chen, S.; Saeed, A.; Liu, Q.; Jiang, Q.; Xu, H.; Xiao, G.G.; Rao, L.; Duo, Y. Macrophages in immunoregulation and therapeutics. Signal Transduct. Target. Ther. 2023, 8, 207. [Google Scholar] [CrossRef] [PubMed]
  80. Faubert, B.; Solmonson, A.; DeBerardinis, R.J. Metabolic reprogramming and cancer progression. Science 2020, 368, eaaw5473. [Google Scholar] [CrossRef] [PubMed]
  81. 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] [PubMed]
  82. Helmlinger, G.; Sckell, A.; Dellian, M.; Forbes, N.S.; Jain, R.K. Acid production in glycolysis-impaired tumors provides new insights into tumor metabolism. Clin. Cancer Res. 2002, 8, 1284–1291. [Google Scholar] [PubMed]
  83. Muri, J.; Kopf, M. The thioredoxin system: Balancing redox responses in immune cells and tumors. Eur. J. Immunol. 2023, 53, e2249948. [Google Scholar] [CrossRef] [PubMed]
  84. Tasdogan, A.; Faubert, B.; Ramesh, V.; Ubellacker, J.M.; Shen, B.; Solmonson, A.; Murphy, M.M.; Gu, Z.; Gu, W.; Martin, M.; et al. Metabolic heterogeneity confers differences in melanoma metastatic potential. Nature 2020, 577, 115–120. [Google Scholar] [CrossRef] [PubMed]
  85. Wang, C.; Zhang, S.; Liu, J.; Tian, Y.; Ma, B.; Xu, S.; Fu, Y.; Luo, Y. Secreted Pyruvate Kinase M2 Promotes Lung Cancer Metastasis through Activating the Integrin Beta1/FAK Signaling Pathway. Cell Rep. 2020, 30, 1780–1797. [Google Scholar] [CrossRef] [PubMed]
  86. Fong, M.Y.; Zhou, W.; Liu, L.; Alontaga, A.Y.; Chandra, M.; Ashby, J.; Chow, A.; O’Connor, S.T.; Li, S.; Chin, A.R.; et al. Breast-cancer-secreted miR-122 reprograms glucose metabolism in premetastatic niche to promote metastasis. Nat. Cell Biol. 2015, 17, 183–194. [Google Scholar] [CrossRef] [PubMed]
  87. Cascone, T.; McKenzie, J.A.; Mbofung, R.M.; Punt, S.; Wang, Z.; Xu, C.; Williams, L.J.; Wang, Z.; Bristow, C.A.; Carugo, A.; et al. Increased Tumor Glycolysis Characterizes Immune Resistance to Adoptive T Cell Therapy. Cell Metab. 2018, 27, 977–987e974. [Google Scholar] [CrossRef] [PubMed]
  88. Mashimo, T.; Pichumani, K.; Vemireddy, V.; Hatanpaa, K.J.; Singh, D.K.; Sirasanagandla, S.; Nannepaga, S.; Piccirillo, S.G.; Kovacs, Z.; Foong, C.; et al. Acetate is a bioenergetic substrate for human glioblastoma and brain metastases. Cell 2014, 159, 1603–1614. [Google Scholar] [CrossRef] [PubMed]
  89. Schild, T.; Low, V.; Blenis, J.; Gomes, A.P. Unique Metabolic Adaptations Dictate Distal Organ-Specific Metastatic Colonization. Cancer Cell 2018, 33, 347–354. [Google Scholar] [CrossRef] [PubMed]
  90. Chen, J.; Lee, H.J.; Wu, X.; Huo, L.; Kim, S.J.; Xu, L.; Wang, Y.; He, J.; Bollu, L.R.; Gao, G.; et al. Gain of glucose-independent growth upon metastasis of breast cancer cells to the brain. Cancer Res. 2015, 75, 554–565. [Google Scholar] [CrossRef] [PubMed]
  91. LeBleu, V.S.; O’Connell, J.T.; Gonzalez Herrera, K.N.; Wikman, H.; Pantel, K.; Haigis, M.C.; de Carvalho, F.M.; Damascena, A.; Domingos Chinen, L.T.; Rocha, R.M.; et al. PGC-1alpha mediates mitochondrial biogenesis and oxidative phosphorylation in cancer cells to promote metastasis. Nat. Cell Biol. 2014, 16, 992–1003. [Google Scholar] [CrossRef] [PubMed]
  92. Stresing, V.; Baltziskueta, E.; Rubio, N.; Blanco, J.; Arriba, M.C.; Valls, J.; Janier, M.; Clezardin, P.; Sanz-Pamplona, R.; Nieva, C.; et al. Peroxiredoxin 2 specifically regulates the oxidative and metabolic stress response of human metastatic breast cancer cells in lungs. Oncogene 2013, 32, 724–735. [Google Scholar] [CrossRef] [PubMed]
  93. 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]
  94. Dupuy, F.; Tabaries, 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] [PubMed]
  95. Loo, J.M.; Scherl, A.; Nguyen, A.; Man, F.Y.; Weinberg, E.; Zeng, Z.; Saltz, L.; Paty, P.B.; Tavazoie, S.F. Extracellular metabolic energetics can promote cancer progression. Cell 2015, 160, 393–406. [Google Scholar] [CrossRef] [PubMed]
  96. Pollari, S.; Kakonen, S.M.; Edgren, H.; Wolf, M.; Kohonen, P.; Sara, H.; Guise, T.; Nees, M.; Kallioniemi, O. Enhanced serine production by bone metastatic breast cancer cells stimulates osteoclastogenesis. Breast Cancer Res. Treat. 2011, 125, 421–430. [Google Scholar] [CrossRef] [PubMed]
  97. Lemma, S.; Di Pompo, G.; Porporato, P.E.; Sboarina, M.; Russell, S.; Gillies, R.J.; Baldini, N.; Sonveaux, P.; Avnet, S. MDA-MB-231 breast cancer cells fuel osteoclast metabolism and activity: A new rationale for the pathogenesis of osteolytic bone metastases. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 3254–3264. [Google Scholar] [CrossRef] [PubMed]
  98. Ma, Q.; Zhang, W.; Wu, K.; Shi, L. The roles of KRAS in cancer metabolism, tumor microenvironment and clinical therapy. Mol. Cancer 2025, 24, 14. [Google Scholar] [CrossRef] [PubMed]
  99. Ying, H.; Kimmelman, A.C.; Lyssiotis, C.A.; Hua, S.; Chu, G.C.; Fletcher-Sananikone, E.; Locasale, J.W.; Son, J.; Zhang, H.; Coloff, J.L.; et al. Oncogenic Kras maintains pancreatic tumors through regulation of anabolic glucose metabolism. Cell 2012, 149, 656–670. [Google Scholar] [CrossRef] [PubMed]
  100. Jiang, P.; Du, W.; Wang, X.; Mancuso, A.; Gao, X.; Wu, M.; Yang, X. p53 regulates biosynthesis through direct inactivation of glucose-6-phosphate dehydrogenase. Nat. Cell Biol. 2011, 13, 310–316. [Google Scholar] [CrossRef] [PubMed]
  101. Liaghat, M.; Ferdousmakan, S.; Mortazavi, S.H.; Yahyazadeh, S.; Irani, A.; Banihashemi, S.; Seyedi Asl, F.S.; Akbari, A.; Farzam, F.; Aziziyan, F.; et al. The impact of epithelial-mesenchymal transition (EMT) induced by metabolic processes and intracellular signaling pathways on chemo-resistance, metastasis, and recurrence in solid tumors. Cell Commun. Signal 2024, 22, 575. [Google Scholar] [CrossRef] [PubMed]
  102. Hua, W.; Ten Dijke, P.; Kostidis, S.; Giera, M.; Hornsveld, M. TGFbeta-induced metabolic reprogramming during epithelial-to-mesenchymal transition in cancer. Cell Mol. Life Sci. 2020, 77, 2103–2123. [Google Scholar] [CrossRef] [PubMed]
  103. Ashton, T.M.; McKenna, W.G.; Kunz-Schughart, L.A.; Higgins, G.S. Oxidative Phosphorylation as an Emerging Target in Cancer Therapy. Clin. Cancer Res. 2018, 24, 2482–2490. [Google Scholar] [CrossRef] [PubMed]
  104. Uslu, C.; Kapan, E.; Lyakhovich, A. Cancer resistance and metastasis are maintained through oxidative phosphorylation. Cancer Lett. 2024, 587, 216705. [Google Scholar] [CrossRef] [PubMed]
  105. Kaufhold, S.; Bonavida, B. Central role of Snail1 in the regulation of EMT and resistance in cancer: A target for therapeutic intervention. J. Exp. Clin. Cancer Res. 2014, 33, 62. [Google Scholar] [CrossRef] [PubMed]
