International Journal of Molecular Medicine is an international journal devoted to molecular mechanisms of human disease.
International Journal of Oncology is an international journal devoted to oncology research and cancer treatment.
Covers molecular medicine topics such as pharmacology, pathology, genetics, neuroscience, infectious diseases, molecular cardiology, and molecular surgery.
Oncology Reports is an international journal devoted to fundamental and applied research in Oncology.
Experimental and Therapeutic Medicine is an international journal devoted to laboratory and clinical medicine.
Oncology Letters is an international journal devoted to Experimental and Clinical Oncology.
Explores a wide range of biological and medical fields, including pharmacology, genetics, microbiology, neuroscience, and molecular cardiology.
International journal addressing all aspects of oncology research, from tumorigenesis and oncogenes to chemotherapy and metastasis.
Multidisciplinary open-access journal spanning biochemistry, genetics, neuroscience, environmental health, and synthetic biology.
Open-access journal combining biochemistry, pharmacology, immunology, and genetics to advance health through functional nutrition.
Publishes open-access research on using epigenetics to advance understanding and treatment of human disease.
An International Open Access Journal Devoted to General Medicine.
Ferroptosis, first described by Dixon in 2012 as an iron-dependent regulated form of cell death, has emerged as a critical determinant of cellular fate across diverse physiological and pathological contexts (1). This unique mode of cell death is primarily driven by phospholipid peroxidation and iron overload, extending from decades of research that recognized the cytotoxic consequences of iron and lipid peroxidation (2,3). The elucidation of antioxidant defense systems in ferroptosis regulation has advanced our integrated understanding of the interplay between iron metabolism and oxidative stress in regulated cell death. Emerging evidence further indicates that ferroptosis significantly contributes to tumor suppression, immune regulation and the maintenance of metabolic homeostasis (4).
Concurrently, lactate has undergone a profound conceptual transformation from a metabolic waste product to a sophisticated signaling molecule. Once considered merely a metabolic waste product associated with hypoxic stress and detrimental effects, subsequent research has revealed that lactate is actively produced and utilized even under aerobic conditions (5). The lactate shuttle hypothesis further revealed its pivotal roles in oxidative substrate transport and cellular signaling via monocarboxylate transporters (MCTs), leading to its recognition as a potent signaling metabolite, particularly through its receptor GPR81 (6-8). subsequent discoveries have revealed the epigenetic functions of lactate through histone lysine lactylation (Kla), a post-translational modification analogous to acetylation and succinylation (9,10). This modification impacts transcriptional regulation and extends to non-histone proteins, thereby regulating enzyme activities (11,12). However, a paradox has emerged: Lactate exhibits opposing effects on ferroptosis, promoting cell death in certain contexts while conferring protection in others, particularly when comparing tumor vs. normal cells. To the best of our knowledge, no relevant articles currently summarize and explain these contradictory observations.
The present review elucidates the molecular mechanisms underlying the lactate-ferroptosis axis, examining how lactate influences ferroptosis through lactylation-independent and -dependent pathways that modulate iron homeostasis, lipid metabolism, redox balance and the immune response. Meanwhile, the 'dual role' refers to the context-dependent, bidirectional regulation of ferroptosis by lactate. In non-tumor tissues, lactate tends to promote ferroptosis, potentially accelerating disease progression. Conversely, in tumor microenvironments, lactate tends to inhibit ferroptosis, facilitating cancer cell survival and therapeutic resistance. Furthermore, the present review highlights the key contextual determinants that may dictate the divergent roles of lactate. Understanding these context-specific mechanisms promises new therapeutic strategies targeting a broad spectrum of diseases ranging from cancer to neurodegeneration.
Unless otherwise specified, the term 'lactate' throughout the present review refers to L-lactate, the predominant enantiomer produced by mammalian lactate dehydrogenase (LDH). By contrast, D-lactate, though present in trace amounts from bacterial metabolism and certain metabolic disorders, has not yet been systematically investigated in ferroptosis contexts.
Ferroptosis is mediated by multiple mechanisms including lipid peroxidation, iron overload and dysfunction of the antioxidant system (Fig. 1). A hallmark of ferroptotic initiation and execution is the upregulated peroxidation of phospholipid-bound polyunsaturated fatty acids (PUFAs) within cellular membranes, a process facilitated by labile iron pools (LIPs) and amplified through autocatalytic radical reactions (13,14). In this context, redox-active iron, principally existing as ferrous (Fe2+) ions, serves as a critical cofactor, driving Fenton reactions that generate reactive oxygen species (ROS). These ROS abstract hydrogen atoms from bis-allylic positions within PUFA chains, producing lipid radicals that react with molecular oxygen to yield lipid peroxyl radicals. This initiates a self-perpetuating cascade of membrane lipid peroxidation, ultimately compromising membrane integrity and triggering cell death (14).
This process of lipid peroxidation is regulated by a complex network of metabolic and enzymatic regulators. Glutathione peroxidase 4 (GPX4) plays a central role in counteracting ferroptosis by reducing membrane lipid hydroperoxides to their corresponding alcohols, utilizing glutathione (GSH) as a reducing substrate (15). Perturbation of this axis, either through direct GPX4 inhibition or GSH depletion via impaired cystine import (for example, via system Xc− inhibition), markedly sensitizes cells to ferroptotic death (1,16). In parallel, ferroptosis suppressor protein 1 (FSP1)/ubiquinol (CoQH2), dihydroorotate dehydrogenase/CoQH2 and GTP cyclohydrolase 1 (GCH1)/tetrahydrobiopterin have been identified as independent systems that scavenge free radicals to exert their antioxidative effects and suppress ferroptosis (17-21).
The execution of the Fenton response is dependent on iron availability. Cellular iron metabolism is exquisitely regulated, with transferrin-mediated uptake, ferritin-based storage and transferrin-transferrin receptor (TFRC)-mediated export collectively maintaining intracellular iron homeostasis (22). Perturbations that expand the LIP, whether through increased iron import, mobilization from stores or diminished export, potentiate ferroptosis by providing increased substrate levels for lipid peroxidation reactions (1,23). In addition, the membrane susceptibility to ferroptotic damage is modulated by its lipidomic composition, with phospholipids enriched in PUFAs, particularly arachidonic acid (AA) and adrenic acid (AdA), being especially prone to peroxidation, thereby rendering membrane vulnerability to ferroptosis (24).
Recently, a growing body of evidence has demonstrated that lactate is not merely a metabolic by-product but also serves as a key energy source and a critical signaling molecule involved in memory formation, neuroprotection, modulation of inflammatory responses, wound healing, ischemic injury repair, as well as tumor growth and metastasis (5,25-27). In this section, a comprehensive overview of the metabolic pathways of lactate is presented and its specific mechanistic roles, with particular emphasis on its dual role in regulating ferroptosis, are examined (Fig. 2).
When cellular energy demands exceed the capacity of aerobic metabolism, as occurs during intense exercise or infection, lactate is generated via glycolysis to serve as an alternative energy source. Under hypoxic conditions, cytoplasmic glucose is metabolized into pyruvate through a series of enzymatic reactions. However, instead of being transported into mitochondria for oxidative metabolism, pyruvate is converted into lactate by LDHA, coupled with NADH/NAD+ interconversion (5). This reaction is reversible: Under sufficient oxygen availability, lactate can be reconverted into pyruvate by LDHB, which is then subsequently oxidized to acetyl-CoA by pyruvate dehydrogenase (PDH) and enters the tricarboxylic acid cycle (TCA) for efficient energy production (5). Furthermore, an electrochemical gradient driving ATP synthesis is created as electrons shuttle through NAD+/NADH and FAD/FADH2 to the electron transport chain (28).
Studies have highlighted that the conversion of glucose to lactate by cells constitutes a tightly regulated metabolic state, which may confer advantages during periods of heightened biosynthetic demand (29,30). By channeling excess pyruvate toward lactate production, proliferating cells effectively prevent cytosolic NADH accumulation and prevent excessive ATP generation. This regulation ensures the continuation of cytosolic glucose metabolism without feedback inhibition from mitochondrial ATP overproduction. Furthermore, glucose-6-phosphate derived from glycolysis can be diverted into branching metabolic pathways, such as the pentose phosphate pathway (PPP), where it is partially oxidized to generate NADPH (28). This NADPH serves as a critical reducing equivalent for fatty acid synthesis and other anabolic processes. Additionally, isotope tracing studies have demonstrated that lactate functions as a major fuel in the TCA cycle, where 13C-lactate labeled TCA intermediates in every tissue of the body, even in tumors (31,32). However, the accumulation of lactate carries significant risks to the human body. Elevated serum lactate levels can result in lactic acidosis, a condition that poses a greater physiological threat compared with other metabolic intermediates (33).
In addition to glycolysis, other biochemical pathways, such as glutamine metabolism, contribute significantly to lactate production in vivo particularly in cancer cells (34). Under the regulation of the oncogene c-Myc, glutamine is metabolized through the TCA cycle to generate pyruvate, which is subsequently converted into lactate by LDH (35). This process provides an alternative and critical source of lactate in rapidly proliferating cells, supporting energy production and biosynthesis.
Hypoxia represents a hallmark of pathological conditions, including tumors, infections, ischemia-reperfusion injury (IRI) and inflammation (36); it occurs when cells experience reduced oxygen availability and activate an adaptive response to cope with it. In response to reduced oxygen availability, cells activate adaptive signaling cascades mediated by hypoxia-inducible factor-1α (HIF-1α). Under this condition, HIF-1α becomes stabilized and subsequently translocates to the nucleus, where it binds to hypoxia-responsive elements within target genes, thereby promoting metabolic programming, notably the upregulation of anaerobic glycolysis (37,38). Under hypoxic conditions, the reduced oxygen availability diminishes the activity of prolyl hydroxylase domain enzymes (PHDs), which typically employ oxygen and α-ketoglutarate as substrates to hydroxylate HIF-1α, thereby targeting it for degradation. Consequently, the decreased PHD activity leads to HIF-1α stabilization (39).
Cellular adaptation to hypoxia involves an enhanced glycolytic flux, primarily mediated by HIF-dependent transcriptional upregulation of genes encoding glucose transporters and glycolytic enzymes. This metabolic shift is accompanied by an active suppression of mitochondrial pyruvate oxidation and respiratory activity (40-43). These biochemical adaptations result in a metabolic reprogramming that promotes lactate production and accumulation, thereby reinforcing the hypoxic cellular response.
LDHA and LDHB constitute the subunits of the catalytically active LDH enzyme, which has long been recognized for its pivotal role in ATP generation and energy homeostasis under both anaerobic and aerobic glycolytic conditions (44). LDHA possesses a higher affinity for pyruvate and preferentially catalyzes its reduction to lactate, thereby sustaining anaerobic glycolysis. By contrast, LDHB catalyzes the reverse reaction (oxidizing lactate to pyruvate) which subsequently fuels mitochondrial oxidative phosphorylation (44). However, accumulating evidence suggests that all LDH isoforms retain the inherent capacity to mediate pyruvate-to-lactate conversion accompanied by NAD+ regeneration, and that LDHA and LDHB can functionally compensate for each other under metabolic stress (44,45). By regulating the NAD+/NADH ratio, mitochondrial function and ROS levels, LDH isoforms indirectly modulate ferroptosis susceptibility. For instance, LDHA has been demonstrated to promote tumor cell survival by mitigating oxidative stress, whereas LDHB deficiency induces mitochondrial dysfunction and oxidative damage, ultimately leading to neurodegeneration in the adult mouse brain (46,47). More recently, non-canonical roles of LDHA and LDHB have been identified, with both isoforms contributing to ferroptosis resistance in IRI and cancer through mechanisms associated with GPX4 activity or GSH availability (48,49).
MCTs serve as the principal mediators of lactate transport across cell membranes. By coupling lactate translocation with proton co-transport, these transporters help maintain acid-base balance and cellular metabolic stability (7). For instance, inhibition of MCT1 disrupts lactate homeostasis and impairs both glycolytic flux and GSH biosynthesis in MYC-driven cancer, resulting in reduced glucose uptake and depletion of ATP, NADPH and GSH (50). Furthermore, extracellular lactate must enter cells through MCTs before contributing to lactylation reactions, which have been demonstrated to modulate ferroptosis through multiple mechanisms (51).
GPR81/HCAR1 is a cell-surface G protein-coupled receptor that recognizes lactate as its endogenous ligand and coordinates metabolic signaling across diverse tissues (8). In adipocytes, GPR81 acts synergistically with insulin to lower intracellular cyclic adenosine monophosphate (cAMP) levels, thereby suppressing postprandial lipolysis (52). Beyond adipose tissue, GPR81 is abundantly expressed in skeletal muscle, kidney, brain, heart and various cancer types (8). Elevated extracellular lactate, a defining feature of the tumor microenvironment, predicts poor outcomes, and high GPR81 expression correlates with poorer survival (8,53,54). Analogous to its role in adipocytes, GPR81 activation in cancer cells decreases intracellular cAMP and suppresses proteinkinase A (PKA) activity, thereby modulating lipid remodeling (55,56).
Elevated lactate concentrations within the tumor microenvironment can activate GPR81 receptors located on the cytoplasmic membrane of cells, subsequently facilitating MCT1-mediated lactate uptake (57,58). Through this pathway, lactate disrupts AMPK signaling, which in turn downregulates sterol regulatory element-binding protein 1 (SREBP1) and its downstream target stearoyl-CoA desaturase 1 (SCD1). Consequently, cells produce more anti-ferroptotic monounsaturated fatty acids (MUFAs), thereby suppressing lipid peroxidation. Simultaneously, long-chain acyl-CoA synthetase 4 (ACSL4) expression decreases, reducing the availability of oxidizable PUFAs, although the exact mechanism underlying this remains unclear. Notably, blocking lactate transport by inhibiting MCT1 or GPR81 promotes ferroptosis primarily through alterations in lipid metabolism rather than through conventional ferroptosis regulators, as evidenced by unchanged GPX4 and FSP1 levels (57). Additionally, other research has demonstrated that intracellular lactate accumulation following MCT inhibition may lead to end-product inhibition of LDH, thereby impairing NAD+ regeneration capacity (59). This sustained disruption of glycolysis can result in the depletion of ATP, NADPH and GSH (50). These findings suggest that therapeutic strategies targeting the function of MCTs could prove effective against both oxidative and hypoxic tumors.
