- Review
- Open access
- Published:
L- and D-Lactate: unveiling their hidden functions in disease and health
Cell Communication and Signaling volume 23, Article number: 134 (2025)
Abstract
Lactate, once considered a mere byproduct of anaerobic metabolism, is now recognized as a critical signaling molecule with diverse roles in physiology and pathology. There are two stereoisomers of lactate: L- and D-lactate. Recent studies have shown that disruptions in these two lactate stereoisomers have distinct effects on health and disease. L-lactate is central to glycolysis and energy transfer through the Cori cycle but also acts as the dominant lactylation isomer induced by glycolysis, influencing metabolism and cell survival. Although less studied, D-lactate is linked to metabolic disorders and plays a role in mitochondrial dysfunction and oxidative stress. This review focuses on both L- and D-lactate and examines their biosynthesis, transport, and expanding roles in physiological and pathological processes, particularly their functions in cancer, immune regulation, inflammation, neurodegeneration and other diseases. Finally, we assess the therapeutic prospects of targeting lactate metabolism, highlighting emerging strategies for intervention in clinical settings. Our review synthesizes the current understanding of L- and D-lactate, offering insights into their potential as targets for therapeutic innovation.
Introduction
Lactate, often characterized as a simple metabolic byproduct, is traditionally viewed as a waste product of anaerobic metabolism. During glycolysis, pyruvate is reduced to lactate by the action of the enzyme lactate dehydrogenase (LDH) [1, 2]. This reduction serves to regenerate positive oxidized nicotinamide adenine dinucleotide (NAD+), allowing glycolysis to proceed in the absence of sufficient oxygen. For decades, lactate was seen primarily as a secondary metabolite, particularly in highly glycolytic tissues such as skeletal muscle, where it accumulates during intense exercise or hypoxia [3,4,5]. This simplistic view of lactate as an energy byproduct has since evolved, with emerging evidence highlighting its crucial roles in metabolic regulation, signaling, and disease processes.
Recent discoveries have dramatically altered the perception of lactate, establishing it as a vital intermediary in metabolism and an important signaling molecule [6, 7]. Lactate exists in two stereoisomeric forms: L-lactate and D-lactate (Fig. 1). L-lactate, the predominant isomer in human metabolism, is produced by LDH-A, which is highly expressed in glycolytic tissues [8]; it is exported from cells via monocarboxylate transporters (MCTs), such as MCT1 and MCT4, facilitating its distribution throughout various tissues, where it can be utilized by more oxidative cells, including cardiomyocytes and neurons [9, 10]. This mechanism forms part of the “lactate shuttle” model, where lactate acts as a fuel source transferred between tissues and cells [11, 12]. L-lactate is reconverted to pyruvate and enters the tricarboxylic acid (TCA) cycle, contributing to oxidative phosphorylation and adenosine triphosphoric acid (ATP) production in oxygen-rich environments [13, 14].
A schematic diagram illustrating the structure isomers of L- and D-Lactate. L- and D-lactate are two stereoisomers of lactate. In the Fischer projection, the hydroxyl group on the chiral carbon atom of D-lactate is located on the right side while the hydroxyl group of L-lactate is located on the left side, which makes the lactylation modification involved by them also produce similar structural differences, generating Lysine D-lactylation (KD−la) and Lysine L-lactylation (KL−la) separately [15, 16]
In normal cellular metabolism, L-lactate production occurs predominantly through anaerobic glycolysis, a pathway activated under hypoxic conditions when oxygen is scarce. In this context, cells rely on glucose metabolism to produce energy without the need for oxygen; hence, the term ‘anaerobic glycolysis.’ In contrast, many cancer cells exhibit an altered metabolic phenotype known as ‘aerobic glycolysis’ or the Warburg effect, in which L-lactate is produced even in the presence of oxygen [5, 17] (Fig. 2). Despite this increased interest in cancer cell metabolism because of the original findings of Warburg [17], it has to be borne in mind that “aerobic glycolysis” is a phenomenon observed in cancer cells in culture [18]. This metabolic reprogramming enables cancer cells to generate both the energy and biosynthetic precursors required for rapid proliferation [19]. Importantly, compared with anaerobic glycolysis, aerobic glycolysis relies on distinct molecular mechanisms [19, 20]; distinguishing between these two metabolic states provides critical insight into the metabolic adaptability of cancer cells and highlights the importance of targeting specific pathways that underpin their metabolic flexibility.
A schematic diagram illustrating the differences between oxidative phosphorylation, anaerobic glycolysis, and aerobic glycolysis (Warburg effect). In oxygen-rich conditions, differentiated tissues metabolize glucose to pyruvate via glycolysis and fully oxidize it in mitochondria through oxidative phosphorylation. When oxygen is limited, cells shift to lactate production (anaerobic glycolysis), which regenerates NAD+ for glycolysis but yields less ATP. Cancer cells and proliferative tissues favor glucose conversion to lactate even with oxygen (aerobic glycolysis), though mitochondria remain functional. Proliferating cells also channel ~ 10% of glucose into biosynthetic pathways for macromolecule synthesis [19]
D-lactate, though less prevalent, is gaining attention because of its unique roles; it is produced primarily through bacterial fermentation in the gastrointestinal tract and, to a lesser extent, in human cells under specific pathological conditions [21, 22]. The role of D-lactate in human physiology has long been overlooked because of its low concentration compared with that of L-lactate. However, increasing evidence suggests that D-lactate has distinct metabolic and signaling roles, particularly in contexts such as gut dysbiosis, mitochondrial dysfunction, and certain metabolic disorders [22,23,24]. Conditions such as short bowel syndrome and disruptions in methylglyoxal metabolism can both lead to the accumulation of D-lactate to toxic levels, resulting in metabolic acidosis and neurological complications [21, 25, 26]. Overall, D-lactate accumulation is typically a consequence of pathological states such as gut disorders and dysbiosis, and its role in normal metabolism is minimal compared with that of L-lactate.
This evolving understanding of the distinct biochemical roles of L- and D-lactate has transformed the reputation of lactate from a mere marker of anaerobic metabolism to a critical regulator of cellular homeostasis, intercellular communication, and immune responses. This paradigm shift highlights the importance of lactate in various pathophysiological settings, including cancer, inflammation, and neurodegenerative diseases, where altered lactate levels and signaling may drive disease progression or present novel therapeutic targets [6, 27, 28]. In this review, we analyze and review the functions of L- and D-lactate and elucidate the mechanisms by which they regulate various pathological and physiological conditions, opening new fields for L- and D-lactate-mediated precision medicine for disease therapy.
Biosynthesis and transport of L- and D-Lactate
Biosynthesis of L- and D-Lactate
L-lactate is the dominant form of lactate in humans and eukaryotes and is catalyzed by LDHA from pyruvate to L-lactate. L-lactate can be further oxidized into pyruvate in the cell and then enter the mitochondria for TCA metabolism. As the dominant isomer in human physiology, L-lactate is produced ubiquitously in most tissues, especially in glycolytic cells such as skeletal muscle and erythrocytes [29, 30]. Additionally, the enzyme LDHA is highly expressed in these tissues, reflecting their reliance on glycolysis for energy production, particularly during periods of oxygen scarcity or high energy demand. Because this process is closely linked to glycolysis, ensuring that NAD+ regeneration maintains ATP production during hypoxia or high energy demand [31,32,33], L-lactate is critical for sustaining metabolic flux under anaerobic conditions [34, 35]. In addition, “aerobic glycolysis” is the major process responsible for L-lactate production, and glutaminolysis is the minor process responsible for L-lactate production in cancer cells [36, 37].
In contrast, D-lactate production in humans and eukaryotes occurs mostly through non-LDH pathways, mainly the methylglyoxal pathway, in abnormal metabolic states (Fig. 3). The methylglyoxal pathway is a branch of glycolysis that converts glucose to methylglyoxal and then to D-lactate [38]. In addition, D-lactate can also be produced by gut microbes (such as Lactobacillus) that breakdown glucose to produce D-lactate, which is absorbed through the gut and into the blood. For example, in short bowel syndrome, increased intestinal permeability and bacterial overgrowth lead to elevated D-lactate levels, which are involved in multiple pathological processes [21, 25]. The human gut microbiota, particularly species of the Firmicutes and Bacteroidetes phyla, are significant producers of D-lactate during carbohydrate fermentation [39]. Additionally, in prokaryotes, D-lactate is the main type of lactate [40]. In addition, the isoform D-LDH is also responsible for the limited production of D-lactate in human tissues, although its physiological role remains less defined than that of L-lactate [41]. These studies revealed that understanding the biosynthesis of D-lactate and its accumulation in pathological states is important for the treatment of diseases.
Overall, L-lactate plays a central role in maintaining metabolic homeostasis under anaerobic conditions [42, 43], whereas the relevance of D-lactate is more pronounced in pathological contexts [44, 45]. Understanding the distinct biosynthetic pathways and roles of these two lactate isomers provides crucial insights into their contributions to both normal physiology and disease states.
Schematic diagram of the biosynthesis of L- and D-Lactate in eukaryotic. L-lactate is produced by glycolysis pathway, D-lactate is produced by glyoxalase pathway. In human physiology, L-lactate is the dominant isomer and D-lactate is produced at much lower levels, primarily by gut microbiota during carbohydrate fermentation. In dysbiosis, lactate-producing bacteria are a significant source of L- and D-lactate. Abbreviations: GAP: Glyceraldehyde 3-phosphate; GAPDH: Glyceraldehyde 3-phosphate dehydrogenase; TCA: Tricarboxylic acid cycle; PDH: Pyruvate dehydrogenase; LDHA: Lactate dehydrogenase A; L-LDH: L-lactate dehydrogenase A with M subunit; D-2-HDH: D-2-hydroxy acid dehydrogenase; DHAP: Dihydroxyacetone phosphate; MGO: Methylglyoxal; GLO1: Glyoxalase 1; GLO2: Glyoxalase 2; mLDH: Mitochondrial lactate dehydrogenase
Transport of L- and D-Lactate
L- and D-lactate share a common transport system, as suggested by substrate competition studies [46]. The role of lactate in cellular metabolism is highly dependent on its efficient transport across membranes, a process mediated by MCT [47, 48]. MCTs are classified as a specific category of transmembrane transporters with the solute carrier family 16 (SLC16) family [49]. Among the 14 identified MCTs, MCT1-4 (also known as Slc16a1, Slc16a7, Slc16a8, and Slc16a3 respectively) are expressed in different tissues and are involved in catalytic proton coupling and the bidirectional transport of monocarboxylic acid [10]. MCT1 and MCT4 are the primary transporters responsible for lactate uptake and export, respectively [50, 51]. MCT1 is expressed predominantly in tissues with high oxidative capacity, such as cardiac and skeletal muscle, where it facilitates the uptake of lactate for oxidation and ATP production. In contrast, MCT4 is expressed in glycolytic tissues and plays a key role in exporting lactate to prevent intracellular acidification during anaerobic glycolysis.