  106. Gurpinar, E.; Vousden, K.H. Hitting cancers’ weak spots: Vulnerabilities imposed by p53 mutation. Trends Cell Biol. 2015, 25, 486–495. [Google Scholar] [CrossRef] [PubMed]
  107. Haupt, S.; Raghu, D.; Haupt, Y. Mutant p53 Drives Cancer by Subverting Multiple Tumor Suppression Pathways. Front. Oncol. 2016, 6, 12. [Google Scholar] [CrossRef] [PubMed]
  108. Peinado, H.; Olmeda, D.; Cano, A. Snail, Zeb and bHLH factors in tumour progression: An alliance against the epithelial phenotype? Nat. Rev. Cancer 2007, 7, 415–428. [Google Scholar] [CrossRef] [PubMed]
  109. Liu, W.; Beck, B.H.; Vaidya, K.S.; Nash, K.T.; Feeley, K.P.; Ballinger, S.W.; Pounds, K.M.; Denning, W.L.; Diers, A.R.; Landar, A.; et al. Metastasis suppressor KISS1 seems to reverse the Warburg effect by enhancing mitochondrial biogenesis. Cancer Res. 2014, 74, 954–963. [Google Scholar] [CrossRef] [PubMed]
  110. Guerrieri, R.A.; Fischer, G.M.; Kircher, D.A.; Joon, A.Y.; Cortez, J.R.; Grossman, A.H.; Hudgens, C.W.; Ledesma, D.A.; Lazcano, R.; Onana, C.Y.; et al. Oxidative Phosphorylation (OXPHOS) Promotes the Formation and Growth of Melanoma Lung and Brain Metastases. bioRxiv 2025. [Google Scholar] [CrossRef] [PubMed]
  111. Tabe, Y.; Konopleva, M. Resistance to energy metabolism—Targeted therapy of AML cells residual in the bone marrow microenvironment. Cancer Drug Resist. 2023, 6, 138–150. [Google Scholar] [CrossRef] [PubMed]
  112. de Beauchamp, L.; Himonas, E.; Helgason, G.V. Mitochondrial metabolism as a potential therapeutic target in myeloid leukaemia. Leukemia 2022, 36, 1–12. [Google Scholar] [CrossRef] [PubMed]
  113. Carracedo, A.; Cantley, L.C.; Pandolfi, P.P. Cancer metabolism: Fatty acid oxidation in the limelight. Nat. Rev. Cancer 2013, 13, 227–232. [Google Scholar] [CrossRef] [PubMed]
  114. Li, M.; Xian, H.C.; Tang, Y.J.; Liang, X.H.; Tang, Y.L. Fatty acid oxidation: Driver of lymph node metastasis. Cancer Cell Int. 2021, 21, 339. [Google Scholar] [CrossRef] [PubMed]
  115. Teuwen, L.A.; Geldhof, V.; Carmeliet, P. How glucose, glutamine and fatty acid metabolism shape blood and lymph vessel development. Dev. Biol. 2019, 447, 90–102. [Google Scholar] [CrossRef] [PubMed]
  116. McDonnell, E.; Crown, S.B.; Fox, D.B.; Kitir, B.; Ilkayeva, O.R.; Olsen, C.A.; Grimsrud, P.A.; Hirschey, M.D. Lipids Reprogram Metabolism to Become a Major Carbon Source for Histone Acetylation. Cell Rep. 2016, 17, 1463–1472. [Google Scholar] [CrossRef] [PubMed]
  117. Yang, H.C.; Stern, A.; Chiu, D.T. G6PD: A hub for metabolic reprogramming and redox signaling in cancer. Biomed. J. 2021, 44, 285–292. [Google Scholar] [CrossRef] [PubMed]
  118. Feng, Q.; Li, X.; Sun, W.; Sun, M.; Li, Z.; Sheng, H.; Xie, F.; Zhang, S.; Shan, C. Targeting G6PD reverses paclitaxel resistance in ovarian cancer by suppressing GSTP1. Biochem. Pharmacol. 2020, 178, 114092. [Google Scholar] [CrossRef] [PubMed]
  119. Liu, W.; Zhao, Y.; Wang, G.; Feng, S.; Ge, X.; Ye, W.; Wang, Z.; Zhu, Y.; Cai, W.; Bai, J.; et al. TRIM22 inhibits osteosarcoma progression through destabilizing NRF2 and thus activation of ROS/AMPK/mTOR/autophagy signaling. Redox Biol. 2022, 53, 102344. [Google Scholar] [CrossRef] [PubMed]
  120. Chen, X.; Chen, S.; Yu, D. Metabolic Reprogramming of Chemoresistant Cancer Cells and the Potential Significance of Metabolic Regulation in the Reversal of Cancer Chemoresistance. Metabolites 2020, 10, 289. [Google Scholar] [CrossRef] [PubMed]
  121. Zaal, E.A.; Wu, W.; Jansen, G.; Zweegman, S.; Cloos, J.; Berkers, C.R. Bortezomib resistance in multiple myeloma is associated with increased serine synthesis. Cancer Metab. 2017, 5, 7. [Google Scholar] [CrossRef] [PubMed]
  122. Van Nyen, T.; Planque, M.; van Wagensveld, L.; Duarte, J.A.G.; Zaal, E.A.; Talebi, A.; Rossi, M.; Korner, P.R.; Rizzotto, L.; Moens, S.; et al. Serine metabolism remodeling after platinum-based chemotherapy identifies vulnerabilities in a subgroup of resistant ovarian cancers. Nat. Commun. 2022, 13, 4578. [Google Scholar] [CrossRef] [PubMed]
  123. Pranzini, E.; Pardella, E.; Muccillo, L.; Leo, A.; Nesi, I.; Santi, A.; Parri, M.; Zhang, T.; Uribe, A.H.; Lottini, T.; et al. SHMT2-mediated mitochondrial serine metabolism drives 5-FU resistance by fueling nucleotide biosynthesis. Cell Rep. 2022, 40, 111233. [Google Scholar] [CrossRef] [PubMed]
  124. Zhang, H.; Wang, X.; Ma, Y.; Zhang, Q.; Liu, R.; Luo, H.; Wang, Z. Review of possible mechanisms of radiotherapy resistance in cervical cancer. Front. Oncol. 2023, 13, 1164985. [Google Scholar] [CrossRef] [PubMed]
  125. Zhang, J.; Zhang, Y.; Mo, F.; Patel, G.; Butterworth, K.; Shao, C.; Prise, K.M. The Roles of HIF-1alpha in Radiosensitivity and Radiation-Induced Bystander Effects Under Hypoxia. Front. Cell Dev. Biol. 2021, 9, 637454. [Google Scholar] [CrossRef]
  126. Read, G.H.; Bailleul, J.; Vlashi, E.; Kesarwala, A.H. Metabolic response to radiation therapy in cancer. Mol. Carcinog. 2022, 61, 200–224. [Google Scholar] [CrossRef] [PubMed]
  127. Fu, S.; Li, Z.; Xiao, L.; Hu, W.; Zhang, L.; Xie, B.; Zhou, Q.; He, J.; Qiu, Y.; Wen, M.; et al. Glutamine Synthetase Promotes Radiation Resistance via Facilitating Nucleotide Metabolism and Subsequent DNA Damage Repair. Cell Rep. 2019, 28, 1136–1143.E4. [Google Scholar] [CrossRef] [PubMed]
  128. Falcone, M.; Uribe, A.H.; Papalazarou, V.; Newman, A.C.; Athineos, D.; Stevenson, K.; Sauve, C.G.; Gao, Y.; Kim, J.K.; Del Latto, M.; et al. Sensitisation of cancer cells to radiotherapy by serine and glycine starvation. Br. J. Cancer 2022, 127, 1773–1786. [Google Scholar] [CrossRef] [PubMed]
  129. Solomou, G.; Finch, A.; Asghar, A.; Bardella, C. Mutant IDH in Gliomas: Role in Cancer and Treatment Options. Cancers 2023, 15, 2883. [Google Scholar] [CrossRef] [PubMed]
  130. Ma, R.; Wu, Y.; Li, S.; Yu, X. Interplay Between Glucose Metabolism and Chromatin Modifications in Cancer. Front. Cell Dev. Biol. 2021, 9, 654337. [Google Scholar] [CrossRef] [PubMed]
  131. Chang, M.M.; Chu, D.T.; Lin, S.C.; Lee, J.S.; Vu, T.D.; Vu, H.T.; Ramasamy, T.S.; Lin, S.P.; Wu, C.C. Enhanced mitochondrial function and delivery from adipose-derived stem cell spheres via the EZH2-H3K27me3-PPARgamma pathway for advanced therapy. Stem Cell Res. Ther. 2025, 16, 129. [Google Scholar] [CrossRef] [PubMed]
  132. Wu, J.; Jing, X.; Du, Q.; Sun, X.; Holgersson, K.; Gao, J.; He, X.; Hosaka, K.; Zhao, C.; Tao, W.; et al. Disruption of the Clock Component Bmal1 in Mice Promotes Cancer Metastasis through the PAI-1-TGF-beta-myoCAF-Dependent Mechanism. Adv. Sci. 2023, 10, e2301505. [Google Scholar] [CrossRef] [PubMed]
