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.
Post-translational modifications (PTMs) increase proteomic functional diversity by covalently adding functional groups, regulating the proteolytic cleavage of subunits or the degradation of entire proteins. These include glycosylation, ubiquitination, small ubiquitin-like modification, acetylation, phosphorylation and palmitoylation, affecting nearly all aspects of cell biology and pathology (1-5). Lactate, once considered a waste product of glucose metabolism, has been revealed in recent years as not only a key circulating energy substrate, but also as a signaling molecule. Notably, it profoundly regulates gene expression and cell fate through a novel PTM, lysine lactylation (Kla) (6-10). In 2019, Zhang et al (11) first reported histone Kla in Nature, inaugurating a new era in lactate biology research. Kla has rapidly emerged as a research hotspot in life sciences, with its mechanisms and pathological significance providing transformative insights in fields such as cancer, immunology and neurological diseases (11-16). Kla involves the transfer of a lactyl group to the amino group of lysine residues, a process regulated by a series of enzymes and metabolites, including lactate, 'writer' enzymes, 'reader' proteins and 'eraser' enzymes (17). Lactate accumulation is the core trigger for Kla, with its level being positively associated with the degree of Kla modification (18,19). Under the action of 'writers', L-lactyl-CoA derived from lactate is transferred to proteins, initiating Kla. Erasers remove L-lactyl-CoA from proteins to terminate Kla, while 'reader' proteins recognize Kla and transduce signals to downstream targets (20-23).
The concept of ferroptosis was first proposed in 2012, referring to a unique form of programmed cell death triggered by iron-dependent lipid peroxidation pathways. It participates in multiple physiological and pathological processes and is prevalent in various diseases (24-28). In recent years, significant progress has been made in understanding the mechanisms of ferroptosis, including iron homeostasis imbalance, lipid peroxidation and the disruption of antioxidant systems (29-33). Furthermore, numerous factors regulate cellular sensitivity to ferroptosis under pathological conditions, thereby influencing disease progression (27). System xCT is an antiporter, with solute carrier family (SLC)7A11 as a key component, responsible for transporting cysteine and glutamate. It increases cellular cysteine uptake, which is converted to glutathione (GSH) under the action of thioredoxin reductase 1. Glutathione (GSH) peroxidase 4 (GPX4) relies on GSH as a substrate to enhance its activity, promoting the conversion of phospholipid hydroperoxides (PL-OOH) to lipid alcohols. Thus, system xCT prevents ferroptosis by reducing PL-OOH accumulation. Erastin and RAS-selective lethal 3 (RSL3, a ferroptosis inducer), as inhibitors of system xCT and GPX4, respectively, promote ferroptosis (34-36). An imbalance in iron homeostasis is another classical pathway in ferroptosis, primarily regulated by a network involving transferrin receptor 1 (TFR1), iron regulatory protein (IRP)1 and IRP2, affecting cellular iron uptake, storage and release (30,37,38). Reactive oxygen species (ROS) generation and phospholipid peroxidation require numerous metabolic enzymes, with iron acting as their catalyst and essential element. Iron-dependent Fenton reactions rapidly amplify PL-OOH and generate various reactive radicals, inducing ferroptosis in cancer cells (39-41). Unrestricted lipid peroxidation is a hallmark of ferroptosis. Acyl-CoA synthetase long-chain family member 4 (ACSL4) is a key enzyme converting polyunsaturated fatty acids (PUFAs) to phospholipids (PUFA-PEs). PUFA-PEs promote intracellular lipid peroxide accumulation under the action of various enzymes. Phospholipids rich in PUFAs in cell membranes and phospholipid peroxidation are considered the direct executors of ferroptosis (42,43).
Lactate is the driver of lactylation and is associated with lactylation levels (11). Lactate production and removal maintain electron flux through specific pathways, including NADH conversion to NAD+ and H+, and lactate dehydrogenase (LDH)-mediated lactate conversion. These reduced coenzymes generate electrons upon oxidation via mitochondrial respiration or lactate fermentation, maintaining redox balance (44,45). Lactate accumulation can also promote intracellular fatty acid synthesis by activating key enzymes and supplying precursors (46,47). Studies have also demonstrated that cellular redox imbalance and fatty acid metabolism are associated with ferroptosis (48,49). Increasing evidence indicates that lactylation regulates ferroptosis either by mediating gene transcription via histone modifications or by directly modifying key ferroptosis enzymes, altering their activity and inducing ferroptosis, thereby influencing disease pathogenesis. Notably, lactylation exhibits significantly higher dynamics than other PTMs, with its level directly regulated by microenvironmental lactate concentration. This characteristic renders Kla a crucial hub connecting cellular metabolic state and epigenetic regulation, particularly in diseases with active glycolysis (e.g., tumors and inflammation). However, a systematic description of the Kla-ferroptosis axis is currently lacking. Therefore, the present review discusses the regulation of ferroptosis by lactylation and summarizes its impact on diseases, including inflammation, degenerative diseases, cancer and ischemia-reperfusion (I/R) injury. The present review aimed to provide insight for future research and disease treatment.
Kla was initially discovered as a process modifying histone lysine residues, thereby altering binding to target gene promoters and activating transcription to exert biological effects. Subsequently, it expanded to non-histone lactylation. In ferroptosis, lactylation primarily involves Kla on histones H3 and H4, binding to promoter regions of target genes to alter transcriptional activity. On the other hand, Kla can directly modify non-histone lysine residues to mediate ferroptosis, demonstrating the multifaceted nature of Kla in regulating ferroptosis. Furthermore, Kla is subject to a complex interplay of upstream signals. A key determinant involves the participating lactyltransferases ('writers') and delactylases. EP300/p300/CREB-binding protein (CBP) are well-established 'writers' of histone lactylation (50-52), while other transferases such as lysine acetyltransferase (KAT), acetyl-CoA acetyltransferase 2 (ACAT2) and N-alpha-acetyltransferase 10 (NAA10) are emerging as modifiers of non-histone substrates (53-55). By contrast, histone deacetylase 1 (HDAC1) and HDAC3 have been identified as possessing delactylase activity (56,57). Their substrate specificity is influenced by expression patterns, spatial localization and interactions with specific factors in different cellular microenvironments. Additionally, metabolic status serves as a fundamental upstream signal. Lactate accumulation can directly drive lactyltransferase-catalyzed lactylation (55,58,59). Local pH in microenvironments, such as tumors, can also influence enzyme kinetics and substrate accessibility (60). Finally, crosstalk with other PTMs creates a regulatory network; for instance, prior acetylation or phosphorylation at or near lysine residues may competitively inhibit or cooperatively promote subsequent lactylation (56). This intricate network, comprising writers, metabolic signals and competitive PTMs, ultimately determines whether lactylation at specific sites on given proteins promotes or suppresses ferroptosis under particular pathophysiological conditions.
The nucleosome, consisting of a core histone octamer (two each of H2A, H2B, H3 and H4) and wrapped DNA, is the fundamental unit of chromatin. Typically, PTMs are key mechanisms dynamically regulating chromatin structure and DNA templating processes such as gene transcription (61,62). Lactylation modifications at histone H3 lysine 18 (H3K18la), lysine 14 (H3K14la) and histone H4 lysine 8 (H4K8la), and lysine 12 (H4K12la) alter histone function by the covalent addition of a lactyl group, mediating epigenetic reprogramming to regulate ferroptosis. The effect is dependent on the modification site, target gene and disease microenvironment (Fig. 1).
H3K18la is the most extensively studied histone lactylation site in ferroptosis. One of its core regulatory targets is the key fatty acid metabolism enzyme, ACSL4; elevated ACSL4 protein levels can induce transient mitochondrial activation, ROS accumulation and lipid peroxidation, ultimately driving ferroptosis. Notably, H3K18la plays context-dependent dual roles in regulating ACSL4. However, despite differing regulatory directions in different diseases, the outcome consistently promotes disease progression. In inflammatory diseases, lactate accumulation upregulates ACSL4 via H3K18la, promoting ferroptosis-mediated disease progression. The mechanisms involved include H3K18la enrichment in promoter regions, which enhances hypoxia-inducible factor (HIF)1A transcriptional activity, thereby transcriptionally activating ACSL4. Additionally, H3K18la can directly bind the ACSL4 gene promoter region to increase transcription (53,54). Simultaneously, in sepsis-mediated lung injury, H3K18la upregulates methyltransferase-like 3 (METTL3) by enhancing its promoter activity, increasing ACSL4 mRNA stability via N6-methyladenosine (m6A) RNA methylation modification and reader protein YTH domain-containing protein 1 dependency, leading to an elevated protein expression of ACSL4 (63). Conversely, in cancer, H3K18la suppresses ACSL4 expression, enhancing ferroptosis resistance and promoting disease progression. In endometriosis, H3K18la binds the promoter of methyltransferase METTL3, promoting its expression, which subsequently stabilizes HIF1A mRNA via m6ARNA methylation and upregulates heme oxygenase 1 (HMOX1), ultimately downregulating ACSL4 (64). H3K18la binding to the promoter region significantly enhances absent in melanoma 2 (AIM2) transcription, downregulating ACSL4 expression by promoting ubiquitin-mediated degradation of transcription factor STAT5B. This leads to decreased cellular lipid peroxidation, reduced GSH depletion and limited ferrous iron (Fe2+) accumulation, ultimately inhibiting ferroptosis (65). Furthermore, H3K18la can influence ferroptosis by regulating other key genes. H3K18la binding promotes transcription of zinc finger protein (ZFP)64, which directly binds the promoter regions of ferroptosis-related genes GTP cyclohydrolase 1 (GCH1) and ferritin heavy chain 1 (FTH1), promoting their transcription. GCH1 reduces ROS accumulation by inhibiting lipid peroxidation, while FTH1 reduces iron-dependent lipid peroxidation triggered by sequestering intracellular free Fe2; both synergistically inhibit ferroptosis (60). The reduced expression of H3K18la specifically decreases ferroptosis driver genes [activating transcription factor (ATF3), ATF4 and ChaC glutathione specific gamma-glutamylcyclotransferase 1 (CHAC1)], thereby inhibiting ferroptosis (66). H3K18la enrichment at the promoter region of the key iron-sulfur cluster synthesis enzyme NSF1 cysteine desulfurase (NFS1) enhances its transcription. NFS1 inhibits lipid peroxidation and ferroptosis by maintaining iron ion homeostasis (67).