Accumulating evidence indicates that lactate metabolism modulates ferroptosis through both lactylation-independent and dependent pathways. Notably, this regulation exhibits context-dependent effects. The lactylation-independent mechanisms, including the iron handling, lipid remodeling, redox regulation and immune responses in ferroptosis, are summarized in this section. Additionally, evidence supporting lactylation-dependent mechanisms has been compiled.
Clinical and experimental evidence has consistently demonstrated a robust association between hyperlactatemia and anemia, indicating a functional interconnection between lactate metabolism and systemic iron homeostasis (60). For instance, in endurance athletes, repeated bouts of exercise induces transient elevations in plasma lactate concentrations, which have been correlated with the development of iron-restrictive anemia in 10-15% of individuals, particularly among those engaging in >10 h of training per week (61). Hepcidin, a peptide hormone predominantly synthesized by hepatocytes, regulates iron homeostasis by binding to ferroportin (FPN), the sole identified cellular iron exporter. This interaction triggers FPN internalization and degradation, consequently restricting iron efflux and leading to the elevation of intracellular iron pools. Concurrently, it limits systemic iron availability by reducing intestinal iron absorption, ultimately resulting in decreased serum iron concentrations (62) (Fig. 3).
Liu et al (63) recently elucidated the molecular mechanism by which lactate regulates hepcidin expression and, consequently, systemic iron homeostasis. Their findings revealed that lactate, transported into cells via MCT1, directly interacts with soluble adenylyl cyclase, resulting in elevated cAMP levels. This, in turn, activates the PKA-Smad1/5/8 signaling cascade, ultimately upregulating hepcidin transcription (63). Further investigation demonstrated that lactate administration in mice induced hepcidin expression, leading to reduced FPN levels. This resulted in increased splenic iron sequestration, diminished duodenal iron absorption and, consequently, decreased serum and tissue iron content accompanied by attenuation of oxidative stress (63,64). Although tissue-specific differences in hepcidin sensitivity, along with variations in critical thresholds, time kinetics and the compensatory capability of cellular antioxidant defense systems, may modulate iron accumulation and ferroptotic outcomes, existing evidence supports a pro-ferroptotic role of hepcidin in Kupffer cells, neurons and hepatocytes (65-67).
Although current evidence supports a sophisticated mechanistic framework linking lactate-mediated hepcidin induction to ferroptosis, the tissue- and cell-type-specific effects of elevated lactate levels induced by diverse physiological and pathophysiological processes on iron homeostasis and ferroptosis throughout the body warrant further investigation.
Iron can be imported into the cell via the TFRC system. In this process, Fe3+-transferrin complexes are internalized through TFRC and eventually trafficked to and liberated within the acidic environment of the lysosome mediated by nuclear receptor coactivator 4 (NCOA4) (2). TFRC silencing and NCOA4 disruption reduce ferroptotic sensitivity by restricting iron retrieval from ferritin and thereby limiting the LIP (68,69). Lactylation of histones or non-histone proteins involved in iron autophagy-related pathways can modulate iron handling, thereby regulating ferroptosis (70-72). Notably, this regulatory effect appears to be context-dependent; for instance, H3K14 lactylation has been demonstrated to promote ferroptosis in endothelial cells exposed to lipopolysaccharide (LPS) by transcriptionally upregulating the TFRC (73). Conversely, lactylation of lysine-specific demethylase 1 (LSD1) in melanoma promotes its interaction with Fos-like antigen 1 (FosL1), resulting in repression of TFRC-mediated iron uptake and conferring resistance to ferroptosis (70). However, at present, there is no direct evidence demonstrating that histone lactylation regulates the TFRC in tumor cells. Notably, ferroptotic susceptibility is determined not only by iron overload but also by the lipid ratio and the capacity of antioxidant defense systems.
Beyond its established roles in energy metabolism and immune modulation within the tumor microenvironment, lactate serves as a crucial metabolic substrate that supports the TCA cycle in major organs under physiological conditions (31). Tracer studies employing 13C-labeled lactate have demonstrated robust incorporation of lactate-derived carbon into TCA intermediates across a wide array of tissues, encompassing both healthy and malignant cells (31,32).
Mechanistically, lactate contributes to the intracellular acetyl-CoA pool through its conversion to pyruvate, followed by subsequent metabolism (74). Acetyl-CoA carboxylase (ACC) then catalyzes the rate-limiting carboxylation of acetyl-CoA to malonyl-CoA, a critical intermediate in fatty acid biosynthesis (75). Fatty acid synthase subsequently utilizes malonyl-CoA to synthesize palmitic acid (C16:0), which can be elongated to stearic acid (C18:0) through the action of elongation-of-very-long-chain-fatty acids 6 (Fig. 4). SCD1 then introduces a double bond at the Δ9 position, converting these saturated fatty acids to palmitoleic and oleic acids, respectively. Malonyl-CoA also plays a vital role in PUFA synthesis, where dietary linoleic acid undergoes sequential desaturation and elongation reactions catalyzed by fatty acid desaturase (FADS)2, ELOVL5 and FADS1, leading to the formation of AA (C20:4, ω-6), which can be further elongated by ELOVL2/4 to generate AdA (C22:4, ω-6) (76). These interconnected pathways highlight the pivotal role of ACC-derived malonyl-CoA in fatty acid metabolism, with ACC inhibition potentially attenuating lipid peroxidation and ferroptosis (77-79). Additionally, lactate metabolism enhances cellular NADPH availability through a glucose-sparing mechanism: When cells use lactate as their primary fuel source, glucose is redirected towards the PPP, which generates NADPH essential for fatty acid biosynthesis (28).
A study by Zhao et al (57) demonstrated that lactate-enriched hepatocellular carcinoma (HCC) cells exhibit enhanced resistance to ferroptotic damage induced by RAS-selective lethal 3 (RSL3) and Erastin. The authors elucidated a mechanism wherein MCT1-mediated lactate uptake promotes ATP production, leading to AMPK deactivation and subsequent upregulation of SREBP1 and SCD1, enhancing the production of MUFAs that confer protection against ferroptosis (57). Corroborating these findings, Yang et al (58) demonstrated that lactate-induced alterations in SCD1/ACSL4 expression and ferroptosis resistance are correlated with lactate production levels in esophageal squamous cell carcinoma (ESCC). The conservation of this lactate/SCD1 regulatory axis across diverse tissue types has been substantiated by multiple studies (80,81). ACSL4 is an enzyme that esterifies CoA into specific PUFAs, such as AA and AdA, contributing to ferroptosis execution by triggering phospholipid peroxidation. In sepsis, lactate promotes ferroptosis via GPR81-mediated upregulation of methyltransferase like 3 (METTL3), which may mediate ACSL4 mRNA stability via N6-methyladenosine (m6A) modification in pulmonary epithelial cells (82).
Although regulating the MUFA/PUFA ratio by lactate through the SCD1/ASCL4 pathway does affect membrane lipid peroxidation sensitivity, this protective effect against ferroptosis may be more highly dependent on the state of the intracellular antioxidant system (83).
Disruptions in redox balance, whether toward excessive oxidation or reduction, are generally deleterious to cellular function. Lactate acts as a redox buffer by modulating the NAD(P)H/NAD(P)+ ratio and serves as a signaling molecule that regulates the activity of antioxidant enzymes involved in ROS metabolism.
The interconversion of lactate and pyruvate, catalyzed by LDHA/B, is closely coupled to the NAD+/NADH redox pair, thereby modulating cellular redox balance in a context-dependent manner (84,85). Elevated lactate concentrations have been demonstrated to increase the NADH/NAD+ ratio, leading to the inhibition of key glycolytic enzymes such as glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and phosphoglycerate dehydrogenase, ultimately suppressing both glycolysis and mitochondrial respiration (85). When the cellular demand for NAD+ to sustain oxidation exceeds the rate of ATP turnover, particularly in cells exhibiting active aerobic glycolysis, NAD+ regeneration becomes a limiting factor under conditions of compromised mitochondrial respiration. In such circumstances, cells preferentially rely on glycolysis, resulting in an elevated NAD+/NADH ratio. Under these conditions, activation of PDH facilitates pyruvate oxidation, attenuates lactate accumulation and restores the NAD+/NADH equilibrium (86). Conversely, when lactate is oxidized to pyruvate and enters mitochondrial oxidative pathways, it can indirectly elevate ROS production through electron leakage from the respiratory chain (87-89). In adipocytes, lactate exposure transiently elevates the NADH/NAD+ ratio, which normalizes after 24 h due to elevated NAD+ levels associated with increased mitochondrial membrane potential and enhanced ROS production (90). Similarly, Jia et al (91) reported that elevated neuronal lactate uptake augments ROS production and mitochondrial oxidative metabolism, thereby disrupting the balance between ROS generation and detoxification, impairing ATP synthesis and ultimately leading to peripheral axonal degeneration. The mitochondrial enzyme nicotinamide nucleotide transhydrogenase further modulates redox status by catalyzing the reversible conversion of NADH and NADP+ into NAD+ and NADPH, the latter serving as a crucial factor for GSH reductase-mediated recycling of reduced GSH (92). Unlike observations in normal cells, lactate uptake in melanoma cells via MCT1 elevates NADH, NADPH, GPX4 and FSP1 levels, thereby conferring resistance to ferroptosis (93). The net effect of lactate on redox status is governed by the balance between reducing equivalents of NADH and ROS production, which may differ between normal and cancer cells due to variations in oxidative phosphorylation activity and NADH-processing pathways. This apparent contradiction may reflect a critical evolutionary adaptation in which cancer cells reprogram lactate metabolism to attain dual benefits (enhanced metabolic flexibility coupled with augmented oxidative stress resistance) thereby facilitating survival under adverse microenvironmental conditions.
Within glucose metabolism, glucose-6-phosphate generated either through glycolysis or via gluconeogenesis from lactate can enter the PPP, where it is partially oxidized to generate NADPH (94,95). Moreover, lactate metabolism contributes to NADPH production through TCA cycle-linked pathways, particularly those catalyzed by malic enzyme 1 (ME1) and isocitrate dehydrogenase 1 (IDH1) (94,96). Similar to NADH, NADPH fulfills dual roles in cellular redox regulation: It is indispensable for antioxidant defense by facilitating GSH reduction yet simultaneously serves as a substrate for NADPH oxidases that generate ROS production (97). Under glucose-deprived conditions, the knockout of IDH1 or ME1 significantly reduces the NADPH/NADP+ and GSH/oxidized GSH ratios, leading to elevated ROS levels and increased cell death in cancer cells (94). Conversely, excessive NADPH accumulation can induce reductive stress, leading to the upregulation of NADPH oxidase 4 through activation of the PI3K/Akt signaling pathway, thereby promoting ROS generation, as observed in osteoarthritis (96).
Collectively, cellular redox homeostasis is maintained by a delicate equilibrium between NAD(P)H-dependent antioxidant defense mechanisms and NAD(P)H-mediated reductive stress. By engaging in NAD(P)H/NAD(P)+ redox cycling, lactate functions as a dynamic redox buffer that modulates cellular responses to oxidative and reductive stress.
Under physiological conditions, lactate indirectly regulates cellular antioxidant systems through modulation of the NAD(P)H/NAD(P)+ redox balance. Specifically, uncontrolled elevated intracellular lactate concentrations disrupt the NADH/NAD+ ratio, thereby impairing the activity of key glycolytic enzymes, particularly GAPDH due to NAD+ depletion (50,57). This metabolic perturbation, coupled with lactate-mediated feedback inhibition of phosphofructokinase-1, ultimately impairs cellular ATP homeostasis and compromises GSH biosynthesis, the predominant intracellular antioxidant (50). GPX4 plays a pivotal role in counteracting ferroptosis. A recent investigation has demonstrated that intracellular lactate accumulation under ischemic conditions can inactivate the AMPK/GPX4 axis to promote myocardial ferroptosis (98). Conversely, emerging evidence indicates that lactate can paradoxically potentiate antioxidant defense mechanisms through alternative signaling cascades. Metabolically reprogrammed lactate has been demonstrated to augment GPX4 expression and confer resistance to ferroptosis via activation of the p38-serum glucocorticoid-regulated kinase 1 signaling axis. This pathway mitigates GPX4 ubiquitination and subsequent degradation in non-small cell lung cancer (NSCLC) cells (99). Furthermore, an additional study has elucidated that lactate activates antioxidant defense and pro-survival pathways, including the unfolded protein response and nuclear factor erythroid 2-related factor 2 (NRF2) signaling cascades, by inducing mild oxidative stress in neuroblastoma cells (100). Notably, lactate has also been shown to alleviate oxidative stress-induced cell death through autophagy activation in retinal pigment epithelial cells (101).
The immunomodulatory properties of lactate significantly influence ferroptosis susceptibility within both inflammatory conditions and the tumor microenvironment, carrying notable therapeutic relevance (102,103).
Sepsis-induced metabolic reprogramming drives increased lactate accumulation and systemic oxidative stress, triggering multiple cell death pathways in both immune and parenchymal cells (82,104). In septic lung injury, elevated lactate levels exacerbate alveolar epithelial cell ferroptosis through lactylated histone H3 lysine 18 (H3K18la)-mediated upregulation of METTL3, which enhances m6A modification of ACSL4, ultimately contributing to the development of acute respiratory distress syndrome (82). Lactate further augments neutrophil functions, including chemotaxis, phagocytosis, oxidative burst and neutrophil extracellular trap formation, via energy provision and PI3K/Akt signaling, which may exacerbate tissue cell ferroptosis (105). Nevertheless, lactate simultaneously exerts immunomodulatory effects in sepsis; it suppresses LPS-induced pro-inflammatory cytokine production in macrophages and promotes M2 polarization through MCTs and HIF-1α activation (106). Moreover, lactate-induced histone lactylation drives macrophages toward a reparative phenotype characterized by reduced pro-inflammatory cytokine expression, thereby potentially constraining excessive inflammation (107,108). This dual role of lactate in modulating ferroptosis and inflammation may account for the inconsistent therapeutic outcomes associated with hypertonic sodium lactate administration in sepsis, as these effects appear to be highly dependent on factors such as timing of intervention, dosage and the specific experimental animal models employed (104,109).