  133. Zhu, Y.; Li, X.; Wang, L.; Hong, X.; Yang, J. Metabolic reprogramming and crosstalk of cancer-related fibroblasts and immune cells in the tumor microenvironment. Front. Endocrinol. 2022, 13, 988295. [Google Scholar] [CrossRef] [PubMed]
  134. Kobayashi, M.; Morinibu, A.; Koyasu, S.; Goto, Y.; Hiraoka, M.; Harada, H. A circadian clock gene, PER2, activates HIF-1 as an effector molecule for recruitment of HIF-1alpha to promoter regions of its downstream genes. FEBS J. 2017, 284, 3804–3816. [Google Scholar] [CrossRef] [PubMed]
  135. Altman, B.J.; Hsieh, A.L.; Sengupta, A.; Krishnanaiah, S.Y.; Stine, Z.E.; Walton, Z.E.; Gouw, A.M.; Venkataraman, A.; Li, B.; Goraksha-Hicks, P.; et al. MYC Disrupts the Circadian Clock and Metabolism in Cancer Cells. Cell Metab. 2015, 22, 1009–1019. [Google Scholar] [CrossRef] [PubMed]
  136. Cazarin, J.; DeRollo, R.E.; Shahidan, S.; Burchett, J.B.; Mwangi, D.; Krishnaiah, S.; Hsieh, A.L.; Walton, Z.E.; Brooks, R.; Mello, S.S.; et al. MYC disrupts transcriptional and metabolic circadian oscillations in cancer and promotes enhanced biosynthesis. PLoS Genet. 2023, 19, e1010904. [Google Scholar] [CrossRef] [PubMed]
  137. Papagiannakopoulos, T.; Bauer, M.R.; Davidson, S.M.; Heimann, M.; Subbaraj, L.; Bhutkar, A.; Bartlebaugh, J.; Vander Heiden, M.G.; Jacks, T. Circadian Rhythm Disruption Promotes Lung Tumorigenesis. Cell Metab. 2016, 24, 324–331. [Google Scholar] [CrossRef] [PubMed]
  138. Liu, T.; Wang, Z.; Ye, L.; Duan, Y.; Jiang, H.; He, H.; Xiao, L.; Wu, Q.; Xia, Y.; Yang, M.; et al. Nucleus-exported CLOCK acetylates PRPS to promote de novo nucleotide synthesis and liver tumour growth. Nat. Cell Biol. 2023, 25, 273–284. [Google Scholar] [CrossRef] [PubMed]
  139. Gong, X.; Tang, H.; Yang, K. PER1 suppresses glycolysis and cell proliferation in oral squamous cell carcinoma via the PER1/RACK1/PI3K signaling complex. Cell Death Dis. 2021, 12, 276. [Google Scholar] [CrossRef] [PubMed]
  140. Gowda, P.; Lathoria, K.; Sharma, S.; Patrick, S.; Umdor, S.B.; Sen, E. Rewiring of Lactate-Interleukin-1beta Autoregulatory Loop with Clock-Bmal1: A Feed-Forward Circuit in Glioma. Mol. Cell Biol. 2021, 41, e0044920. [Google Scholar] [CrossRef] [PubMed]
  141. Yuan, P.; Yang, T.; Mu, J.; Zhao, J.; Yang, Y.; Yan, Z.; Hou, Y.; Chen, C.; Xing, J.; Zhang, H.; et al. Circadian clock gene NPAS2 promotes reprogramming of glucose metabolism in hepatocellular carcinoma cells. Cancer Lett. 2020, 469, 498–509. [Google Scholar] [CrossRef] [PubMed]
  142. Hojo, H.; Enya, S.; Arai, M.; Suzuki, Y.; Nojiri, T.; Kangawa, K.; Koyama, S.; Kawaoka, S. Remote reprogramming of hepatic circadian transcriptome by breast cancer. Oncotarget 2017, 8, 34128–34140. [Google Scholar] [CrossRef] [PubMed]
  143. Masri, S.; Papagiannakopoulos, T.; Kinouchi, K.; Liu, Y.; Cervantes, M.; Baldi, P.; Jacks, T.; Sassone-Corsi, P. Lung Adenocarcinoma Distally Rewires Hepatic Circadian Homeostasis. Cell 2016, 165, 896–909. [Google Scholar] [CrossRef] [PubMed]
  144. Ghosh, R.; Samanta, P.; Sarkar, R.; Biswas, S.; Saha, P.; Hajra, S.; Bhowmik, A. Targeting HIF-1alpha by Natural and Synthetic Compounds: A Promising Approach for Anti-Cancer Therapeutics Development. Molecules 2022, 27, 5192. [Google Scholar] [CrossRef] [PubMed]
  145. Li, H.; Liu, Y.; Xue, Z.; Zhang, L.; Ruan, X.; Yang, J.; Fan, Z.; Zhao, H.; Cao, Y.; Chen, G.; et al. Adamantaniline Derivatives Target ATP5B to Inhibit Translation of Hypoxia Inducible Factor-1alpha. Adv. Sci. 2023, 10, e2301071. [Google Scholar] [CrossRef] [PubMed]
  146. Han, H.J.; Sivaraman, A.; Kim, M.; Min, K.H.; Song, M.E.; Choi, Y.; Choi, W.J.; Han, H.K.; Han, J.; Jang, J.P.; et al. HIF-1alpha inhibition by MO-2097, a novel chiral-free benzofuran targeting hnRNPA2B1. J. Adv. Res. 2024, 64, 67–81. [Google Scholar] [CrossRef] [PubMed]
  147. Schmidt, K.T.; Chau, C.H.; Strope, J.D.; Huitema, A.D.R.; Sissung, T.M.; Price, D.K.; Figg, W.D. Antitumor Activity of NLG207 (Formerly CRLX101) in Combination with Enzalutamide in Preclinical Prostate Cancer Models. Mol. Cancer Ther. 2021, 20, 915–924. [Google Scholar] [CrossRef] [PubMed]
  148. Hwang, S.J.; Jung, Y.; Song, Y.S.; Park, S.; Park, Y.; Lee, H.J. Enhanced anti-angiogenic activity of novel melatonin-like agents. J. Pineal Res. 2021, 71, e12739. [Google Scholar] [CrossRef] [PubMed]
  149. He, K.; Tao, F.; Lu, Y.; Fang, M.; Huang, H.; Zhou, Y. The Role of HK2 in Tumorigenesis and Development: Potential for Targeted Therapy with Natural Products. Int. J. Med. Sci. 2025, 22, 790–805. [Google Scholar] [CrossRef] [PubMed]
  150. Han, J.H.; Lee, E.J.; Park, W.; Ha, K.T.; Chung, H.S. Natural compounds as lactate dehydrogenase inhibitors: Potential therapeutics for lactate dehydrogenase inhibitors-related diseases. Front. Pharmacol. 2023, 14, 1275000. [Google Scholar] [CrossRef] [PubMed]
  151. 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]
  152. Miller, Z.A.; Muthuswami, S.; Mueller, A.; Ma, R.Z.; Sywanycz, S.M.; Naik, A.; Huang, L.; Brody, R.M.; Diab, A.; Carey, R.M.; et al. GLUT1 inhibitor BAY-876 induces apoptosis and enhances anti-cancer effects of bitter receptor agonists in head and neck squamous carcinoma cells. Cell Death Discov. 2024, 10, 339. [Google Scholar] [CrossRef] [PubMed]
  153. Yang, H.; Zhang, M.Z.; Sun, H.W.; Chai, Y.T.; Li, X.; Jiang, Q.; Hou, J. A Novel Microcrystalline BAY-876 Formulation Achieves Long-Acting Antitumor Activity Against Aerobic Glycolysis and Proliferation of Hepatocellular Carcinoma. Front. Oncol. 2021, 11, 783194. [Google Scholar] [CrossRef] [PubMed]
  154. Liu, Y.; Cao, Y.; Zhang, W.; Bergmeier, S.; Qian, Y.; Akbar, H.; Colvin, R.; Ding, J.; Tong, L.; Wu, S.; et al. A small-molecule inhibitor of glucose transporter 1 downregulates glycolysis, induces cell-cycle arrest, and inhibits cancer cell growth in vitro and in vivo. Mol. Cancer Ther. 2012, 11, 1672–1682. [Google Scholar] [CrossRef] [PubMed]
  155. Noble, R.A.; Bell, N.; Blair, H.; Sikka, A.; Thomas, H.; Phillips, N.; Nakjang, S.; Miwa, S.; Crossland, R.; Rand, V.; et al. Inhibition of monocarboxyate transporter 1 by AZD3965 as a novel therapeutic approach for diffuse large B-cell lymphoma and Burkitt lymphoma. Haematologica 2017, 102, 1247–1257. [Google Scholar] [CrossRef] [PubMed]
  156. Guan, X.; Bryniarski, M.A.; Morris, M.E. In Vitro and In Vivo Efficacy of the Monocarboxylate Transporter 1 Inhibitor AR-C155858 in the Murine 4T1 Breast Cancer Tumor Model. AAPS J. 2018, 21, 3. [Google Scholar] [CrossRef] [PubMed]