H3K14la represents lactylation at another H3 site, K14, and is a clear signal promoting ferroptosis. H3K14la transcriptionally activates transferrin receptor (TFRC) and suppresses ferroportin SLC40A1 expression by binding the promoter regions of ferroptosis-related genes, leading to intracellular iron overload, lipid peroxidation and mitochondrial damage, ultimately triggering endothelial cell ferroptosis (52).
H4K8la represents lactylation at histone H4 K8. Its core mechanism involves tight coupling with glycolysis. 6-Phosphofructo-2-kinase/fructose-2,6-biphosphatase 3 (PFKFB3), a key glycolytic rate-limiting enzyme, is stabilized by deubiquitinase ubiquitin C-terminal hydrolase L1 (UCHL1), cleaving its K48-linked ubiquitin chains, promoting lactate production. Increased lactate levels further induce H4K8la modification, forming a positive feedback loop that activates UCHL1 and glycolytic-related genes [e.g., LDHA, pyruvate kinase M2 (PKM), MYC and Yes-associated protein 1], amplifying the inhibition of ferroptosis (68).
H4K12la, catalyzed by lactyltransferase p300 at histone H4 lysine 12, is enriched at the glutamate-cysteine ligase catalytic subunit (GCLC) promoter region, enhancing glutathione synthesis, inhibiting lipid peroxide accumulation and ultimately blocking the ferroptosis pathway (51).
Beyond histones, Kla precisely regulates cell fate by directly modifying key ferroptosis enzymes. This regulation operates through three core mechanisms: By directly targeting key ferroptosis effectors, remodeling cellular metabolic pathways and reprogramming the epigenetic landscape, collectively constituting a dynamic hub for ferroptosis regulation in response to metabolic stress (Fig. 2).
At the level of the direct targeting of key molecules, lactylation acts directly on core ferroptosis proteins, altering their stability or functional activity to regulate cellular antioxidant capacity and iron metabolism, thereby influencing ferroptosis progression. GPX4 is a key inhibitor of oxidative stress-induced ferroptosis. Lactate-driven lactylation at lysines 218 and 228 (K218/K228) of GPX4 significantly reduces its protein stability, impairing its ability to clear lipid peroxides and enhancing ferroptosis sensitivity (69,70). FTH1 alters intracellular iron homeostasis by influencing iron storage and release, and iron imbalance is a major trigger for ferroptosis (71,72). Tau protein lactylation at K677 upregulates nuclear receptor coactivator 4 (NCOA4) expression by activating p38 phosphorylation signaling, subsequently lowering FTH1 levels and inducing ferroptosis (73). Conversely, alpha-crystallin B chain (CRYAB), via lactylation, directly binds and stabilizes the FTH1 protein to enhance ferroptosis resistance (74).
Metabolic reprogramming constitutes the second major mechanism by which lactylation regulates ferroptosis, driving lipid peroxidation by intervening in mitochondrial energy metabolism and lipid synthesis pathways. The lactylation of the mitochondrial phosphoenolpyruvate carboxykinase 2 (PCK2) protein at K100 competitively inhibits E3 ubiquitin ligase Parkin-mediated ubiquitination and degradation of mitochondrial 3-oxoacyl-ACP synthase (OXSM), the rate-limiting enzyme of mitochondrial fatty acid synthesis (mtFAS). This promotes mtFAS pathway metabolic reprogramming, exacerbating ferroptosis (75). The inhibition of lactylation at K241 of malate dehydrogenase 2 (MDH2) maintains MDH2 catalytic activity, sustaining the tricarboxylic acid (TCA) cycle and blocking ferroptosis (76). Lactylation at K82 of sterol regulatory element-binding protein (SREBP) pathway regulator in Golgi (SPRING) protein activates the SREBP2-driven mevalonate pathway, ultimately conferring ferroptosis resistance to cancer cells through enhanced antioxidant capacity (53).
Epigenetic reprogramming forms the third pillar of lactylation-mediated ferroptosis regulation. It dynamically affects the transcription and mRNA stability of ferroptosis-related genes by modifying epigenetic regulators, such as RNA methylation enzymes and histone modifiers. In the epitranscriptome, m6A methylation bidirectionally regulates ferroptosis by increasing mRNA stability. On the one hand, lactylation of the m6A writer, METTL3, increases m6A methylation on TFRC mRNA, promoting TFRC mRNA stability, accelerating cellular iron uptake and inducing ferroptosis (77). De-lactylation at K412 of HDAC1 activates the transcription of the m6A erasers, fat mass and obesity-associated protein (FTO) and AlkB homolog 5 (ALKBH5), suppressing m6A modification on ferroptosis suppressor protein 1 (FSP1) mRNA and accelerating its degradation, thereby enhancing ferroptosis sensitivity (56). Conversely, lactylation at K240 of protein arginine methyltransferase 5 inhibits the transcription of the m6A eraser ALKBH5, leading to increased m6A modification and enhanced stability of solute carrier family 7 member 11 (SLC7A11) mRNA, conferring resistance to ferroptosis (78). m5C methylation is another way to regulate mRNA stability. Lactylation at K508 of NOP2/Sun RNA methyltransferase 2 (NSUN2) catalyzes m5C formation, maintaining GCLC mRNA stability, promoting GCLC-dependent glutathione synthesis and avoiding ferroptosis in cancer cells within the acidic tumor microenvironment (55). Furthermore, in epigenomics, lactylation can regulate gene transcription through key proteins beyond histones. For instance, lactylation at K503 of lysine-specific demethylase 1 (LSD1) enhances the formation of a stable complex with transcription factor FosL1, specifically enriching at the ferroportin receptor TFRC promoter region to suppress TFRC transcription, reduce iron uptake and enhance ferroptosis resistance (79).
As an emerging metabolic-epigenetic regulatory mechanism, Kla plays a core role in reshaping cell fate by targeting ferroptosis in inflammatory injury, neurodegenerative diseases, cancer and I/R injury, altering disease outcomes through actions on key targets. This section integrates the latest research progress on Kla-mediated ferroptosis in diseases, discusses the underlying pathological mechanisms and explores translational therapeutic potential (Table I).
Lactylation drives pro-ferroptotic pathways promoting disease progression. Previous studies have suggested that the inhibition of ferroptosis holds broad prospects for the prevention and treatment of inflammatory diseases. Lactate, the end product of glycolysis, accumulates in the extracellular environment, and the resulting acidosis is a hallmark of inflammatory diseases. Lactate in the microenvironment promotes inflammatory disease progression via Kla-mediated ferroptosis, rendering the inhibition of glycolysis and ferroptosis a promising therapeutic approach (7,80-82). Sepsis is a major cause of mortality in infectious diseases, and among organs susceptible to its harmful effects, the lungs are most frequently affected (83,84). Research indicates that lactylation-mediated ferroptosis promotes sepsis-associated lung injury. In sepsis-associated acute respiratory distress syndrome, H3K14la in pulmonary vascular endothelial cells transcriptionally activates TFRC and suppresses ferroportin SLC40A1, leading to iron overload and mitochondrial-damaging ferroptosis, promoting disease progression. Oxamate, an LDHA inhibitor reducing lactate production, lowers H3K14la modification levels, subsequently downregulating TFRC transcription, reducing iron uptake and alleviating endothelial cell ferroptosis (52). Sepsis-induced lung injury further reveals the core role of the H3K18la-METTL3-ACSL4 axis, inducing mitochondrial fatty acid oxidation-dependent ferroptosis; the METTL3 inhibitor, STM2457, effectively blocks this pathway (63). Severe acute pancreatitis (SAP) and non-alcoholic steatohepatitis (NASH) are non-infectious inflammatory diseases. In SAP, aberrant glycolysis causes lactate accumulation, enhancing HIF1A promoter activity via histone H3K18 lactylation, which transcriptionally activates the ACSL4/lysophosphatidylcholine acyltransferase 3/arachidonate lipoxygenase pathway, inducing pancreatic acinar cell ferroptosis and ultimately exacerbating organ damage. As previously demonstrated, combining a glycolysis inhibitor with a ferroptosis inducer (Ferrostatin-1) significantly alleviates organ damage (85). Previous research has also confirmed that Qing Xia Jie Yi Formula granules alleviate acute pancreatitis by inhibiting the glycolysis pathway (86). In NASH, irisin (TEC) recruits HDAC1 via tRF-31R9J to clear H3K18la, suppressing ferroptosis driver genes such as ATF3/CHAC1, protecting hepatocytes from ferroptosis and providing a novel intervention target for metabolic liver disease (66). These findings establish the 'lactylation-pro-ferroptosis gene activation' axis as a hub in the inflammatory cascade. The inhibition of glycolysis and lactylation may thus be a strategy for the treatment of infectious and inflammatory diseases.