In summary, both proton (MCT1-4) and sodium-dependent transporters (Slc5a8 and Slc5a12, also known as sMCT1, sMCT2 respectively) medicated the transport system [7, 52, 53]. These transporters are promiscuous and can act on a variety of substrates, including pyruvate or ketone bodies [7, 52]. Disruptions in lactate transport are implicated in various diseases, such as cancer and cardiovascular diseases [47, 54,55,56]. In cancer, the overexpression of MCT4 in glycolytic tumor cells contributes to the formation of an acidic microenvironment, promoting tumor growth, immune evasion, and metastasis [47]. Similarly, in conditions such as type 2 diabetes, MCT4 expression is upregulated, and lactate transport homeostasis is disrupted, leading to oxidative stress and inflammatory responses that exacerbate myocardial damage [56].Collectively, these transporters maintain lactate homeostasis, ensuring the use of lactate as both fuel and a signaling molecule and understanding the regulation of MCTs is crucial for developing interventions aimed at restoring lactate homeostasis in disease contexts where lactate dynamics are disrupted. Targeting lactate transport could offer therapeutic potential in treating such conditions.
The expanding roles of L- and D-Lactate in cellular metabolism and energy homeostasis
L-lactate functions as both metabolic fuel and a signaling molecule; it is produced during anaerobic glycolysis and shuttled to oxidative tissues such as the heart and brain, where it is reconverted to pyruvate for entry into the TCA cycle, supporting ATP production [6, 13, 29, 57]. This lactate shuttle is crucial for maintaining energy homeostasis, especially during periods of high metabolic demand, such as exercise or hypoxia [58,59,60]. Moreover, lactate supports metabolic flexibility, enabling cells to shift between carbohydrate and lipid metabolism depending on oxygen and substrate availability [61, 62]. In addition to its metabolic role, L-lactate modulates key signaling pathways linked to cell growth, survival, and immune responses [63, 64]; it can regulate the cell cycle and proliferation by interacting with the anaphase-promoting complex APC/C [65]. Some proposed mechanisms suggest that the entry of L-lactate into neurons alters the ATP/ADP ratio or affects the cellular redox state [66]. Additionally, L-lactate exerts an immunomodulatory effect by promoting the differentiation of regulatory T cells (Tregs), thereby contributing to immune suppression within the tumor microenvironment [67]. This immunomodulatory function allows lactate to contribute to both tumor progression and immune evasion, highlighting its complexity as both a substrate and a regulatory signaling molecule in health and disease.
L-lactate, once known primarily as an energy metabolite, is now recognized as a potent signaling molecule through the Gi/o-protein coupled receptor GPR81 (also known as HCAR1). As an extracellular messenger, L-lactate activates GPR81 via both autocrine and paracrine pathways. In cancer, lactate produced by tumor cells activates GPR81 on the same cells in an autocrine loop and on immune cells, endothelial cells, and adipocytes in the tumor microenvironment through paracrine signaling, facilitating immune evasion and chemoresistance [68,69,70,71,72]. Notably, GPR81 activation by lactate is independent of proton (H+)- or MCT-mediated import and does not require the conversion of lactate to pyruvate, indicating that L-lactate is its specific ligand [72]. This unique mechanism is supported by studies showing that lactate signaling through GPR81 works synergistically with insulin to lower intracellular cAMP levels and inhibit lipolysis, suggesting a key role in metabolic regulation [73]. In addition to its role in cancer, GPR81 is expressed in neurons, astrocytes, and endothelial cells across the brain, retina, and blood‒brain barrier, as well as in adipose and liver tissues, where it influences lipolysis and gluconeogenesis, thus impacting systemic energy balance [74,75,76]. In summary, the lactate/GPR81 axis supports tumor angiogenesis, immune evasion, and progression, playing a dual role in both tumor biology and metabolic regulation. This pathway presents a compelling target for the development of novel anticancer therapies aimed at disrupting the role of lactate in promoting tumor survival and progression.
Compared with L-lactate, D-lactate plays a distinct role in certain metabolic and pathological contexts. Excessive D-lactate is associated with D-lactic acidosis, which is characterized by metabolic acidosis and neurological symptoms such as confusion and ataxia [21, 24, 25]. While D-lactate production in human tissues is limited, its accumulation in disease states highlights its potential systemic impact. Recent studies suggest that D-lactate contributes to mitochondrial dysfunction and oxidative stress [77,78,79]. Specifically, elevated D-lactate levels can impair mitochondrial respiration by inhibiting key enzymes in the electron transport chain (ETC), leading to the increased production of reactive oxygen species (ROS) and reduced ATP generation [79, 80]. This mitochondrial disruption is linked to metabolic disorders such as type 2 diabetes and may exacerbate insulin resistance and cellular stress [81]. Although the precise mechanisms remain to be fully elucidated, the role of D-lactate in mitochondrial regulation and redox balance suggests that it plays a significant role in metabolic dysfunction under specific pathological states.
L-and D-Lactate and diseases
Cancer metabolism
L-lactate is central to cancer metabolism, and increasing evidence indicates that increased glucose metabolism and reprogramming toward aerobic glycolysis are hallmarks of cancer cells, meeting their energy needs for continuous proliferation [82]. The production of L-lactate, particularly through the Warburg effect, i.e., when cancer cells preferentially rely on “aerobic glycolysis”, even in the presence of oxygen, in cancer cells, increases [83, 84]. This excess lactate contributes to an acidic tumor microenvironment (TME) that facilitates cancer cell invasion, metastasis, and immune evasion by suppressing cytotoxic T cells and natural killer (NK) cells [85,86,87]. Lactate-driven lactylation has emerged as a critical factor in DNA repair and chemoresistance. For example, the lactylation of Nibrin (NBS1) at K388 is necessary for the formation of the MRE11-Rad50-NBS1 (MRN) complex, which is involved in homologous recombination. Decreasing lactate levels via LDHA reduces NBS1 lactylation, impairing DNA repair and reducing chemoresistance [88]. Additionally, the lactylation of MRE11 at K673 serves as a vital response to DNA damage, further contributing to tumor resistance [89]. Alanyl-tRNA synthetase (AARS1) functions as both a lactate sensor and a lactyltransferase, mediating the lactylation of proteins such as p53, which promotes tumorigenesis [90]. Recent studies have indicated that AARS1, AARS2, and AlaRS in Escherichia coli can sense L-lactate accumulation, driving widespread proteome lactylation [91]. Additionally, the lactylation of proteins such as p53, adenylate kinase 2 (AK2), protein kinase R-like ER kinase (PERK) and nucleolin drives tumorigenesis and immune suppression [92,93,94,95]. In colorectal cancer, KAT8-catalyzed lactylation promotes protein synthesis via eukaryotic translation elongation factor 1 alpha 2 (eEF1A2), enhancing carcinogenesis [96]. Similarly, in prostate cancer and ocular melanoma, lactylation influences metabolic reprogramming and tumor progression [97, 98]. Additionally, L-lactate-producing lactic acid bacteria present at various body sites can alter tumor metabolism and contribute to therapeutic resistance, suggesting that the modulation of these bacteria could be a novel therapeutic strategy [84]. Collectively, these findings highlight the potential of targeting lactylation as a means to increase the efficacy of immunotherapy in cancer treatment.
D-lactate levels are significantly increased in various cancers, particularly gastrointestinal cancers, where altered gut microbiota increases D-lactate production [99,100,101]. D-lactate plays a vital role in regulating cancer cell ferroptosis and macrophage activity, although its exact functions in cancer metabolism are still being elucidated. Elevated D-lactate levels contribute to mitochondrial dysfunction, oxidative stress, and metabolic acidosis, highlighting the involvement of D-lactate in the complex metabolic landscape of tumors [44, 102, 103]. In esophageal squamous carcinoma, Lv et al. [104]. demonstrated that cyclin-dependent kinase 7 (CDK7) phosphorylates nuclear Yes-associated protein 1 (YAP) at S127 and S397, increasing its transcriptional activity and upregulating D-lactate dehydrogenase (LDHD) expression. This pathway helps cancer stem cells evade D-lactate-induced ferroptosis and generates pyruvate to meet the energetic demands of their high self-renewal capacity, suggesting a potential metabolic checkpoint for therapeutic targeting in esophageal squamous carcinoma [105]. Moreover, synthetic D-lactate dimers and organometallic redox catalysts have shown promise in inhibiting cancer cell proliferation [106, 107]. D-lactate, similar to the L-lactate, can enhance DNA repair and influence the resistance of cervical carcinoma cells to anticancer drugs through histone deacetylase inhibition and HCAR1 activation [108]. These findings suggest a critical role for D-lactate in cancer metabolism and present opportunities for therapeutic interventions that target lactate dynamics to disrupt tumor progression.
Immune modulation and inflammation
L-lactate plays a pivotal role in modulating immune responses, particularly in the TME, where its accumulation suppresses cytotoxic T cells and NK cells, which are key players in antitumor immunity [84, 109, 110]. Additionally, L-lactate enhances Treg activity, fostering an immunosuppressive environment. In KRAS-mutant tumors, lactate-driven histone lactylation activates the transcription of circATXN7, a circular RNA that interacts with NF-κB, sequestering it in the cytoplasm and promoting immune escape through increased T cell sensitivity to activation-induced cell death [111]. Lactate-driven lactylation of membrane-organizing extension spike protein (MOESIN) in Tregs further supports tumor immunosuppression via TGF-β signaling, and the inhibition of lactate production via LDH inhibitors diminishes this effect [112]. Furthermore, MCT4-driven metabolic reprogramming in cancer-associated fibroblasts (CAFs) represents a targetable vulnerability in breast cancer [109]. Lactate-mediated modifications extend beyond histone lactylation. For example, METTL16 lactylation promotes cuproptosis in gastric cancer, whereas METTL3-induced m6A RNA modifications in colorectal tumors highlight the regulatory role of lactate in immune responses [113, 114]. Raychaudhuri et al.. further reported that H3K18la and H3K9la dynamically regulate CD8+ T cell transcription by influencing promoters and enhancers. Specifically, H3K18la is linked to inflammatory responses and glycolysis during T cell activation, whereas H3K9la is associated with oxidative phosphorylation and fatty acid metabolism in memory T cells, underscoring the role of lactate in linking metabolism and epigenetics to CD8+ T cell function [115].
In inflammatory diseases, lactate-induced histone lactylation in macrophages contributes to lung fibrosis by promoting the expression of profibrotic mediators [116,117,118]. In conditions such as IBD and rheumatoid arthritis, the B cell adapter for PI3K (BCAP) suppresses glycogen synthase kinase 3 beta (GSK3β) and FOXO1, influencing lactate production, histone modifications, and macrophage polarization [119]. Lactate also induces SMCT2 expression in CD4+ T cells, driving lactate uptake and increasing IL-17 production via PKM2/STAT3 signaling, suggesting that targeting SMCT2 could help alleviate chronic inflammatory disorders [120]. While D-lactate is less prevalent, it contributes to proinflammatory responses, particularly gut dysbiosis and increased intestinal permeability [121, 122]. Elevated D-lactate levels activate macrophages, triggering TNF-α and IL-6 production and driving systemic inflammation [123, 124]. Cortney et al.. reported that D-lactate elevation and IL-10 in prosthetic joint infections play a role in the immune response, whereas gut-derived D-lactate supports Kupffer cells in pathogen clearance [125, 126]. Furthermore, D-lactate modulates macrophage polarization, shifting from an anti-inflammatory M2 phenotype to a proinflammatory M1 phenotype, as observed in hepatocellular carcinoma [125, 126]. In IBD, elevated D-lactate is correlated with disease severity and systemic inflammation, suggesting that D-lactate is a potential biomarker for IBD management [127, 128]. Collectively, these findings underscore the dual roles of L- and D-lactate in immune modulation and inflammation, with L-lactate primarily promoting immunosuppression in cancer and D-lactate driving proinflammatory responses. Future research into their crosstalk could reveal novel therapeutic targets for immune and inflammatory diseases.