  157. Benjamin, D.; Robay, D.; Hindupur, S.K.; Pohlmann, J.; Colombi, M.; El-Shemerly, M.Y.; Maira, S.M.; Moroni, C.; Lane, H.A.; Hall, M.N. Dual Inhibition of the Lactate Transporters MCT1 and MCT4 Is Synthetic Lethal with Metformin due to NAD+ Depletion in Cancer Cells. Cell Rep. 2018, 25, 3047–3058.e4. [Google Scholar] [CrossRef] [PubMed]
  158. Esquea, E.M.; Ciraku, L.; Young, R.G.; Merzy, J.; Talarico, A.N.; Ahmed, N.N.; Karuppiah, M.; Ramesh, A.; Chatoff, A.; Crispim, C.V.; et al. Selective and brain-penetrant ACSS2 inhibitors target breast cancer brain metastatic cells. Front. Pharmacol. 2024, 15, 1394685. [Google Scholar] [CrossRef] [PubMed]
  159. Sharabi, K.; Lin, H.; Tavares, C.D.J.; Dominy, J.E.; Camporez, J.P.; Perry, R.J.; Schilling, R.; Rines, A.K.; Lee, J.; Hickey, M.; et al. Selective Chemical Inhibition of PGC-1alpha Gluconeogenic Activity Ameliorates Type 2 Diabetes. Cell 2017, 169, 148–160.E15. [Google Scholar] [CrossRef] [PubMed]
  160. Xiang, Y.; Fang, B.; Liu, Y.; Yan, S.; Cao, D.; Mei, H.; Wang, Q.; Hu, Y.; Guo, T. SR18292 exerts potent antitumor effects in multiple myeloma via inhibition of oxidative phosphorylation. Life Sci. 2020, 256, 117971. [Google Scholar] [CrossRef] [PubMed]
  161. Szeliga, M.; Rola, R. Conoidin A, a Covalent Inhibitor of Peroxiredoxin 2, Reduces Growth of Glioblastoma Cells by Triggering ROS Production. Cells 2023, 12, 1934. [Google Scholar] [CrossRef] [PubMed]
  162. Udumula, M.P.; Rashid, F.; Singh, H.; Pardee, T.; Luther, S.; Bhardwaj, T.; Anjaly, K.; Piloni, S.; Hijaz, M.; Gogoi, R.; et al. Targeting mitochondrial metabolism with CPI-613 in chemoresistant ovarian tumors. J. Ovarian Res. 2024, 17, 226. [Google Scholar] [CrossRef] [PubMed]
  163. Arlt, B.; Mastrobuoni, G.; Wuenschel, J.; Astrahantseff, K.; Eggert, A.; Kempa, S.; Deubzer, H.E. Inhibiting PHGDH with NCT-503 reroutes glucose-derived carbons into the TCA cycle, independently of its on-target effect. J. Enzyme Inhib. Med. Chem. 2021, 36, 1282–1289. [Google Scholar] [CrossRef] [PubMed]
  164. Leung, P.; Pickarski, M.; Zhuo, Y.; Masarachia, P.J.; Duong, L.T. The effects of the cathepsin K inhibitor odanacatib on osteoclastic bone resorption and vesicular trafficking. Bone 2011, 49, 623–635. [Google Scholar] [CrossRef] [PubMed]
  165. Ahmadi, S.E.; Rahimi, S.; Zarandi, B.; Chegeni, R.; Safa, M. MYC: A multipurpose oncogene with prognostic and therapeutic implications in blood malignancies. J. Hematol. Oncol. 2021, 14, 121. [Google Scholar] [CrossRef] [PubMed]
  166. Boike, L.; Cioffi, A.G.; Majewski, F.C.; Co, J.; Henning, N.J.; Jones, M.D.; Liu, G.; McKenna, J.M.; Tallarico, J.A.; Schirle, M.; et al. Discovery of a Functional Covalent Ligand Targeting an Intrinsically Disordered Cysteine within MYC. Cell Chem. Biol. 2021, 28, 4–13e17. [Google Scholar] [CrossRef] [PubMed]
  167. Perkins, S.N.; Hursting, S.D.; Haines, D.C.; James, S.J.; Miller, B.J.; Phang, J.M. Chemoprevention of spontaneous tumorigenesis in nullizygous p53-deficient mice by dehydroepiandrosterone and its analog 16alpha-fluoro-5-androsten-17-one. Carcinogenesis 1997, 18, 989–994. [Google Scholar] [CrossRef] [PubMed]
  168. Schwartz, A.G. Dehydroepiandrosterone, Cancer, and Aging. Aging Dis. 2022, 13, 423–432. [Google Scholar] [CrossRef] [PubMed]
  169. Li, Y.; Zheng, F.; Zhang, Y.; Lin, Z.; Yang, J.; Han, X.; Feng, Y.; Pei, X.; Li, F.; Liu, Q.; et al. Targeting glucose-6-phosphate dehydrogenase by 6-AN induces ROS-mediated autophagic cell death in breast cancer. FEBS J. 2023, 290, 763–779. [Google Scholar] [CrossRef] [PubMed]
  170. Zemnou, C.T. In silico Comparative Study of the Anti-Cancer Potential of Inhibitors of Glucose-6-Phosphate Dehydrogenase Enzyme Using ADMET Analysis, Molecular Docking, and Molecular Dynamic Simulation. Adv. Theor. Simul. 2025, 8, 2400757. [Google Scholar] [CrossRef]
  171. Gupta, T.; Mondal, A.K.; Pani, I.; Chattopadhyay, K.; Pal, S.K. Elucidating liquid crystal-aqueous interface for the study of cholesterol-mediated action of a beta-barrel pore forming toxin. Soft Matter 2022, 18, 5293–5301. [Google Scholar] [CrossRef] [PubMed]
  172. Zheng, X.; Huang, C.H.; Lin, Y.Q.; Han, B.F.; Chen, Y.Z.; Li, C.J.; Li, J.W.; Ding, Y.Y.; Song, X.C.; Wang, W.; et al. Generation of inactivated IL2RG and RAG1 monkeys with severe combined immunodeficiency using base editing. Signal Transduct. Tar. 2023, 8, 327. [Google Scholar] [CrossRef] [PubMed]
  173. Rashmi, R.; Huang, X.; Floberg, J.M.; Elhammali, A.E.; McCormick, M.L.; Patti, G.J.; Spitz, D.R.; Schwarz, J.K. Radioresistant Cervical Cancers Are Sensitive to Inhibition of Glycolysis and Redox Metabolism. Cancer Res. 2018, 78, 1392–1403. [Google Scholar] [CrossRef] [PubMed]
  174. Zhao, Y.; Liu, H.; Liu, Z.; Ding, Y.; Ledoux, S.P.; Wilson, G.L.; Voellmy, R.; Lin, Y.; Lin, W.; Nahta, R.; et al. Overcoming trastuzumab resistance in breast cancer by targeting dysregulated glucose metabolism. Cancer Res. 2011, 71, 4585–4597. [Google Scholar] [CrossRef] [PubMed]
  175. Fiorillo, M.; Sotgia, F.; Sisci, D.; Cappello, A.R.; Lisanti, M.P. Mitochondrial “power” drives tamoxifen resistance: NQO1 and GCLC are new therapeutic targets in breast cancer. Oncotarget 2017, 8, 20309–20327. [Google Scholar] [CrossRef] [PubMed]
  176. Thompson, R.M.; Dytfeld, D.; Reyes, L.; Robinson, R.M.; Smith, B.; Manevich, Y.; Jakubowiak, A.; Komarnicki, M.; Przybylowicz-Chalecka, A.; Szczepaniak, T.; et al. Glutaminase inhibitor CB-839 synergizes with carfilzomib in resistant multiple myeloma cells. Oncotarget 2017, 8, 35863–35876. [Google Scholar] [CrossRef] [PubMed]
  177. Tao, L.; Yu, H.; Liang, R.; Jia, R.; Wang, J.; Jiang, K.; Wang, Z. Rev-erbalpha inhibits proliferation by reducing glycolytic flux and pentose phosphate pathway in human gastric cancer cells. Oncogenesis 2019, 8, 57. [Google Scholar] [CrossRef] [PubMed]
  178. Chu, G.; Zhou, X.; Hu, Y.; Shi, S.; Yang, G. Rev-erbalpha Inhibits Proliferation and Promotes Apoptosis of Preadipocytes through the Agonist GSK4112. Int. J. Mol. Sci. 2019, 20, 4524. [Google Scholar] [CrossRef] [PubMed]
  179. D’Amore, C.; Borgo, C.; Sarno, S.; Salvi, M. Role of CK2 inhibitor CX-4945 in anti-cancer combination therapy—Potential clinical relevance. Cell Oncol. 2020, 43, 1003–1016. [Google Scholar] [CrossRef] [PubMed]
  180. Cheng, J.; Zhang, R.; Xu, Z.; Ke, Y.; Sun, R.; Yang, H.; Zhang, X.; Zhen, X.; Zheng, L.T. Early glycolytic reprogramming controls microglial inflammatory activation. J. Neuroinflamm. 2021, 18, 129. [Google Scholar] [CrossRef] [PubMed]