Ferroptosis is a critical factor in degenerative diseases and its pharmacological inhibition is a therapeutic target for these conditions (87-89). Lactylation is a novel mechanism regulating ferroptosis, and modulating this process provides new hope for treatment. In degenerative diseases, one disease-promoting mechanism involves lactylation-mediated iron overload to trigger ferroptosis by regulating key genes in iron homeostasis. Research on Alzheimer's disease has found that tau protein de-lactylation (K677R mutation) inhibits p38 phosphorylation, downregulates NCOA4 and reduces ferritinophagy-dependent free iron release, thereby alleviating neuronal ferroptosis and improving cognitive function (73). Breakthrough research in osteoporosis (OP) revealed significantly downregulated CRYAB/FTH1 expression in patients with OP. CRYAB de-lactylation promotes ferroptosis in bone marrow-derived mesenchymal stem cells and inhibits osteogenic differentiation (74). These studies highlight the impact of lactylation on cell survival and disease progression via iron homeostasis regulation. Promoting ACSL4 expression via lactylation is another key mechanism regulating ferroptosis in degenerative diseases. LDHB promotes H3K18la and ACSL4 expression by increasing microenvironment lactylation levels, inducing chondrocyte ferroptosis and further promoting osteoarthritis development; inhibiting glycolysis alleviates osteoarthritis (90). During inflammation-induced human intervertebral disc degeneration, the shift from aerobic to anaerobic metabolism produces lactate that not only promotes H3K18la and ACSL4 transcription, but also increases ACSL4 lactylation at K412. Concurrently, lactate reduces the expression of sirtuin 3 (a Kla eraser), suppressing the clearance of ACSL4 lactylation. These factors increase cellular ferroptosis and promote disease progression (91).
Ferroptosis is recognized as a cause of reperfusion injury and organ failure, and the development of ferroptosis modulators provides novel opportunities for the treatment of I/R injury (92). Lactate accumulation is a hallmark of ischemia, and lactate-driven ferroptosis is often a key factor in disease progression. The inhibition of lactylation-mediated ferroptosis has been shown to exert positive effects in various studies on diseases. Research indicates that lactate accumulation upregulates METTL3 expression via EP300-mediated H3K18la, subsequently accelerating SLC7A11 mRNA degradation in an m6A modification- and YTH N6-methyladenosine RNA binding protein F2-dependent manner, ultimately inducing macrophage ferroptosis and promoting coronary heart disease progression (50). In myocardial I/R, enhanced glycolysis and lactate accumulation promote the lactylation of GPX4 at K218 and K228, destabilizing GPX4 and enhancing cardiomyocyte sensitivity to ferroptosis, exacerbating injury (70). Dexmedetomidine provides a novel cardioprotective strategy by inhibiting lactate production, blocking MDH2 K241 lactylation, restoring TCA cycle function and upregulating GPX4 (76). In intestinal ischemia, mitochondrial alanyl-tRNA synthetase 2 epigenetically upregulates ACSL4 transcriptional activity by enhancing H3K18la, driving lipid peroxidation and ferroptosis to worsen intestinal injury (93). In mice with intestinal I/R injury and intestinal epithelial cells subjected to hypoxia-reoxygenation, the nuclear translocation of pyruvate kinase M2 (PKM2) dimers promotes the expression of glycolytic key enzymes (GLUT1/enolase 1/LDHA) and lactate accumulation. Lactate drives histone modifier p300 to specifically catalyze H4K12la, enriching at the promoter region to significantly enhance high-mobility group box 1 (HMGB1) transcription. This disrupts the ferroptosis regulatory balance (ACSL4↑/GPX4↓), promoting intestinal damage. Targeting PKM2 dimerization (ML265) or HMGB1 (Gly-cyrrhizin) blocks this pathway and alleviates tissue injury (59). Research on liver I/R has revealed that lactate activates PCK2 K100 lactylation, stabilizing mitochondrial fatty acid synthase OXSM, and promoting lipid peroxidation and ferroptosis. KAT8 inhibitors or adeno-associated virus-short hairpin RNA targeting PCK2 significantly reduce liver injury (75). Intracerebral hemorrhage causes lactate accumulation and an acidic microenvironment within neurons. Lactate, mediated by histone acetyltransferases p300/CBP, specifically promotes lactylation at the H3K14 site. H3K14la functions as an epigenetic mark, directly inhibiting the transcription of the calcium pump plasma membrane Ca2+ ATPase 2 gene, hindering intracellular calcium efflux, causing calcium overload and exacerbating neuronal ferroptosis. The elevated lactate microenvironment induces the lactylation of METTL3 protein, increasing m6A methylation on TFRC mRNA, accelerating cellular iron uptake, leading to Fe2+ accumulation, elevated ROS and malondialdehyde levels, and ultimately inducing ferroptosis (77,94). In diabetic nephropathy, a high-lactate microenvironment inhibits the E3 ubiquitin ligase activity of tripartite motif-containing protein 65 (TRIM65) via p300-mediated lactylation at K206, preventing the degradation of its two key substrates, glycolytic kinase pyruvate dehydrogenase kinase 4 and iron regulatory protein iron-responsive element-binding protein 2. This promotes glycolytic flux and induces iron overload and lipid peroxidation, activating ferroptosis. Targeting the TRIM65-K206 lactylation site or reducing lactate production via sodium-glucose cotransporter 2 inhibitors restores the dual inhibition of ferroptosis and glycolysis by TRIM65 (58). However, lactate accumulation is not always detrimental to tissue injury. In spinal cord injury, deubiquitinase UCHL1 stabilizes the key glycolytic enzyme PFKFB3 by cleaving its K48-linked ubiquitin chains, enhancing glycolysis and lactate production in astrocytes. The produced lactate serves as an energy substrate for neurons and inhibits neuronal ferroptosis. Furthermore, elevated lactate levels induce H4K8la in astrocytes, promoting UCHL1 and glycolytic gene (e.g., PKM and LDHB) transcription, forming a positive feedback loop that amplifies the protective effect (62).
Under both hypoxic and aerobic conditions, tumor cells prefer glycolysis over oxidative phosphorylation as their primary energy source to meet rapid growth and proliferation demands, resulting in high glycolytic flux, massive lactate production and an acidic tumor microenvironment (94-98). Lactate accumulation in the tumor microenvironment promotes tumor cell invasion, metastasis and drug resistance by facilitating lactylation (12,79,99-101). Lactylation regulates key ferroptosis molecules through diverse targets, serving not only as a survival strategy for tumor cells to adapt to the microenvironment, but also as a crucial mechanism to evade therapeutic pressure.
Lactate is considered the initiator of lactylation-regulated ferroptosis. Therefore, reducing lactate production by inhibiting key glycolytic enzymes can suppress tumor growth. Aberrant glycolysis in lung cancer cells causes lactate accumulation, inducing histone H3K18 lactylation at the AIM2 gene promoter region, significantly enhancing AIM2 transcription. This promotes the ubiquitin-mediated degradation of the transcription factor STAT5B, downregulating its downstream target gene ACSL4 and ultimately inhibiting ferroptosis. The natural compound Shikonin can target and inhibit PKM2 (a key glycolytic enzyme) and AIM2 expression, reversing histone lactylation-mediated AIM2 activation (65). Cancer-associated fibroblasts enhance histone H3K18 lactylation by secreting lactate, activating ZFP64, which promotes GCH1 and FTH1 transcription. GCH1 reduces ROS accumulation by inhibiting lipid peroxidation, and FTH1 reduces iron-dependent lipid peroxidation triggered by sequestering free Fe2+; both synergistically inhibit ferroptosis, enhancing tumor cell drug resistance. The lactate production inhibitor Oxamate can reverse this resistant phenotype (60). ZNF207 increases lactate levels by upregulating LDHA expression, inducing lactylation at lysine 67 (K67 lactylation) of antioxidant protein peroxiredoxin 1 (PRDX1). Lactylated PRDX1 binds the transcription factor nuclear factor erythroid 2-related factor 2 (NRF2) and promotes its nuclear translocation, activating the NRF2 antioxidant pathway and upregulating ferroptosis inhibitor genes, such as SLC7A11, GPX4 and HMOX1, ultimately causing drug resistance in hepatocellular carcinoma (HCC) cells. Knockdown of ZNF207, blocking PRDX1 lactylation at K67, or inhibition of NRF2 activity, restore regorafenib sensitivity and induce ferroptosis (102).