The lactate-ferroptosis axis influences anti-tumor immunity through multifaceted mechanisms that reshape the tumor microenvironment. Lactate accumulation within the tumor microenvironment suppresses effector T cell proliferation and cytotoxic activity via GPR81-mediated signaling and metabolic competition, while simultaneously polarizing tumor-associated macrophages toward an immunosuppressive M2 phenotype that promotes tumor progression (8). Notably, ferroptosis induction in cancer cells can trigger immunogenic cell death (ICD), leading to the release of damage-associated molecular patterns that activate dendritic cells and stimulate antitumor T cell responses (110). However, lactate-mediated ferroptosis resistance through SCD1 upregulation and enhanced antioxidant capacity diminishes this immunogenic potential, effectively creating an immune-evasive tumor phenotype (57,58). Recent evidence demonstrates that targeting the lactate-ferroptosis axis can reprogram the immunosuppressive tumor microenvironment. Specifically, the combination of MCT4 inhibition with ferroptosis inducers not only depletes lactate accumulation but also enhances CD8+ T cell infiltration and ferroptosis-driven ICD, thereby synergistically improving checkpoint blockade efficacy (111). This immunometabolic reprogramming represents a promising approach to convert 'cold' tumors into 'hot' tumors that are more responsive to immunotherapy.
Kla was initially identified as an enzymatically catalyzed post-translational modification, wherein lactyl groups derived from lactate are covalently attached to lysine residues (9). Originally identified on histones, Kla has been indicated to accumulate at gene promoters in response to diverse stimuli, including hypoxia, interferon-γ, LPS exposure and bacterial infections. This modification directly modulates transcriptional activity and gene expression, thereby establishing a mechanistic link between cellular metabolism and transcriptional regulation (107,108). Subsequent investigations have expanded the functional repertoire of Kla, demonstrating its presence on non-histone proteins, particularly metabolic enzymes (11,12). The lactylation of these enzymes regulates cellular metabolism by modulating enzymatic activity, notably through feedback mechanisms that regulate glycolytic flux (108,112). However, the precise enzymatic machinery responsible for Kla deposition and removal remains incompletely characterized and the relative contributions of enzymatic vs. non-enzymatic lactylation pathways warrant further clarification.
Nuclear Kla is predominantly observed at H3K18, with p300-mediated H3K18la at gene promoters serving as a key determinant of transcriptional regulation. For instance, H3K18la enrichment at the METTL3 promoter upregulates METTL3 expression, subsequently augmenting m6A modification of ACSL4 mRNA (82). This cascade stabilizes ACSL4 transcripts, elevates ACSL4 protein levels and promotes mitochondrial ROS accumulation, ultimately driving ferroptosis in alveolar epithelial cells during sepsis (82). Furthermore, H3K18la facilitates ACSL4 expression through direct promoter engagement and activation of the HIF-1α signaling pathway (113,114). Under hypoxic conditions, elevated lactate levels stimulate H3K18la at the HIF-1α promoter, leading to upregulation of HIF-1α expression (115). This signaling cascade subsequently elevates ACSL4 expression, driving lipid peroxidation and ferroptosis through the ACSL4/lysophosphatidylcholine acyltransferase 3/arachidonate lipoxygenase 15 axis, as demonstrated in both in vitro and in vivo models of severe acute pancreatitis (114). Paradoxically, Kla modifications can also confer cytoprotective effects in specific cancer contexts. In triple-negative breast cancer, cancer-associated fibroblasts exhibit elevated H3K18la levels, which enhance zinc finger protein 64 expression and subsequently activate the transcription of GCH1 and FTH. This cascade facilitates iron sequestration and protects cells from doxorubicin-induced ferroptosis (71). Similarly, in colorectal cancer stem cells, p300-mediated H4K12la upregulates glutamate-cysteine ligase catalytic subunit (GCLC), thereby expanding the GSH pool and conferring ferroptosis resistance (116). Furthermore, H3K18la enhances the transcriptional activity of NFS1 cysteine desulfurase, a cysteine desulfurase essential for iron-sulfur cluster biosynthesis, consequently reducing the susceptibility of HCC to ferroptosis following microwave ablation (117). These findings underscore the notable tissue-specific and context-dependent nature of histone Kla function. However, the molecular determinants underlying this functional dichotomy remain poorly understood, and the systematic frameworks capable of predicting whether Kla will promote or suppress ferroptosis within specific cellular contexts are still lacking.
Beyond its role in chromatin regulation, Kla also influences the activity and stability of various non-histone proteins, including key metabolic and RNA-modifying enzymes that regulate ferroptosis. Notably, lactate-primed lysine acetyltransferase 8 directly lactylates mitochondrial phosphoenolpyruvate carboxykinase 2 at K100, thereby enhancing its kinase activity and reprogramming mitochondrial fatty-acid synthesis to promote ferroptotic processes (118). Additionally, Kla of METTL3 stabilizes the protein and promotes m6A modifications on ACSL4 and TFRC transcripts, thereby accelerating ferroptosis in PC12 cells (72). In the context of Alzheimer's disease, reduced lactylation of tau at K677 inhibits ferroptosis by disrupting ferritinophagy, resulting in dysregulated iron metabolism and increased resistance to cell death (119). Similarly, decreased lactylation of malate dehydrogenase 2 at K241, coupled with reduced lactate production, leads to elevated levels of GSH and GPX4, thereby alleviating ferroptosis and improving mitochondrial function in myocardial IRI (120). By contrast, lactylation of LSD1 in melanoma promotes its interaction with FosL1, resulting in repression of TFRC-mediated iron uptake and thereby conferring resistance to ferroptosis (70). Similarly, lactylation of NOP2/Sun RNA methyltransferase 2 enhances its catalytic activity and stabilizes GCLC mRNA via m5C modifications, leading to increased intracellular GSH levels and ferroptosis resistance in gastric cancer cells (121). These findings underscore the versatility of Kla as a regulatory mechanism capable of either promoting or inhibiting ferroptotic cell death through modulation of enzymatic activity.
The bidirectional influence of Kla on ferroptosis is of significant therapeutic interest. In cancer, Kla-driven modulation of ferroptosis has been implicated in developing resistance to chemotherapy and radiation. For example, evodiamine has been found to inhibit histone lactylation at the HIF-1α promoter, suppressing angiogenesis and programmed death-ligand 1 expression while inducing ferroptosis in prostate cancer cells (122). This highlights the potential of targeting Kla 'writers', 'erasers' or 'readers' to selectively induce ferroptosis in cancer cells, thereby enhancing the efficacy of cancer therapies while minimizing damage to normal tissues. Conversely, Kla manipulation could also hold promise in treating degenerative diseases and ischemic injury by reducing ferroptotic damage (123). However, the clinical translation of Kla-targeted therapies faces notable challenges, including the lack of specific inhibitors, potential off-target effects and the complex tissue-specific functions of lactylation.
In summary, Kla represents a critical metabolic regulator integrating cellular metabolic status with ferroptotic outcomes. Although the context-dependent, bidirectional modulation of ferroptosis by Kla presents promising therapeutic avenues, notable barriers remain between current mechanistic insights and clinical implementation. This underscores the need for more rigorous investigations into tissue-specific regulatory networks and the development of targeted therapeutic strategies.
The seemingly contradictory roles of lactate in ferroptosis (its capacity to both promote and inhibit this regulated form of cell death) highlight the complex interplay among cellular metabolism, redox homeostasis and iron regulation. In the previous section, the factors that determine whether lactate functions in a pro-ferroptotic or anti-ferroptotic manner within specific cellular contexts were systematically examined (Fig. 5). In pathological conditions such as sepsis (73,82), neurodegeneration (91,119), osteoarthritis (96), pancreatitis (114), IRI (120), intracerebral hemorrhage (72) and adipocytes browning (90), lactate promotes ferroptosis. Conversely, in various malignancies, including HCC (57,117), ESCC (58), melanoma (70), NSCLC (99), neuroblastoma (100), breast cancer (71), colorectal cancer (116), gastric cancer (121) and prostate cancer (122), lactate suppresses ferroptosis. Table I summarizes recent in vitro studies examining the role of lactate in ferroptosis regulation. In this section, three potential mechanisms underlying this context-dependent regulation are discussed: Differences in metabolic enzyme expression, pH homeostatic capacity and antioxidant defense systems (Fig. 6).
Cancer cells characterized by the Warburg effect demonstrate elevated expression of MCT1 and MCT4, which promote enhanced lactate uptake and consequent metabolic reprogramming. Notably, identical lactate concentrations exert diametrically contrasting effects on ferroptosis susceptibility in malignant lung cancer cells compared with their normal epithelial counterparts, highlighting fundamental differences in lactate metabolism between these cellular contexts (82,99). Although LDHA has been extensively characterized as a tumor survival factor through its role in mitigating ROS, targeted depletion via small interfering RNA or pharmacological inhibition (GSK2837808A or R-GNE-140) fails to sensitize A549 cells to ferroptosis inducers, including RSL3 or erastin (49). This observation necessitates the existence of LDHA-independent anti-ferroptosis mechanisms. In KRAS-driven NSCLC, LDHB has been identified as a key regulator of GSH-dependent ferroptosis resistance through STAT1 signaling, although the precise molecular mechanisms underlying LDHB-mediated STAT1 activation remain incompletely elucidated (49). A recent investigation propose that LDHB mediates a complex three-step reaction cycle, wherein reducing equivalents are transferred from lactate to reduced CoQH2, with NAD+ serving as a cycling cofactor (124). The markedly elevated NAD+ concentrations observed in tumor cells compared with normal tissues may provide enhanced substrate availability for LDHB-mediated lactate oxidation cycles (125). Given that CoQH2 serves as an alternative anti-ferroptosis system operating in parallel with GPX4, the differential metabolic environments between malignant and normal cells likely contribute to the divergent effects of lactate on ferroptosis sensitivity observed across these cellular contexts. Nevertheless, this proposed mechanism lacks direct biochemical validation and relies heavily on circumstantial evidence from functional studies. Despite these limitations, this finding provides an important avenue for subsequent research.
The differential expression of MCTs between normal and malignant cells provides a mechanistic framework for understanding lactate-mediated modulation of ferroptosis susceptibility. Normal cells predominantly express MCT1, which confers a relatively limited capacity for lactate transport (126). By contrast, cancer cells typically co-express MCT1 and MCT4, facilitating highly efficient lactate efflux that preserves intracellular pH homeostasis even within the acidic tumor microenvironment (127). This enhanced acid-extruding capacity enables tumor cells to sustain intracellular pH at neutral or slightly alkaline levels, potentially exceeding those observed in normal cells under comparable conditions (128). The pH dependence of Fe2+-catalyzed lipid peroxidation constitutes a critical determinant of ferroptotic sensitivity, proceeding efficiently under acidic conditions but being markedly impaired at neutral or basic pH due to reduced Fe2+ solubility. Consequently, the efficacy of ferroptosis-inducing therapies may be limited in alkaline cytoplasmic environments (129). However, the proposed lactate-acidification-ferroptosis axis warrants careful evaluation in light of emerging contradictory evidence. An early study by Jackson and Halestrap (130) documented lactate-induced intracellular acidification in rat hepatocytes, whereas Bozzo et al (131) observed only modest pH reductions following 5 mM lactate treatment in mouse neurons. These findings suggest that normal cells possess notable buffering capacity against lactate-induced acidification, challenging the assumption that lactate uniformly acidifies the cytoplasm across diverse cell types. Moreover, LDHA-mediated lactate accumulation has been reported to promote ferroptosis resistance in a pH-dependent manner in tumors, a mechanism that may involve inhibition of Piezo1 in an acidic environment (132-134). Accordingly, the mechanistic link between lactate-mediated pH changes and iron-dependent ferroptosis remains inadequately characterized.
Tumor cells have developed intricate adaptive mechanisms to evade ferroptosis, a key tumor-suppressive process. In response to the heightened oxidative stress associated with malignant transformation, cancer cells activate a comprehensive antioxidant defense pathway that, paradoxically, shields them from ferroptotic cell death (4). At the core of this protective adaptation lies the upregulation of cystine/glutamate antiporter solute carrier family 7 member 11 (SLC7A11), driven by the inactivation of key tumor suppressors including TP53, BRCA1 associated protein 1 and alternate reading frame (16,135). This dysregulation fundamentally alters the cellular redox balance, conferring ferroptosis resistance while simultaneously promoting tumor proliferation and survival. The oncogenic KRAS signaling cascade further amplifies this effect by directly upregulating SLC7A11 expression, thereby establishing a robust anti-ferroptotic defense that is particularly pronounced in lung adenocarcinoma progression (136). Beyond cystine import, the ferroptosis evasion machinery encompasses the upregulation of critical antioxidant enzymes, notably GSH and GPX4, which are consistently upregulated across multiple tumor types (137,138). Complementing these classical antioxidant systems, cancer cells also exploit radical-trapping antioxidant mechanisms mediated by FSP1 and GCH1, both of which are frequently upregulated in diverse malignancies and markedly contribute to ferroptosis resistance (21,139). Furthermore, NRF2 activation, a hallmark of numerous cancer types, serves as both a driver of tumor progression and a coordinator of therapy resistance. Through its transcriptional control of ferroptosis-regulatory components, including SLC7A11, GPX4 and FSP1, NRF2 creates a unified resistance program that simultaneously promotes tumor progression and confers therapeutic resistance (140).
Notably, this mechanism of ferroptosis evasion may be closely associated with altered lactate metabolism in tumor cells. While lactate has been reported to promote ferroptosis in normal cellular contexts by modulating iron homeostasis and lipid peroxidation, tumor cells appear to exploit lactate signaling pathways to reinforce their anti-ferroptotic defenses. This metabolic rewiring suggests that the Warburg effect and ferroptosis evasion may be mechanistically linked, representing complementary adaptive strategies that collectively support malignant transformation and tumor progression.
The lactate-ferroptosis axis does not function independently but rather integrates with other forms of programmed cell death, constituting a complex regulatory network that modulates therapeutic outcomes. Growing evidence suggests that lactate metabolism exerts regulatory control over multiple cell death modalities, with the dominant pathway contingent upon the specific cellular context and magnitude of stress stimuli.
In cancer cells, increased lactate concentrations have been reported to suppress apoptosis through HIF-1α-mediated upregulation of anti-apoptotic proteins, including BCL-2 and survivin, while concurrently promoting ferroptosis resistance via SCD1 upregulation (58,141). This coordinated suppression of apoptotic and ferroptotic pathways confers a metabolic survival advantage under stress conditions. Conversely, in certain therapeutic contexts, lactate depletion strategies have been demonstrated to induce a synergistic activation of both apoptotic and ferroptotic pathways. For instance, MCT1 inhibition in glycolytic tumors induces a metabolic catastrophe that triggers caspase-dependent apoptosis alongside GSH depletion-mediated ferroptosis, thereby yielding enhanced antitumor efficacy relative to the activation of either pathway alone (142). A study has demonstrated that GPX4 inhibition can simultaneously activate the caspase-8-mediated apoptotic pathway and ferroptosis, suggesting shared upstream signaling mechanisms (143). Furthermore, the tumor suppressor p53 serves as a critical node connecting these pathways as it can promote ferroptosis via SLC7A11 repression while simultaneously regulating apoptotic gene expression (144).