  181. Cal, M.; Matyjaszczyk, I.; Litwin, I.; Augustyniak, D.; Ogorek, R.; Ko, Y.; Ulaszewski, S. The Anticancer Drug 3-Bromopyruvate Induces DNA Damage Potentially Through Reactive Oxygen Species in Yeast and in Human Cancer Cells. Cells 2020, 9, 1161. [Google Scholar] [CrossRef] [PubMed]
  182. Park, J.H.; Kundu, A.; Lee, S.H.; Jiang, C.; Lee, S.H.; Kim, Y.S.; Kyung, S.Y.; Park, S.H.; Kim, H.S. Specific Pyruvate Kinase M2 Inhibitor, Compound 3K, Induces Autophagic Cell Death through Disruption of the Glycolysis Pathway in Ovarian Cancer Cells. Int. J. Biol. Sci. 2021, 17, 1895–1908. [Google Scholar] [CrossRef] [PubMed]
  183. Le, A.; Cooper, C.R.; Gouw, A.M.; Dinavahi, R.; Maitra, A.; Deck, L.M.; Royer, R.E.; Vander Jagt, D.L.; Semenza, G.L.; Dang, C.V. Inhibition of lactate dehydrogenase A induces oxidative stress and inhibits tumor progression. Proc. Natl. Acad. Sci. USA 2010, 107, 2037–2042. [Google Scholar] [CrossRef] [PubMed]
  184. Zhao, S.S.; Liu, J.; Wu, Q.C.; Zhou, X.L. Lactate regulates pathological cardiac hypertrophy via histone lactylation modification. J. Cell Mol. Med. 2024, 28, e70022. [Google Scholar] [CrossRef] [PubMed]
  185. Forteza, M.J.; Berg, M.; Edsfeldt, A.; Sun, J.; Baumgartner, R.; Kareinen, I.; Casagrande, F.B.; Hedin, U.; Zhang, S.; Vuckovic, I.; et al. Pyruvate dehydrogenase kinase regulates vascular inflammation in atherosclerosis and increases cardiovascular risk. Cardiovasc. Res. 2023, 119, 1524–1536. [Google Scholar] [CrossRef] [PubMed]
  186. Tambe, Y.; Terado, T.; Kim, C.J.; Mukaisho, K.I.; Yoshida, S.; Sugihara, H.; Tanaka, H.; Chida, J.; Kido, H.; Yamaji, K.; et al. Antitumor activity of potent pyruvate dehydrogenase kinase 4 inhibitors from plants in pancreatic cancer. Mol. Carcinog. 2019, 58, 1726–1737. [Google Scholar] [CrossRef] [PubMed]
  187. Bodeker, K.L.; Smith, B.J.; Berg, D.J.; Chandrasekharan, C.; Sharif, S.; Fei, N.; Vollstedt, S.; Brown, H.; Chandler, M.; Lorack, A.; et al. A randomized trial of pharmacological ascorbate, gemcitabine, and nab-paclitaxel for metastatic pancreatic cancer. Redox Biol. 2024, 77, 103375. [Google Scholar] [CrossRef] [PubMed]
  188. Chang, Y.C.; Chiou, J.; Yang, Y.F.; Su, C.Y.; Lin, Y.F.; Yang, C.N.; Lu, P.J.; Huang, M.S.; Yang, C.J.; Hsiao, M. Therapeutic Targeting of Aldolase A Interactions Inhibits Lung Cancer Metastasis and Prolongs Survival. Cancer Res. 2019, 79, 4754–4766. [Google Scholar] [CrossRef] [PubMed]
  189. Curcio, C.; Mucciolo, G.; Roux, C.; Brugiapaglia, S.; Scagliotti, A.; Guadagnin, G.; Conti, L.; Longo, D.; Grosso, D.; Papotti, M.G.; et al. PI3Kgamma inhibition combined with DNA vaccination unleashes a B-cell-dependent antitumor immunity that hampers pancreatic cancer. J. Exp. Clin. Cancer Res. 2024, 43, 157. [Google Scholar] [CrossRef] [PubMed]
  190. Cappello, P.; Rolla, S.; Chiarle, R.; Principe, M.; Cavallo, F.; Perconti, G.; Feo, S.; Giovarelli, M.; Novelli, F. Vaccination with ENO1 DNA prolongs survival of genetically engineered mice with pancreatic cancer. Gastroenterology 2013, 144, 1098–1106. [Google Scholar] [CrossRef] [PubMed]
  191. Lin, X.; Xu, J.; Ma, C.; Wang, Z.; Yang, X.; Song, X.; Zheng, H. Metformin and 2-Deoxy-d-glucose synergistically inhibit the viability of thyroid papillary carcinoma with BRAF mutation by ROS-dependent PI3K/AKT signaling pathway. Anticancer. Agents Med. Chem. 2023, 12, 357. [Google Scholar] [CrossRef]
  192. Qin, H.; Zheng, G.; Li, Q.; Shen, L. Metabolic reprogramming induced by DCA enhances cisplatin sensitivity through increasing mitochondrial oxidative stress in cholangiocarcinoma. Front. Pharmacol. 2023, 14, 1128312. [Google Scholar] [CrossRef] [PubMed]
  193. LaMoia, T.E.; Shulman, G.I. Cellular and Molecular Mechanisms of Metformin Action. Endocr. Rev. 2021, 42, 77–96. [Google Scholar] [CrossRef] [PubMed]
  194. Wei, H.; Li, Z.; Qian, K.; Du, W.; Ju, L.; Shan, D.; Yu, M.; Fang, Y.; Zhang, Y.; Xiao, Y.; et al. Unveiling the association between HMG-CoA reductase inhibitors and bladder cancer: A comprehensive analysis using Mendelian randomization, animal models, and transcriptomics. Pharmacogenomics J. 2024, 24, 24. [Google Scholar] [CrossRef] [PubMed]
  195. Pouyssegur, J.; Marchiq, I.; Parks, S.K.; Durivault, J.; Zdralevic, M.; Vucetic, M. ‘Warburg effect’ controls tumor growth, bacterial, viral infections and immunity—Genetic deconstruction and therapeutic perspectives. Semin. Cancer Biol. 2022, 86, 334–346. [Google Scholar] [CrossRef] [PubMed]
  196. Silva, A.; Antunes, B.; Batista, A.; Pinto-Ribeiro, F.; Baltazar, F.; Afonso, J. In Vivo Anticancer Activity of AZD3965: A Systematic Review. Molecules 2021, 27, 181. [Google Scholar] [CrossRef] [PubMed]
  197. Sasaki, S.; Futagi, Y.; Ideno, M.; Kobayashi, M.; Narumi, K.; Furugen, A.; Iseki, K. Effect of diclofenac on SLC16A3/MCT4 by the Caco-2 cell line. Drug Metab. Pharmacokinet. 2016, 31, 218–223. [Google Scholar] [CrossRef] [PubMed]
  198. Vander Linden, C.; Corbet, C.; Bastien, E.; Martherus, R.; Guilbaud, C.; Petit, L.; Wauthier, L.; Loriot, A.; De Smet, C.; Feron, O. Therapy-induced DNA methylation inactivates MCT1 and renders tumor cells vulnerable to MCT4 inhibition. Cell Rep. 2021, 35, 109202. [Google Scholar] [CrossRef] [PubMed]
  199. Yadav, D.; Yadav, A.; Bhattacharya, S.; Dagar, A.; Kumar, V.; Rani, R. GLUT and HK: Two primary and essential key players in tumor glycolysis. Semin. Cancer Biol. 2024, 100, 17–27. [Google Scholar] [CrossRef] [PubMed]
  200. Reckzeh, E.S.; Karageorgis, G.; Schwalfenberg, M.; Ceballos, J.; Nowacki, J.; Stroet, M.C.M.; Binici, A.; Knauer, L.; Brand, S.; Choidas, A.; et al. Inhibition of Glucose Transporters and Glutaminase Synergistically Impairs Tumor Cell Growth. Cell Chem. Biol. 2019, 26, 1214–1228.E25. [Google Scholar] [CrossRef] [PubMed]
  201. 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] [PubMed]
  202. Billiard, J.; Dennison, J.B.; Briand, J.; Annan, R.S.; Chai, D.; Colon, M.; Dodson, C.S.; Gilbert, S.A.; Greshock, J.; Jing, J.; et al. Quinoline 3-sulfonamides inhibit lactate dehydrogenase A and reverse aerobic glycolysis in cancer cells. Cancer Metab. 2013, 1, 19. [Google Scholar] [CrossRef] [PubMed]
  203. Mohammad Nezhady, M.A.; Chemtob, S. 3-OBA Is Not an Antagonist of GPR81. Front. Pharmacol. 2021, 12, 803907. [Google Scholar] [CrossRef] [PubMed]
  204. Sun, Y.; Mu, G.; Xue, Z.; Wang, S.; Li, X.; Han, M. Targeting cholesterol homeostasis through inhibiting SREBP-2: An Achilles’ heel for glioblastoma. Neuro Oncol. 2023, 25, 2100–2102. [Google Scholar] [CrossRef] [PubMed]
  205. Alzahrani, A.S. PI3K/Akt/mTOR inhibitors in cancer: At the bench and bedside. Semin. Cancer Biol. 2019, 59, 125–132. [Google Scholar] [CrossRef] [PubMed]