The inhibition of the lactylation of key ferroptosis proteins is considered an effective anti-cancer strategy, as it can promote ferroptosis and suppress tumor growth. Lactate maintains GCLC mRNA stability via the NAA10-mediated lactylation of NSUN2 K508, evading cancer cell ferroptosis in the acidic tumor microenvironment (55). The re-activation of glycolysis causes lactate re-accumulation, inducing lactylation at lysine 503 of LSD1, blocking TRIM21-mediated LSD1 degradation and altering its genomic localization. The inhibition of H3K4me2 modification suppresses TFRC transcription, reducing cellular iron uptake, inhibiting ferroptosis and promoting the survival of resistant cells. Inhibition of LSD1 activates ferroptosis and effectively suppresses tumor growth (79). HDAC inhibitors reduce the lactylation of HDAC1 at K412, which in turn activates the transcription of the m6A demethylases, FTO and ALKBH5 via H3K27ac. This reduces m6A modification on ferroptosis suppressor protein FSP1 mRNA, accelerating its degradation and ultimately enhancing lipid peroxide accumulation and sensitivity to ferroptosis. Vorinostat (SAHA) and Trichostatin A (TSA) both significantly reduce HDAC1 K412 lactylation and inhibit tumor growth (56). As previously demonstrated, following incomplete microwave ablation, a sublethal temperature zone forms in the residual tumor area, triggering enhanced glycolysis and lactate accumulation in HCC cells, specifically upregulating H3K18la. This enhances NFS1 transcription, suppressing lipid peroxidation and ferroptosis by maintaining iron ion homeostasis, thereby enhancing the metastatic capacity of HCC cells. NFS1 deletion combined with oxaliplatin (OXA) synergistically inhibits HCC metastasis and overcomes resistance to OXA (67). Evodiamine inhibits monocarboxylate transporter 4 (MCT4) function to block lactate signaling, directly reducing H3K18la to restore Semaphorin 3A transcription and suppress programmed cell death ligand 1. It also downregulates GPX4, inducing ferroptosis, and ultimately inhibiting tumor growth (103).
Lactate enhances the resistance of tumor cells to ferroptosis through lactyltransferase-mediated lactylation. Therefore, the inhibition of lactyltransferase activity to promote ferroptosis is a potential strategy for cancer therapy. Driven by AKT hyperactivation, phosphorylated PCK1 interacts with LDHA, enhancing glycolytic activity and lactate production, promoting the KAT7-mediated lactylation of SPRING at K82. This subsequently activates the SREBP2-driven mevalonate pathway, promoting resistance to ferroptosis (53). Lactyltransferase p300 catalyzes H4K12la, which transcriptionally activates GCLC, enhancing glutathione synthesis, inhibiting lipid peroxide accumulation and ultimately blocking the ferroptosis pathway. The inhibition of p300, LDHA or GCLC enhances sensitivity to chemotherapy (51). KRAS G12D mutation enhances LDHA phosphorylation via the MEK/ERK pathway, promoting glycolysis and lactate generation. This induces the ACAT2-mediated lactylation of glutamate-cysteine ligase modifier subunit (GCLM) at K34, elevating GCL activity, increasing GSH synthesis and conferring ferroptosis resistance to KRASG 12D-mutant cancer cells. Inhibiting ACAT2 or mutating GCLM-K34R restores the sensitivity of cancer cells to ferroptosis inducers (e.g., erastin/RSL3) and significantly suppresses tumor growth (54).
As an emerging epigenetic regulatory mechanism, the precise detection of lactylation is crucial for elucidating its functions in physiological and pathological processes. The detection of lactylation primarily relies on mass spectrometry and specific antibody-based detection. Mass spectrometry, particularly liquid chromatography-tandem mass spectrometry, serves as the core method, enabling the highly sensitive identification of specific sites and quantification of modification levels. This method involves digesting protein samples with proteases to generate a mixture of peptides. These peptides are then separated by liquid chromatography and are subjected to precise mass detection and sequence analysis via mass spectrometry. This allows for the unbiased identification of specific lactylation sites and their absolute or relative quantification. However, challenges include potential misidentification due to the similar chemical properties of lactylation and acetylation, and the need for higher-resolution instruments due to low modification abundance (11,100,104-107). Specific antibody-based detection utilizes antibodies targeting specific modification sites. Techniques such as western blot analysis for quantification, immunofluorescence for observing intranuclear distribution and co-immunoprecipitation for exploring protein interactions enable the in situ and intuitive observation of the subcellular localization and tissue distribution characteristics of lactylation. This approach provides advantages, such as operational simplicity and high sensitivity; however, it is associated with risks of non-specific binding and the limited availability of commercial antibodies (108-112). The comprehensive application of these methods provides key technical support for elucidating the role of lactylation in tumorigenesis and development, laying the foundation for its clinical translation.
The intricate interplay between lactylation and ferroptosis has positioned the Kla-ferroptosis axis as an attractive novel therapeutic frontier across a broad spectrum of human diseases. Although basic research continues to uncover the enzymatic regulation, site-specific functions and context-dependent outcomes of this axis, its translational potential demands focused exploration. Looking ahead, clinical development targeting this axis may proceed through the following strategies: First, directly targeting lactylation regulatory mechanisms holds significant promise. Developing inhibitors of specific lactyltransferases ('writers') or activators of delactylases ('erasers') could enable precise control over pro-ferroptotic or anti-ferroptotic lactylation events (17,20). For example, p300/CBP inhibitors (e.g., SGC-CBP30, A-485) and KAT8 inhibitors (KAT8-IN-1/MC4033) may disrupt lactylation that drives disease progression (51,75). Conversely, deacetylase inhibitors (SAHA, TSA and TEC) can modulate lactylation and sensitize cells to ferroptosis, highlighting the therapeutic feasibility of targeting 'erasers' (56,66). Second, regulating lactate metabolism provides a broader, yet effective strategy with which to indirectly modulate the entire Kla-ferroptosis axis (7,101). Inhibiting key glycolytic enzymes (Oxamate, 2-DG, Shikonin and Simvastatin) or disrupting lactate transport via monocarboxylate transporters (MCTs), e.g., using Syrosingopine to inhibit MCT4, can effectively deplete lactate, the fundamental driver of lactylation (60,65,103). This approach is particularly attractive in tumors and inflammatory microenvironments characterized by a high glycolytic flux (12,82).
To date, to the best of our knowledge, no clinical studies have been designed with 'direct regulation of protein lactylation' as the primary endpoint. However, some drugs targeting lactylation-regulating enzymes and lactate metabolism-related proteins have entered clinical trials. Among them, the most advanced is CCS1477 (inobrodib). It is a highly effective p300/CBP bromodomain inhibitor. Its phase I/IIa clinical trials in advanced solid tumors and hematological malignancies have been initiated (NCT04068597) (113,114). In patients with metastatic castration-resistant prostate cancer (mCRPC), CCS1477 monotherapy was shown to reduce the expression of key oncogenic protein MYC and proliferation marker Ki-67 in tumor tissues (113). In trials for relapsed/refractory acute myeloid leukemia and multiple myeloma, the drug demonstrated the ability to induce leukemia cell differentiation and significantly reduce myeloma-related serum/urine biomarkers. Notably, in the hematological malignancy cohort, about one-third of heavily pretreated patients responded to CCS1477 monotherapy. Some of these patients had progression-free survival exceeding 12 months (114). Another potent p300/CBP bromodomain inhibitor, FT-7051, has also shown early clinical progress. Early results from an open-label phase I trial in patients with mCRPC indicated that its monotherapy achieved the effective exposure threshold predicted by pharmacokinetic/pharmacodynamic models. However, safety concerns exist. Approximately 80% of patients experienced mild to moderate treatment-related adverse events, including hyperglycemia (115). NEO2734 (EP31670) is a dual bromodomain inhibitor targeting both p300/CBP and bromodomain and extra-terminal motif proteins. Due to its broader mechanism, it has entered a phase I trial. This trial aims to treat patients with advanced cancer, including those with metastatic nuclear protein in testis midline carcinoma and refractory mCRPC (NCT05488548); the results are not yet available. Furthermore, drug development targeting HDAC has advanced significantly. Currently, five HDAC inhibitors are approved for clinical use in hematological malignancies. These include Vorinostat and Romidepsin for cutaneous T-cell lymphoma, Belinostat for peripheral T-cell lymphoma, Panobinostat for relapsed/refractory multiple myeloma and Tucidinostat for peripheral T-cell lymphoma (116-119). In solid tumors, HDAC inhibitor monotherapy shows limited efficacy. However, combination strategies with other agents hold promise. For example, the selective HDAC inhibitors Tucidinostat and Entinostat combined with endocrine therapy (e.g., Exemestane) improved progression-free survival and the objective response rate in hormone receptor-positive breast cancer trials (120,121). Combining HDAC inhibitors with immune checkpoint inhibitors also demonstrated synergistic anti-tumor activity in various solid tumors, such as melanoma and colorectal cancer (122,123). Challenges and failures exist in this field. Some trial results did not meet primary endpoints. For instance, Mocetinostat monotherapy was ineffective in urothelial carcinoma. Entinostat combined with Atezolizumab did not significantly prolong progression-free survival in advanced triple-negative breast cancer (124). Additionally, HDAC inhibitor applications are expanding into non-oncological areas. Romidepsin is being explored for HIV latency reversal (125), and Givinostat and Ricolinostat show potential in clinical studies for Duchenne muscular dystrophy and diabetic neuropathic pain, respectively (126,127). Overall, further development of HDAC inhibitors in solid tumors still faces toxicity issues. Future efforts should focus on optimizing combination strategies and developing more subtype-selective inhibitors to enhance their therapeutic value.