The interplay between lactate, autophagy and ferroptosis is notably intricate. Lactate has been reported to induce protective autophagy in retinal pigment epithelial cells, mitigating oxidative stress and cell death through AMPK activation (101,145). However, in the context of ferroptosis, selective autophagy pathways such as ferritinophagy (autophagic degradation of ferritin) and lipophagy (degradation of lipid droplets) can paradoxically facilitate ferroptosis by elevating the LIP and releasing PUFAs for peroxidation (146,147). Studies have demonstrated that lactate-induced autophagy activation can dictate cellular fate between survival and ferroptotic death based on the iron-handling capacity and antioxidant reserve of specific cell types (101,119,148). In Alzheimer's disease, diminished tau lactylation suppresses ferritinophagy, resulting in iron accumulation and altered susceptibility to ferroptosis, highlighting the complex interplay among lactate metabolism, autophagy and iron homeostasis (119). Additionally, clockophagy (selective degradation of aryl hydrocarbon receptor nuclear translocator-like protein 1, a core clock gene) has been reported to enhance ferroptosis sensitivity by disrupting the circadian regulation of lipid metabolism and antioxidant defenses (149).
As the role of lactate in disease progression becomes increasingly evident, strategies aimed at modulating lactate metabolism are receiving growing attention in conditions including neurodegenerative disorders, IRI, sepsis and cancer. These approaches encompass systemic administration of lactate-enriched solutions to provide metabolic support, as well as interventions aimed at inhibiting lactate production or transport as well as promoting its depletion to disrupt metabolic symbiosis in tumors, highlighting the dual therapeutic potential of lactate (150,151).
Increasing evidence implicates ferroptosis in the pathogenesis of major neurodegenerative disorders, including Alzheimer's disease, Parkinson's disease, Huntington's disease, multiple sclerosis and amyotrophic lateral sclerosis (152). In these chronic conditions, persistent elevations in lactate are associated with oxidative stress and disrupted iron homeostasis, collectively promoting ferroptotic neuronal death (153-155). Similarly, lactate accumulation during prolonged ischemia or late reperfusion has detrimental effects. Sun et al (98) demonstrated that intracellular lactate overload under prolonged ischemia inactivates the AMPK/NRF2/GPX4 protective axis, thereby promoting myocardial ferroptosis and exacerbating cardiac injury. By contrast, lactate exhibits striking neuroprotective properties in acute ischemic stroke. Intraventricular administration following reperfusion markedly reduces infarct volume and ameliorates neurological deficits (156). This protective action is attributed to the role of lactate as an alternative energy substrate, whereby its conversion to pyruvate enables mitochondrial oxidation once oxygen is restored (157). However, lactate accumulated during the ischemic phase itself, when oxygen tension is absent, cannot fuel oxidative metabolism. Instead, it drives protein lactylation in ischemic tissues, a post-translational modification that exacerbates cellular injury (158).
The neuroprotective effects of exogenous lactate are highly dependent on both dose and timing. Following oxygen-glucose deprivation, 4 mM lactate markedly attenuates hippocampal neuronal death, whereas 20 mM exerts neurotoxic effects (159). Optimal neuroprotection is observed at ~10 mM, with concentrations below this threshold insufficient to alleviate the metabolic crisis associated with cerebral ischemia (160). Similarly, perfusion of isolated mouse hearts with 20 mmol/l L-lactate for only 15 min at reperfusion onset reduced infarct size from 44.74 to 21.46% (161). Clinical and preclinical studies suggest that lactate-enriched solutions confer notable benefits in traumatic brain injury and myocardial ischemia, including reductions in cognitive deficits, improvements in cerebral blood flow and attenuation of reperfusion injuries through mechanisms such as anti-inflammatory effects and provision of alternative energy substrates (156,162). Moreover, lactate-enriched solutions exhibit promising effects by producing a positive inotropic response in both healthy individuals and patients with acute heart failure, and they may further mitigate reperfusion injuries following myocardial ischemia (163-165). Collectively, these findings underscore the context-dependent duality of lactate: It acts as a metabolic burden in chronic neurodegeneration and hypoxic conditions; however, it serves as a therapeutic asset during reperfusion when oxidative capacity is restored.
Elevated serum lactate levels serve as a prognostic indicator in sepsis, with higher concentrations correlating with increased mortality (166). A mechanistic study has further implicated lactate in ferroptosis-mediated lung injury during sepsis progression (82). However, Besnier et al (104) reported that hypertonic sodium lactate solutions protected against cardiac dysfunction, microcirculatory impairment and vascular leakage in septic animals, while simultaneously attenuating inflammation and promoting ketogenesis. This paradox, wherein lactate exhibits context-dependent anti-inflammatory effects while promoting ferroptosis, represents a significant challenge for its clinical translation.
Given its role in catalyzing the conversion of pyruvate to lactate in glycolytic cancer cells and the observation that LDHA deficiency results in only relatively mild symptoms, such as exertional myopathy, LDHA represents a pivotal and safe therapeutic target (150,167). Inhibitors, including gossypol (AT-101) and its derivative FX-11, have demonstrated efficacy in preclinical and early clinical studies, while other molecules, such as galloflavin, oxamate, quinoline derivatives and N-hydroxyindole-based compounds, demonstrate potential in suppressing tumor progression (168-173). Despite these promising findings, challenges, including isoform-specific expression (LDHA vs. LDHB) and tumor metabolic plasticity, have limited their broader application (174). Targeting other metabolic enzymes, including hexokinase 2 and PDH kinase 1, using agents such as 2-deoxy-D-glucose and dichloroacetate (DCA), has demonstrated potential in reducing lactate production and suppressing tumor growth (175,176).
Nanomedicines, owing to their nanoscale size and multi-functional design, facilitate precise tumor targeting, controlled drug release and enhanced bioavailability, thereby offering innovative solutions to modulate tumor lactate metabolism and amplify antitumor efficacy (150). For example, Zhang et al (177) developed PMVL, a lonidamine-loaded nanoplatform that enhances ferroptosis and immune activation through dual inhibition of glycolysis and the PPP, effectively reducing lactate production.
Lactate oxidase provides a direct strategy by catalyzing lactate oxidation to pyruvate and H2O2, reducing lactate levels, exacerbating tumor hypoxia and increasing oxidative stress (150). This approach enhances antitumor immune responses and activates hypoxia-sensitive prodrugs. A recent study reported that an engineered biohybrid of DH5α Escherichia coli with hypoxia-inducible lactate oxidase and iron-doped zeolitic imidazolate framework-8 nanoparticles enables targeted lactate depletion, immune activation and ferroptosis, significantly inhibiting tumor growth and metastasis (178). However, systemic administration of lactate oxidase carries the risk of off-target H2O2 toxicity, underscoring the need for tumor-specific delivery systems to improve safety and therapeutic outcomes.
Inhibiting MCT-mediated lactate transport offers another avenue to disrupt tumor metabolic symbiosis. MCT1 inhibition compels cancer cells to compete for glucose, thereby inducing apoptosis in hypoxic cancer cells, while MCT4 inhibition triggers intracellular acidosis under hypoxic conditions (142). Early inhibitors (hydroxycinnamate and lonidamine) lacked isoform selectivity, limiting their clinical potential (179,180). However, next-generation inhibitors, such as AZD3965, AR-C155858, SR13800 and VB124, demonstrate improved specificity, with AZD3965 progressing to phase I trials (111,181-183). Additionally, CD147-targeting therapies, including anti-CD147 antibodies, modulate MCT1/4 surface expression, although off-target effects remain a concern (184). Certain statin drugs demonstrate MCT4 inhibitory activity, with lipophilic statins exhibiting greater potency compared with their hydrophilic counterparts (185,186). Recently, Chen et al (187) developed a folic acid-decorated, manganese dioxide-coated mesoporous silica nanoparticle to co-deliver fluvastatin sodium (MCT4 inhibitor) and metformin, effectively targeting tumor lactate metabolism by promoting lactate production and inhibiting lactate efflux, thereby exacerbating intracellular acidosis and inducing cancer cell death. This nanomedicine demonstrated enhanced antitumor efficacy, suppressed tumor cell migration and suppressed metastasis by disrupting the MCT4-mediated lactate shuttling in breast cancer models. Targeting lactate metabolism has therefore emerged as a promising therapeutic approach in oncology, and the aforementioned strategies are summarized in Table II.
Therapeutic approaches aimed at modulating lactate metabolism have demonstrated some potential in regulating ferroptosis in neurodegenerative diseases, IRI and sepsis. However, their effectiveness is critically context-dependent, varying according to the metabolic characteristics of the specific tissues and cell types (48,188,189). Approaches that suppress lactate production, such as using PDK inhibitors (DCA and thiamine), have been demonstrated to alleviate organ damage and mitigate ferroptosis in sepsis by reversing the pathological Warburg effect-like state, thereby improving mitochondrial function and reducing ROS accumulation (189,190). Conversely, in myocardial IRI, activation of LDHA has been found to phosphorylate and stabilize the ferroptosis-suppressing enzyme GPX4 via its kinase activity, suggesting that direct inhibition of LDHA could be detrimental (48).
Enhancing lactate clearance has also emerged as a promising strategy in sepsis, as increased lactate elimination has been positively correlated with improved clinical outcomes in patients with sepsis (191). Beyond merely correcting metabolic acidosis, this approach also restricts lactate-driven histone lactylation, which influences the epigenetic regulation of TFRC expression and ferroptosis susceptibility (73,192). By contrast, controlled utilization of lactate, such as lactate post-conditioning, has been demonstrated to exert neuroprotective effects in cerebral IRI (156).
Inhibition of MCT-mediated lactate shuttling has demonstrated cell-type-specific therapeutic potential. For instance, blocking lactate efflux via MCT4, using agents such as VB124, can redirect pyruvate toward mitochondrial oxidation, thereby conferring cardioprotection in myocardial IRI (193). Meanwhile, MCT1 inhibitors, including AZD3965, have been revealed to regulate immune responses and significantly reduce mortality in septic mice by promoting neutrophil apoptosis (194). However, MCT1-mediated lactate transport remains crucial for sustaining energy homeostasis and preserving tissue health in neurons and pulmonary epithelial cells under comparable conditions (195,196).
In summary, modulation of lactate metabolism represents a promising avenue for managing ferroptosis in disease. Nevertheless, future translational studies must focus on developing drugs with high tissue specificity and isoform-selective activity to account for the metabolic duality inherent to lactate biology.
Regulated cell death and metabolic homeostasis constitute fundamental determinants of cellular development and growth. In the present review, comprehensive analysis revealed a particularly intriguing phenomenon: Lactate exhibits diametrically contrasting effects on ferroptosis, promoting cell death in normal cells while conferring protection in tumor cells. This apparent paradox, largely attributed to context-dependent mechanisms, has remained incompletely understood.
Several critical research gaps remain to be addressed. The molecular mechanism underlying lactate-induced downregulation of ACSL4 via MCT1 remains incompletely elucidated, although accumulating evidence suggests a potential involvement of the lactate-induced STAT3 signaling pathway (197,198). Notably, MCT1 blockade in HCC cells promotes ferroptosis without observable alterations in GPX4 and FSP1 expression levels, while lactate uptake in melanoma cells via MCT1 elevates NADH, NADPH, GPX4 and FSP1. This paradox may be explained by enzymatic dysfunction rather than transcriptional downregulation. In neurodegenerative contexts, the notable metabolic differences between immature and mature neurons indicate that the lactate-ferroptosis axis during neurogenesis is largely unexplored, with potential implications for neurodevelopmental disorders (199-201). Similarly, in IRI and sepsis, studies lack systematic investigation of temporal exposure windows, a critical limitation given that prolonged lactate exposure may trigger adaptive responses or cumulative toxic effects. While Wu et al (202) recently addressed lactate-regulated ferroptosis in cancer contexts and acknowledged the context-dependent nature of this phenomenon, their analysis did not elucidate the mechanistic determinants underlying this differential regulation. The present review addresses this knowledge gap by identifying three key context-dependent mechanisms governing the dual role of lactate: Metabolic enzyme expression patterns, pH homeostasis capacity and antioxidant defense capabilities. Additionally, the present review explored the therapeutic potential of lactate in non-tumor diseases and elucidated its crosstalk with other cell death pathways. Although lactylation modifications have been demonstrated to regulate ferroptosis across diverse experimental models, it remains unclear whether these mechanisms are restricted to the specific disease contexts and models studied or can be universally extended to all cellular systems.
The therapeutic implications are profound yet complex. Strategies aimed at disrupting lactate-mediated ferroptosis resistance in tumors while reducing lactate-mediated ferroptosis in sepsis may offer selective therapeutic windows. However, in IRI, the dual nature of lactate, providing a protective effect through energy supply while simultaneously promoting ferroptotic damage, poses significant challenges for its clinical application. In the nervous system, interventions harnessing the neuroprotective properties of lactate while mitigating pro-ferroptotic effects offer novel approaches (203).
Future research directions must prioritize three key directions: First, investigating the triggering factors of lactate-ferroptosis interactions across diverse biological contexts to enable the design of tissue-specific lactate-targeting therapeutic interventions; second, developing spatiotemporally resolved analytical techniques to identify predictive biomarkers that can distinguish between different ferroptotic responses; and third, leveraging advanced imaging technologies, metabolomics approaches and novel biosensors to provide essential tools for tracking lactate-ferroptosis dynamics in vivo.
In conclusion, the lactate-ferroptosis network represents a sophisticated cellular regulatory mechanism that integrates metabolic status with cell fate determination. Translating these mechanistic insights into targeted therapeutic strategies across diverse disease contexts, from cancer to neurodegeneration, constitutes a pivotal future challenge, requiring interdisciplinary collaboration and innovative experimental approaches.
Not applicable.
QY and HY discussed the structure of the article and completed the manuscript; YK, JH and LY provided technical and material support and completed the manuscript. XL provided suggestions for revision. All authors read and approved the final version of the manuscript. Data authentication is not applicable.