Neurodegenerative diseases
In the brain, L-lactate is produced by neurons and astrocytes and serves as an auxiliary energy source for neurons [66]. Recent findings indicate that low-density lipoprotein receptor-related protein 1 (LRP1) regulates the lactylation of ADP-ribosylation factor 1 (ARF1), influencing astrocyte–neuron mitochondrial translocation, a mechanism that may protect against ischemic organ damage and highlights the role of LRP1 in astrocytic mitochondrial transport [129]. In early Alzheimer’s disease (AD), excessive glycolysis leads to intracellular lactate accumulation, triggering excessive histone lactylation, amplifying glycolysis-related gene expression, and creating a feedback loop that exacerbates glucose metabolism in microglia [130]. Additionally, L-lactate stabilizes NDRG family member 2 (NDRG2), inhibiting TNFα expression and promoting neuroprotection postischemia [31]. Impaired oxidative phosphorylation in AD further accelerates disease progression via the isocitrate dehydrogenase 3 noncatalytic subunit beta (IDH3β)-lactate-paired box 6 (PAX6) feedback loop [131]. Lactylation also supports neural development, where lactylation enhances the expression of genes vital for neural crest cell differentiation [132]. These findings underscore the complex role of L-lactate in maintaining microglial and neuronal homeostasis, with implications for neurodegenerative disease progression.
Although less explored in neurodegeneration, D-lactate is associated with D-lactic acidosis, a condition observed in patients with malabsorption syndromes such as short bowel syndrome [44, 133, 134]. Excess D-lactate entering the bloodstream leads to neurological symptoms, including encephalopathy [134, 135]. The differential metabolism of methylglyoxal and D-lactate in cancer and AD cells may help explain the “eternal youth” of cancer cells and the “premature death” of AD neurons [26]. Most neurodegenerative diseases attributed to D-lactate result from life-threatening D-lactic acidosis, a rare complication occurring primarily in patients with malabsorption due to surgically altered gastrointestinal tract anatomy, such as in short bowel syndrome or following bariatric surgery. D-lactic acidosis is characterized by the rapid development of neurological symptoms [44, 103]. While most neurodegenerative effects of D-lactate are linked to D-lactic acidosis in patients with gastrointestinal surgeries, its connection to neurological disorders suggests a potential role for D-lactate as a biomarker for neurodegenerative conditions, warranting further investigation.
Other diseases
The involvement of L-lactate in cardiovascular disease is gaining recognition [55, 136]. For example, the reduced lactylation of α-MHC-k1897 contributes to heart failure [137]. In vascular health, nuclear receptor subfamily 4 group A member 3 (NR4A3)-mediated histone lactylation has emerged as a novel metabolome‒epigenome signaling cascade implicated in arterial calcification [138]. Additionally, L-lactate is crucial in early development. Using the FiLa lactate probe, researchers observed lactate enrichment in the nuclei of two-cell embryos an determined that it was critical for preimplantation development; the inhibition of lactate production halted blastocyst formation, underscoring the role of lactate in embryogenesis [139]. In infectious disease, the upregulation of the expression of protein S6 kinase 2 (S6K2), which is linked to lactylation, enhances susceptibility to white spot syndrome virus in shrimp, indicating the role of lactylation in the pathogen response [140]. The lactylation of Yin Yang-1 (YY1) in microglia also promotes retinal neovascularization, identifying the lactate/p300/YY1/FGF2 axis as a potential retinopathy therapy target [141].
In angiogenesis, L-lactate promotes blood vessel formation in multiple contexts. Physical exercise stimulates VEGFA expression and cerebral angiogenesis via GPR81 in brain endothelial cells [70]. This angiogenic effect extends to wound healing, where macrophage-derived VEGF facilitates new vessel formation, involving both aerobic and anaerobic glycolysis in hypoxic, damaged cells, as well as immune-cell activation and proliferation [70, 142]. In tumors, L-lactate activates endothelial cells and induces angiogenesis through HIF-dependent and HIF-independent pathways, which rely on MCT1-mediated lactate import and prolyl hydroxylate (PHD) inhibition [143, 144]. Recently, a distinct HIF-independent mechanism in which accumulated lactate during hypoxia binds to NDRG3, preventing its degradation and further promoting angiogenesis, was identified [145]. Together, these findings underscore the potential of L-lactate as a therapeutic target in cardiovascular and developmental disorders.
D-lactate, produced by gut bacteria, has been implicated in cardiovascular disease, as its translocation into systemic circulation may impact cardiovascular health, suggesting that gut barrier-targeted therapies may prevent cardiovascular events [146]. Elevated D-lactate levels aid in the diagnosis of acute appendicitis, with the potential to differentiate types of appendicitis in pediatric patients [147]. In cystic fibrosis, D-lactate is correlated with clinical severity and pancreatic insufficiency, as measured by the Shwachman–Kulczycki score [148]. Elevated D-lactate levels have also been observed in type 1 and type 2 diabetes, where gut dysbiosis and increased, intestinal permeability may contribute to systemic D-lactate elevation [149,150,151]. This correlation with diabetic complications, possibly through elevated methylglyoxal levels, suggests that D-lactate could serve as a diagnostic marker; however, further studies are needed to clarify its role in various conditions.
Summary, it is worth noting that the regulation of diseases by L- and D-lactate exists within a complex network of interplay and mutual influence in diseases (Fig. 4). The impact of L- and D-lactate on inflammation could serve as a driving factor for the tumorigenesis. In addition, inflammation driven by L- and D-lactate, often linked to infectious diseases. Within this intricate network, L- and D-lactate emerges as a critical regulator, influencing intracellular signaling, gene expression, intercellular communication, and immune responses, thereby shaping disease progression. The diverse effects of L- and D-lactate underscore the need for a holistic perspective in disease research and treatment. Understanding its mechanisms across various diseases could enable the development of precise, targeted therapies with improved efficacy and fewer side effects.
Lactate shuttle hypothesis
Intracellular and intercellular shuttle of lactate
Targeting pathological lactate levels, regulating lactate transport, and modulating disease-related lactate levels offers promising therapeutic strategies for different lactate-related diseases in the future. One prominent approach for targeting lactate metabolism involves the inhibition of MCTs, which regulate lactate transport across cell membranes. The “lactate shuttle” hypothesis highlights the pivotal role of lactate in the transport of substrates for oxidation and gluconeogenesis, as well as in cellular signaling pathways [5]. Brooks initially proposed that lactate shuttling can take two forms: intracellular and intercellular [5]. Intracellular shuttling primarily occurs between the cell membrane and mitochondria [152]. When lactate is transported in cells, there are a variety of mechanisms, with the lactate/H+ cotransporter-mediated shuttle facilitating the transmembrane movement of both L- and D-lactate. In particular, MCTs are recognized as key mediators of L-lactate transport [153]. Although the H+ cotransport mechanism of D-lactate resembles that of L-lactate, the specific proteins mediating D-lactate transport remain unidentified. In addition to H⁺ cotransport, D-lactate transport within cells involves both D-lactate/pyruvate and D-lactate/malate reverse transporters. The D-lactate/pyruvate transporter shuttles D-lactate from the cytoplasm to the mitochondrial membrane while moving pyruvate in the reverse direction. The D-lactate/malate transporter, potentially localized within the inner mitochondrial membrane, transfers D-pyruvate (generated after mitochondrial D-LDH oxidation) into the mitochondria and exports malate to the cytoplasm [8, 77]. The intracellular lactate shuttle hypothesis challenges traditional perspectives by revising the concept of lactate oxidation as an inherent cellular function. Furthermore, L-lactate production during exercise serves as an adaptive cellular response, with lactate acting as a signaling molecule that underpins the enhancement of lactate oxidation capacity through training [154].
Intercellular L-lactate shuttling mainly occurs between cells that depend on glycolysis and those engaged in oxidative respiration, such as glycolytic and oxidative muscle fibers, skeletal and cardiac muscles, astrocytes and neurons [155, 156]. According to Brooks et al.., lactate is produced and accumulates rapidly in muscle cells at the beginning of exercise. Subsequently, part of this lactate is transported to neighboring tissues, where it is internalized and oxidized by adjacent cells, while the remainder enters the bloodstream and is transported to organs such as the heart, liver, and kidneys, where it serves as a substrate for oxidative energy production and gluconeogenesis [3]. In coculture systems of cardiomyocytes and fibroblasts, lactate production in fibroblasts increases, with MCT1 translocating to the myocardial membrane to facilitate lactate uptake [157]. Similarly, in the brain, lactate produced by astrocytes is taken up by neurons, contributing to energy metabolism via conversion to pyruvate and acetyl-CoA, which also plays a role in regulating fatty acid synthesis [158]. Impaired lactate shuttling from glial cells to neurons disrupts brain metabolism, leading to neurodegeneration akin to that observed in AD [159]. This shuttling mechanism is also present in the kidney, where lactate is produced by proximal tubule cells and consumed by distal tubule cells [160]. Recently, lactate shuttling within the TME has emerged as a novel area of interest in cancer biology. Lactate not only meets the energy requirements of stromal cells but also regulates their function through its role as a signaling molecule, with activated stromal cells subsequently supporting tumor progression [161]. In summary, lactate shuttling underscores the multifaceted roles of lactate in energy metabolism and signal transduction, offering new therapeutic insights into the functions of lactate in both health and disease.
Histone lactylation
Histone lactylation and enzymes
The discovery of histone lactylation by Zhao’s and Galligan’s team both supported the role of lactate in gene regulation [16, 57, 162]. They reported that both KL-la and KD-la are predominantly modified on histones and that KL-la is the primary lactylation isomer on histones [57]. As a newly identified type of posttranslational modification, histone lactylation links lactate metabolism to transcriptional control and which also directly links the metabolic and epigenetic roles of lactate. Also, recent findings have revealed the enzymatic processes underlying this newly identified histone modification driven by L- or D-lactate, identifying key regulatory factors and their role in gene regulation and expression [104, 163, 164] (Fig. 5). Among these factors, p300 was the first protein identified with a “writer” function for histone lactylation. In HEK293T cells, p300 overexpression led to a slight increase in histone Kla levels, whereas p300 silencing in HCT116 and HEK293T cells resulted in a reduction in Kla, confirming that p300 is a potential Kla writer [16]. Similarly, in a study of myocardial infarction, lysine acetyltransferase 2 A (or GCN5) was shown to catalyze histone H3K18 lactylation [165], and both p300 and GCN5 were shown to be members of the HAT family. Additionally, HBO1, a newly identified lysine acetyltransferase, has been found to regulate histone Kla both in vitro and intracellularly, with a preference for catalyzing H3K9la [166]. Recent studies have identified AARS1/2 as conserved intracellular sensors of L-lactate that are capable of catalyzing the lactylation of nuclear proteins, thereby acting as lactyltransferases that modulate pathways such as the cGAS, P53, and YAP pathways, which influence tumor progression and immune responses. However, whether AARS1/2 also serve as “writers” of histone lactylation remains uncertain, necessitating further research to validate and elucidate this process [90, 167, 168].
