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.
Lactate, once regarded as a metabolic waste product of glycolysis, is now considered a multifunctional molecule that promotes cancer progression. Notably, the Warburg effect is a hallmark of cancer metabolism; even under oxygen-rich conditions, tumor cells still preferentially metabolize glucose through glycolysis and tend to ‘ferment’ glucose into lactic acid, leading to an accumulation of lactic acid within the tumor microenvironment (TME) (1). In addition to serving a role in central carbon metabolism, lactate regulates tumor immunity, antiviral responses and endoplasmic reticulum (ER)-mitochondrial magnesium ion dynamics (2), and contributes to pathological processes such as cancer progression (3). Lactate, as a signaling molecule, regulates immune evasion, angiogenesis and therapeutic resistance (4,5).
The discovery of lactylation, a lactate-derived post-translational modification (PTM) of lysine residues, has revealed a direct mechanistic link between glycolytic flux and cellular regulation. Lactate, as the substrate for this modification, catalyzes the addition of lactate groups to both histone and non-histone proteins (6). This process links metabolic disorders to epigenetic reprogramming, alterations in chromatin structure, transcriptional programs and protein-protein interactions, thereby promoting tumorigenesis (7). For example, M1-polarized macrophages rely on aerobic glycolysis to drive histone lactylation (8), whereas B-cell adapter for PI3K facilitates their transition to a reparative macrophage phenotype via lactylation (9). Lactate enhances the activity of pyruvate kinase M2 (PKM2) by lactylating its K62 residues, inhibiting the Warburg effect and promoting the transformation of inflammatory macrophages (iNOS+ CD68+ cells) into reparative macrophages (ARG1+ CD68+ positive cells) (10). In non-small cell lung cancer (NSCLC), hypoxia-induced long noncoding RNA-AC020978 amplifies glycolysis by promoting PKM2 nuclear translocation and hypoxia-inducible factor-1α (HIF-1α) activation (11). Furthermore, lactate directly inhibits CD8+ T cells, natural killer T (NKT) cells, dendritic cells and macrophages, while enhancing regulatory T (Treg) cell stability and function, promoting immunosuppression (12).
Researchers have also demonstrated the notable role of lactylation in various types of cancer via numerous mechanisms (Table I; Fig. 1). For example, pan-cancer analyses have revealed upregulated lactylation-related genes (such as CREBBP and EP300) and their association with kidney renal papillary cell carcinoma and hepatocellular carcinoma (HCC) (13). Furthermore, hypoxia-glycolysis-lactylation-related genes predict gastric cancer (GC) progression (14), whereas lactate-driven lactylation in colorectal cancer (CRC) is associated with proliferation, metastasis and immune evasion (15). Lactylation serves a role in tumorigenesis, including in tumor cell proliferation, invasion, metastasis, DNA damage repair and immune cell killing, and also promotes therapeutic resistance through autophagy activation, drug efflux pumps and DNA damage repair (16).
The present review discusses the latest progress in lactic acid biology, highlighting its dual role as a driving factor in the pathogenesis of cancer and in therapeutic sensitivity. The review also explores the impact of lactylation on metabolic reprogramming, immune evasion and therapeutic resistance, and discusses new strategies for precision oncology using this pathway.
Lactylation critically regulates glycolytic flux in cancer. Digestive system cancers account for >25% of global cancer cases (17), with lactylation emerging as a key regulator of immunosuppression and metabolic reprogramming in the TME. Yang et al (18) identified 9,256 non-histone lysine lactylation (Kla) sites in hepatitis B virus-associated HCC, including adenylate kinase 2 lactylation at K28, which may drive metastasis by disrupting metabolic regulation.