Not applicable.
Not applicable.
The authors declare that they have no competing interests.
During the preparation of this work, artificial intelligence tools (Claude 4.5 Sonnet) to improve the readability and language of the manuscript or to generate images, and subsequently, the authors revised and edited the content produced by the artificial intelligence tools as necessary, taking full responsibility for the ultimate content of the present manuscript.
Not applicable.
This work was supported by National Natural Science Foundation of China (grant no. 824B2063).
|
Dixon SJ, Lemberg KM, Lamprecht MR, Skouta R, Zaitsev EM, Gleason CE, Patel DN, Bauer AJ, Cantley AM, Yang WS, et al: Ferroptosis: An iron-dependent form of nonapoptotic cell death. Cell. 149:1060–1072. 2012. View Article : Google Scholar : PubMed/NCBI | |
|
Dixon SJ and Olzmann JA: The cell biology of ferroptosis. Nat Rev Mol Cell Biol. 25:424–442. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Mishima E, Nakamura T, Doll S, Proneth B, Fedorova M, Pratt DA, Friedmann Angeli JP, Dixon SJ, Wahida A and Conrad M: Recommendations for robust and reproducible research on ferroptosis. Nat Rev Mol Cell Biol. 26:615–630. 2025. View Article : Google Scholar : PubMed/NCBI | |
|
Zhou Q, Meng Y, Li D, Yao L, Le J, Liu Y, Sun Y, Zeng F, Chen X and Deng G: Ferroptosis in cancer: From molecular mechanisms to therapeutic strategies. Signal Transduct Target Ther. 9:552024. View Article : Google Scholar : PubMed/NCBI | |
|
Li X, Yang Y, Zhang B, Lin X, Fu X, An Y, Zou Y, Wang JX, Wang Z and Yu T: Lactate metabolism in human health and disease. Signal Transduct Target Ther. 7:3052022. View Article : Google Scholar : PubMed/NCBI | |
|
Brooks GA: The science and translation of lactate shuttle theory. Cell Metab. 27:757–785. 2018. View Article : Google Scholar : PubMed/NCBI | |
|
Felmlee MA, Jones RS, Rodriguez-Cruz V, Follman KE and Morris ME: Monocarboxylate transporters (SLC16): Function, regulation, and role in health and disease. Pharmacol Rev. 72:466–485. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Brown TP and Ganapathy V: Lactate/GPR81 signaling and proton motive force in cancer: Role in angiogenesis, immune escape, nutrition, and Warburg phenomenon. Pharmacol Ther. 206:1074512020. View Article : Google Scholar | |
|
Zhang D, Tang Z, Huang H, Zhou G, Cui C, Weng Y, Liu W, Kim S, Lee S, Perez-Neut M, et al: Metabolic regulation of gene expression by histone lactylation. Nature. 574:575–580. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Hua S, Jeong HN, Dimapasoc LM, Kang I, Han C, Choi JS, Lebrilla CB and An HJ: Isomer-specific LC/MS and LC/MS/MS profiling of the mouse serum N-glycome revealing a number of novel sialylated N-glycans. Anal Chem. 85:4636–4643. 2013. View Article : Google Scholar : PubMed/NCBI | |
|
Li H, Sun L, Gao P and Hu H: Lactylation in cancer: Current understanding and challenges. Cancer Cell. 42:1803–1807. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Liu YQ, Yang Q and He GW: Post-translational acylation of proteins in cardiac hypertrophy. Nat Rev Cardiol. 22:944–960. 2025. View Article : Google Scholar : PubMed/NCBI | |
|
Conrad M and Pratt DA: The chemical basis of ferroptosis. Nat Chem Biol. 15:1137–1147. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Liang D, Minikes AM and Jiang X: Ferroptosis at the intersection of lipid metabolism and cellular signaling. Mol Cell. 82:2215–2227. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Yang WS, SriRamaratnam R, Welsch ME, Shimada K, Skouta R, Viswanathan VS, Cheah JH, Clemons PA, Shamji AF, Clish CB, et al: Regulation of ferroptotic cancer cell death by GPX4. Cell. 156:317–331. 2014. View Article : Google Scholar : PubMed/NCBI | |
|
Badgley MA, Kremer DM, Maurer HC, DelGiorno KE, Lee HJ, Purohit V, Sagalovskiy IR, Ma A, Kapilian J, Firl CEM, et al: Cysteine depletion induces pancreatic tumor ferroptosis in mice. Science. 368:85–89. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Bersuker K, Hendricks JM, Li Z, Magtanong L, Ford B, Tang PH, Roberts MA, Tong B, Maimone TJ, Zoncu R, et al: The CoQ oxidoreductase FSP1 acts parallel to GPX4 to inhibit ferroptosis. Nature. 575:688–692. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Doll S, Freitas FP, Shah R, Aldrovandi M, da Silva MC, Ingold I, Goya Grocin A, Xavier da Silva TN, Panzilius E, Scheel CH, et al: FSP1 is a glutathione-independent ferroptosis suppressor. Nature. 575:693–698. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Mao C, Liu X, Zhang Y, Lei G, Yan Y, Lee H, Koppula P, Wu S, Zhuang L, Fang B, et al: DHODH-mediated ferroptosis defence is a targetable vulnerability in cancer. Nature. 593:586–590. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Wang D, Liang W, Huo D, Wang H, Wang Y, Cong C, Zhang C, Yan S, Gao M, Su X, et al: SPY1 inhibits neuronal ferroptosis in amyotrophic lateral sclerosis by reducing lipid peroxidation through regulation of GCH1 and TFR1. Cell Death Differ. 30:369–382. 2023. View Article : Google Scholar : | |
|
Kraft VAN, Bezjian CT, Pfeiffer S, Ringelstetter L, Müller C, Zandkarimi F, Merl-Pham J, Bao X, Anastasov N, Kössl J, et al: GTP cyclohydrolase 1/tetrahydrobiopterin counteract ferroptosis through lipid remodeling. ACS Cent Sci. 6:41–53. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Kappler A, Bryce C, Mansor M, Lueder U, Byrne JM and Swanner ED: An evolving view on biogeochemical cycling of iron. Nat Rev Microbiol. 19:360–374. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Cañeque T, Baron L, Müller S, Carmona A, Colombeau L, Versini A, Solier S, Gaillet C, Sindikubwabo F, Sampaio JL, et al: Activation of lysosomal iron triggers ferroptosis in cancer. Nature. 642:492–500. 2025. View Article : Google Scholar : PubMed/NCBI | |
|
Magtanong L, Ko PJ and Dixon SJ: Emerging roles for lipids in non-apoptotic cell death. Cell Death Differ. 23:1099–1109. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Schwörer S, Vardhana SA and Thompson CB: Cancer metabolism drives a stromal regenerative response. Cell Metab. 29:576–591. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Won W, Bhalla M, Lee JH and Lee CJ: Astrocytes as key regulators of neural signaling in health and disease. Annu Rev Neurosci. 48:251–276. 2025. View Article : Google Scholar : PubMed/NCBI | |
|
Certo M, Tsai CH, Pucino V, Ho PC and Mauro C: Lactate modulation of immune responses in inflammatory versus tumour microenvironments. Nat Rev Immunol. 21:151–161. 2021. View Article : Google Scholar | |
|
Pavlova NN and Thompson CB: The emerging hallmarks of cancer metabolism. Cell Metab. 23:27–47. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Rabinowitz JD and Enerbäck S: Lactate: The ugly duckling of energy metabolism. Nat Metab. 2:566–571. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Wang Y and Patti GJ: The Warburg effect: A signature of mitochondrial overload. Trends Cell Biol. 33:1014–1020. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Hui S, Ghergurovich JM, Morscher RJ, Jang C, Teng X, Lu W, Esparza LA, Reya T, Zhan L, Yanxiang Guo J, et al: Glucose feeds the TCA cycle via circulating lactate. Nature. 551:115–118. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Faubert B, Li KY, Cai L, Hensley CT, Kim J, Zacharias LG, Yang C, Do QN, Doucette S, Burguete D, et al: Lactate metabolism in human lung tumors. Cell. 171:358–371.e9. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Wang Y, Huang Y, Yang J, Zhou FQ, Zhao L and Zhou H: Pyruvate is a prospective alkalizer to correct hypoxic lactic acidosis. Mil Med Res. 5:132018.PubMed/NCBI | |
|
DeBerardinis RJ, Mancuso A, Daikhin E, Nissim I, Yudkoff M, Wehrli S and Thompson CB: Beyond aerobic glycolysis: Transformed cells can engage in glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis. Proc Natl Acad Sci USA. 104:19345–19350. 2007. View Article : Google Scholar : PubMed/NCBI | |
|
Dang CV: MYC on the path to cancer. Cell. 149:22–35. 2012. View Article : Google Scholar : PubMed/NCBI | |
|
Ivashkiv LB: The hypoxia-lactate axis tempers inflammation. Nat Rev Immunol. 20:85–86. 2020. View Article : Google Scholar : | |
|
Taylor CT and Scholz CC: The effect of HIF on metabolism and immunity. Nat Rev Nephrol. 18:573–587. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Wang GL and Semenza GL: General involvement of hypoxia-inducible factor 1 in transcriptional response to hypoxia. Proc Natl Acad Sci USA. 90:4304–4308. 1993. View Article : Google Scholar : PubMed/NCBI | |
|
Silagi ES, Schipani E, Shapiro IM and Risbud MV: The role of HIF proteins in maintaining the metabolic health of the intervertebral disc. Nat Rev Rheumatol. 17:426–439. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Iyer NV, Kotch LE, Agani F, Leung SW, Laughner E, Wenger RH, Gassmann M, Gearhart JD, Lawler AM, Yu AY and Semenza GL: Cellular and developmental control of O2 homeostasis by hypoxia-inducible factor 1 alpha. Genes Dev. 12:149–162. 1998. View Article : Google Scholar : PubMed/NCBI | |
|
Seagroves TN, Ryan HE, Lu H, Wouters BG, Knapp M, Thibault P, Laderoute K and Johnson RS: Transcription factor HIF-1 is a necessary mediator of the pasteur effect in mammalian cells. Mol Cell Biol. 21:3436–3444. 2001. View Article : Google Scholar : PubMed/NCBI | |
|
Papandreou I, Cairns RA, Fontana L, Lim AL and Denko NC: HIF-1 mediates adaptation to hypoxia by actively downregulating mitochondrial oxygen consumption. Cell Metab. 3:187–197. 2006. View Article : Google Scholar : PubMed/NCBI | |
|
Kim JW, Tchernyshyov I, Semenza GL and Dang CV: HIF-1-mediated expression of pyruvate dehydrogenase kinase: A metabolic switch required for cellular adaptation to hypoxia. Cell Metab. 3:177–185. 2006. View Article : Google Scholar : PubMed/NCBI | |
|
Claps G, Faouzi S, Quidville V, Chehade F, Shen S, Vagner S and Robert C: The multiple roles of LDH in cancer. Nat Rev Clin Oncol. 19:749–762. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Ždralević M, Brand A, Di Ianni L, Dettmer K, Reinders J, Singer K, Peter K, Schnell A, Bruss C, Decking SM, et al: Double genetic disruption of lactate dehydrogenases A and B is required to ablate the 'Warburg effect' restricting tumor growth to oxidative metabolism. J Biol Chem. 293:15947–15961. 2018. View Article : Google Scholar | |
|
Kim EY, Chung TW, Han CW, Park SY, Park KH, Jang SB and Ha KT: A novel lactate dehydrogenase inhibitor, 1-(Phenylseleno)-4-(Trifluoromethyl) benzene, suppresses tumor growth through apoptotic cell death. Sci Rep. 9:39692019. View Article : Google Scholar : PubMed/NCBI | |
|
Park JS, Saeed K, Jo MH, Kim MW, Lee HJ, Park CB, Lee G and Kim MO: LDHB deficiency promotes mitochondrial dysfunction mediated oxidative stress and neurodegeneration in adult mouse brain. Antioxidants (Basel). 11:2612022. View Article : Google Scholar : PubMed/NCBI | |
|
Yin M, Li S, Liu M, Zhu W, Chen Y, Qiu W, Li Q, Li Y, Chen J, Zhou Y, et al: GUCY1A1-LDHA axis suppresses ferroptosis in cardiac ischemia-reperfusion injury. Circ Res. 137:986–1005. 2025. View Article : Google Scholar : PubMed/NCBI | |
|
Zhao L, Deng H, Zhang J, Zamboni N, Yang H, Gao Y, Yang Z, Xu D, Zhong H, van Geest G, et al: Lactate dehydrogenase B noncanonically promotes ferroptosis defense in KRAS-driven lung cancer. Cell Death Differ. 32:632–645. 2025. View Article : Google Scholar : | |
|
Doherty JR, Yang C, Scott KE, Cameron MD, Fallahi M, Li W, Hall MA, Amelio AL, Mishra JK, Li F, et al: Blocking lactate export by inhibiting the Myc target MCT1 disables glycolysis and glutathione synthesis. Cancer Res. 74:908–920. 2014. View Article : Google Scholar : | |
|
Wang N, Wang W, Wang X, Mang G, Chen J, Yan X, Tong Z, Yang Q, Wang M, Chen L, et al: Histone lactylation boosts reparative gene activation post-myocardial infarction. Circ Res. 131:893–908. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Ahmed K, Tunaru S, Tang C, Müller M, Gille A, Sassmann A, Hanson J and Offermanns S: An autocrine lactate loop mediates insulin-dependent inhibition of lipolysis through GPR81. Cell Metab. 11:311–319. 2010. View Article : Google Scholar : PubMed/NCBI | |
|
Roland CL, Arumugam T, Deng D, Liu SH, Philip B, Gomez S, Burns WR, Ramachandran V, Wang H, Cruz-Monserrate Z and Logsdon CD: Cell surface lactate receptor GPR81 is crucial for cancer cell survival. Cancer Res. 74:5301–5310. 2014. View Article : Google Scholar : PubMed/NCBI | |
|