In addition to “writers”, HDAC1-3, HDAC8 and SIRT1-3 act as “erasers” of histone lactylation. Christian A. Olsen’s group was the first to systematically explore the role of class I histone deacetylases (HDAC1-3) as histone lysine delactylases, demonstrating their ability to reverse lactylation [169]. HDAC3 was found to regulate lactylation at multiple sites, including H4K5. Compared with other sirtuins, Sirtuin 3 (SIRT3) shows particularly strong erasure activity at the H4K16la site [170]. Although few studies have identified “readers” of histone lactylation, Hu et al.. reported that Brg1 (or Smarca4) binds to H3K18la. Proteomic analysis revealed that Brg1 was enriched at pluripotency and epithelial junction gene promoters during iPSC reprogramming, making it the first recognized histone lactylation reader. Further research is needed to fully characterize the potential readers involved in this process, and more studies need to focus on histone lactylation driven by D- or L-lactate [171]. More studies are needed to identify potential readers of histone lactylation. Modulating the ‘writers’ and ‘erasers’ of histone lactylation represents a promising therapeutic approach, with emerging strategies offering potential for clinical applications targeting lactate metabolism.
Conclusions and perspectives
Future research on D-and L-Lactate
Once considered mere metabolic byproducts, L- and D-lactate are now considered important regulators of cellular homeostasis, immune function, cancer, and neurodegenerative diseases. While the roles of L-lactate in energy metabolism and immune regulation have been established, the influence of L-lactate on brain function, cardiovascular health, and immune responses warrants further investigation [172, 173]. As the understanding of L-lactate functions deepens, it becomes equally important to explore the unique and potentially tissue-specific roles of D-lactate.
D-lactate, which is primarily produced by lactic acid bacteria, plays a significant role in diseases associated with gut microbiota imbalances. For example, the accumulation of D-lactate has been linked to pathological processes along the gut‒liver axis, suggesting its potential as a therapeutic target. Modulating D-lactate activity in this axis could provide novel treatment strategies. Additionally, recent studies have shown that D-lactate can influence immune function by regulating M2 macrophages and reshaping the immunosuppressive TME in hepatocellular carcinoma [123], highlighting the need to investigate its specific roles in cancer and TME modulation. The impact of D-lactate also extends to mitochondrial dysfunction, oxidative stress, and metabolic diseases such as type 2 diabetes and neurodegenerative disorders. Further research into these areas could uncover new insights into the broader physiological roles of D-lactate. Additionally, lactate-related regulatory enzymes, such as lactyl-CoA synthetases, and their substrates (both L- and D-lactate) remain understudied [13, 29].
Interestingly, lactylation may not be the only modification with stereoisomeric forms. For example, β-hydroxybutyrylation could theoretically occur as D- and L-isomers generated from their respective CoA precursors [174]. This possibility raises questions about the distinct regulatory roles of D-lactyl-CoA and L-lactyl-CoA. Understanding how these enzymes orchestrate histone L- and D-lactylation in physiological contexts will be essential for elucidating the regulatory potential of lactate. Future research on these mechanisms may reveal broader physiological impacts of lactate, advancing therapeutic possibilities across diverse disease contexts.
Crosstalk between lactylation and other modifications
Emerging evidence has revealed distinct L-lactate and D-lactate modifications at specific protein sites, although it remains uncertain whether both can cooccur at the same or different sites within a single protein. Additionally, the functional interplay between these two modifications, whether synergistic or antagonistic, remains unexplored. Investigating this crosstalk could provide critical insights into the regulatory role of lactylation in disease mechanisms, particularly in cancer, the immune response, and inflammation. Future studies may pave the way for novel therapies that target these modifications to address diseases with dysregulated lactate signaling.
Histone lactylation has recently been identified as a significant epigenetic marker that is regulated by glycolysis and primarily manifests as K-lactylation [57]. This finding suggests possible crosstalk with other histone modifications, such as phosphorylation and acetylation, creating an interconnected network that responds to cellular metabolic states. As shown in Fig. 2, differentiated tissues rely on oxidative phosphorylation and anaerobic glycolysis, whereas tumor cells exhibit aerobic glycolysis (the Warburg effect), leading to increased lactate production. This metabolic shift not only fuels energy demand but also provides substrates such as ATP and lactate for phosphorylation and lactylation, respectively. These metabolites can influence chromatin structure through cooperative or competitive interactions between modifications, directly impacting gene expression. Investigating these interactions may clarify how metabolic changes shape the epigenome in various contexts, particularly in cancer, where metabolism and epigenetic regulation are tightly linked. Such studies could reveal novel therapeutic targets that address metabolic and epigenetic dysregulation in tumors and other diseases.
Moreover, many types of acylation, including lactylation, share common enzymatic pathways and are recognized by overlapping ‘readers’ that facilitate downstream effects. For example, SIRT2 acts as an ‘eraser’ by removing multiple acyl marks from both histone and nonhistone proteins [175,176,177,178,179], and several metabolic enzymes implicated in disease pathogenesis are known to modulate more than one modification [180]. Histone acetyl transferase (HATs) family, which catalyzes the transfer of the acetyl group from acetyl-CoA to histones, including P300/CBP, MOF, GCN5 and TIP60 are also reported catalyzes the transfer of lactate groups from lactate-CoA to histones in different disease and disease states [16, 88, 165, 181, 182]. Therefore, understanding the dynamic interactions among various lactylation modifications and their crosstalk with other acylation modifications, particularly their competition for regulatory enzymes, may be crucial for fully elucidating disease mechanisms. This knowledge could pave the way for therapeutic strategies that specifically target these acylation pathways, offering new possibilities for precision medicine.
Therapeutic targets of L- and D-Lactate
Key enzymes and transporters involved in lactate homeostasis, such as glycolytic enzymes, MCTs, LDH and mitochondrial pyruvate carriers, have emerged as promising therapeutic targets [6]. Future lactylation-targeted therapeutic approaches will likely focus on developing specific inhibitors and activators that precisely modulate the activity of lactylation-related enzymes, thereby influencing disease progression. Notably, MCT expression is clinically associated with cancer metastasis and poor prognosis [47, 183, 184], highlighting the need to clarify the roles of MCT1 and MCT4 in cancer cell survival and drug resistance. A variety of MCT inhibitors, including cyanoacetic acid, coumarin, flavone, and indole derivatives, have been explored [185], yet only a few have advanced to clinical trials due to challenges with side effects and potential off-target effects. For example, AZD3965, an MCT1 inhibitor that entered phase 1 trials in 2013 under Cancer Research UK’s Centre for Drug Development, was generally well tolerated, although common side effects included nausea and fatigue, with dose-limiting cardiac toxicity observed at 20 mg [186].
While MCT4 inhibitors are less developed, recent advancements are encouraging. AZD0095, a selective MCT4 inhibitor, has demonstrated antitumor effects in mouse models when combined with VEGFR inhibitors [187]. However, its efficacy in clinical settings remains to be established, underscoring the need for further trials. Exploring the therapeutic potential of targeting lactate metabolism could significantly advance cancer treatment, particularly by addressing challenges related to drug resistance and metastasis. As research progresses, metabolic modulation may emerge as a critical strategy for innovative cancer therapies.
In addition to MCTs, LDH inhibition represents a promising avenue for anticancer strategies. LDH, composed of various isozymes formed by A and B subunits, includes a recently discovered isozyme, LDH-X, in mature human testes and sperm [188]. Despite extensive research, developing LDH inhibitors with drug-like properties remains challenging. Many inhibitors, such as 2-amino-5-aryl-pyrazine and 3,6-disubstituted dihydropyrones, show potential, yet high LDH abundance and increased expression in tumors indicate that effective inhibition may require high doses, limiting therapeutic viability [189]. To enhance efficacy, combining LDH inhibitors with compounds that target gene expression may offer a solution. Unlike enzyme inhibitors, which often impact both LDH-A and LDH-B, expression inhibitors can selectively target LDH-A through distinct gene regulatory mechanisms. This dual approach, which addresses both LDH expression and activity, could improve outcomes in LDH-targeted therapies. Research into these combined mechanisms may reveal new therapeutic pathways for cancer and other diseases.
In conclusion, as research into D- and L-lactate has progressed, it has become increasingly clear that these molecules play pivotal roles beyond simple metabolism, acting as regulators of cellular homeostasis, immune function, and gene expression. This review highlights the distinct and complementary functions of D- and L-lactate across physiological and pathological landscapes, revealing their potential as dynamic modulators within diverse biological systems. To fully elucidate the breadth of the biological effects of lactate, future studies must integrate lactate metabolism into the broader context of metabolic and signaling networks, recognizing it as a fundamental component of cellular regulation. The advancement of sophisticated analytical and computational methods will be key in unraveling the complexities of these stereoisomers, enabling the discovery of novel therapeutic targets, particularly for diseases such as cancer, metabolic syndromes, and neurodegenerative disorders. Furthermore, understanding the nuanced role of lactylation and its interplay with other acyl modifications is important for opening entirely new avenues in both basic research and translational science.
Data availability
No datasets were generated or analysed during the current study.
Abbreviations
- LDH:
-
Lactate dehydrogenase
- NAD:
-
Nicotinamide adenine dinucleotide
- MCT:
-
Monocarboxylate transporter
- TCA:
-
Tricarboxylic acid cycle
- ATP:
-
Adenosine triphosphate
- LDHA:
-
Lactate dehydrogenase A
- LDHB:
-
Lactate dehydrogenase B
- GAP:
-
Glyceraldehyde 3-phosphate
- GAPDH:
-
Glyceraldehyde 3-phosphate dehydrogenase
- PDH:
-
Pyruvate dehydrogenase
- L-LDH:
-
L-lactate dehydrogenase A with M subunit
- D-2-HDH:
-
D-2-hydroxy acid dehydrogenase
- DHAP:
-
Dihydroxyacetone phosphate
- MGO:
-
Methylglyoxal
- GLO1:
-
Glyoxalase 1
- GLO2:
-
Glyoxalase 2
- mLDH:
-
Mitochondrial lactate dehydrogenase
- ADP:
-
Adenosine diphosphate
- HCA1:
-
Hydroxycarboxylic acid receptor 1
- Treg:
-
Regulatory T cell
- ETC:
-
Electron transport chain
- ROS:
-
Reactive oxygen species
- NK:
-
Natural killer cells
- AARS:
-
Alanyl-tRNA synthetase
- PERK:
-
R-like ER kinase
- TME:
-
Tumor microenvironment
- AK2:
-
adenylate kinase 2
- eEFiA2:
-
Eukaryotic translation elongation factor 1 alpha 2
- CDK7:
-
cyclin-dependent kinase 7
- YAP:
-
Yes-associated protein 1
- LDHD:
-
D-lactate dehydrogenase D
- MOESIN:
-
Membrane-organizing extension spike protein
- IBD:
-
Inflammatory Bowel Disease
- UC:
-
Ulcerative colitis
- CD:
-
Crohn’s disease
- AD:
-
Alzheimer’s disease
- DM:
-
Type 2 diabetes mellitus
- CAF:
-
Cancer-associated fibroblasts
- METTL:
-
Methyltransferase
- BCAP:
-
B cell adapter for PI3K
- GSK3β:
-
Glycogen synthase kinase 3 beta
- FOXO1:
-
Forkhead box1
- PHD:
-
Prolyl hydroxylate
- TNF:
-
α-Tumor Necrosis Factorα
- ARF1:
-
ADP-ribosylation factor 1
- IDH3β:
-
Isocitrate dehydrogenase 3 noncatalytic subunit beta
- PAX6:
-
Paired box 6
- S6K2:
-
Protein S6 kinase 2
- VEGF:
-
Vascular endothelial growth factor
- HIF:
-
Hypoxia inducible factor
- HAT:
-
Histone acetyl transferase
- GCN5:
-
Lysine acetyltransferase 2 A
- HDAC:
-
Class I histone deacetylases
- SIRT:
-
Sirtuin
- SMARCA4:
-
SWI/SNF Related, matrix associated, actin dependent regulator of chromatin, subfamily A, member 4
References
Ferguson BS, Rogatzki MJ, Goodwin ML, Kane DA, Rightmire Z, Gladden LB. Lactate metabolism: historical context, prior misinterpretations, and current Understanding. Eur J Appl Physiol. 2018;118:691–728.