Research on lactate metabolism-related proteins remains limited. Glycolysis is the primary source of lactate. 2-deoxy-D-glucose (2-DG) is a broad-spectrum glycolysis inhibitor. Intranasal administration of 3.5% 2-DG was safe and well-tolerated in healthy volunteers. Systemic exposure was almost negligible. Local nasal concentrations reached effective anti-viral levels in vitro (NCT05314933) (128). LDHA is a key enzyme for lactate production in glycolysis. Its inhibitors show extensive pre-clinical evidence across multiple cancers. They can inhibit glycolysis, reduce lactate and acidification, suppress invasion and metastasis, and enhance chemo/radiotherapy sensitivity. However, most findings are still at the pre-clinical stage (129). For lactate transmembrane transport inhibition, only one MCT-targeting drug, AZD3965 (an MCT1 inhibitor), has entered clinical trials. It aims to evaluate safety and preliminary efficacy in patients with advanced cancer (NCT01791595). Preliminary results indicate manageable safety (130-132). Specific MCT4 inhibitors are mostly in pre-clinical development. However, dual inhibition or combination with LDH/GLUT inhibitors appears more promising. This approach may sustainably reduce lactate supply, improve the acidic microenvironment and restore sensitivity to ferroptosis (133).
In summary, clinical interventions directly regulating lactylation are still emerging. However, drugs targeting its key nodes have established a considerable clinical and translational foundation. Based on metabolic resistance mechanisms and solid tumor treatment experience, it may be recommended to prioritize combination therapies that integrate metabolic-epigenetic-ferroptosis signaling. This should be accompanied by biomarker stratification and optimized administration strategies, such as intranasal delivery for high lesion exposure and low systemic exposure. These steps will help accelerate the clinical translation of the Kla-ferroptosis axis.
Notably, the cross-regulatory mechanisms between lactylation and other programmed cell death pathways are increasingly becoming a focus of research. Beyond ferroptosis, Kla may modulate processes such as necroptosis, apoptosis and autophagy by modifying key proteins, forming a complex network governing cell fate decisions (134). For instance, preliminary studies suggest that lactate accumulation can influence the activity of receptor interacting serine/threonine kinase 3, a core necroptosis protein, via lactylation, thereby modulating cell death patterns in inflammatory diseases (135). In tumor models, Kla may promote cell survival and resistance to therapy by inhibiting apoptotic signaling pathways (e.g., BCL-2 family proteins) or enhancing the stability of autophagy-related proteins (e.g., light chain 3) (136,137). Such cross-regulation not only reveals the central role of metabolic-epigenetic networks in multiple death pathways, but also provides new insight for developing combination therapies targeting lactylation. Future studies are required to systematically elucidate the site-specific functions of Kla in necroptosis, apoptosis and autophagy, and clarify the dynamic interactions among these pathways to more comprehensively evaluate the therapeutic potential of targeting Kla.
However, the field still faces multiple challenges: Structural similarities between lactylation-regulating enzymes and other acyltransferases/deacetylases necessitate highly specific drugs to avoid off-target effects on PTMs such as acetylation; the dual role of the Kla-ferroptosis axis in diseases (e.g., promoting injury in sepsis while exerting protection in spinal cord injury) calls for cell- and microenvironment-specific targeted delivery systems; notably, dynamic detection technologies for lactylation remain underdeveloped, requiring novel in vivo imaging probes to facilitate patient stratification and treatment monitoring (108,112,138). Further studies are thus required to focus on designing highly specific modulators, constructing intelligent delivery systems and validating dynamic lactylation biomarkers. This may broaden the understanding of cell fate regulation mechanisms and may provide novel paradigms for the treatment of cancer, as well as inflammatory, degenerative and ischemic diseases.
Not applicable.
All authors contributed to the study conception and design. Material preparation, data collection and analysis were performed by ZL, YZ and SZ. The first draft of the manuscript was written by ZL and KW. All authors commented on previous versions of the manuscript. Data authentication is not applicable. All authors read and approved the final manuscript.
Not applicable.
Not applicable.
The authors declare that they have no competing interests.
During the preparation of this work, the authors used Grammarly (app.grammarly.com) for the purpose of checking and correcting spelling, grammar, and punctuation errors.
|
Kla |
lysine lactylation |
|
PTMs |
post-translational modifications |
|
GSH |
glutathione |
|
GPX4 |
glutathione peroxidase 4 |
|
ROS |
reactive oxygen species |
|
ACSL4 |
acyl-CoA synthetase long-chain family member 4 |
|
PUFAs |
polyunsaturated fatty acids |
|
LDH |
lactate dehydrogenase |
|
HIF1A |
hypoxia-inducible factor 1-alpha |
|
METTL3 |
methyltransferase-like 3 |
|
m6A |
N6-methyladenosine |
|
FTH1 |
ferritin heavy chain 1 |
|
PFKFB3 |
6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 3 |
|
NCOA4 |
nuclear receptor coactivator 4 |
|
FSP1 |
ferroptosis suppressor protein 1 |
|
HDAC1 |
histone deacetylase 1 |
|
SLC7A11 |
solute carrier family 7 member 11 |
|
IREB2 |
iron-responsive element-binding protein 2 |
|
NRF2 |
nuclear factor erythroid 2-related factor 2 |
|
HMOX1 |
heme oxygenase 1 |
|
HMGB1 |
high-mobility group box 1 |
|
HDACi |
HDAC inhibitor |
|
MCT4 |
monocarboxylate transporter 4 |
The figures were created with BioGDP (https://biogdp.com/) (139).
No funding was received.
|
Miao C, Huang Y, Zhang C, Wang X, Wang B, Zhou X, Song Y, Wu P, Chen ZS and Feng Y: Post-translational modifications in drug resistance. Drug Resist Updat. 78:1011732025. View Article : Google Scholar | |
|
Hirano A, Fu YH and Ptáček LJ: The intricate dance of post-translational modifications in the rhythm of life. Nat Struct Mol Biol. 23:1053–1060. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Tharuka MDN, Courelli AS and Chen Y: Immune regulation by the SUMO family. Nat Rev Immunol. 25:608–620. 2025. View Article : Google Scholar : PubMed/NCBI | |
|
Zhong Q, Xiao X, Qiu Y, Xu Z, Chen C, Chong B, Zhao X, Hai S, Li S, An Z and Dai L: Protein posttranslational modifications in health and diseases: Functions, regulatory mechanisms, and therapeutic implications. MedComm. 4:e2612023. View Article : Google Scholar : PubMed/NCBI | |
|
Wang Y, Hu J, Wu S, Fleishman JS, Li Y, Xu Y, Zou W, Wang J, Feng Y, Chen J and Wang H: Targeting epigenetic and posttranslational modifications regulating ferroptosis for the treatment of diseases. Signal Transduct Target Ther. 8:4492023. 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 | |
|
Brooks GA: The science and translation of lactate shuttle theory. Cell Metab. 27:757–785. 2018. View Article : Google Scholar : PubMed/NCBI | |
|
Hui S, Ghergurovich JM, Morscher RJ, Jang C, Teng X, Lu W, Esparza LA, Reya T, Le Zhan, Yanxiang Guo J, et al: Glucose feeds the TCA cycle via circulating lactate. Nature. 551:115–118. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Brooks GA, Osmond AD, Arevalo JA, Duong JJ, Curl CC, Moreno-Santillan DD and Leija RG: Lactate as a myokine and exerkine: drivers and signals of physiology and metabolism. J Appl Physiol (1985). 134:529–548. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Yao S, Chai H, Tao T, Zhang L, Yang X, Li X, Yi Z, Wang Y, An J, Wen G, et al: Role of lactate and lactate metabolism in liver diseases (Review). Int J Mol Med. 54:592024. View Article : Google Scholar : PubMed/NCBI | |
|
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 | |
|
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 | |
|
Chen J, Huang Z, Chen Y, Tian H, Chai P, Shen Y, Yao Y, Xu S, Ge S and Jia R: Lactate and lactylation in cancer. Signal Transduct Target Ther. 10:382025. View Article : Google Scholar : PubMed/NCBI | |
|
Chen L, Huang L, Gu Y, Cang W, Sun P and Xiang Y: Lactate-lactylation hands between metabolic reprogramming and immunosuppression. Int J Mol Sci. 23:119432022. View Article : Google Scholar : PubMed/NCBI | |