Lee YJ, Shin KJ, Park SA, Park KS, Park S, Heo K, Seo YK, Noh DY, Ryu SH and Suh PG: G-protein-coupled receptor 81 promotes a malignant phenotype in breast cancer through angiogenic factor secretion. Oncotarget. 7:70898–70911. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Feng J, Yang H, Zhang Y, Wei H, Zhu Z, Zhu B, Yang M, Cao W, Wang L and Wu Z: Tumor cell-derived lactate induces TAZ-dependent upregulation of PD-L1 through GPR81 in human lung cancer cells. Oncogene. 36:5829–5839. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Luo M, Zhu J, Ren J, Tong Y, Wang L, Ma S and Wang J: Lactate increases tumor malignancy by promoting tumor small extracellular vesicles production via the GPR81-cAMP-PKA-HIF-1α axis. Front Oncol. 12:10365432022. View Article : Google Scholar | |
|
Zhao Y, Li M, Yao X, Fei Y, Lin Z, Li Z, Cai K, Zhao Y and Luo Z: HCAR1/MCT1 regulates tumor ferroptosis through the lactate-mediated AMPK-SCD1 activity and its therapeutic implications. Cell Rep. 33:1084872020. View Article : Google Scholar : PubMed/NCBI | |
|
Yang P, Li H, Sun M, Guo X, Liao Y, Hu M, Ye P and Liu R: Zinc deficiency drives ferroptosis resistance by lactate production in esophageal squamous cell carcinoma. Free Radic Biol Med. 213:512–522. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Benjamin D, Robay D, Hindupur SK, Pohlmann J, Colombi M, El-Shemerly MY, Maira SM, Moroni C, Lane HA and Hall MN: Dual inhibition of the lactate transporters MCT1 and MCT4 Is synthetic lethal with metformin due to NAD+ depletion in cancer cells. Cell Rep. 25:3047–3058.e4. 2018. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang S, Liu W, Ganz T and Liu S: Exploring the relationship between hyperlactatemia and anemia. Trends Endocrinol Metab. 35:300–307. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Sim M, Garvican-Lewis LA, Cox GR, Govus A, McKay AKA, Stellingwerff T and Peeling P: Iron considerations for the athlete: A narrative review. Eur J Appl Physiol. 119:1463–1478. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Nemeth E, Tuttle MS, Powelson J, Vaughn MB, Donovan A, Ward DM, Ganz T and Kaplan J: Hepcidin regulates cellular iron efflux by binding to ferroportin and inducing its internalization. Science. 306:2090–2093. 2004. View Article : Google Scholar : PubMed/NCBI | |
|
Liu W, Zhang S, Li Q, Wu Y, Jia X, Feng W, Li Z, Shi Y, Hou Q, Ma J, et al: Lactate modulates iron metabolism by binding soluble adenylyl cyclase. Cell Metab. 35:1597–1612.e6. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Liu W, Wu Y, Wei H, Ma J, Feng W, Yang Q, Zhang S, Ganz T and Liu S: Lactate administration improves laboratory parameters in murine models of iron overload. Blood. 143:1045–1049. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang J, Wang Y, Fan M, Guan Y, Zhang W, Huang F, Zhang Z, Li X, Yuan B, Liu W, et al: Reactive oxygen species regulation by NCF1 governs ferroptosis susceptibility of Kupffer cells to MASH. Cell Metab. 36:1745–1763.e6. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Davaanyam D, Lee H, Seol SI, Oh SA, Kim SW and Lee JK: HMGB1 induces hepcidin upregulation in astrocytes and causes an acute iron surge and subsequent ferroptosis in the postischemic brain. Exp Mol Med. 55:2402–2416. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Davaanyam D, Seol SI, Oh SA, Lee H and Lee JK: Hepatocyte activation and liver injury following cerebral ischemia promote HMGB1-mediated hepcidin upregulation in hepatocytes and regulation of systemic iron levels. Exp Mol Med. 56:2171–2183. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Gao M, Monian P, Quadri N, Ramasamy R and Jiang X: Glutaminolysis and transferrin regulate ferroptosis. Mol Cell. 59:298–308. 2015. View Article : Google Scholar : PubMed/NCBI | |
|
Hou W, Xie Y, Song X, Sun X, Lotze MT, Zeh HJ III, Kang R and Tang D: Autophagy promotes ferroptosis by degradation of ferritin. Autophagy. 12:1425–1428. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Li A, Gong Z, Long Y, Li Y, Liu C, Lu X, Li Q, He X, Lu H, Wu K, et al: Lactylation of LSD1 is an acquired epigenetic vulnerability of BRAFi/MEKi-resistant melanoma. Dev Cell. 60:1974–1990.e11. 2025. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang K, Guo L, Li X, Hu Y and Luo N: Cancer-associated fibroblasts promote doxorubicin resistance in triple-negative breast cancer through enhancing ZFP64 histone lactylation to regulate ferroptosis. J Transl Med. 23:2472025. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang L, Wang X, Che W, Zhou S and Feng Y: METTL3 silenced inhibited the ferroptosis development via regulating the TFRC levels in the intracerebral hemorrhage progression. Brain Res. 1811:1483732023. View Article : Google Scholar : PubMed/NCBI | |
|
Gong F, Zheng X, Xu W, Xie R, Liu W, Pei L, Zhong M, Shi W, Qu H, Mao E, et al: H3K14la drives endothelial dysfunction in sepsis-induced ARDS by promoting SLC40A1/transferrin-mediated ferroptosis. MedComm. 6:e700492025. View Article : Google Scholar : PubMed/NCBI | |
|
Martinez-Outschoorn UE, Peiris-Pagés M, Pestell RG, Sotgia F and Lisanti MP: Cancer metabolism: A therapeutic perspective. Nat Rev Clin Oncol. 14:11–31. 2017. View Article : Google Scholar | |
|
Zheng J and Conrad M: Ferroptosis: When metabolism meets cell death. Physiol Rev. 105:651–706. 2025. View Article : Google Scholar | |
|
de Kivit S, Mensink M, Kostidis S, Derks RJE, Zaal EA, Heijink M, Verleng LJ, de Vries E, Schrama E, Blomberg N, et al: Immune suppression by human thymus-derived effector Tregs relies on glucose/lactate-fueled fatty acid synthesis. Cell Rep. 43:1146812024. View Article : Google Scholar : PubMed/NCBI | |
|
Li C, Dong X, Du W, Shi X, Chen K, Zhang W and Gao M: LKB1-AMPK axis negatively regulates ferroptosis by inhibiting fatty acid synthesis. Signal Transduct Target Ther. 5:1872020. View Article : Google Scholar : PubMed/NCBI | |
|
Hardie DG, Ross FA and Hawley SA: AMPK: A nutrient and energy sensor that maintains energy homeostasis. Nat Rev Mol Cell Biol. 13:251–262. 2012. View Article : Google Scholar : PubMed/NCBI | |
|
Song X, Liu J, Kuang F, Chen X, Zeh HJ III, Kang R, Kroemer G, Xie Y and Tang D: PDK4 dictates metabolic resistance to ferroptosis by suppressing pyruvate oxidation and fatty acid synthesis. Cell Rep. 34:1087672021. View Article : Google Scholar : PubMed/NCBI | |
|
Sondermeijer BM, Battjes S, van Dijk TH, Ackermans MT, Serlie MJ, Nieuwdorp M, Groen AK, Dallinga-Thie GM and Stroes ES: Lactate increases hepatic secretion of VLDL-triglycerides in humans. Atherosclerosis. 228:443–450. 2013. View Article : Google Scholar : PubMed/NCBI | |
|
Fan Z, Ye M, Liu D, Zhou W, Zeng T, He S and Li Y: Lactate drives the ESM1-SCD1 axis to inhibit the antitumor CD8+ T-cell response by activating the Wnt/β-catenin pathway in ovarian cancer cells and inducing cisplatin resistance. Int Immunopharmacol. 137:1124612024. View Article : Google Scholar | |
|
Wu D, Spencer CB, Ortoga L, Zhang H and Miao C: Histone lactylation-regulated METTL3 promotes ferroptosis via m6A-modification on ACSL4 in sepsis-associated lung injury. Redox Biol. 74:1031942024. View Article : Google Scholar : PubMed/NCBI | |
|
Pope LE and Dixon SJ: Regulation of ferroptosis by lipid metabolism. Trends Cell Biol. 33:1077–1087. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Tilton WM, Seaman C, Carriero D and Piomelli S: Regulation of glycolysis in the erythrocyte: Role of the lactate/pyruvate and NAD/NADH ratios. J Lab Clin Med. 118:146–152. 1991.PubMed/NCBI | |
|
Quinn WJ III, Jiao J, TeSlaa T, Stadanlick J, Wang Z, Wang L, Akimova T, Angelin A, Schäfer PM, Cully MD, et al: Lactate limits T cell proliferation via the NAD(H) redox state. Cell Rep. 33:1085002020. View Article : Google Scholar : PubMed/NCBI | |
|
Luengo A, Li Z, Gui DY, Sullivan LB, Zagorulya M, Do BT, Ferreira R, Naamati A, Ali A, Lewis CA, et al: Increased demand for NAD+ relative to ATP drives aerobic glycolysis. Mol Cell. 81:691–707.e6. 2021. View Article : Google Scholar | |
|
Corkey BE and Deeney JT: The redox communication network as a regulator of metabolism. Front Physiol. 11:5677962020. View Article : Google Scholar : PubMed/NCBI | |
|
Yang S and Lian G: ROS and diseases: Role in metabolism and energy supply. Mol Cell Biochem. 467:1–12. 2020. View Article : Google Scholar : | |
|
Young A, Oldford C and Mailloux RJ: Lactate dehydrogenase supports lactate oxidation in mitochondria isolated from different mouse tissues. Redox Biol. 28:1013392020. View Article : Google Scholar | |
|
Bauzá-Thorbrügge M, Peris E, Zamani S, Micallef P, Paul A, Bartesaghi S, Benrick A and Wernstedt Asterholm I: NRF2 is essential for adaptative browning of white adipocytes. Redox Biol. 68:1029512023. View Article : Google Scholar : PubMed/NCBI | |
|
Jia L, Liao M, Mou A, Zheng Q, Yang W, Yu Z, Cui Y, Xia X, Qin Y, Chen M and Xiao B: Rheb-regulated mitochondrial pyruvate metabolism of Schwann cells linked to axon stability. Dev Cell. 56:2980–2994.e6. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Rydström J: Mitochondrial NADPH, transhydrogenase and disease. Biochim Biophys Acta. 1757:721–726. 2006. View Article : Google Scholar : PubMed/NCBI | |
|
Lin W, Lu X, Yang H, Huang L, Huang W, Tang Y, Liu S, Wang H and Zhang Y: Metabolic heterogeneity protects metastatic mucosal melanomas cells from ferroptosis. Int J Mol Med. 50:1242022. View Article : Google Scholar : PubMed/NCBI | |
|
Ying M, You D, Zhu X, Cai L, Zeng S and Hu X: Lactate and glutamine support NADPH generation in cancer cells under glucose deprived conditions. Redox Biol. 46:1020652021. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang W, Guo C, Jiang K, Ying M and Hu X: Quantification of lactate from various metabolic pathways and quantification issues of lactate isotopologues and isotopmers. Sci Rep. 7:84892017. View Article : Google Scholar : PubMed/NCBI | |
|
Huang YF, Wang G, Ding L, Bai ZR, Leng Y, Tian JW, Zhang JZ, Li YQ, Ahmad, Qin YH, et al: Lactate-upregulated NADPH-dependent NOX4 expression via HCAR1/PI3K pathway contributes to ROS-induced osteoarthritis chondrocyte damage. Redox Biol. 67:1028672023. View Article : Google Scholar : PubMed/NCBI | |
|
Taylor JP and Tse HM: The role of NADPH oxidases in infectious and inflammatory diseases. Redox Biol. 48:1021592021. View Article : Google Scholar : PubMed/NCBI | |
|
Sun F, He Y, Yang Z, Xu G, Wang R, Juan Z and Sun X: Propofol pretreatment inhibits ferroptosis and alleviates myocardial ischemia-reperfusion injury through the SLC16A13-AMPK-GPX4 pathway. Biomed Pharmacother. 179:1173452024. View Article : Google Scholar : PubMed/NCBI | |
|
Cheng F, Dou J, Yang Y, Sun S, Chen R, Zhang Z, Wei H, Li J and Wu Z: Drug-induced lactate confers ferroptosis resistance via p38-SGK1-NEDD4L-dependent upregulation of GPX4 in NSCLC cells. Cell Death Discov. 9:1652023. View Article : Google Scholar : PubMed/NCBI | |
|
Tauffenberger A, Fiumelli H, Almustafa S and Magistretti PJ: Lactate and pyruvate promote oxidative stress resistance through hormetic ROS signaling. Cell Death Dis. 10:6532019. View Article : Google Scholar : PubMed/NCBI | |
|
Zou GP, Wang T, Xiao JX, Wang XY, Jiang LP, Tou FF, Chen ZP, Qu XH and Han XJ: Lactate protects against oxidative stress-induced retinal degeneration by activating autophagy. Free Radic Biol Med. 194:209–219. 2023. View Article : Google Scholar | |
|
Ma M, Zhang Y, Pu K and Tang W: Nanomaterial-enabled metabolic reprogramming strategies for boosting antitumor immunity. Chem Soc Rev. 54:653–714. 2025. View Article : Google Scholar | |
|
Lei G, Zhuang L and Gan B: Targeting ferroptosis as a vulnerability in cancer. Nat Rev Cancer. 22:381–396. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Besnier E, Coquerel D, Kouadri G, Clavier T, Favory R, Duburcq T, Lesur O, Bekri S, Richard V, Mulder P and Tamion F: Hypertonic sodium lactate improves microcirculation, cardiac function, and inflammation in a rat model of sepsis. Crit Care. 24:3542020. View Article : Google Scholar : PubMed/NCBI | |
|
Zeng M, Niu Y, Huang J and Deng L: Advances in neutrophil extracellular traps and ferroptosis in sepsis-induced cardiomyopathy. Front Immunol. 16:15903132025. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang T, Chen L, Kueth G, Shao E, Wang X, Ha T, Williams DL, Li C, Fan M and Yang K: Lactate's impact on immune cells in sepsis: Unraveling the complex interplay. Front Immunol. 15:14834002024. View Article : Google Scholar : PubMed/NCBI | |
|