Brooks GA. Lactate as a fulcrum of metabolism. Redox Biol. 2020;35:101454.
Brooks GA. Lactate shuttles in nature. Biochem Soc Trans. 2002;30:258–64.
Sun W, Jia M, Feng Y, Cheng X. Lactate is a Bridge linking Glycolysis and autophagy through lactylation. Autophagy. 2023;19:3240–1.
Brooks GA. The science and translation of lactate shuttle theory. Cell Metab. 2018;27:757–85.
Rabinowitz JD, Enerback S. Lactate: the ugly duckling of energy metabolism. Nat Metab. 2020;2:566–71.
Certo M, Llibre A, Lee W, Mauro C. Understanding lactate sensing and signalling. Trends Endocrinol Metab. 2022;33:722–35.
Li X, Yang Y, Zhang B, Lin X, Fu X, An Y, Zou Y, Wang JX, Wang Z, Yu T. Lactate metabolism in human health and disease. Signal Transduct Target Ther. 2022;7:305.
Halestrap AP, Price NT. The proton-linked monocarboxylate transporter (MCT) family: structure, function and regulation. Biochem J. 1999;343:281–99.
Halestrap AP. The SLC16 gene family - structure, role and regulation in health and disease. Mol Aspects Med. 2013;34:337–49.
Brooks GA. Cell-cell and intracellular lactate shuttles. J Physiol. 2009;587:5591–600.
Brooks GA. Role of the heart in lactate shuttling. Front Nutr. 2021;8:663560.
Adeva M, Gonzalez-Lucan M, Seco M, Donapetry C. Enzymes involved in l-lactate metabolism in humans. Mitochondrion. 2013;13:615–29.
Cuozzo F, Viloria K, Shilleh AH, Nasteska D, Frazer-Morris C, Tong J, Jiao Z, Boufersaoui A, Marzullo B, Rosoff DB, et al. LDHB contributes to the regulation of lactate levels and basal insulin secretion in human pancreatic Β cells. Cell Rep. 2024;43:114047.
Zhang D, Gao J, Zhu Z, Mao Q, Xu Z, Singh PK, Rimayi CC, Moreno-Yruela C, Xu S, Li G, et al. Lysine L-lactylation is the dominant lactylation isomer induced by Glycolysis. Nat Chem Biol. 2025;21:91–9.
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. 2019;574:575–80.
Koppenol WH, Bounds PL, Dang CV. Otto Warburg’s contributions to current concepts of cancer metabolism. Nat Rev Cancer. 2011;11:325–37.
Johar D, Elmehrath AO, Khalil RM, Elberry MH, Zaky S, Shalabi SA, Bernstein LH. Protein networks linking Warburg and reverse Warburg effects to cancer cell metabolism. BioFactors. 2021;47:713–28.
Vander Heiden MG, Cantley LC, Thompson CB. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science. 2009;324:1029–33.
DeBerardinis RJ, Chandel NS. Fundamentals of cancer metabolism. Sci Adv. 2016;2:e1600200.
Remund B, Yilmaz B, Sokollik C. D-Lactate: implications for Gastrointestinal diseases. Child (Basel). 2023;10.
Levitt MD, Levitt DG. Quantitative evaluation of D-Lactate pathophysiology: new insights into the mechanisms involved and the many areas in need of further investigation. Clin Exp Gastroenterol. 2020;13:321–37.
Cai X, Ng CC, Jones O, Fung TS, Ryu K, Li D, Thompson CB. Lactate activates the mitochondrial electron transport chain independent of its metabolism. bioRxiv 2023.
Thurn JR, Pierpont GL, Ludvigsen CW, Eckfeldt JH. D-lactate encephalopathy. Am J Med. 1985;79:717–21.
Kowlgi NG, Chhabra L. D-lactic acidosis: an underrecognized complication of short bowel syndrome. Gastroenterol Res Pract. 2015;2015:476215.
de Bari L, Atlante A, Armeni T, Kalapos MP. Synthesis and metabolism of Methylglyoxal, S-D-lactoylglutathione and D-lactate in cancer and Alzheimer’s disease. Exploring the crossroad of eternal youth and premature aging. Ageing Res Rev. 2019;53:100915.
Hall MM, Rajasekaran S, Thomsen TW, Peterson AR. Lactate: friend or foe. Pm R. 2016;8:S8–15.
Philp A, Macdonald AL, Watt PW. Lactate–a signal coordinating cell and systemic function. J Exp Biol. 2005;208:4561–75.
Cai X, Ng CP, Jones O, Fung TS, Ryu KW, Li D, Thompson CB. Lactate activates the mitochondrial electron transport chain independently of its metabolism. Mol Cell. 2023;83:3904–e39203907.
Emhoff CA, Messonnier LA, Horning MA, Fattor JA, Carlson TJ, Brooks GA. Gluconeogenesis and hepatic glycogenolysis during exercise at the lactate threshold. J Appl Physiol (1985). 2013;114:297–306.
Xu J, Ji T, Li G, Zhang H, Zheng Y, Li M, Ma J, Li Y, Chi G. Lactate attenuates astrocytic inflammation by inhibiting ubiquitination and degradation of NDRG2 under oxygen-glucose deprivation conditions. J Neuroinflammation. 2022;19:314.
Navas LE, Carnero A. NAD(+) metabolism, stemness, the immune response, and cancer. Signal Transduct Target Ther. 2021;6:2.
Chen W, Wei L, Zhang Y, Shi D, Ren W, Zhang Z, Wang J, Shao W, Liu X, Chen C, Gao Q. Involvement of the two L-lactate dehydrogenase in development and pathogenicity in fusarium graminearum. Curr Genet. 2019;65:591–605.
Amin FM, Aristeidou S, Baraldi C, Czapinska-Ciepiela EK, Ariadni DD, Di Lenola D, Fenech C, Kampouris K, Karagiorgis G, Braschinsky M, Linde M. The association between migraine and physical exercise. J Headache Pain. 2018;19:83.
Cerexhe L, Easton C, Macdonald E, Renfrew L, Sculthorpe N. Blood lactate concentrations during rest and exercise in people with multiple sclerosis: A systematic review and meta-analysis. Mult Scler Relat Disord. 2022;57:103454.
Bhutia YD, Ganapathy V. Glutamine transporters in mammalian cells and their functions in physiology and cancer. Biochim Biophys Acta. 2016;1863:2531–9.
DeBerardinis RJ, Mancuso A, Daikhin E, Nissim I, Yudkoff M, Wehrli S, 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 U S A. 2007;104:19345–50.
Allaman I, Belanger M, Magistretti PJ. Methylglyoxal, the dark side of Glycolysis. Front Neurosci. 2015;9:23.
Mayeur C, Gratadoux JJ, Bridonneau C, Chegdani F, Larroque B, Kapel N, Corcos O, Thomas M, Joly F. Faecal D/L lactate ratio is a metabolic signature of microbiota imbalance in patients with short bowel syndrome. PLoS ONE. 2013;8:e54335.
Kaur C, Sharma S, Hasan MR, Pareek A, Singla-Pareek SL, Sopory SK. Characteristic variations and similarities in biochemical, molecular, and functional properties of glyoxalases across prokaryotes and eukaryotes. Int J Mol Sci. 2017;18.
Jin S, Chen X, Yang J, Ding J. Lactate dehydrogenase D is a general dehydrogenase for D-2-hydroxyacids and is associated with D-lactic acidosis. Nat Commun. 2023;14:6638.
Kraut JA, Madias NE. Lactic acidosis. N Engl J Med. 2014;371:2309–19.
Salon JE. Lactic acidosis. N Engl J Med. 2015;372:1076.
Petersen C. D-lactic acidosis. Nutr Clin Pract. 2005;20:634–45.
Ewaschuk JB, Naylor JM, Zello GA. D-lactate in human and ruminant metabolism. J Nutr. 2005;135:1619–25.
Matin A, Konings WN. Transport of lactate and succinate by membrane vesicles of Escherichia coli, Bacillus subtilis and a pseudomonas species. Eur J Biochem. 1973;34:58–67.
Payen VL, Mina E, Van Hee VF, Porporato PE, Sonveaux P. Monocarboxylate transporters in cancer. Mol Metab. 2020;33:48–66.
Núñez MF, Kwon O, Wilson TH, Aguilar J, Baldoma L, Lin EC. Transport of L-Lactate, D-Lactate, and glycolate by the LldP and glca membrane carriers of Escherichia coli. Biochem Biophys Res Commun. 2002;290:824–9.
Fan C, Davison PA, Habgood R, Zeng H, Decker CM, Gesell Salazar M, Lueangwattanapong K, Townley HE, Yang A, Thompson IP, et al. Chromosome-free bacterial cells are safe and programmable platforms for synthetic biology. Proc Natl Acad Sci U S A. 2020;117:6752–61.
Kobayashi M, Narumi K, Furugen A, Iseki K. Transport function, regulation, and biology of human monocarboxylate transporter 1 (hMCT1) and 4 (hMCT4). Pharmacol Ther. 2021;226:107862.
Liu T, Han S, Yao Y, Zhang G. Role of human monocarboxylate transporter 1 (hMCT1) and 4 (hMCT4) in tumor cells and the tumor microenvironment. Cancer Manag Res. 2023;15:957–75.
Halestrap AP, Wilson MC. The monocarboxylate transporter family–role and regulation. IUBMB Life. 2012;64:109–19.
Srinivas SR, Gopal E, Zhuang L, Itagaki S, Martin PM, Fei YJ, Ganapathy V, Prasad PD. Cloning and functional identification of slc5a12 as a sodium-coupled low-affinity transporter for monocarboxylates (SMCT2). Biochem J. 2005;392:655–64.
Bovenzi CD, Hamilton J, Tassone P, Johnson J, Cognetti DM, Luginbuhl A, Keane WM, Zhan T, Tuluc M, Bar-Ad V et al. Prognostic Indications of Elevated MCT4 and CD147 across Cancer Types: A Meta-Analysis. Biomed Res Int. 2015;2015:242437.
Ouyang J, Wang H, Huang J. The role of lactate in cardiovascular diseases. Cell Commun Signal. 2023;21:317.