Abundant immune cell infiltration, particularly macrophages, and increased genetic instability are traits associated with high lactylation scores. Lactylation score can be used to predict the malignant progression of GC and immune escape (19). Hypoxia-induced mitochondrial alanyl-tRNA synthetase AARS2 drives lactylation of pyruvate dehydrogenase (PDH) complex components (PDHA1-K336 and CPT2-K457/8), suppressing oxidative phosphorylation (OXPHOS) by limiting pyruvate and acetyl-CoA influx (20). AARS1, upregulated in GC, catalyzes lactate-AMP formation, promoting lactylation of p53 at lysine 120/139. This disrupts the DNA binding and transcriptional activity of p53, coupling metabolic rewiring to proteomic changes that fuel tumorigenesis (21). AARS1 also activates the YAP-TEAD complex via nuclear lactylation, forming a Hippo pathway feedback loop that accelerates GC progression and is associated with poor prognosis (22). In addition, G-protein-coupled receptor 37 (GPR37) activates the Hippo pathway, upregulating lactate dehydrogenase A (LDHA) to enhance glycolysis and H3K18 lactylation (H3K18la). This promotes CXCL1/CXCL5 secretion, remodeling the TME to favor metastasis. By contrast, GPR37 depletion suppresses the progression of liver metastasis in CRC (23). Lactate also stabilizes β-catenin via hypoxia-induced lactylation, activating Wnt signaling and CRC proliferation (24).
Nucleolin lactylation promotes MAPK-activated death domain protein translation and ERK activation, which is associated with poor intrahepatic cholangiocarcinoma (iCCA) prognosis (25). Non-histone lactylation sites in oral squamous cell carcinoma (OSCC) cells, such as hnRNPA1, SF3A1, hnRNPU and SLU7, are linked to glycolytic dysregulation and tumorigenesis, highlighting the role of Kla in pathogenesis (26). In lung adenocarcinoma (LUAD), basic leucine zipper and W2 domains 2 (BZW2) enhances glycolysis and isocitrate dehydrogenase 3 (IDH3G) lactylation. Combined inhibition of BZW2 and glycolysis [such as 2-deoxy-D-glucose (2-DG)] suppresses tumor growth (27).
CircXRN2 is downregulated in bladder cancer (BCa) tissues and suppresses tumor growth by binding to the speckle-type POZ protein (SPOP) degron. This interaction blocks SPOP-mediated degradation of large tumor suppressor kinase 1, activating the Hippo signaling pathway to suppress tumor progression driven by H3K18la (28). Upregulation of zinc finger E-box binding homeobox 1 (ZEB1) in BCa is counteracted by phosphofructokinase-1 (PFK-1), and PFK-1 inhibits glycolysis through the lactylation of ZEB1 and suppresses its malignant effects, including cell proliferation, migration and invasion (29). During neuroendocrine prostate cancer (NEPC) development, ZEB1 transcriptionally regulates the expression of several key glycolytic enzymes, thereby predisposing tumor cells to utilize glycolysis for energy metabolism. Lactate accumulation enhances histone lactylation, increasing chromatin accessibility and cellular plasticity (including neural gene expression), thereby promoting NEPC progression (30).
In BCa, the enhanced aerobic glycolysis rate supports c-Myc expression through histone lactylation at the promoter level. c-Myc further upregulates serine/arginine splicing factor 10 (SRSF10) through transcription to drive the selective splicing of MDM4 and Bcl-x in BCa cells. Restricting the activity of key glycolytic enzymes may affect the c-Myc/SRSF10 axis to reduce the proliferation of BCa cells (31). Potassium channel subfamily K member 1, which is upregulated in BCa, activates LDHA to enhance glycolysis and H3K18la. LDHA upregulation creates a feed-forward loop, reducing tumor cell adhesion and promoting invasion/metastasis (32). Deficiency in the Numb/Parkin pathway in prostate cancer (PCa) or LUAD induces metabolic reprogramming, resulting in a significant increase in the production of lactate acid, which subsequently leads to the upregulation of histone lactylation and transcription of neuroendocrine-associated genes (33). Aerobic glycolysis in anaplastic thyroid cancer (ATC) increases overall protein lactylation by increasing cellular lactate availability, and H4K12la activates the expression of multiple genes necessary for ATC proliferation. Furthermore, the oncogene BRAFV600E enhances glycolysis to reprogram the cellular lactylation landscape, resulting in H4K12la-driven gene transcription and cell cycle dysregulation. Notably, treatment with a BRAFV600E inhibitor, combined with blocking cell emulsification, can inhibit ATC progression in an 8505c xenograft tumor mouse model (34).