Irizarry-Caro RA, McDaniel MM, Overcast GR, Jain VG, Troutman TD and Pasare C: TLR signaling adapter BCAP regulates inflammatory to reparatory macrophage transition by promoting histone lactylation. Proc Natl Acad Sci USA. 117:30628–30638. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Hu Y, He Z, Li Z, Wang Y, Wu N, Sun H, Zhou Z, Hu Q and Cong X: Lactylation: The novel histone modification influence on gene expression, protein function, and disease. Clin Epigenetics. 16:722024. View Article : Google Scholar : PubMed/NCBI | |
|
Su F, Xie K, He X, Orbegozo D, Hosokawa K, Post EH, Donadello K, Taccone FS, Creteur J and Vincent JL: The harmful effects of hypertonic sodium lactate administration in hyperdynamic septic shock. Shock. 46:663–671. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Tang R, Xu J, Zhang B, Liu J, Liang C, Hua J, Meng Q, Yu X and Shi S: Ferroptosis, necroptosis, and pyroptosis in anticancer immunity. J Hematol Oncol. 13:1102020. View Article : Google Scholar : PubMed/NCBI | |
|
Fang Y, Liu W, Tang Z, Ji X, Zhou Y, Song S, Tian M, Tao C, Huang R, Zhu G, et al: Monocarboxylate transporter 4 inhibition potentiates hepatocellular carcinoma immunotherapy through enhancing T cell infiltration and immune attack. Hepatology. 77:109–123. 2023. View Article : Google Scholar | |
|
Hagihara H, Shoji H, Otabi H, Toyoda A, Katoh K, Namihira M and Miyakawa T: Protein lactylation induced by neural excitation. Cell Rep. 37:1098202021. View Article : Google Scholar : PubMed/NCBI | |
|
Sun K, Shi Y, Yan C, Wang S, Han L, Li F, Xu X, Wang Y, Sun J, Kang Z and Shi J: Glycolysis-derived lactate induces ACSL4 expression and lactylation to activate ferroptosis during intervertebral disc degeneration. Adv Sci (Weinh). 12:e24161492025. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang T, Huang X, Feng S and Shao H: Lactate-dependent HIF1A transcriptional activation exacerbates severe acute pancreatitis through the ACSL4/LPCAT3/ALOX15 pathway induced ferroptosis. J Cell Biochem. 126:e306872025. View Article : Google Scholar | |
|
Wu Q, You L, Nepovimova E, Heger Z, Wu W, Kuca K and Adam V: Hypoxia-inducible factors: Master regulators of hypoxic tumor immune escape. J Hematol Oncol. 15:772022. View Article : Google Scholar : PubMed/NCBI | |
|
Deng J, Li Y, Yin L, Liu S, Li Y, Liao W, Mu L, Luo X and Qin J: Histone lactylation enhances GCLC expression and thus promotes chemoresistance of colorectal cancer stem cells through inhibiting ferroptosis. Cell Death Dis. 16:1932025. View Article : Google Scholar : PubMed/NCBI | |
|
Huang J, Xie H, Li J, Huang X, Cai Y, Yang R, Yang D, Bao W, Zhou Y, Li T and Lu Q: Histone lactylation drives liver cancer metastasis by facilitating NSF1-mediated ferroptosis resistance after microwave ablation. Redox Biol. 81:1035532025. View Article : Google Scholar : PubMed/NCBI | |
|
Yuan J, Yang M, Wu Z, Wu J, Zheng K, Wang J, Zeng Q, Chen M, Lv T, Shi Y, et al: The lactate-primed KAT8-PCK2 axis exacerbates hepatic ferroptosis during ischemia/reperfusion injury by reprogramming OXSM-dependent mitochondrial fatty acid synthesis. Adv Sci (Weinh). 12:e24141412025. View Article : Google Scholar | |
|
An X, He J, Xie P, Li C, Xia M, Guo D, Bi B, Wu G, Xu J, Yu W and Ren Z: The effect of tau K677 lactylation on ferritinophagy and ferroptosis in Alzheimer's disease. Free Radic Biol Med. 224:685–706. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
She H, Hu Y, Zhao G, Du Y, Wu Y, Chen W, Li Y, Wang Y, Tan L, Zhou Y, et al: Dexmedetomidine ameliorates myocardial ischemia-reperfusion injury by inhibiting MDH2 lactylation via regulating metabolic reprogramming. Adv Sci (Weinh). 11:e24094992024. View Article : Google Scholar : PubMed/NCBI | |
|
Niu K, Chen Z, Li M, Ma G, Deng Y, Zhang J, Wei D, Wang J and Zhao Y: NSUN2 lactylation drives cancer cell resistance to ferroptosis through enhancing GCLC-dependent glutathione synthesis. Redox Biol. 79:1034792025. View Article : Google Scholar : PubMed/NCBI | |
|
Yu Y, Huang X, Liang C and Zhang P: Evodiamine impairs HIF1A histone lactylation to inhibit Sema3A-mediated angiogenesis and PD-L1 by inducing ferroptosis in prostate cancer. Eur J Pharmacol. 957:1760072023. View Article : Google Scholar : PubMed/NCBI | |
|
Xiong J, Ge X, Pan D, Zhu Y, Zhou Y, Gao Y, Wang H, Wang X, Gu Y, Ye W, et al: Metabolic reprogramming in astrocytes prevents neuronal death through a UCHL1/PFKFB3/H4K8la positive feedback loop. Cell Death Differ. 32:1214–1230. 2025. View Article : Google Scholar : PubMed/NCBI | |
|
Deng H, Zhao L, Ge H, Gao Y, Fu Y, Lin Y, Masoodi M, Losmanova T, Medová M, Ott J, et al: Ubiquinol-mediated suppression of mitochondria-associated ferroptosis is a targetable function of lactate dehydrogenase B in cancer. Nat Commun. 16:25972025. View Article : Google Scholar : PubMed/NCBI | |
|
Moreira JD, Hamraz M, Abolhassani M, Bigan E, Pérès S, Paulevé L, Nogueira ML, Steyaert JM and Schwartz L: The redox status of cancer cells supports mechanisms behind the warburg effect. Metabolites. 6:332016. View Article : Google Scholar : PubMed/NCBI | |
|
Halestrap AP and Wilson MC: The monocarboxylate transporter family-role and regulation. IUBMB Life. 64:109–119. 2012. View Article : Google Scholar | |
|
Singh M, Afonso J, Sharma D, Gupta R and Kumar V, Rani R, Baltazar F and Kumar V: Targeting monocarboxylate transporters (MCTs) in cancer: How close are we to the clinics? Semin Cancer Biol. 90:1–14. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Hosonuma M and Yoshimura K: Association between pH regulation of the tumor microenvironment and immunological state. Front Oncol. 13:11755632023. View Article : Google Scholar : PubMed/NCBI | |
|
Wang X, Zhao Y, Hu Y, Fei Y, Zhao Y, Xue C, Cai K, Li M and Luo Z: Activatable biomineralized nanoplatform remodels the intracellular environment of multidrug-resistant tumors for enhanced ferroptosis/apoptosis therapy. Small. 17:e21022692021. View Article : Google Scholar : PubMed/NCBI | |
|
Jackson VN and Halestrap AP: The kinetics, substrate, and inhibitor specificity of the monocarboxylate (lactate) transporter of rat liver cells determined using the fluorescent intracellular pH indicator, 2',7'-bis(carboxyethyl)-5(6)-carboxyfluorescein. J Biol Chem. 271:861–868. 1996. View Article : Google Scholar : PubMed/NCBI | |
|
Bozzo L, Puyal J and Chatton JY: Lactate modulates the activity of primary cortical neurons through a receptor-mediated pathway. PLoS One. 8:e717212013. View Article : Google Scholar : PubMed/NCBI | |
|
Yang Z, Su W, Wei X, Qu S, Zhao D, Zhou J, Wang Y, Guan Q, Qin C, Xiang J, et al: HIF-1α drives resistance to ferroptosis in solid tumors by promoting lactate production and activating SLC1A1. Cell Rep. 42:1129452023. View Article : Google Scholar | |
|
Bae C, Sachs F and Gottlieb PA: Protonation of the human PIEZO1 ion channel stabilizes inactivation. J Biol Chem. 290:5167–5173. 2015. View Article : Google Scholar : PubMed/NCBI | |
|
Jin C, Zhang DP, Lin Z, Lin YZ, Shi YF, Dong XY, Jin MQ, Song FQ, Du ST, Feng YZ, et al: Piezo1-mediated ferroptosis delays wound healing in aging mice by regulating the transcriptional activity of SLC7A11 through activating transcription factor 3. Research (Wash D C). 8:07182025.PubMed/NCBI | |
|
Koppula P, Zhang Y, Zhuang L and Gan B: Amino acid transporter SLC7A11/xCT at the crossroads of regulating redox homeostasis and nutrient dependency of cancer. Cancer Commun (Lond). 38:122018.PubMed/NCBI | |
|
Hu K, Li K, Lv J, Feng J, Chen J, Wu H, Cheng F, Jiang W, Wang J, Pei H, et al: Suppression of the SLC7A11/glutathione axis causes synthetic lethality in KRAS-mutant lung adenocarcinoma. J Clin Invest. 130:1752–1766. 2020. View Article : Google Scholar : | |
|
Harris IS, Treloar AE, Inoue S, Sasaki M, Gorrini C, Lee KC, Yung KY, Brenner D, Knobbe-Thomsen CB, Cox MA, et al: Glutathione and thioredoxin antioxidant pathways synergize to drive cancer initiation and progression. Cancer Cell. 27:211–222. 2015. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang L, Hobeika CS, Khabibullin D, Yu D, Filippakis H, Alchoueiry M, Tang Y, Lam HC, Tsvetkov P, Georgiou G, et al: Hypersensitivity to ferroptosis in chromophobe RCC is mediated by a glutathione metabolic dependency and cystine import via solute carrier family 7 member 11. Proc Natl Acad Sci USA. 119:e21228401192022. View Article : Google Scholar : PubMed/NCBI | |
|
Koppula P, Lei G, Zhang Y, Yan Y, Mao C, Kondiparthi L, Shi J, Liu X, Horbath A, Das M, et al: A targetable CoQ-FSP1 axis drives ferroptosis- and radiation-resistance in KEAP1 inactive lung cancers. Nat Commun. 13:22062022. View Article : Google Scholar : PubMed/NCBI | |
|
Rojo de la Vega M, Chapman E and Zhang DD: NRF2 and the hallmarks of cancer. Cancer Cell. 34:21–43. 2018. View Article : Google Scholar : PubMed/NCBI | |
|
Rawat SG, Tiwari RK, Jaiswara PK, Gupta VK, Sonker P, Vishvakarma NK, Kumar S, Pathak C, Gautam V and Kumar A: Phosphodiesterase 5 inhibitor sildenafil potentiates the antitumor activity of cisplatin by ROS-mediated apoptosis: A role of deregulated glucose metabolism. Apoptosis. 27:606–618. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Kobayashi M, Narumi K, Furugen A and Iseki K: Transport function, regulation, and biology of human monocarboxylate transporter 1 (hMCT1) and 4 (hMCT4). Pharmacol Ther. 226:1078622021. View Article : Google Scholar : PubMed/NCBI | |
|
Ding Y, Chen X, Liu C, Ge W, Wang Q, Hao X, Wang M, Chen Y and Zhang Q: Identification of a small molecule as inducer of ferroptosis and apoptosis through ubiquitination of GPX4 in triple negative breast cancer cells. J Hematol Oncol. 14:192021. View Article : Google Scholar | |
|
Jiang L, Kon N, Li T, Wang SJ, Su T, Hibshoosh H, Baer R and Gu W: Ferroptosis as a p53-mediated activity during tumour suppression. Nature. 520:57–62. 2015. View Article : Google Scholar : PubMed/NCBI | |
|
Zhou L, Mo Y, Zhang H, Zhang M, Xu J and Liang S: Role of AMPK-regulated autophagy in retinal pigment epithelial cell homeostasis: A review. Medicine (Baltimore). 103:e389082024. View Article : Google Scholar : PubMed/NCBI | |
|
Chen X, Yu C, Kang R, Kroemer G and Tang D: Cellular degradation systems in ferroptosis. Cell Death Differ. 28:1135–1148. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Gao M, Monian P, Pan Q, Zhang W, Xiang J and Jiang X: Ferroptosis is an autophagic cell death process. Cell Res. 26:1021–1032. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Zhou B, Liu J, Kang R, Klionsky DJ, Kroemer G and Tang D: Ferroptosis is a type of autophagy-dependent cell death. Semin Cancer Biol. 66:89–100. 2020. View Article : Google Scholar | |
|
Yang M, Chen P, Liu J, Zhu S, Kroemer G, Klionsky DJ, Lotze MT, Zeh HJ, Kang R and Tang D: Clockophagy is a novel selective autophagy process favoring ferroptosis. Sci Adv. 5:eaaw22382019. View Article : Google Scholar : PubMed/NCBI | |
|
Chen J, Zhu Y, Wu C and Shi J: Engineering lactate-modulating nanomedicines for cancer therapy. Chem Soc Rev. 52:973–1000. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Annoni F, Peluso L, Gouvêa Bogossian E, Creteur J, Zanier ER and Taccone FS: Brain protection after anoxic brain injury: Is lactate supplementation helpful? Cells. 10:17142021. View Article : Google Scholar : PubMed/NCBI | |
|
Fei Y and Ding Y: The role of ferroptosis in neurodegenerative diseases. Front Cell Neurosci. 18:14759342024. View Article : Google Scholar : PubMed/NCBI | |
|
Pan RY, He L, Zhang J, Liu X, Liao Y, Gao J, Liao Y, Yan Y, Li Q, Zhou X, et al: Positive feedback regulation of microglial glucose metabolism by histone H4 lysine 12 lactylation in Alzheimer's disease. Cell Metab. 34:634–648.e6. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Devos D, Labreuche J, Rascol O, Corvol JC, Duhamel A, Guyon Delannoy P, Poewe W, Compta Y, Pavese N, Růžička E, et al: Trial of deferiprone in Parkinson's disease. N Engl J Med. 387:2045–2055. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Chen L, Shen Q, Liu Y, Zhang Y, Sun L, Ma X, Song N and Xie J: Homeostasis and metabolism of iron and other metal ions in neurodegenerative diseases. Signal Transduct Target Ther. 10:312025. View Article : Google Scholar : PubMed/NCBI | |
|
Berthet C, Lei H, Thevenet J, Gruetter R, Magistretti PJ and Hirt L: Neuroprotective role of lactate after cerebral ischemia. J Cereb Blood Flow Metab. 29:1780–1789. 2009. View Article : Google Scholar : PubMed/NCBI | |
|