Ma XM, Geng K, Wang P, Jiang Z, Law BY, Xu Y. MCT4-dependent lactate transport: a novel mechanism for cardiac energy metabolism injury and inflammation in type 2 diabetes mellitus. Cardiovasc Diabetol. 2024;23:96.
Zhang D, Gao J, Zhu Z, Mao Q, Xu Z, Singh PK, Rimayi CC, Moreno-Yruela C, Xu S, Li G et al. Lysine L-lactylation is the dominant lactylation isomer induced by Glycolysis. Nat Chem Biol. 2024.
Lee S, Choi Y, Jeong E, Park J, Kim J, Tanaka M, Choi J. Physiological significance of elevated levels of lactate by exercise training in the brain and body. J Biosci Bioeng. 2023;135:167–75.
Brooks GA, Osmond AD, Arevalo JA, Duong JJ, Curl CC, Moreno-Santillan DD, Leija RG. Lactate as a myokine and exerkine: drivers and signals of physiology and metabolism. J Appl Physiol (1985). 2023;134:529–48.
Poole DC, Rossiter HB, Brooks GA, Gladden LB. The anaerobic threshold: 50 + years of controversy. J Physiol. 2021;599:737–67.
San-Millán I, Brooks GA. Assessment of metabolic flexibility by means of measuring blood lactate, fat, and carbohydrate oxidation responses to exercise in professional endurance athletes and Less-Fit individuals. Sports Med. 2018;48:467–79.
Monsorno K, Buckinx A, Paolicelli RC. Microglial metabolic flexibility: emerging roles for lactate. Trends Endocrinol Metab. 2022;33:186–95.
Jin J, Qiu S, Wang P, Liang X, Huang F, Wu H, Zhang B, Zhang W, Tian X, Xu R, et al. Cardamonin inhibits breast cancer growth by repressing HIF-1α-dependent metabolic reprogramming. J Exp Clin Cancer Res. 2019;38:377.
Carreno-Florez GP, Kocak BR, Hendricks MR, Melvin JA, Mar KB, Kosanovich J, Cumberland RL, Delgoffe GM, Shiva S, Empey KM, et al. Interferon signaling drives epithelial metabolic reprogramming to promote secondary bacterial infection. PLoS Pathog. 2023;19:e1011719.
Liu W, Wang Y, Bozi LHM, Fischer PD, Jedrychowski MP, Xiao H, Wu T, Darabedian N, He X, Mills EL, et al. Lactate regulates cell cycle by remodelling the anaphase promoting complex. Nature. 2023;616:790–7.
Mosienko V, Teschemacher AG, Kasparov S. Is L-lactate a novel signaling molecule in the brain? J Cereb Blood Flow Metab. 2015;35:1069–75.
MaruYama T, Kobayashi S, Shibata H, Chen W, Owada Y. Curcumin analog GO-Y030 boosts the efficacy of anti-PD-1 cancer immunotherapy. Cancer Sci. 2021;112:4844–52.
Guo S, Zhou J, Lou P, Weng L, Ye X, Guo J, Liu H, Ma R. Potentiated effects of lactate receptor GPR81 on immune microenvironment in breast cancer. Mol Carcinog. 2023;62:1369–77.
Yang K, Xu J, Fan M, Tu F, Wang X, Ha T, Williams DL, Li C. Lactate suppresses macrophage Pro-Inflammatory response to LPS stimulation by Inhibition of YAP and NF-kappaB activation via GPR81-Mediated signaling. Front Immunol. 2020;11:587913.
Brown TP, Ganapathy V. Lactate/GPR81 signaling and proton motive force in cancer: role in angiogenesis, immune escape, nutrition, and Warburg phenomenon. Pharmacol Ther. 2020;206:107451.
Sun Z, Han Y, Song S, Chen T, Han Y, Liu Y. Activation of GPR81 by lactate inhibits oscillatory shear stress-induced endothelial inflammation by activating the expression of KLF2. IUBMB Life. 2019;71:2010–9.
Offermanns S, Colletti SL, Lovenberg TW, Semple G, Wise A, AP IJ. International union of basic and clinical pharmacology. LXXXII: nomenclature and classification of Hydroxy-carboxylic acid receptors (GPR81, GPR109A, and GPR109B). Pharmacol Rev. 2011;63:269–90.
Ahmed K, Tunaru S, Tang C, Muller M, Gille A, Sassmann A, Hanson J, Offermanns S. An autocrine lactate loop mediates insulin-dependent Inhibition of lipolysis through GPR81. Cell Metab. 2010;11:311–9.
Yan J, Xie J, Xu S, Guo Y, Ji K, Li C, Gao H, Zhao L. Fibroblast growth factor 21 protects the liver from apoptosis in a type 1 diabetes mouse model via regulating L-lactate homeostasis. Biomed Pharmacother. 2023;168:115737.
Harun-Or-Rashid M, Inman DM. Reduced AMPK activation and increased HCAR activation drive anti-inflammatory response and neuroprotection in glaucoma. J Neuroinflammation. 2018;15:313.
Lauritzen KH, Morland C, Puchades M, Holm-Hansen S, Hagelin EM, Lauritzen F, Attramadal H, Storm-Mathisen J, Gjedde A, Bergersen LH. Lactate receptor sites link neurotransmission, neurovascular coupling, and brain energy metabolism. Cereb Cortex. 2014;24:2784–95.
Adeva-Andany M, Lopez-Ojen M, Funcasta-Calderon R, Ameneiros-Rodriguez E, Donapetry-Garcia C, Vila-Altesor M, Rodriguez-Seijas J. Comprehensive review on lactate metabolism in human health. Mitochondrion. 2014;17:76–100.
Pal A, Grossmann D, Glaß H, Zimyanin V, Günther R, Catinozzi M, Boeckers TM, Sterneckert J, Storkebaum E, Petri S et al. Glycolic acid and D-lactate-putative products of DJ-1-restore neurodegeneration in FUS - and SOD1-ALS. Life Sci Alliance. 2024;7.
de Bari L, Scirè A, Minnelli C, Cianfruglia L, Kalapos MP, Armeni T. Interplay among oxidative stress, Methylglyoxal pathway and S-Glutathionylation. Antioxid (Basel). 2020;10.
Quiroga J, Alarcón P, Ramírez MF, Manosalva C, Teuber S, Carretta MD, Burgos RA. d-lactate-induced etosis in cattle polymorphonuclear leucocytes is dependent on the release of mitochondrial reactive oxygen species and the PI3K/Akt/HIF-1 and GSK-3β pathways. Dev Comp Immunol. 2023;145:104728.
Chou CK, Lee YT, Chen SM, Hsieh CW, Huang TC, Li YC, Lee JA. Elevated urinary D-lactate levels in patients with diabetes and microalbuminuria. J Pharm Biomed Anal. 2015;116:65–70.
Enzo E, Santinon G, Pocaterra A, Aragona M, Bresolin S, Forcato M, Grifoni D, Pession A, Zanconato F, Guzzo G, et al. Aerobic Glycolysis tunes YAP/TAZ transcriptional activity. EMBO J. 2015;34:1349–70.
Jiang S, Chen X, Lin J, Huang P. Lactate-Oxidase-Instructed Cancer diagnosis and therapy. Adv Mater. 2023;35:e2207951.
Colbert LE, El Alam MB, Wang R, Karpinets T, Lo D, Lynn EJ, Harris TA, Elnaggar JH, Yoshida-Court K, Tomasic K, et al. Tumor-resident Lactobacillus iners confer chemoradiation resistance through lactate-induced metabolic rewiring. Cancer Cell. 2023;41:1945–62. e1911.
Wang ZH, Peng WB, Zhang P, Yang XP, Zhou Q. Lactate in the tumour microenvironment: from immune modulation to therapy. EBioMedicine. 2021;73:103627.
Cheng CS, Tan HY, Wang N, Chen L, Meng Z, Chen Z, Feng Y. Functional Inhibition of lactate dehydrogenase suppresses pancreatic adenocarcinoma progression. Clin Transl Med. 2021;11:e467.
de Bari L, Atlante A. Including the mitochondrial metabolism of L-lactate in cancer metabolic reprogramming. Cell Mol Life Sci. 2018;75:2763–76.
Chen H, Li Y, Li H, Chen X, Fu H, Mao D, Chen W, Lan L, Wang C, Hu K, et al. NBS1 lactylation is required for efficient DNA repair and chemotherapy resistance. Nature. 2024;631:663–9.
Chen Y, Wu J, Zhai L, Zhang T, Yin H, Gao H, Zhao F, Wang Z, Yang X, Jin M, et al. Metabolic regulation of homologous recombination repair by MRE11 lactylation. Cell. 2024;187:294–e311221.
Zong Z, Xie F, Wang S, Wu X, Zhang Z, Yang B, Zhou F. Alanyl-tRNA synthetase, AARS1, is a lactate sensor and lactyltransferase that lactylates p53 and contributes to tumorigenesis. Cell. 2024;187:2375–e23922333.
Li H, Liu C, Li R, Zhou L, Ran Y, Yang Q, Huang H, Lu H, Song H, Yang B et al. AARS1 and AARS2 sense L-lactate to regulate cGAS as global lysine lactyltransferases. Nature 2024.
Yang Z, Yan C, Ma J, Peng P, Ren X, Cai S, Shen X, Wu Y, Zhang S, Wang X, et al. Lactylome analysis suggests lactylation-dependent mechanisms of metabolic adaptation in hepatocellular carcinoma. Nat Metab. 2023;5:61–79.
Yang L, Niu K, Wang J, Shen W, Jiang R, Liu L, Song W, Wang X, Zhang X, Zhang R, et al. Nucleolin lactylation contributes to intrahepatic cholangiocarcinoma pathogenesis via RNA splicing regulation of MADD. J Hepatol. 2024;81:651–66.
De Leo A, Ugolini A, Yu X, Scirocchi F, Scocozza D, Peixoto B, Pace A, D’Angelo L, Liu JKC, Etame AB, et al. Glucose-driven histone lactylation promotes the immunosuppressive activity of monocyte-derived macrophages in glioblastoma. Immunity. 2024;57:1105–e11231108.
Li F, Si W, Xia L, Yin D, Wei T, Tao M, Cui X, Yang J, Hong T, Wei R. Positive feedback regulation between Glycolysis and histone lactylation drives oncogenesis in pancreatic ductal adenocarcinoma. Mol Cancer. 2024;23:90.
Xie B, Zhang M, Li J, Cui J, Zhang P, Liu F, Wu Y, Deng W, Ma J, Li X, et al. KAT8-catalyzed lactylation promotes eEF1A2-mediated protein synthesis and colorectal carcinogenesis. Proc Natl Acad Sci U S A. 2024;121:e2314128121.
Wang D, Du G, Chen X, Wang J, Liu K, Zhao H, Cheng C, He Y, Jing N, Xu P, et al. Zeb1-controlled metabolic plasticity enables remodeling of chromatin accessibility in the development of neuroendocrine prostate cancer. Cell Death Differ. 2024;31:779–91.
Yu J, Chai P, Xie M, Ge S, Ruan J, Fan X, Jia R. Histone lactylation drives oncogenesis by facilitating m(6)A reader protein YTHDF2 expression in ocular melanoma. Genome Biol. 2021;22:85.
Busing JD, Fouladi F, Bulik-Sullivan EC, Carroll IM, Fodor AA, Thomsen KF, Gulati AS, Nicholson MR. Gut microbial changes following fecal microbiota transplantation for D-Lactic acidosis in two children. JPGN Rep. 2023;4:e319.