Autophagy and glycolysis are highly conserved biological processes involving physiological and pathological cellular activities. A number of core autophagy proteins undergo lactylation during cancer (35). Under nutrient deprivation, lactate-mediated lactylation of autophagy proteins (such as VPS34, ULK1) enhances autophagic flux, promoting tumor cell survival (36). In lung cancer and GC, lactate mediates the lactylation of PIK3C3/VPS34 at lysine 356 and lysine 781 through acyltransferase KAT5/TIP60. The lactylation of PIK3C3/VPS34 enhances the binding of PIK3C3/VPS34 to beclin 1, ATG14 and UVRAG, increases the lipid kinase activity of PIK3C3/VPS34, promotes macroautophagy/autophagy and advances the endolysome degradation pathway (37).
In the TME, tumor cells export lactate via monocarboxylate transporter 4 (MCT4), while oxidative tumor cells import it through MCT1 to fuel OXPHOS (38). Although glycolysis dominates in most types of cancer, OXPHOS remains active in malignancies such as leukemia and lymphoma, enabling metabolic flexibility (39). Glucose fuels glycolysis, which feeds into the TCA cycle and OXPHOS in oxidative cells, or sustains lactate fermentation in hypoxic regions (40). In SW480 colon cancer cells, lactylation targets glycolytic enzymes (such as PFKP-K688 and ALDOA-K147), potentially establishing negative feedback loops to modulate glycolysis (41). In pancreatic adenocarcinoma cells, under glucose deprivation, lactate enhances glutaminase 1-mediated glutamine metabolism and NMNAT1 lactylation, suppressing p38 MAPK and promoting survival (42). Solute carrier family 16 member 1 lactylation is critical for pancreatic ductal adenocarcinoma (PDAC) progression, and targeting this axis may inhibit tumor growth (43). In NSCLC, metabolic dysregulation drives lactate accumulation, which replenishes the TCA cycle while suppressing glucose uptake and glycolysis. Lactate downregulates glycolytic enzymes [such as hexokinase (HK)-1 and PKM] and upregulates TCA enzymes (including SDHA and IDH3G) via histone lactylation at their promoters, enhancing NSCLC proliferation and migration (44).
Lactate regulates Mg2+ flux between the ER and mitochondria, impacting glycolysis and mitochondrial metabolism in lung cancer (45). Exogenous lactate stabilizes insulin-like growth factor 1 receptor (IGF1R) via lactylation, enhancing extracellular acidification and metabolic reprogramming. LDHA deficiency disrupts this axis, suggesting the role of lactylation in IGF1R-mediated proliferation and invasion (46,47). In glioblastoma (GBM), lactate drives tumor progression via metabolic reprogramming (48). However, in uveal melanoma (UM) cells, the targeting of lactate and hydroxy-carboxylic acid receptor 1 (HCAR1) serves the opposite role. Lactate (20 mM) has been shown to cause a marked reduction in tumor cell proliferation and migration, and can shift cell metabolism toward OXPHOS. Clinically, elevated levels of the lactate transporters MCT4 and HCAR1 are associated with spindle-cell UM subtypes. In addition, lactate reshapes metabolic reprogramming and induces the quiescence phenotype through upregulating HCAR1 and MCT4 in UM cells, suggesting lactate metabolism may serve as a prognostic marker of UM progression (49). Clear cell renal cell carcinoma (ccRCC) is a challenging subtype of kidney cancer that often complicates patient prognosis due to factors such as postoperative recurrence or late diagnosis. Yang et al (50) further characterized the interplay of lactylation with m6A modifications, pinpointing 3-hydroxyisobutyryl-CoA hydrolase (HIBCH) as a key regulator. HIBCH, downregulated in ccRCC, may connect histone lactylation to mitochondrial energy metabolism and modulate the tumor immune microenvironment, influencing therapeutic response.
The hypoxic microenvironment resulting from reduced local blood flow in pancreatic cancer (PC) triggers a shift in glycolytic-dependent energy metabolism by stabilizing HIF-1α (51). Elevated H3K18la in PDAC activates TTK and BUB1B transcription, upregulating P300 to enhance glycolysis. TTK phosphorylates LDHA at Y239, amplifying lactate and H3K18la levels (52). Nuclear and spindle-associated protein 1 (NUSAP1) binds c-Myc and HIF-1α to upregulate LDHA. Lactate stabilizes NUSAP1 via Kla, forming a feed-forward loop (NUSAP1-LDHA-lactate-NUSAP1) that drives metastasis (53). In esophageal cancer (EC), hypoxia not only enhances the expression of serine hydroxymethyltransferase 2 (SHMT2) protein but also initiates the lactylation of SHMT2 protein and enhances its stability, thus accelerating the malignant progression of EC (54).