|
Liu J, Zhao F and Qu Y: Lactylation: A novel post-translational modification with clinical implications in CNS Diseases. Biomolecules. 14:11752024. View Article : Google Scholar : PubMed/NCBI | |
|
Li R, Yang Y, Wang H, Zhang T, Duan F, Wu K, Yang S, Xu K, Jiang X and Sun X: Lactate and lactylation in the brain: Current progress and perspectives. Cell Mol Neurobiol. 43:2541–2555. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Peng X and Du J: Histone and non-histone lactylation: Molecular mechanisms, biological functions, diseases, and therapeutic targets. Mol Biomed. 6:382025. View Article : Google Scholar : PubMed/NCBI | |
|
Certo M, Llibre A, Lee W and Mauro C: Understanding lactate sensing and signalling. Trends Endocrinol Metab. 33:722–735. 2022. 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 | |
|
Shang S, Liu J and Hua F: Protein acylation: Mechanisms, biological functions and therapeutic targets. Signal Transduct Target Ther. 7:3962022. View Article : Google Scholar : PubMed/NCBI | |
|
Fan Z, Liu Z, Zhang N, Wei W, Cheng K, Sun H and Hao Q: Identification of SIRT3 as an eraser of H4K16la. iScience. 26:1077572023. View Article : Google Scholar : PubMed/NCBI | |
|
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. 8:eabi66962022. View Article : Google Scholar : PubMed/NCBI | |
|
Hu X, Huang X, Yang Y, Sun Y, Zhao Y, Zhang Z, Qiu D, Wu Y, Wu G and Lei L: Dux activates metabolism-lactylation-MET network during early iPSC reprogramming with Brg1 as the histone lactylation reader. Nucleic Acids Res. 52:5529–5548. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
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 | |
|
Zheng J and Conrad M: Ferroptosis: When metabolism meets cell death. Physiol Rev. 105:651–706. 2025. View Article : Google Scholar | |
|
Wang Z, Wu C, Yin D and Do K: Ferroptosis: Mechanism and role in diabetes-related cardiovascular diseases. Cardiovasc Diabetol. 24:602025. View Article : Google Scholar : PubMed/NCBI | |
|
Zhao L, Zhou X, Xie F and Zhang L, Yan H, Huang J, Zhang C, Zhou F, Chen J and Zhang L: Ferroptosis in cancer and cancer immunotherapy. Cancer Commun (Lond). 42:88–116. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Tang D, Chen X, Kang R and Kroemer G: Ferroptosis: Molecular mechanisms and health implications. Cell Res. 31:107–125. 2021. View Article : Google Scholar : | |
|
Jiang X, Stockwell BR and Conrad M: Ferroptosis: Mechanisms, biology and role in disease. Nat Rev Mol Cell Biol. 22:266–282. 2021. View Article : Google Scholar : PubMed/NCBI | |
|
Rochette L, Dogon G, Rigal E, Zeller M, Cottin Y and Vergely C: Lipid peroxidation and iron metabolism: Two corner stones in the homeostasis control of ferroptosis. Int J Mol Sci. 24:4492022. View Article : Google Scholar | |
|
Alves F, Lane D, Nguyen TPM, Bush AI and Ayton S: In defence of ferroptosis. Signal Transduct Target Ther. 10:22025. View Article : Google Scholar : PubMed/NCBI | |
|
Chen X, Li J, Kang R, Klionsky DJ and Tang D: Ferroptosis: Machinery and regulation. Autophagy. 17:2054–2081. 2021. View Article : Google Scholar : | |
|
Stockwell BR: Ferroptosis turns 10: Emerging mechanisms, physiological functions, and therapeutic applications. Cell. 185:2401–2421. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Li FJ, Long HZ, Zhou ZW, Luo HY, Xu SG and Gao LC: System Xc -/GSH/GPX4 axis: An important antioxidant system for the ferroptosis in drug-resistant solid tumor therapy. Front Pharmacol. 13:9102922022. View Article : Google Scholar | |
|
Du Y and Guo Z: Recent progress in ferroptosis: Inducers and inhibitors. Cell Death Discov. 8:5012022. View Article : Google Scholar : PubMed/NCBI | |
|
Pal D and Sandur SK: Novel role of peroxiredoxin 6 in ferroptosis: Bridging selenium transport with enzymatic functions. Free Radic Biol Med. 238:611–620. 2025. View Article : Google Scholar : PubMed/NCBI | |
|
Ru Q, Li Y, Chen L, Wu Y, Min J and Wang F: Iron homeostasis and ferroptosis in human diseases: mechanisms and therapeutic prospects. Signal Transduct Target Ther. 9:2712024. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang Y, Zou L, Li X, Guo L, Hu B, Ye H and Liu Y: SLC40A1 in iron metabolism, ferroptosis, and disease: A review. WIREs Mech Dis. 16:e16442024. View Article : Google Scholar : PubMed/NCBI | |
|
Zheng D, Liu J, Piao H, Zhu Z, Wei R and Liu K: ROS-triggered endothelial cell death mechanisms: Focus on pyroptosis, parthanatos, and ferroptosis. Front Immunol. 13:10392412022. View Article : Google Scholar : PubMed/NCBI | |
|
Liu J, Kang R and Tang D: Signaling pathways and defense mechanisms of ferroptosis. FEBS J. 289:7038–7050. 2022. View Article : Google Scholar | |
|
Wang B, Wang Y, Zhang J, Hu C, Jiang J, Li Y and Peng Z: ROS-induced lipid peroxidation modulates cell death outcome: mechanisms behind apoptosis, autophagy, and ferroptosis. Arch Toxicol. 97:1439–1451. 2023. 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 | |
|
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 | |
|
Ying W: NAD+/NADH and NADP+/NADPH in cellular functions and cell death: Regulation and biological consequences. Antioxid Redox Signal. 10:179–206. 2008. View Article : Google Scholar | |
|
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 | |
|
Reddy A, Winther S, Tran N, Xiao H, Jakob J, Garrity R, Smith A, Ordonez M, Laznik-Bogoslavski D, Rothstein JD, et al: Monocarboxylate transporters facilitate succinate uptake into brown adipocytes. Nat Metab. 6:567–577. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Ding K, Liu C, Li L, Yang M, Jiang N, Luo S and Sun L: Acyl-CoA synthase ACSL4: An essential target in ferroptosis and fatty acid metabolism. Chin Med J (Engl). 136:2521–2537. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Lan H, Gao Y, Zhao Z, Mei Z and Wang F: Ferroptosis: Redox imbalance and hematological tumorigenesis. Front Oncol. 12:8346812022. View Article : Google Scholar : PubMed/NCBI | |
|
Chen J, Liu Z, Yue Z, Tan Q, Yin H, Wang H, Chen Z, Zhu Y and Zheng J: EP300-mediated H3K18la regulation of METTL3 promotes macrophage ferroptosis and atherosclerosis through the m6A modification of SLC7A11. Biochim Biophys Acta Gen Subj. 1869:1308382025. 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 | |
|
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 | |
|
Zhu J, Xiong Y, Zhang Y, Liang H, Cheng K, Lu Y, Cai G, Wu Y, Fan Y, Chen X, et al: Simvastatin overcomes the pPCK1-pLDHA-SPRINGlac axis-mediated ferroptosis and chemo-immunotherapy resistance in AKT-hyperactivated intrahepatic cholangiocarcinoma. Cancer Commun (Lond). 45:1038–1071. 2025. View Article : Google Scholar : PubMed/NCBI | |
|
Chen Y, Yan Q, Ruan S, Cui J, Li Z, Zhang Z, Yang J, Fang J, Liu S, Huang S, et al: GCLM lactylation mediated by ACAT2 promotes ferroptosis resistance in KRASG12D-mutant cancer. Cell Rep. 44:1157742025. View Article : Google Scholar | |
|
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 | |
|
Yang Z, Su W, Zhang Q, Niu L, Feng B, Zhang Y, Huang F, He J, Zhou Q, Zhou X, et al: Lactylation of HDAC1 confers resistance to ferroptosis in colorectal cancer. Adv Sci (Weinh). 12:e24088452025. View Article : Google Scholar : PubMed/NCBI | |
|
Liu J, Li Y, Ma R, Chen Y, Wang J, Zhang L, Wang B, Zhang Z, Huang L, Zhang H, et al: Cold atmospheric plasma drives USP49/HDAC3 axis mediated ferroptosis as a novel therapeutic strategy in endometrial cancer via reinforcing lactylation dependent p53 expression. J Transl Med. 23:4422025. View Article : Google Scholar : PubMed/NCBI | |
|
Yang G, Liu X, Li Y, Li L, Xiang J, Liang Z, Jiang M and Yang S: TRIM65 as a key regulator of ferroptosis and glycolysis in lactate-driven renal tubular injury and diabetic kidney disease. Cell Rep. 44:1160912025. View Article : Google Scholar : PubMed/NCBI | |
|
Liu Z, Zhou Y, Li M, Xiao Z, Zhao Z and Li Y: Dimeric PKM2 induces ferroptosis from intestinal ischemia/reperfusion in mice by histone H4 lysine 12 lactylation-mediated HMGB1 transcription activation through the lactic acid/p300 axis. Biochim Biophys Acta Mol Basis Dis. 1871:1679982025. 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 | |
|
Lai WKM and Pugh BF: Understanding nucleosome dynamics and their links to gene expression and DNA replication. Nat Rev Mol Cell Biol. 18:548–562. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Millán-Zambrano G, Burton A, Bannister AJ and Schneider R: Histone post-translational modifications-cause and consequence of genome function. Nat Rev Genet. 23:563–580. 2022. 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 | |