Wang ZL, Pang SJ, Zhang KW, Li PY, Li PG, Yang C. Dietary vitamin A modifies the gut microbiota and intestinal tissue transcriptome, impacting intestinal permeability and the release of inflammatory factors, thereby influencing Aβ pathology. Front Nutr. 2024;11:1367086.
Xu R, Tan C, He Y, Wu Q, Wang H, Yin J. Dysbiosis of gut microbiota and Short-Chain fatty acids in encephalitis: A Chinese pilot study. Front Immunol. 2020;11:1994.
Lorenz I, Gentile A. D-lactic acidosis in neonatal ruminants. Vet Clin North Am Food Anim Pract. 2014;30:317–31.
Fabian E, Kramer L, Siebert F, Högenauer C, Raggam RB, Wenzl H, Krejs GJ. D-lactic acidosis - case report and review of the literature. Z Gastroenterol. 2017;55:75–82.
Lv X, Lv Y, Dai X. Lactate, histone lactylation and cancer hallmarks. Expert Rev Mol Med. 2023;25:e7.
Pan L, Feng F, Wu J, Fan S, Han J, Wang S, Yang L, Liu W, Wang C, Xu K. Demethylzeylasteral targets lactate by inhibiting histone lactylation to suppress the tumorigenicity of liver cancer stem cells. Pharmacol Res. 2022;181:106270.
Dikshit A, Lu J, Ford AE, Degan S, Jin YJ, Sun H, Nichols A, Salama AKS, Beasley G, Gooden D, Zhang JY. Potential utility of synthetic D-Lactate polymers in skin Cancer. JID Innov. 2021;1:100043.
Bridgewater HE, Bolitho EM, Romero-Canelón I, Sadler PJ, Coverdale JPC. Targeting cancer lactate metabolism with synergistic combinations of synthetic catalysts and monocarboxylate transporter inhibitors. J Biol Inorg Chem. 2023;28:345–53.
Wagner W, Ciszewski WM, Kania KD. L- and D-lactate enhance DNA repair and modulate the resistance of cervical carcinoma cells to anticancer drugs via histone deacetylase Inhibition and hydroxycarboxylic acid receptor 1 activation. Cell Commun Signal. 2015;13:36.
Affinito A, Quintavalle C, Chianese RV, Roscigno G, Fiore D, D’Argenio V, Thomas G, Savarese A, Ingenito F, Cocca L, et al. MCT4-driven CAF-mediated metabolic reprogramming in breast cancer microenvironment is a vulnerability targetable by miR-425-5p. Cell Death Discov. 2024;10:140.
Feichtinger RG, Lang R. Targeting L-Lactate Metabolism to Overcome Resistance to Immune Therapy of Melanoma and Other Tumor Entities. J Oncol. 2019;2019:2084195.
Zhou C, Li W, Liang Z, Wu X, Cheng S, Peng J, Zeng K, Li W, Lan P, Yang X, et al. Mutant KRAS-activated circATXN7 fosters tumor Immunoescape by sensitizing tumor-specific T cells to activation-induced cell death. Nat Commun. 2024;15:499.
Gu J, Zhou J, Chen Q, Xu X, Gao J, Li X, Shao Q, Zhou B, Zhou H, Wei S, et al. Tumor metabolite lactate promotes tumorigenesis by modulating MOESIN lactylation and enhancing TGF-β signaling in regulatory T cells. Cell Rep. 2022;39:110986.
Sun L, Zhang Y, Yang B, Sun S, Zhang P, Luo Z, Feng T, Cui Z, Zhu T, Li Y, et al. Lactylation of METTL16 promotes Cuproptosis via m(6)A-modification on FDX1 mRNA in gastric cancer. Nat Commun. 2023;14:6523.
Xiong J, He J, Zhu J, Pan J, Liao W, Ye H, Wang H, Song Y, Du Y, Cui B, et al. Lactylation-driven METTL3-mediated RNA m(6)A modification promotes immunosuppression of tumor-infiltrating myeloid cells. Mol Cell. 2022;82:1660–e16771610.
Raychaudhuri D, Singh P, Chakraborty B, Hennessey M, Tannir AJ, Byregowda S, Natarajan SM, Trujillo-Ocampo A, Im JS, Goswami S. Histone lactylation drives CD8(+) T cell metabolism and function. Nat Immunol. 2024;25:2140–51.
Xie Y, Hu H, Liu M, Zhou T, Cheng X, Huang W, Cao L. The role and mechanism of histone lactylation in health and diseases. Front Genet. 2022;13:949252.
Cui H, Xie N, Banerjee S, Ge J, Jiang D, Dey T, Matthews QL, Liu RM, Liu G. Lung myofibroblasts promote macrophage profibrotic activity through Lactate-induced histone lactylation. Am J Respir Cell Mol Biol. 2021;64:115–25.
Reynolds MB, Klein B, McFadden MJ, Judge NK, Navarrete HE, Michmerhuizen BC, Awad D, Schultz TL, Harms PW, Zhang L, et al. Type I interferon governs immunometabolic checkpoints that coordinate inflammation during Staphylococcal infection. Cell Rep. 2024;43:114607.
Irizarry-Caro RA, McDaniel MM, Overcast GR, Jain VG, Troutman TD, Pasare C. TLR signaling adapter BCAP regulates inflammatory to reparatory macrophage transition by promoting histone lactylation. Proc Natl Acad Sci U S A. 2020;117:30628–38.
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. 2019;30:1055–e10741058.
Tang Y, Zhu Y, He H, Peng Y, Hu P, Wu J, Sun W, Liu P, Xiao Y, Xu X, Wei M. Gut dysbiosis and intestinal barrier dysfunction promotes IgA nephropathy by increasing the production of Gd-IgA1. Front Med (Lausanne). 2022;9:944027.
Camara-Lemarroy CR, Silva C, Greenfield J, Liu WQ, Metz LM, Yong VW. Biomarkers of intestinal barrier function in multiple sclerosis are associated with disease activity. Mult Scler. 2020;26:1340–50.
Han S, Bao X, Zou Y, Wang L, Li Y, Yang L, Liao A, Zhang X, Jiang X, Liang D, et al. d-lactate modulates M2 tumor-associated macrophages and remodels immunosuppressive tumor microenvironment for hepatocellular carcinoma. Sci Adv. 2023;9:eadg2697.
Yan Y, Li X, Yang Q, Zhang H, Hettinga K, Li H, Chen W. Dietary d-Lactate intake facilitates inflammatory resolution by modulating M1 macrophage polarization. Mol Nutr Food Res. 2022;66:e2200196.
Heim CE, Bosch ME, Yamada KJ, Aldrich AL, Chaudhari SS, Klinkebiel D, Gries CM, Alqarzaee AA, Li Y, Thomas VC, et al. Lactate production by Staphylococcus aureus biofilm inhibits HDAC11 to reprogramme the host immune response during persistent infection. Nat Microbiol. 2020;5:1271–84.
McDonald B, Zucoloto AZ, Yu IL, Burkhard R, Brown K, Geuking MB, McCoy KD. Programing of an intravascular immune firewall by the gut microbiota protects against pathogen dissemination during infection. Cell Host Microbe. 2020;28:660–e668664.
Yang C, Zhang X, Wang S, Huo X, Wang J. Small intestinal bacterial overgrowth and evaluation of intestinal barrier function in patients with ulcerative colitis. Am J Transl Res. 2021;13:6605–10.
Song WB, Lv YH, Zhang ZS, Li YN, Xiao LP, Yu XP, Wang YY, Ji HL, Ma L. Soluble intercellular adhesion molecule-1, D-lactate and Diamine oxidase in patients with inflammatory bowel disease. World J Gastroenterol. 2009;15:3916–9.
Zhou J, Zhang L, Peng J, Zhang X, Zhang F, Wu Y, Huang A, Du F, Liao Y, He Y et al. Astrocytic LRP1 enables mitochondria transfer to neurons and mitigates brain ischemic stroke by suppressing ARF1 lactylation. Cell Metab. 2024;36:2054–2068.e2014.
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. 2022;34:634–e648636.
Wang X, Liu Q, Yu HT, Xie JZ, Zhao JN, Fang ZT, Qu M, Zhang Y, Yang Y, Wang JZ. A positive feedback Inhibition of isocitrate dehydrogenase 3β on paired-box gene 6 promotes Alzheimer-like pathology. Signal Transduct Target Ther. 2024;9:105.
Merkuri F, Rothstein M, Simoes-Costa M. Histone lactylation couples cellular metabolism with developmental gene regulatory networks. Nat Commun. 2024;15:90.
Morovic P, Gonzalez Moreno M, Trampuz A, Karbysheva S. In vitro evaluation of microbial D- and L-lactate production as biomarkers of infection. Front Microbiol. 2024;15:1406350.
Oh MS, Phelps KR, Traube M, Barbosa-Saldivar JL, Boxhill C, Carroll HJ. D-lactic acidosis in a man with the short-bowel syndrome. N Engl J Med. 1979;301:249–52.
Hingorani AD, Chan NN. D-lactate encephalopathy. Lancet. 2001;358:1814.
Zhu W, Guo S, Sun J, Zhao Y, Liu C. Lactate and lactylation in cardiovascular diseases: current progress and future perspectives. Metabolism. 2024;158:155957.
Zhang N, Zhang Y, Xu J, Wang P, Wu B, Lu S, Lu X, You S, Huang X, Li M, et al. α-myosin heavy chain lactylation maintains sarcomeric structure and function and alleviates the development of heart failure. Cell Res. 2023;33:679–98.
Ma W, Jia K, Cheng H, Xu H, Li Z, Zhang H, Xie H, Sun H, Yi L, Chen Z, et al. Orphan nuclear receptor NR4A3 promotes vascular calcification via histone lactylation. Circ Res. 2024;134:1427–47.
Li J, Hou W, Zhao Q, Han W, Cui H, Xiao S, Zhu L, Qu J, Liu X, Cong W, et al. Lactate regulates major zygotic genome activation by H3K18 lactylation in mammals. Natl Sci Rev. 2024;11:nwad295.
Zhang Y, Zhang X. Virus-Induced histone lactylation promotes virus infection in crustacean. Adv Sci (Weinh). 2024;11:e2401017.
Wang X, Fan W, Li N, Ma Y, Yao M, Wang G, He S, Li W, Tan J, Lu Q, Hou S. YY1 lactylation in microglia promotes angiogenesis through transcription activation-mediated upregulation of FGF2. Genome Biol. 2023;24:87.
Constant JS, Feng JJ, Zabel DD, Yuan H, Suh DY, Scheuenstuhl H, Hunt TK, Hussain MZ. Lactate elicits vascular endothelial growth factor from macrophages: a possible alternative to hypoxia. Wound Repair Regen. 2000;8:353–60.
Vegran F, Boidot R, Michiels C, Sonveaux P, Feron O. Lactate influx through the endothelial cell monocarboxylate transporter MCT1 supports an NF-kappaB/IL-8 pathway that drives tumor angiogenesis. Cancer Res. 2011;71:2550–60.
Sonveaux P, Vegran F, Schroeder T, Wergin MC, Verrax J, Rabbani ZN, De Saedeleer CJ, Kennedy KM, Diepart C, Jordan BF, et al. Targeting lactate-fueled respiration selectively kills hypoxic tumor cells in mice. J Clin Invest. 2008;118:3930–42.