Angiogenic mimicry provides tumor cells with a blood supply independent of endothelial cells (55). MAPK 6 pseudogene 4 stabilizes Krüppel-like factor 15, which transcriptionally activates LDHA. LDHA-mediated lactylation of vascular endothelial growth factor receptor 2 (VEGFR2) and VE-cadherin enhances their expression, fostering GBM cell proliferation, migration and angiogenic mimicry. However, the specific lactylation sites in VEGFR2 and VE-cadherin involved in GBM still require further exploration (56). FK506-binding protein 10 (FKBP10) binds LDHA, enhancing LDHA-Y10 phosphorylation to amplify the Warburg effect and histone lactylation. By contrast, HIF-2α, a driver of angiogenesis and redox balance, suppresses FKBP10 transcription. Combining HIF-2α inhibitors (such as PT2385) with FKBP10 targeting enhances antitumor efficacy, particularly in low-FKBP10 ccRCC (57). Lactate import via MCT1 stabilizes HIF-1α under normoxia through HIF-1α lactylation, perpetuating KIAA1199 expression. By contrast, silencing KIAA1199 restores 3A semaphoring (sema)3A and inhibits angiogenesis (58). Lactate also stabilizes HIF-1α, promoting discoidin, CUB, and LCCL domain-containing protein 1 (DCBLD1) transcription. DCBLD1-K172 lactylation inhibits ubiquitination, blocking autophagy-mediated degradation of G6PD to fuel the pentose phosphate pathway (PPP). This activates the PPP, fueling cervical cancer cell migration, invasion and growth (59).
G6PD K45 is lactylated during G6PD-mediated antioxidant stress. In primary human keratinocytes and HPV-negative cervical cancer C33A cell lines ectopically expressing HPV16 E6, G6PD K45A is not lactylated, and the transduction of G6PD K45A has been shown to increase the levels of glutathione and NADPH, and can accordingly decrease the levels of reactive oxygen species. In vivo, 6-aminonicotinamide inhibits the activity of the G6PD enzyme or the re-expression of G6PD K45T, thereby suppressing tumor proliferation (60). The non-metabolizable glucose analog 2-DG and oxamate treatment have been reported to decrease the levels of lactylation to inhibit proliferation and migration, induce apoptosis and arrest the cell cycle of EC cells. In addition, a study on EC xenograft tumor model mice and EC cells confirmed that H3K18la may upregulate the deubiquitinase ubiquitin-specific protease (USP)39, stabilizing phosphoglycerate kinase 1 to activate the PI3K/AKT/HIF-1α axis and to enhance glycolysis (61).
Lactylation orchestrates the progression of cancer through extensive epigenetic reprogramming, modulating both histone and non-histone targets to alter gene expression, RNA processing and cellular phenotypes.
Histone lactylation directly modulates chromatin dynamics. In GC, H3K18la activates vascular cell adhesion molecule 1, promoting GC cell proliferation and migration via AKT/mTOR signaling, and recruits immunosuppressive mesenchymal stem cells through CXCL1 upregulation, thereby promoting immune suppression and tumor metastasis (62). In CRC, tumor-derived lactate inhibits retinoic acid receptor γ in macrophages, elevating H3K18la and IL-6/STAT3 signaling to polarize macrophages toward a pro-tumorigenic state (63). Histone lactylation (such as H4K8la, H4K16la) is suppressed by liver kinase B1, which induces senescence by inhibiting telomerase reverse transcriptase (64). Furthermore, lactate in the TME of LUAD reduces the transcription of solute carrier family 25, member 29 (SLC25A29) in endothelial cells. Specifically, the increase of H3K14la and H3K18la in the SLC25A29 promoter region reduces the transcription of SLC25A29, which affects the proliferation, migration and apoptosis of endothelial cells (65). The NF-κB pathway promotes the Warburg effect, thereby inducing the lactylation of H3 histone and increasing the expression of LINC01127. The enhanced expression of LINC01127 promotes the self-renewal of GBM cells and directly guides POLR2A to the MAP4K4 promoter region to regulate the expression of MAP4K4, thereby activating the JNK pathway and ultimately regulating the self-renewal of GBM cells (66). Inactivation of the von Hippel-Lindau (VHL) tumor suppressor triggers histone lactylation, activating platelet-derived growth factor receptor β (PDGFRβ) transcription; PDGFRβ signaling reciprocally enhances histone lactylation, creating an oncogenic feedback loop (67).