|
Liang Z, Liu J, Gou Y, Wang H, Li Z, Cao Y, Zhang H, Bai R and Zhang Z: Elevated histone lactylation mediates ferroptosis resistance in endometriosis through the METTL3-Regulated HIF1A/HMOX1 signaling pathway. Adv Sci (Weinh). 12:e082202025. View Article : Google Scholar : PubMed/NCBI | |
|
Wu S, Liu M, Wang X and Wang S: The histone lactylation of AIM2 influences the suppression of ferroptosis by ACSL4 through STAT5B and promotes the progression of lung cancer. FASEB J. 39:e703082025. View Article : Google Scholar : PubMed/NCBI | |
|
Zhu J, Wu X, Mu M, Zhang Q and Zhao X: TEC-mediated tRF-31R9J regulates histone lactylation and acetylation by HDAC1 to suppress hepatocyte ferroptosis and improve non-alcoholic steatohepatitis. Clin Epigenetics. 17:92025. 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 | |
|
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 | |
|
Liu J, Tang D and Kang R: Targeting GPX4 in ferroptosis and cancer: chemical strategies and challenges. Trends Pharmacol Sci. 45:666–670. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Wang Y, Yue Q, Song X, Du W and Liu R: Hypoxia/reoxygenation-induced glycolysis mediates myocardial ischemia-reperfusion injury through promoting the lactylation of GPX4. J Cardiovasc Transl Res. 18:762–774. 2025. View Article : Google Scholar : PubMed/NCBI | |
|
Di Sanzo M, Quaresima B, Biamonte F, Palmieri C and Faniello MC: FTH1 pseudogenes in cancer and cell metabolism. Cells. 9:25542020. View Article : Google Scholar : PubMed/NCBI | |
|
Muhoberac BB and Vidal R: Iron, ferritin, Hereditary ferritinopathy, and neurodegeneration. Front Neurosci. 13:11952019. 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 | |
|
Tian B, Li X, Li W, Shi Z, He X, Wang S, Zhu X, Shi N, Li Y, Wan P and Zhu C: CRYAB suppresses ferroptosis and promotes osteogenic differentiation of human bone marrow stem cells via binding and stabilizing FTH1. Aging (Albany NY). 16:8965–8979. 2024. 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 | |
|
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 | |
|
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 | |
|
Qu S, Feng B, Xing M, Qiu Y, Ma L, Yang Z, Ji Y, Huang F, Wang Y, Zhou J, et al: PRMT5 K240lac confers ferroptosis resistance via ALKBH5/SLC7A11 axis in colorectal cancer. Oncogene. 44:2814–2830. 2025. 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 | |
|
Deng L, He S, Guo N, Tian W, Zhang W and Luo L: Molecular mechanisms of ferroptosis and relevance to inflammation. Inflamm Res. 72:281–299. 2023. View Article : Google Scholar | |
|
Chen Y, Fang ZM, Yi X, Wei X and Jiang DS: The interaction between ferroptosis and inflammatory signaling pathways. Cell Death Dis. 14:2052023. View Article : Google Scholar : PubMed/NCBI | |
|
Fang Y, Li Z, Yang L, Li W, Wang Y, Kong Z, Miao J, Chen Y, Bian Y and Zeng L: Emerging roles of lactate in acute and chronic inflammation. Cell Commun Signal. 22:2762024. View Article : Google Scholar : PubMed/NCBI | |
|
Qiao X, Yin J, Zheng Z, Li L and Feng X: Endothelial cell dynamics in sepsis-induced acute lung injury and acute respiratory distress syndrome: pathogenesis and therapeutic implications. Cell Commun Signal. 22:2412024. View Article : Google Scholar : PubMed/NCBI | |
|
Li W, Li D, Chen Y, Abudou H, Wang H, Cai J, Wang Y, Liu Z, Liu Y and Fan H: Classic signaling pathways in alveolar injury and repair involved in sepsis-induced ALI/ARDS: New research progress and prospect. Dis Markers. 2022:63623442022.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 | |
|
Han X, Bao J, Ni J, Li B, Song P, Wan R, Wang X, Hu G and Chen C: Qing Xia Jie Yi Formula granules alleviated acute pancreatitis through inhibition of M1 macrophage polarization by suppressing glycolysis. J Ethnopharmacol. 325:1177502024. View Article : Google Scholar : PubMed/NCBI | |
|
Wang Y, Wu S, Li Q, Sun H and Wang H: Pharmacological inhibition of ferroptosis as a therapeutic target for neurodegenerative diseases and strokes. Adv Sci (Weinh). 10:e23003252023. View Article : Google Scholar : PubMed/NCBI | |
|
Sheng W, Liao S, Wang D, Liu P and Zeng H: The role of ferroptosis in osteoarthritis: Progress and prospects. Biochem Biophys Res Commun. 733:1506832024. View Article : Google Scholar : PubMed/NCBI | |
|
Wang J, Chen T and Gao F: Mechanism and application prospect of ferroptosis inhibitors in improving osteoporosis. Front Endocrinol (Lausanne). 15:14926102024. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang Y, Zhao CY, Zhou Z, Li CC and Wang Q: The effect of lactate dehydrogenase B and its mediated histone lactylation on chondrocyte ferroptosis during osteoarthritis. J Orthop Surg. 20:4932025. View Article : Google Scholar | |
|
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 | |
|
Li X, Ma N, Xu J, Zhang Y, Yang P, Su X, Xing Y, An N, Yang F, Zhang G, et al: Targeting ferroptosis: pathological mechanism and treatment of ischemia-reperfusion injury. Oxid Med Cell Longev. 2021:15879222021. View Article : Google Scholar : PubMed/NCBI | |
|
Dong W, Huang SX, Qin ML and Pan Z: Mitochondrial alanyl-tRNA synthetase 2 mediates histone lactylation to promote ferroptosis in intestinal ischemia-reperfusion injury. World J Gastrointest Surg. 17:1067772025. View Article : Google Scholar : PubMed/NCBI | |
|
Sun T, Zhang JN, Lan T, Shi L, Hu L, Yan L, Wei C, Hei L, Wu W, Luo Z, et al: H3K14 lactylation exacerbates neuronal ferroptosis by inhibiting calcium efflux following intracerebral hemorrhagic stroke. Cell Death Dis. 16:5532025. View Article : Google Scholar : PubMed/NCBI | |
|
Apostolova P and Pearce EL: Lactic acid and lactate: Revisiting the physiological roles in the tumor microenvironment. Trends Immunol. 43:969–977. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Sun Q, Wu J, Zhu G, Li T, Zhu X, Ni B, Xu B, Ma X and Li J: Lactate-related metabolic reprogramming and immune regulation in colorectal cancer. Front Endocrinol. 13:10899182023. View Article : Google Scholar | |
|
Zhang W, Xia M, Li J, Liu G, Sun Y, Chen X and Zhong J: Warburg effect and lactylation in cancer: Mechanisms for chemoresistance. Mol Med. 31:1462025. View Article : Google Scholar : PubMed/NCBI | |
|
Liao M, Yao D, Wu L, Luo C, Wang Z, Zhang J and Liu B: Targeting the Warburg effect: A revisited perspective from molecular mechanisms to traditional and innovative therapeutic strategies in cancer. Acta Pharm Sin B. 14:953–1008. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Boedtkjer E and Pedersen SF: The acidic tumor microenvironment as a driver of cancer. Annu Rev Physiol. 82:103–126. 2020. View Article : Google Scholar | |
|
Wang T, Ye Z, Li Z, Jing DS, Fan GX, Liu MQ, Zhuo QF, Ji SR, Yu XJ, Xu XW and Qin Y: Lactate-induced protein lactylation: A bridge between epigenetics and metabolic reprogramming in cancer. Cell Prolif. 56:e134782023. View Article : Google Scholar : PubMed/NCBI | |
|
Sun Y, Wang H, Cui Z, Yu T, Song Y, Gao H, Tang R, Wang X, Li B, Li W and Wang Z: Lactylation in cancer progression and drug resistance. Drug Resist Updat. 81:1012482025. View Article : Google Scholar : PubMed/NCBI | |
|
Yang T, Zhang S, Nie K, Cheng C, Peng X, Huo J and Zhang Y: ZNF207-driven PRDX1 lactylation and NRF2 activation in regorafenib resistance and ferroptosis evasion. Drug Resist Updat. 82:1012742025. 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 | |
|
Tong Q, Huang C, Tong Q and Zhang Z: Histone lactylation: A new frontier in laryngeal cancer research (Review). Oncol Lett. 30:4212025. View Article : Google Scholar : PubMed/NCBI | |
|
Bao C, Ma Q, Ying X, Wang F, Hou Y, Wang D, Zhu L, Huang J and He C: Histone lactylation in macrophage biology and disease: From plasticity regulation to therapeutic implications. EBioMedicine. 111:1055022025. View Article : Google Scholar : | |
|
Bao Q, Wan N, He Z, Cao J, Yuan W, Hao H and Ye H: Subcellular proteomic mapping of lysine lactylation. J Am Soc Mass Spectrom. 35:3221–3232. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Zhao W, Zhang J, Chen K, Yuan J, Zhai L and Tan M: Mass spectrometry-based characterization of histone post-translational modification. Curr Opin Chem Biol. 88:1026222025. View Article : Google Scholar : PubMed/NCBI | |
|