Lee DC, Sohn HA, Park ZY, Oh S, Kang YK, Lee KM, Kang M, Jang YJ, Yang SJ, Hong YK, et al. A lactate-induced response to hypoxia. Cell. 2015;161:595–609.
Zhou X, Li J, Guo J, Geng B, Ji W, Zhao Q, Li J, Liu X, Liu J, Guo Z, et al. Gut-dependent microbial translocation induces inflammation and cardiovascular events after ST-elevation myocardial infarction. Microbiome. 2018;6:66.
Davidovics ZH, Vance K, Etienne N, Hyams JS. Fecal transplantation successfully treats recurrent D-Lactic acidosis in a child with short bowel syndrome. JPEN J Parenter Enter Nutr. 2017;41:896–7.
Więcek S, Chudek J, Woś H, Bożentowicz-Wikarek M, Kordys-Darmolinska B, Grzybowska-Chlebowczyk U. Serum level of D-Lactate in patients with cystic fibrosis: preliminary data. Dis Markers. 2018;2018:5940893.
Hasegawa H, Fukushima T, Lee JA, Tsukamoto K, Moriya K, Ono Y, Imai K. Determination of serum D-lactic and L-lactic acids in normal subjects and diabetic patients by column-switching HPLC with pre-column fluorescence derivatization. Anal Bioanal Chem. 2003;377:886–91.
Beisswenger PJ, Howell SK, Touchette AD, Lal S, Szwergold BS. Metformin reduces systemic Methylglyoxal levels in type 2 diabetes. Diabetes. 1999;48:198–202.
Lu J, Zello GA, Randell E, Adeli K, Krahn J, Meng QH. Closing the anion Gap: contribution of D-lactate to diabetic ketoacidosis. Clin Chim Acta. 2011;412:286–91.
Cruz RS, de Aguiar RA, Turnes T, Penteado Dos Santos R, de Oliveira MF, Caputo F. Intracellular shuttle: the lactate aerobic metabolism. Scientific World Journal. 2012;2012:420984.
Bosshart PD, Kalbermatter D, Bonetti S, Fotiadis D. Mechanistic basis of L-lactate transport in the SLC16 solute carrier family. Nat Commun. 2019;10:2649.
Li J, Xia Y, Xu H, Xiong R, Zhao Y, Li P, Yang T, Huang Q, Shan F. Activation of brain lactate receptor GPR81 aggravates exercise-induced central fatigue. Am J Physiol Regul Integr Comp Physiol. 2022;323:R822–31.
Hashimoto T, Hussien R, Cho HS, Kaufer D, Brooks GA. Evidence for the mitochondrial lactate oxidation complex in rat neurons: demonstration of an essential component of brain lactate shuttles. PLoS ONE. 2008;3:e2915.
Sahlin K, Fernstrom M, Svensson M, Tonkonogi M. No evidence of an intracellular lactate shuttle in rat skeletal muscle. J Physiol. 2002;541:569–74.
Gizak A, McCubrey JA, Rakus D. Cell-to-cell lactate shuttle operates in heart and is important in age-related heart failure. Aging. 2020;12:3388–406.
Liu L, MacKenzie KR, Putluri N, Maletic-Savatic M, Bellen HJ. The glia-Neuron lactate shuttle and elevated ROS promote lipid synthesis in neurons and lipid droplet accumulation in glia via APOE/D. Cell Metab. 2017;26:719–e737716.
Sun Y, Wang Y, Chen ST, Chen YJ, Shen J, Yao WB, Gao XD, Chen S. Modulation of the Astrocyte-Neuron lactate shuttle system contributes to neuroprotective action of fibroblast growth factor 21. Theranostics. 2020;10:8430–45.
Sheikh-Hamad D. Hints for a kidney lactate shuttle and lactomone. Am J Physiol Ren Physiol. 2021;320:F1028–9.
Niu D, Wu Y, Lei Z, Zhang M, Xie Z, Tang S. Lactic acid, a driver of tumor-stroma interactions. Int Immunopharmacol. 2022;106:108597.
Gaffney DO, Jennings EQ, Anderson CC, Marentette JO, Shi T, Schou Oxvig AM, Streeter MD, Johannsen M, Spiegel DA, Chapman E, et al. Non-enzymatic lysine lactoylation of glycolytic enzymes. Cell Chem Biol. 2020;27:206–e213206.
Gong H, Zhong H, Cheng L, Li LP, Zhang DK. Post-translational protein lactylation modification in health and diseases: a double-edged sword. J Transl Med. 2024;22:41.
Rong Y, Dong F, Zhang G, Tang M, Zhao X, Zhang Y, Tao P, Cai H. The crosstalking of lactate-Histone lactylation and tumor. Proteom Clin Appl. 2023;17:e2200102.
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. 2022;131:893–908.
Niu Z, Chen C, Wang S, Lu C, Wu Z, Wang A, Mo J, Zhang J, Han Y, Yuan Y, et al. HBO1 catalyzes lysine lactylation and mediates histone H3K9la to regulate gene transcription. Nat Commun. 2024;15:3561.
Li H, Liu C, Li R, Zhou L, Ran Y, Yang Q, Huang H, Lu H, Song H, Yang B, et al. AARS1 and AARS2 sense L-lactate to regulate cGAS as global lysine lactyltransferases. Nature. 2024;634:1229–37.
Ju J, Zhang H, Lin M, Yan Z, An L, Cao Z, Geng D, Yue J, Tang Y, Tian L et al. The alanyl-tRNA synthetase AARS1 moonlights as a lactyltransferase to promote YAP signaling in gastric cancer. J Clin Invest. 2024;134.
Moreno-Yruela C, Zhang D, Wei W, Bæk M, Liu W, Gao J, Danková D, Nielsen AL, Bolding JE, Yang L, et al. Class I histone deacetylases (HDAC1-3) are histone lysine delactylases. Sci Adv. 2022;8:eabi6696.
Fan Z, Liu Z, Zhang N, Wei W, Cheng K, Sun H, Hao Q. Identification of SIRT3 as an eraser of H4K16la. iScience. 2023;26:107757.
Hu X, Huang X, Yang Y, Sun Y, Zhao Y, Zhang Z, Qiu D, Wu Y, Wu G, Lei L. Dux activates metabolism-lactylation-MET network during early iPSC reprogramming with Brg1 as the histone lactylation reader. Nucleic Acids Res. 2024;52:5529–48.
Yu Y, Yang W, Bilotta AJ, Zhao X, Cong Y, Li Y. L-lactate promotes intestinal epithelial cell migration to inhibit colitis. FASEB J. 2021;35:e21554.
Llorente-Folch I, Rueda CB, Perez-Liebana I, Satrustegui J, Pardo B. L-Lactate-Mediated neuroprotection against Glutamate-Induced excitotoxicity requires ARALAR/AGC1. J Neurosci. 2016;36:4443–56.
Moreno-Yruela C, Baek M, Monda F, Olsen CA. Chiral posttranslational modification to lysine epsilon-Amino groups. Acc Chem Res. 2022;55:1456–66.
Bao X, Wang Y, Li X, Li XM, Liu Z, Yang T, Wong CF, Zhang J, Hao Q, Li XD. Identification of ‘erasers’ for lysine crotonylated histone marks using a chemical proteomics approach. Elife. 2014;3.
Bhaskar A, Kumar S, Khan MZ, Singh A, Dwivedi VP, Nandicoori VK. Host Sirtuin 2 as an immunotherapeutic target against tuberculosis. Elife. 2020;9.
Delaney K, Tan M, Zhu Z, Gao J, Dai L, Kim S, Ding J, He M, Halabelian L, Yang L, et al. Histone lysine methacrylation is a dynamic post-translational modification regulated by HAT1 and SIRT2. Cell Discov. 2021;7:122.
Huang H, Lin S, Garcia BA, Zhao Y. Quantitative proteomic analysis of histone modifications. Chem Rev. 2015;115:2376–418.
Ntorla A, Burgoyne JR. The regulation and function of histone crotonylation. Front Cell Dev Biol. 2021;9:624914.
Shang S, Liu J, Hua F. Protein acylation: mechanisms, biological functions and therapeutic targets. Signal Transduct Target Ther. 2022;7:396.
Zhao S, Zhang X, Li H. Beyond histone acetylation-writing and erasing histone acylations. Curr Opin Struct Biol. 2018;53:169–77.
Liu X, Wei W, Liu Y, Yang X, Wu J, Zhang Y, Zhang Q, Shi T, Du JX, Zhao Y, et al. MOF as an evolutionarily conserved histone crotonyltransferase and transcriptional activation by histone acetyltransferase-deficient and crotonyltransferase-competent CBP/p300. Cell Discov. 2017;3:17016.
Payen VL, Hsu MY, Radecke KS, Wyart E, Vazeille T, Bouzin C, Porporato PE, Sonveaux P. Monocarboxylate transporter MCT1 promotes tumor metastasis independently of its activity as a lactate transporter. Cancer Res. 2017;77:5591–601.
De Saedeleer CJ, Porporato PE, Copetti T, Perez-Escuredo J, Payen VL, Brisson L, Feron O, Sonveaux P. Glucose deprivation increases monocarboxylate transporter 1 (MCT1) expression and MCT1-dependent tumor cell migration. Oncogene. 2014;33:4060–8.
Puri S, 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. 2020;199:112393.
Benyahia Z, Blackman M, Hamelin L, Zampieri LX, Capeloa T, Bedin ML, Vazeille T, Schakman O, Sonveaux P. In vitro and in vivo characterization of MCT1 inhibitor AZD3965 confirms preclinical safety compatible with breast Cancer treatment. Cancers (Basel). 2021;13.
Goldberg FW, Kettle JG, Lamont GM, Buttar D, Ting AKT, McGuire TM, Cook CR, Beattie D, Morentin Gutierrez P, Kavanagh SL, et al. Discovery of clinical candidate AZD0095, a selective inhibitor of monocarboxylate transporter 4 (MCT4) for oncology. J Med Chem. 2023;66:384–97.
Gupta GS. LDH-C4: a target with therapeutic potential for cancer and contraception. Mol Cell Biochem. 2012;371:115–27.
Bononi G, Di Bussolo V, Tuccinardi T, Minutolo F, Granchi C. A patent review of lactate dehydrogenase inhibitors (2014-present). Expert Opin Ther Pat. 2024;34:1121–35.
Funding
This study is supported by the National Natural Science Foundation of China (No.82201841; No.U23A20420), Shanxi Basic Research Program (No. 202203021212369) and Science Research Start⁃up Fund for Doctor of Shanxi Medical University (No. SD2209;No. XD2114) and Shanxi Province Higher Education “Billion Project” Science and Technology Guidance Project (No. BYBLD006).
Author information
Authors and Affiliations
Contributions
J.L. and P.M. wrote the main manuscript and original draft and Z.L. and J.X. edited the manuscript. All authors reviewed the manuscript.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Additional information
Publisher’s note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Li, J., Ma, P., Liu, Z. et al. L- and D-Lactate: unveiling their hidden functions in disease and health. Cell Commun Signal 23, 134 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12964-025-02132-z
Received:
Accepted:
Published:
DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12964-025-02132-z