  206. Liu, X.; Guo, B.; Li, Q.; Nie, J. mTOR in metabolic homeostasis and disease. Exp. Cell Res. 2024, 441, 114173. [Google Scholar] [CrossRef] [PubMed]
  207. Xie, Z.; Wang, J.; Liu, M.; Chen, D.; Qiu, C.; Sun, K. CC-223 blocks mTORC1/C2 activation and inhibits human hepatocellular carcinoma cells in vitro and in vivo. PLoS ONE 2017, 12, e0173252. [Google Scholar] [CrossRef] [PubMed]
  208. Li, D.; Yi, J.; Han, G.; Qiao, L. MALDI-TOF Mass Spectrometry in Clinical Analysis and Research. ACS Meas. Sci. Au 2022, 2, 385–404. [Google Scholar] [CrossRef] [PubMed]
  209. Venkateswaran, S.V.; Kreuzaler, P.; Maclachlan, C.; McMahon, G.; Greenidge, G.; Collinson, L.; Bunch, J.; Yuneva, M. A multimodal imaging pipeline to decipher cell-specific metabolic functions and tissue microenvironment dynamics. Nat. Protoc. 2025, 20, 1678–1699. [Google Scholar] [CrossRef] [PubMed]
  210. Andersen, M.K.; Hoiem, T.S.; Claes, B.S.R.; Balluff, B.; Martin-Lorenzo, M.; Richardsen, E.; Krossa, S.; Bertilsson, H.; Heeren, R.M.A.; Rye, M.B.; et al. Spatial differentiation of metabolism in prostate cancer tissue by MALDI-TOF MSI. Cancer Metab. 2021, 9, 9. [Google Scholar] [CrossRef] [PubMed]
  211. Krestensen, K.K.; Hendriks, T.F.E.; Grgic, A.; Derweduwe, M.; De Smet, F.; Heeren, R.M.A.; Cuypers, E. Molecular Profiling of Glioblastoma Patient-Derived Single Cells Using Combined MALDI-MSI and MALDI-IHC. Anal. Chem. 2025, 97, 3846–3854. [Google Scholar] [CrossRef] [PubMed]
  212. Hao, X.; Gu, H.; Chen, C.; Huang, D.; Zhao, Y.; Xie, L.; Zou, Y.; Shu, H.S.; Zhang, Y.; He, X.; et al. Metabolic Imaging Reveals a Unique Preference of Symmetric Cell Division and Homing of Leukemia-Initiating Cells in an Endosteal Niche. Cell Metab. 2019, 29, 950–965.E6. [Google Scholar] [CrossRef] [PubMed]
  213. Diaz-Garcia, C.M.; Mongeon, R.; Lahmann, C.; Koveal, D.; Zucker, H.; Yellen, G. Neuronal Stimulation Triggers Neuronal Glycolysis and Not Lactate Uptake. Cell Metab. 2017, 26, 361–374e364. [Google Scholar] [CrossRef] [PubMed]
  214. Fan, H.; Demirci, U.; Chen, P. Emerging organoid models: Leaping forward in cancer research. J. Hematol. Oncol. 2019, 12, 142. [Google Scholar] [CrossRef] [PubMed]
  215. Ko, J.; Ahn, J.; Kim, S.; Lee, Y.; Lee, J.; Park, D.; Jeon, N.L. Tumor spheroid-on-a-chip: A standardized microfluidic culture platform for investigating tumor angiogenesis. Lab. Chip 2019, 19, 2822–2833. [Google Scholar] [CrossRef] [PubMed]
  216. Ayuso, J.M.; Rehman, S.; Virumbrales-Munoz, M.; McMinn, P.H.; Geiger, P.; Fitzgerald, C.; Heaster, T.; Skala, M.C.; Beebe, D.J. Microfluidic tumor-on-a-chip model to evaluate the role of tumor environmental stress on NK cell exhaustion. Sci. Adv. 2021, 7, eabc2331. [Google Scholar] [CrossRef] [PubMed]
  217. Vlachogiannis, G.; Hedayat, S.; Vatsiou, A.; Jamin, Y.; Fernandez-Mateos, J.; Khan, K.; Lampis, A.; Eason, K.; Huntingford, I.; Burke, R.; et al. Patient-derived organoids model treatment response of metastatic gastrointestinal cancers. Science 2018, 359, 920–926. [Google Scholar] [CrossRef] [PubMed]
  218. Loong, H.H.; Wong, A.M.; Chan, D.T.; Cheung, M.S.; Chow, C.; Ding, X.; Chan, A.K.; Johnston, P.A.; Lau, J.Y.; Poon, W.S.; et al. Patient-derived tumor organoid predicts drugs response in glioblastoma: A step forward in personalized cancer therapy? J. Clin. Neurosci. 2020, 78, 400–402. [Google Scholar] [CrossRef] [PubMed]
  219. Wang, Y.; Wang, D.; Zhang, Y.; Xu, H.; Shen, L.; Cheng, J.; Xu, X.; Tan, H.; Chen, X.; Li, J. Tumor Microenvironment-Adaptive Nanoplatform Synergistically Enhances Cascaded Chemodynamic Therapy. Bioact. Mater. 2023, 22, 239–253. [Google Scholar] [CrossRef] [PubMed]
  220. Dang, J.; Ye, H.; Li, Y.; Liang, Q.; Li, X.; Yin, L. Multivalency-assisted membrane-penetrating siRNA delivery sensitizes photothermal ablation via inhibition of tumor glycolysis metabolism. Biomaterials 2019, 223, 119463. [Google Scholar] [CrossRef] [PubMed]
  221. Zhang, Y.H.; Qiu, W.X.; Zhang, M.; Zhang, L.; Zhang, X.Z. MnO2 Motor: A Prospective Cancer-Starving Therapy Promoter. ACS Appl. Mater. Interfaces 2018, 10, 15030–15039. [Google Scholar] [CrossRef] [PubMed]
  222. Chen, W.H.; Luo, G.F.; Lei, Q.; Hong, S.; Qiu, W.X.; Liu, L.H.; Cheng, S.X.; Zhang, X.Z. Overcoming the Heat Endurance of Tumor Cells by Interfering with the Anaerobic Glycolysis Metabolism for Improved Photothermal Therapy. ACS Nano 2017, 11, 1419–1431. [Google Scholar] [CrossRef] [PubMed]
  223. Wang, X.; Wang, L.; Hao, Q.; Cai, M.; Wang, X.; An, W. Harnessing glucose metabolism with nanomedicine for cancer treatment. Theranostics 2024, 14, 6831–6882. [Google Scholar] [CrossRef] [PubMed]
  224. Choudhury, S.; Moret, M.; Salvy, P.; Weilandt, D.; Hatzimanikatis, V.; Miskovic, L. Reconstructing Kinetic Models for Dynamical Studies of Metabolism using Generative Adversarial Networks. Nat. Mach. Intell. 2022, 4, 710–719. [Google Scholar] [CrossRef] [PubMed]
  225. Antoniewicz, M.R.; Stephanopoulos, G.; Kelleher, J.K. Evaluation of regression models in metabolic physiology: Predicting fluxes from isotopic data without knowledge of the pathway. Metabolomics 2006, 2, 41–52. [Google Scholar] [CrossRef] [PubMed]
  226. Pun, F.W.; Ozerov, I.V.; Zhavoronkov, A. AI-powered therapeutic target discovery. Trends Pharmacol. Sci. 2023, 44, 561–572. [Google Scholar] [CrossRef] [PubMed]
  227. Azad, R.K.; Shulaev, V. Metabolomics technology and bioinformatics for precision medicine. Brief. Bioinform. 2019, 20, 1957–1971. [Google Scholar] [CrossRef] [PubMed]
  228. Liang, L.; Rasmussen, M.H.; Piening, B.; Shen, X.; Chen, S.; Rost, H.; Snyder, J.K.; Tibshirani, R.; Skotte, L.; Lee, N.C.; et al. Metabolic Dynamics and Prediction of Gestational Age and Time to Delivery in Pregnant Women. Cell 2020, 181, 1680–1692.E15. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Spatiotemporal heterogeneity of tumor glucose metabolism. This schematic illustrates the dynamic interplay of spatial and temporal heterogeneity in tumor glucose metabolism, driven by microenvironmental gradients (e.g., oxygen, pH, metabolite). Spatially, tumors exhibit metabolic zonation: a hypoxic core characterized by severe oxygen deprivation, decreased pH, HIF-1α stabilization, and lactate accumulation, predominantly reliant on aerobic glycolysis (the Warburg effect); contrasted with an oxygen-rich rim characterized by adequate oxygen supply and relatively higher pH compared to the core, dependent on OXPHOS for efficient energy production. The metabolic symbiosis between these regions is mediated by lactate shuttling via monocarboxylate transporters (MCT1/MCT4), promoting tumor survival and invasion. Additionally, tumor-immune metabolic crosstalk further modulates the microenvironment. Temporally, heterogeneity arises from distinct tumor progression stages (e.g., glycolysis-driven early growth vs. mitochondrial reactivation in metastasis), therapy-induced metabolic plasticity, and intrinsic metabolic rhythms. Tumor spatiotemporal adaptability facilitates therapeutic evasion by dynamically reshaping the metabolic landscape, contributing to treatment resistance and relapse. Elucidating these mechanisms is crucial for advancing precision oncology strategies to counteract compensatory adaptations and intratumoral heterogeneity, such as spatially targeted metabolic inhibitors or rational combination therapies. HIF-1α, hypoxia-inducible factor 1 alpha; OXPHOS, oxidative phosphorylation; MCT1/MCT4, monocarboxylate transporter 1/4. Graphical Elements from BioGDP (https://biogdp.com) [1].
Figure 1. Spatiotemporal heterogeneity of tumor glucose metabolism. This schematic illustrates the dynamic interplay of spatial and temporal heterogeneity in tumor glucose metabolism, driven by microenvironmental gradients (e.g., oxygen, pH, metabolite). Spatially, tumors exhibit metabolic zonation: a hypoxic core characterized by severe oxygen deprivation, decreased pH, HIF-1α stabilization, and lactate accumulation, predominantly reliant on aerobic glycolysis (the Warburg effect); contrasted with an oxygen-rich rim characterized by adequate oxygen supply and relatively higher pH compared to the core, dependent on OXPHOS for efficient energy production. The metabolic symbiosis between these regions is mediated by lactate shuttling via monocarboxylate transporters (MCT1/MCT4), promoting tumor survival and invasion. Additionally, tumor-immune metabolic crosstalk further modulates the microenvironment. Temporally, heterogeneity arises from distinct tumor progression stages (e.g., glycolysis-driven early growth vs. mitochondrial reactivation in metastasis), therapy-induced metabolic plasticity, and intrinsic metabolic rhythms. Tumor spatiotemporal adaptability facilitates therapeutic evasion by dynamically reshaping the metabolic landscape, contributing to treatment resistance and relapse. Elucidating these mechanisms is crucial for advancing precision oncology strategies to counteract compensatory adaptations and intratumoral heterogeneity, such as spatially targeted metabolic inhibitors or rational combination therapies. HIF-1α, hypoxia-inducible factor 1 alpha; OXPHOS, oxidative phosphorylation; MCT1/MCT4, monocarboxylate transporter 1/4. Graphical Elements from BioGDP (https://biogdp.com) [1].
Ijms 26 06901 g001
Figure 2. Metabolic reprogramming across tumor progression and therapeutic intervention. This schematic illustrates dynamic metabolic adaptations during three critical phases: (1) Primary tumor expansion, characterized by enhanced glycolysis (GLUT1/HK2 upregulation driving lactate accumulation), PPP activation (G6PD upregulation leading to increased G6PD-mediated NADPH production for antioxidant defense), and enhanced suppression of OXPHOS via upregulation of PDK1 and its inhibition of PDC; (2) Metastatic adaptation, featuring niche-specific metabolic shifts: bone metastasis (PTHrP-TGF-β/calcium signaling driving osteoclast-mediated glycolysis and lactate production), brain metastasis (ACSS2-facilitated efficient ketone utilization via β-hydroxybutyrate-to-acetyl-CoA conversion to promote OXPHOS), lung metastasis (OXPHOS enhancement and CPT1A-dependent fatty acid oxidation in an oxygen-rich microenvironment), liver metastasis (MCT1/LDHB-mediated lactate scavenging to fuel OXPHOS via TCA cycle anaplerosis), and CTCs (CPT1A-driven fatty acid β-oxidation supporting anoikis resistance and metastasis); (3) Therapeutic resistance, where initial glycolysis suppression gives way to resistance mechanisms, including glycolytic rebound causing T-cell exhaustion and NK cell suppression, PRC-1α-mediated OXPHOS compensation promoting tumor cell survival, and G6PD/PPP-driven antioxidant defense via upregulated G6PD activation. GLUT1, glucose transporter 1; HK2, hexokinase 2; PDK1, pyruvate dehydrogenase kinase 1; PDC, pyruvate dehydrogenase complex; OXPHOS, oxidative phosphorylation; PPP, phosphorylative pentose; G6PD, glucose-6-phosphate dehydrogenase; NADPH, nicotinamide adenine dinucleotide phosphate; PTHrP, parathyroid hormone-related protein; ACSS2, acyl-CoA synthetase short-chain family member 2; CTC, circulating tumor cell. Graphical Elements from BioGDP (https://biogdp.com) [1].
Figure 2. Metabolic reprogramming across tumor progression and therapeutic intervention. This schematic illustrates dynamic metabolic adaptations during three critical phases: (1) Primary tumor expansion, characterized by enhanced glycolysis (GLUT1/HK2 upregulation driving lactate accumulation), PPP activation (G6PD upregulation leading to increased G6PD-mediated NADPH production for antioxidant defense), and enhanced suppression of OXPHOS via upregulation of PDK1 and its inhibition of PDC; (2) Metastatic adaptation, featuring niche-specific metabolic shifts: bone metastasis (PTHrP-TGF-β/calcium signaling driving osteoclast-mediated glycolysis and lactate production), brain metastasis (ACSS2-facilitated efficient ketone utilization via β-hydroxybutyrate-to-acetyl-CoA conversion to promote OXPHOS), lung metastasis (OXPHOS enhancement and CPT1A-dependent fatty acid oxidation in an oxygen-rich microenvironment), liver metastasis (MCT1/LDHB-mediated lactate scavenging to fuel OXPHOS via TCA cycle anaplerosis), and CTCs (CPT1A-driven fatty acid β-oxidation supporting anoikis resistance and metastasis); (3) Therapeutic resistance, where initial glycolysis suppression gives way to resistance mechanisms, including glycolytic rebound causing T-cell exhaustion and NK cell suppression, PRC-1α-mediated OXPHOS compensation promoting tumor cell survival, and G6PD/PPP-driven antioxidant defense via upregulated G6PD activation. GLUT1, glucose transporter 1; HK2, hexokinase 2; PDK1, pyruvate dehydrogenase kinase 1; PDC, pyruvate dehydrogenase complex; OXPHOS, oxidative phosphorylation; PPP, phosphorylative pentose; G6PD, glucose-6-phosphate dehydrogenase; NADPH, nicotinamide adenine dinucleotide phosphate; PTHrP, parathyroid hormone-related protein; ACSS2, acyl-CoA synthetase short-chain family member 2; CTC, circulating tumor cell. Graphical Elements from BioGDP (https://biogdp.com) [1].