Sun Y, Chen Y and Peng T: A bioorthogonal chemical reporter for the detection and identification of protein lactylation. Chem Sci. 13:6019–6027. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang X, Liu Y, Rekowski MJ and Wang N: Lactylation of tau in human Alzheimer's disease brains. Alzheimers Dement. 21:e144812025. View Article : Google Scholar : PubMed/NCBI | |
|
Chu X, Di C, Chang P, Li L, Feng Z, Xiao S, Yan X, Xu X, Li H, Qi R, et al: Lactylated Histone H3K18 as a potential biomarker for the diagnosis and predicting the severity of septic shock. Front Immunol. 12:7866662022. View Article : Google Scholar : PubMed/NCBI | |
|
Nuñez R, Sidlowski PFW, Steen EA, Wynia-Smith SL, Sprague DJ, Keyes RF and Smith BC: The TRIM33 bromodomain recognizes histone lysine lactylation. ACS Chem Biol. 19:2418–2428. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Wang Y, Fan J, Meng X, Shu Q, Wu Y, Chu GC, Ji R, Ye Y, Wu X, Shi J, et al: Development of nucleus-targeted histone-tail-based photoaffinity probes to profile the epigenetic interactome in native cells. Nat Commun. 16:4152025. View Article : Google Scholar : PubMed/NCBI | |
|
Crabb S, Plummer R, Greystoke A, Carter L, Pacey S, Walter H, Coyle VM, Knurowski T, Clegg K, Ashby F, et al: 560TiP a phase I/IIa study to evaluate the safety and efficacy of CCS1477, a first in clinic inhibitor of p300/CBP, as monotherapy in patients with selected molecular alterations. Ann Oncol. 32:S6172021. View Article : Google Scholar | |
|
Nicosia L, Spencer GJ, Brooks N, Ciceri F, Wiseman DH, Pegg N, West W, Knurowski T, Frese K, Clegg K, et al: Therapeutic targeting of EP300/CBP by bromodomain inhibition in hematologic malignancies. Cancer Cell. 41:2136–2153.e13. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Armstrong AJ, Gordon MS, Reimers MA, Sedkov A, Lipford K, Snavely-Merhaut J, Kumar S, Guichard SM and Shore N: The courage study: A first-in-human phase 1 study of the CBP/p300 inhibitor FT-7051 in men with metastatic castration-resistant prostate cancer. J Clin Oncol. 39:TPS50852021. View Article : Google Scholar | |
|
Marks PA and Breslow R: Dimethyl sulfoxide to vorinostat: Development of this histone deacetylase inhibitor as an anti-cancer drug. Nat Biotechnol. 25:84–90. 2007. View Article : Google Scholar : PubMed/NCBI | |
|
Campbell P and Thomas CM: Belinostat for the treatment of relapsed or refractory peripheral T-cell lymphoma. J Oncol Pharm Pract. 23:143–147. 2017. View Article : Google Scholar | |
|
Tzogani K, van Hennik P, Walsh I, De Graeff P, Folin A, Sjöberg J, Salmonson T, Bergh J, Laane E, Ludwig H, et al: EMA review of panobinostat (farydak) for the treatment of adult patients with relapsed and/or refractory multiple myeloma. Oncologist. 23:631–636. 2018. View Article : Google Scholar : | |
|
Lu X, Ning Z, Li Z, Cao H and Wang X: Development of chidamide for peripheral T-cell lymphoma, the first orphan drug approved in China. Intractable Rare Dis Res. 5:185–191. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Yardley DA, Ismail-Khan RR, Melichar B, Lichinitser M, Munster PN, Klein PM, Cruickshank S, Miller KD, Lee MJ and Trepel JB: Randomized phase II, double-blind, placebo-controlled study of exemestane with or without entinostat in postmenopausal women with locally recurrent or metastatic estrogen receptor-positive breast cancer progressing on treatment with a nonsteroidal aromatase inhibitor. J Clin Oncol. 31:2128–2135. 2013. View Article : Google Scholar : PubMed/NCBI | |
|
Jiang Z, Li W, Hu X, Zhang Q, Sun T, Cui S, Wang S, Ouyang Q, Yin Y, Geng C, et al: Tucidinostat plus exemestane for postmenopausal patients with advanced, hormone receptor-positive breast cancer (ACE): A randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol. 20:806–815. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Kitir B, Maolanon AR, Ohm RG, Colaço AR, Fristrup P, Madsen AS and Olsen CA: Chemical editing of macrocyclic natural products and kinetic profiling reveal slow, tight-binding histone deacetylase inhibitors with picomolar affinities. Biochemistry. 56:5134–5146. 2017. View Article : Google Scholar : PubMed/NCBI | |
|
Whitehead L, Dobler MR, Radetich B, Zhu Y, Atadja PW, Claiborne T, Grob JE, McRiner A, Pancost MR, Patnaik A, et al: Human HDAC isoform selectivity achieved via exploitation of the acetate release channel with structurally unique small molecule inhibitors. Bioorg Med Chem. 19:4626–4634. 2011. View Article : Google Scholar : PubMed/NCBI | |
|
Wang Y, Stowe RL, Pinello CE, Tian G, Madoux F, Li D, Zhao LY, Li JL, Wang Y, Wang Y, et al: Identification of histone deacetylase inhibitors with benzoylhydrazide scaffold that selectively inhibit class I histone deacetylases. Chem Biol. 22:273–284. 2015. View Article : Google Scholar : PubMed/NCBI | |
|
Rosás-Umbert M, Ruiz-Riol M, Fernández MA, Marszalek M, Coll P, Manzardo C, Cedeño S, Miró JM, Clotet B, Hanke T, et al: In vivo effects of romidepsin on T-cell activation, apoptosis and function in the BCN02 HIV-1 kick&kill clinical trial. Front Immunol. 11:4182020. View Article : Google Scholar | |
|
Sandoná M, Consalvi S, Tucciarone L, Puri PL and Saccone V: HDAC inhibitors for muscular dystrophies: Progress and prospects. Expert Opin Orphan Drugs. 4:125–127. 2016. View Article : Google Scholar | |
|
Wang X, Shen X, Xu Y, Xu S, Xia F, Zhu B, Liu Y, Wang W, Wu H and Wang F: The etiological changes of acetylation in peripheral nerve injury-induced neuropathic hypersensitivity. Mol Pain. 14:17448069187984082018. View Article : Google Scholar : PubMed/NCBI | |
|
Bhatt AN, Shenoy S, Munjal S, Chinnadurai V, Agarwal A, Vinoth Kumar A, Shanavas A, Kanwar R and Chandna S: 2-deoxy-d-glucose as an adjunct to standard of care in the medical management of COVID-19: A proof-of-concept and dose-ranging randomised phase II clinical trial. BMC Infect Dis. 22:6692022. View Article : Google Scholar | |
|
Sharma D, Singh M and Rani R: Role of LDH in tumor glycolysis: Regulation of LDHA by small molecules for cancer therapeutics. Semin Cancer Biol. 87:184–195. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Halford SER, Walter H, McKay P, Townsend W, Linton K, Heinzmann K, Dragoni I, Brotherton L, Veal G, Siskos A, et al: Phase I expansion study of the first-in-class monocarboxylate transporter 1 (MCT1) inhibitor AZD3965 in patients with diffuse large B-cell lymphoma (DLBCL) and burkitt lymphoma (BL). J Clin Oncol. 39:31152021. View Article : Google Scholar | |
|
Halford SER, Jones P, Wedge S, Hirschberg S, Katugampola S, Veal G, Payne G, Bacon C, Potter S, Griffin M, et al: A first-in-human first-in-class (FIC) trial of the monocarboxylate transporter 1 (MCT1) inhibitor AZD3965 in patients with advanced solid tumours. J Clin Oncol. 35:25162017. View Article : Google Scholar | |
|
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 | |
|
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 | |
|
Zhang W, Shan G, Bi G, Hu Z, Yi Y, Zeng D, Lin Z and Zhan C: Lactylation and regulated cell death. Biochim Biophys Acta Mol Cell Res. 1872:1199272025. View Article : Google Scholar : PubMed/NCBI | |
|
Gong T, Wang QD, Loughran PA, Li YH, Scott MJ, Billiar TR, Liu YT and Fan J: Mechanism of lactic acidemia-promoted pulmonary endothelial cells death in sepsis: role for CIRP-ZBP1-PANoptosis pathway. Mil Med Res. 11:712024.PubMed/NCBI | |
|
Xu H, Li L, Wang S, Wang Z, Qu L, Wang C and Xu K: Royal jelly acid suppresses hepatocellular carcinoma tumorigenicity by inhibiting H3 histone lactylation at H3K9la and H3K14la sites. Phytomedicine Int J Phytother Phytopharm. 118:1549402023. | |
|
Li W, Zhou C, Yu L, Hou Z, Liu H, Kong L, Xu Y, He J, Lan J, Ou Q, et al: Tumor-derived lactate promotes resistance to bevacizumab treatment by facilitating autophagy enhancer protein RUBCNL expression through histone H3 lysine 18 lactylation (H3K18la) in colorectal cancer. Autophagy. 20:114–130. 2024. View Article : Google Scholar : | |
|
Sun Y, Chen Y, Xu Y, Zhang Y, Lu M, Li M, Zhou L and Peng T: Genetic encoding of ε-N-L-lactyllysine for detecting delactylase activity in living cells. Chem Commun (Camb). 58:8544–8547. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Jiang S, Li H, Zhang L, Mu W, Zhang Y, Chen T, Wu J, Tang H, Zheng S, Liu Y, et al: Generic diagramming platform (GDP): A comprehensive database of high-quality biomedical graphics. Nucleic Acids Res. 53(D1): D1670–D1676. 2025. View Article : Google Scholar : |