Cerina M, Levers M, Keller JM and Frega M: Neuroprotective role of lactate in a human in vitro model of the ischemic penumbra. Sci Rep. 14:79732024. View Article : Google Scholar : PubMed/NCBI | |
|
Xiong XY, Pan XR, Luo XX, Wang YF, Zhang XX, Yang SH, Zhong ZQ, Liu C, Chen Q, Wang PF, et al: Astrocyte-derived lactate aggravates brain injury of ischemic stroke in mice by promoting the formation of protein lactylation. Theranostics. 14:4297–4317. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Cai M, Wang H, Song H, Yang R, Wang L, Xue X, Sun W and Hu J: Lactate Is answerable for brain function and treating brain diseases: Energy substrates and signal molecule. Front Nutr. 9:8009012022. View Article : Google Scholar : PubMed/NCBI | |
|
Liu Y, Hu P, Cheng H, Xu F and Ye Y: The impact of glycolysis on ischemic stroke: From molecular mechanisms to clinical applications. Front Neurol. 16:15143942025. View Article : Google Scholar : PubMed/NCBI | |
|
Cler M, Perez-Amodio S, Valls-Lacalle L, Martinez E, Barba I, Ganse GFS, Engel E and Rodriguez-Sinovas A: Abstract 4146509: L-lactic acid reduces infarct size after ischemia in isolated mouse hearts through acidosis, MCT1-mediated uptake, and metabolic reprogramming. Circulation. 150(Suppl1): A41465092024. View Article : Google Scholar | |
|
Berthet C, Castillo X, Magistretti PJ and Hirt L: New evidence of neuroprotection by lactate after transient focal cerebral ischaemia: Extended benefit after intracerebroventricular injection and efficacy of intravenous administration. Cerebrovasc Dis. 34:329–335. 2012. View Article : Google Scholar : PubMed/NCBI | |
|
Nalos M, Kholodniak E, Smith L, Orde S, Ting I, Slama M, Seppelt I, McLean AS and Huang S: The comparative effects of 3% saline and 0.5M sodium lactate on cardiac function: A randomised, crossover study in volunteers. Crit Care Resusc. 20:124–130. 2018.PubMed/NCBI | |
|
Nalos M, Leverve X, Huang S, Weisbrodt L, Parkin R, Seppelt I, Ting I and Mclean A: Half-molar sodium lactate infusion improves cardiac performance in acute heart failure: A pilot randomised controlled clinical trial. Crit Care. 18:R482014. View Article : Google Scholar : PubMed/NCBI | |
|
Koyama T: Lactated Ringer's solution for preventing myocardial reperfusion injury. Int J Cardiol Heart Vasc. 15:1–8. 2017.PubMed/NCBI | |
|
Nolt B, Tu F, Wang X, Ha T, Winter R, Williams DL and Li C: Lactate and immunosuppression in sepsis. Shock. 49:120–125. 2018. View Article : Google Scholar : | |
|
Doherty JR and Cleveland JL: Targeting lactate metabolism for cancer therapeutics. J Clin Invest. 123:3685–3692. 2013. View Article : Google Scholar : PubMed/NCBI | |
|
Sonpavde G, Matveev V, Burke JM, Caton JR, Fleming MT, Hutson TE, Galsky MD, Berry WR, Karlov P, Holmlund JT, et al: Randomized phase II trial of docetaxel plus prednisone in combination with placebo or AT-101, an oral small molecule Bcl-2 family antagonist, as first-line therapy for metastatic castration-resistant prostate cancer. Ann Oncol. 23:1803–1808. 2012. View Article : Google Scholar | |
|
Wu J, Gu X, Zhang J, Mi Z, He Z, Dong Y, Ge W, Ghimire K, Rong P, Wang W and Ma X: 4-OI protects MIN6 cells from oxidative stress injury by reducing LDHA-mediated ROS generation. Biomolecules. 12:12362022. View Article : Google Scholar : PubMed/NCBI | |
|
Zhao Z, Han F, Yang S, Wu J and Zhan W: Oxamate-mediated inhibition of lactate dehydrogenase induces protective autophagy in gastric cancer cells: Involvement of the Akt-mTOR signaling pathway. Cancer Lett. 358:17–26. 2015. View Article : Google Scholar | |
|
Farabegoli F, Vettraino M, Manerba M, Fiume L, Roberti M and Di Stefano G: Galloflavin, a new lactate dehydrogenase inhibitor, induces the death of human breast cancer cells with different glycolytic attitude by affecting distinct signaling pathways. Eur J Pharm Sci. 47:729–738. 2012. View Article : Google Scholar : PubMed/NCBI | |
|
Granchi C, Roy S, Giacomelli C, Macchia M, Tuccinardi T, Martinelli A, Lanza M, Betti L, Giannaccini G, Lucacchini A, et al: Discovery of N-hydroxyindole-based inhibitors of human lactate dehydrogenase isoform A (LDH-A) as starvation agents against cancer cells. J Med Chem. 54:1599–1612. 2011. View Article : Google Scholar : PubMed/NCBI | |
|
Laganá G, Barreca D, Calderaro A and Bellocco E: Lactate dehydrogenase inhibition: Biochemical relevance and therapeutical potential. Curr Med Chem. 26:3242–3252. 2019. View Article : Google Scholar | |
|
Ippolito L, Morandi A, Giannoni E and Chiarugi P: Lactate: A metabolic driver in the tumour landscape. Trends Biochem Sci. 44:153–166. 2019. View Article : Google Scholar | |
|
Raez LE, Papadopoulos K, Ricart AD, Chiorean EG, Dipaola RS, Stein MN, Rocha Lima CM, Schlesselman JJ, Tolba K, Langmuir VK, et al: A phase I dose-escalation trial of 2-deoxy-D-glucose alone or combined with docetaxel in patients with advanced solid tumors. Cancer Chemother Pharmacol. 71:523–530. 2013. View Article : Google Scholar | |
|
Sutendra G and Michelakis ED: Pyruvate dehydrogenase kinase as a novel therapeutic target in oncology. Front Oncol. 3:382013. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang Y, Du X, He Z, Gao S, Ye L, Ji J, Yang X and Zhai G: A vanadium-based nanoplatform synergizing ferroptotic-like therapy with glucose metabolism intervention for enhanced cancer cell death and antitumor immunity. ACS Nano. 17:11537–11556. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Li F, Zhu P, Zheng B, Lu Z, Fang C, Fu Y and Li X: A customized biohybrid presenting cascade responses to tumor microenvironment. Adv Mater. 36:e24049012024. View Article : Google Scholar : PubMed/NCBI | |
|
Deng X, Zhu Y, Dai Z, Liu Q, Song Z, Liu T, Huang Y and Chen H: A bimetallic nanomodulator to reverse immunosuppression via sonodynamic-ferroptosis and lactate metabolism modulation. Small. 20:e24045802024. View Article : Google Scholar : PubMed/NCBI | |
|
Nancolas B, Guo L, Zhou R, Nath K, Nelson DS, Leeper DB, Blair IA, Glickson JD and Halestrap AP: The anti-tumour agent lonidamine is a potent inhibitor of the mitochondrial pyruvate carrier and plasma membrane monocarboxylate transporters. Biochem J. 473:929–936. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Halford S, Veal GJ, Wedge SR, Payne GS, Bacon CM, Sloan P, Dragoni I, Heinzmann K, Potter S, Salisbury BM, et al: A phase I dose-escalation study of AZD3965, an oral monocarboxylate transporter 1 inhibitor, in patients with advanced cancer. Clin Cancer Res. 29:1429–1439. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Puri S and Juvale K: Monocarboxylate transporter 1 and 4 inhibitors as potential therapeutics for treating solid tumours: A review with structure-activity relationship insights. Eur J Med Chem. 199:1123932020. View Article : Google Scholar : PubMed/NCBI | |
|
Khan A, Valli E, Lam H, Scott DA, Murray J, Hanssen KM, Eden G, Gamble LD, Pandher R, Flemming CL, et al: Targeting metabolic activity in high-risk neuroblastoma through Monocarboxylate transporter 1 (MCT1) inhibition. Oncogene. 39:3555–3570. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Kirk P, Wilson MC, Heddle C, Brown MH, Barclay AN and Halestrap AP: CD147 is tightly associated with lactate transporters MCT1 and MCT4 and facilitates their cell surface expression. EMBO J. 19:3896–3904. 2000. View Article : Google Scholar : PubMed/NCBI | |
|
Zhu J, Cai H, Xu C, Wang W, Song X, Li B, Shen Y and Dong X: Acidity-responsive nanoreactors destructed 'Warburg effect' for toxic-acidosis and starvation synergistic therapy. Small. 19:e23040582023. View Article : Google Scholar | |
|
Kobayashi M, Otsuka Y, Itagaki S, Hirano T and Iseki K: Inhibitory effects of statins on human monocarboxylate transporter 4. Int J Pharm. 317:19–25. 2006. View Article : Google Scholar : PubMed/NCBI | |
|
Chen ZX, Liu MD, Guo DK, Zou MZ, Wang SB, Cheng H, Zhong Z and Zhang XZ: A MSN-based tumor-targeted nanoplatform to interfere with lactate metabolism to induce tumor cell acidosis for tumor suppression and anti-metastasis. Nanoscale. 12:2966–2972. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Minhas PS, Jones JR, Latif-Hernandez A, Sugiura Y, Durairaj AS, Wang Q, Mhatre SD, Uenaka T, Crapser J, Conley T, et al: Restoring hippocampal glucose metabolism rescues cognition across Alzheimer's disease pathologies. Science. 385:eabm61312024. View Article : Google Scholar : PubMed/NCBI | |
|
Zeng Z, Huang Q, Mao L, Wu J, An S, Chen Z and Zhang W: The pyruvate dehydrogenase complex in sepsis: Metabolic regulation and targeted therapy. Front Nutr. 8:7831642021. View Article : Google Scholar : | |
|
Ryoo SM and Kim WY: Clinical applications of lactate testing in patients with sepsis and septic shock. J Emerg Crit Care Med. 2:142018. View Article : Google Scholar | |
|
Wei Y, Zhuang J, Li J, Wang Z, Wang J, Zhang X and Leng J: Lactate trajectories and outcomes in patients with sepsis in the intensive care unit: Group-based trajectory modeling. Front Public Health. 13:16102202025. View Article : Google Scholar : PubMed/NCBI | |
|
Liu S, Yang T, Jiang Q, Zhang L, Shi X, Liu X and Li X: Lactate and lactylation in sepsis: A comprehensive review. J Inflamm Res. 17:4405–4417. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Visker JR, Cluntun AA, Velasco-Silva JN, Eberhardt DR, Cedeño-Rosario L, Shankar TS, Hamouche R, Ling J, Kwak H, Hillas JY, et al: Enhancing mitochondrial pyruvate metabolism ameliorates ischemic reperfusion injury in the heart. JCI Insight. 9:e1809062024. View Article : Google Scholar : PubMed/NCBI | |
|
Fei M, Zhang H, Meng F, An G, Tang J, Tong J, Xiong L, Liu Q and Li C: Enhanced lactate accumulation upregulates PD-L1 expression to delay neutrophil apoptosis in sepsis. VIEW. 5:202300532024. View Article : Google Scholar | |
|
Suzuki A, Stern SA, Bozdagi O, Huntley GW, Walker RH, Magistretti PJ and Alberini CM: Astrocyte-neuron lactate transport is required for long-term memory formation. Cell. 144:810–823. 2011. View Article : Google Scholar : PubMed/NCBI | |
|
Lottes RG, Newton DA, Spyropoulos DD and Baatz JE: Lactate as substrate for mitochondrial respiration in alveolar epithelial type II cells. Am J Physiol Lung Cell Mol Physiol. 308:L953–L961. 2015. View Article : Google Scholar : PubMed/NCBI | |
|
Brown CW, Amante JJ, Goel HL and Mercurio AM: The α6β4 integrin promotes resistance to ferroptosis. J Cell Biol. 216:4287–4297. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Pucino V, Certo M, Bulusu V, Cucchi D, Goldmann K, Pontarini E, Haas R, Smith J, Headland SE, Blighe K, et al: Lactate buildup at the site of chronic inflammation promotes disease by inducing CD4+ T cell metabolic rewiring. Cell Metab. 30:1055–1074.e8. 2019. View Article : Google Scholar | |
|
Lei P, Walker T and Ayton S: Neuroferroptosis in health and diseases. Nat Rev Neurosci. 26:497–511. 2025. View Article : Google Scholar : PubMed/NCBI | |
|
Zheng X, Boyer L, Jin M, Mertens J, Kim Y, Ma L, Ma L, Hamm M, Gage FH and Hunter T: Metabolic reprogramming during neuronal differentiation from aerobic glycolysis to neuronal oxidative phosphorylation. Elife. 5:e133742016. View Article : Google Scholar : PubMed/NCBI | |
|
Bittar PG, Charnay Y, Pellerin L, Bouras C and Magistretti PJ: Selective distribution of lactate dehydrogenase isoenzymes in neurons and astrocytes of human brain. J Cereb Blood Flow Metab. 16:1079–1089. 1996. View Article : Google Scholar : PubMed/NCBI | |
|
Wu N, Wei X, Yu S, Yang L and Zhang X: Lactate in ferroptosis regulation: A new perspective on tumor progression and therapy. Pharmacol Res. 218:1078412025. View Article : Google Scholar : PubMed/NCBI | |
|
Kennedy L, Glesaaen ER, Palibrk V, Pannone M, Wang W, Al-Jabri A, Suganthan R, Meyer N, Austbø ML, Lin X, et al: Lactate receptor HCAR1 regulates neurogenesis and microglia activation after neonatal hypoxia-ischemia. Elife. 11:e764512022. View Article : Google Scholar : PubMed/NCBI | |
|
Hu L, Huang S, Chen G, Li B, Li T, Lin M, Huang Y, Xiao Z, Shuai X and Su Z: Nanodrugs incorporating LDHA siRNA inhibit M2-like polarization of TAMs and amplify autophagy to assist oxaliplatin chemotherapy against colorectal cancer. ACS Appl Mater Interfaces. 14:31625–31633. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Lu H, Liang B, Hu A, Zhou H, Jia C, Aji A, Chen Q, Ma Y, Cui W, Jiang L and Dong J: Engineered biomimetic cancer cell membrane nanosystems trigger gas-immunometabolic therapy for spinal-metastasized tumors. Adv Mater. 37:e24126552025. View Article : Google Scholar |