Ijms 26 06901 g002
Figure 3. Molecular crosstalk between circadian rhythm disruption and tumor metabolic reprogramming. The core circadian oscillator, the CLOCK-BMAL1 heterodimer, positively regulates PER gene transcription. The accumulated PER protein forms an inhibitory feedback loop by suppressing CLOCK-BMAL1 activity, thereby maintaining a 24-h circadian rhythm (gray arrow). PER recruits HIF-1α, and the hypoxic tumor microenvironment stabilizes HIF-1α, enabling cooperative suppression of CLOCK–BMAL1 heterodimer activity. This dual repression concurrently enhances glycolysis and impairs mitochondrial respiration. Conversely, oncogenic metabolic reprogramming involves MYC-driven transcriptional repression of BMAL1 by REV-ERBα, leading to the upregulation of glycolytic enzymes (HK2 and GLUT1) and enhanced glycolysis (purple pathway). Simultaneously, the ZNF704–SIN3A complex destabilizes PER2, exacerbating circadian disruption and reinforcing metabolic reprogramming. CLOCK, circadian locomotor output cycles kaput; BMAL1, brain and muscle ARNT-like 1; PER, period circadian regulator; HIF-1α, hypoxia-inducible factor 1 alpha; REV-ERBα, nuclear receptor subfamily 1 group D member 1; HK2, hexokinase 2; GLUT1, glucose transporter 1; ZNF704, zinc finger protein 704; SIN3A, SIN3 transcription regulator family member A; PER2, period circadian regulator 2.
Figure 3. Molecular crosstalk between circadian rhythm disruption and tumor metabolic reprogramming. The core circadian oscillator, the CLOCK-BMAL1 heterodimer, positively regulates PER gene transcription. The accumulated PER protein forms an inhibitory feedback loop by suppressing CLOCK-BMAL1 activity, thereby maintaining a 24-h circadian rhythm (gray arrow). PER recruits HIF-1α, and the hypoxic tumor microenvironment stabilizes HIF-1α, enabling cooperative suppression of CLOCK–BMAL1 heterodimer activity. This dual repression concurrently enhances glycolysis and impairs mitochondrial respiration. Conversely, oncogenic metabolic reprogramming involves MYC-driven transcriptional repression of BMAL1 by REV-ERBα, leading to the upregulation of glycolytic enzymes (HK2 and GLUT1) and enhanced glycolysis (purple pathway). Simultaneously, the ZNF704–SIN3A complex destabilizes PER2, exacerbating circadian disruption and reinforcing metabolic reprogramming. CLOCK, circadian locomotor output cycles kaput; BMAL1, brain and muscle ARNT-like 1; PER, period circadian regulator; HIF-1α, hypoxia-inducible factor 1 alpha; REV-ERBα, nuclear receptor subfamily 1 group D member 1; HK2, hexokinase 2; GLUT1, glucose transporter 1; ZNF704, zinc finger protein 704; SIN3A, SIN3 transcription regulator family member A; PER2, period circadian regulator 2.
Ijms 26 06901 g003
Table 2. Metabolic adaptations underlying acquired resistance to cancer therapies.
Table 2. Metabolic adaptations underlying acquired resistance to cancer therapies.
Therapeutic ContextResistance MechanismKey Metabolic AlterationsAssociated Biomarkers/Pathways
Gefitinib-Resistant EGFR-Mutant NSCLC [35]Enhanced GlycolysisUpregulation of GLUT1, HK2, LDHAIncreased lactate production
Vemurafenib-Resistant BRAF-Mutant Melanoma [36]Enhanced GlycolysisUpregulation of GLUT1/3, HK2 via MEK/ERK signalingElevated glucose uptake and glycolytic flux
Trastuzumab-Resistant HER2+ Breast Cancer [37]Enhanced GlycolysisPFKFB3-mediated fructose-2,6-bisphosphate synthesisIncreased glycolytic activity
Osimertinib-Resistant NSCLC [38]Mitochondrial ReprogrammingEnhanced OXPHOSIncreased OXPHOS activity
BRAF Inhibitor-Resistant Melanoma [39]Mitochondrial BiogenesisPGC-1α-mediated increase in mitochondrial density and oxidative capacityElevated cristae density, increased OXPHOS
Immunotherapy Resistance (e.g., PD-1 Inhibition) [40]Metabolic Reprogramming Linked to Immune EvasionSelection for clones with enhanced OXPHOS, co-expression of TIM-3Link between metabolic reprogramming and immune evasion
Note: EGFR, epidermal growth factor receptor; NSCLC, non-small cell lung cancer; GLUT1, glucose transporter 1; HK2, hexokinase 2; LDHA, lactate dehydrogenase A; BRAF, B-Raf proto-oncogene, serine/threonine kinase; HER2, human epidermal growth factor receptor 2; PFKFB3, 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 3; OXPHOS, oxidative phosphorylation; PGC-1α, peroxisome proliferator-activated receptor gamma coactivator 1-alpha; PD-1, programmed cell death protein 1; TIM-3, T-cell immunoglobulin and mucin-domain containing-3.
Table 3. Hypoxia-induced metabolites function as epigenetic modulators.
Table 3. Hypoxia-induced metabolites function as epigenetic modulators.
MetaboliteTarget EnzymesEpigenetic EffectsKey Mechanisms
Lactate
[51,52]
HDACs, GPR81Increased H3K18 lactylationLactate promotes the epithelial–mesenchymal transition of liver cancer cells via TWIST1 lactylation.
Succinate
[53]
PHDs, KDMsDNA hypermethylationSuccinic acid stabilizes HIF-1α activity by inhibiting its degradation. Succinate promotes angiogenesis by inducing the expression of VEGF.
Fumarate
[56]
TETs, PHDsHistone succinylationThe accumulation of fumaric acid results in pseudohypoxia by stabilizing the transcription factor HIF-1α and DNA hypermethylation, and inhibiting enzymes of the 2OGDD family.
2-HG
[57,58]
FTO, ATP synthaseDecreased m6A modificationSelective inhibition of the m6A epitranscriptomic regulator FTO attenuates growth in IDH1wt glioma.
Note: HDACs, histone deacetylases; GPR81, G-protein-coupled receptor 81; TWIST1, twist family bHLH transcription factor 1; PHD, prolyl hydroxylase; KDM, histone lysine demethylase; HIF-1α, hypoxia-inducible factor 1 alpha; VEGF, vascular endothelial growth factor; TET, ten–eleven translocation methylcytosine dioxygenases; 2OGDD, 2-oxoglutarate-dependent dioxygenases; 2-HG, 2-hydroxyglutarate; FTO, fat mass and obesity-associated protein; ATP, adenosine triphosphate; IDH1wt, isocitrate dehydrogenase 1 wild type.
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

Chai, X.; Tao, Q.; Li, L. Spatiotemporal Heterogeneity of Tumor Glucose Metabolism Reprogramming: From Single-Cell Mechanisms to Precision Interventions. Int. J. Mol. Sci. 2025, 26, 6901. https://doi.org/10.3390/ijms26146901

AMA Style

Chai X, Tao Q, Li L. Spatiotemporal Heterogeneity of Tumor Glucose Metabolism Reprogramming: From Single-Cell Mechanisms to Precision Interventions. International Journal of Molecular Sciences. 2025; 26(14):6901. https://doi.org/10.3390/ijms26146901

Chicago/Turabian Style

Chai, Xiaoxue, Qian Tao, and Lili Li. 2025. "Spatiotemporal Heterogeneity of Tumor Glucose Metabolism Reprogramming: From Single-Cell Mechanisms to Precision Interventions" International Journal of Molecular Sciences 26, no. 14: 6901. https://doi.org/10.3390/ijms26146901

APA Style

Chai, X., Tao, Q., & Li, L. (2025). Spatiotemporal Heterogeneity of Tumor Glucose Metabolism Reprogramming: From Single-Cell Mechanisms to Precision Interventions. International Journal of Molecular Sciences, 26(14), 6901. https://doi.org/10.3390/ijms26146901

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