The lactylation and delactylation of non-histone proteins are closely related to the pathogenesis of HCC. Sirtuin (SIRT)3 deficiency in HCC permits the accumulation of cyclin (CCN)E2 lactylation, whereas honokiol-mediated SIRT3 reactivation delactylates CCNE2-K348 to suppress HCC cell proliferation (68). Furthermore, CENPA-K124 lactylation enhances the interaction between CENPA and the transcription factor YY1, driving CCND1 and neuropilin 2 expression to promote HCC progression (69). In CRC, KAT8-catalyzed lactylation of eEF1A2K408 has been shown to result in boosted translation elongation and enhanced protein synthesis, which contribute to tumorigenesis (70). Furthermore, H3K18la has been reported to increase METTL3 expression in tumor-infiltrating myeloid cells. Notably, METTL3 mediates m6A modification on Jak1 mRNA in tumor-infiltrating myeloid cells, and the m6A-YTHDF1 axis enhances JAK1 protein translation efficiency, subsequent phosphorylation of STAT3 and immunosuppression (71). In ocular melanoma, histone lactylation promotes tumorigenesis by increasing the expression of YTHDF2. YTHDF2 recognizes m6A-modified PER1 and TP53 mRNA, promotes their degradation and accelerates the occurrence of ocular melanoma (72). Histone lactylation enhances the expression of AlkB homolog 3 (ALKBH3) by removing the m1A methylation of SP100A. ALKBH3 lactylation facilitates promyelocytic leukemia protein nuclear condensate formation, bridging m1A modification to metabolic reprogramming (73). Elevated NOP2/Sun RNA methyltransferase family member 2 (NSUN2), an m5C methyltransferase, and Y-box binding protein 1 (YBX1), an m5C methylation recognition enzyme, drive m5C modification of ENO1 mRNA in CRC, promoting lactate production. This lactate then induces NSUN2 expression via histone H3K18 lactylation and enhances RNA binding of NSUN2 through its own lactylation (K356), creating a feed-forward loop linking metabolism and epigenetics (74). Furthermore, the chemotherapeutic agent gemcitabine (GEM) increases histone acetylation but reduces histone lactylation, suggesting cross-talk between PTMs in the mechanism of GEM (75). These mechanisms collectively establish lactylation as a master regulator of cancer epigenetics, influencing angiogenesis, therapy resistance and metabolic adaptation.
Lactylation orchestrates immunosuppression within the TME by reprogramming immune cell function and recruitment. In HCC, dysregulation of histone acetyltransferase EP300 and histone deacetylase (HDAC)1-3 alters immune cell infiltration (including B cells) and predicts poor prognosis when HDAC1/2 are upregulated (76). Glucose transporter 3 (GLUT3) levels are significantly increased in GC tissues. By contrast, after knocking down GLUT3, the levels of LDHA, L-lactylation, H3K9, H3K18 and H3K56 are markedly reduced. Notably, in GLUT3-knockdown cell lines, upregulation of LDHA reverses the lactylation and epithelial-mesenchymal transition functional phenotypes, while Kla is enriched in GC tissues and predicts poor survival, underscoring its role as a prognostic biomarker (77,78). Claudin-9 promotes glycolytic metabolism in GC cells by activating the PI3K/AKT/HIF-1α signaling pathway, resulting in increased lactate production. Lactate, as a glycolytic metabolite, can enhance the lactylation of programmed death ligand 1 (PD-L1) and improve its stability. This modified PD-L1 will inhibit the antitumor immune response of CD8+ T cells, thereby enhancing GC cell immune evasion and promoting the progression of GC (79). STAT5-induced lactate accumulation promotes nuclear translocation of E3 binding protein, increases lactylation of the PD-L1 promoter, and subsequently induces PD-L1 transcription, driving immunosuppression in acute myeloid leukemia (80).
The gut microbiome further amplifies immunosuppression through lactylation. In CRC, RIG-I lactylation induced by Escherichia coli inhibits the recruitment of NF-κB to the NLRP3 promoter, suppresses inflammasome activity, impairs CD8+ T-cell function and simultaneously enhances Treg cell-mediated tolerance (81). Similarly, gram-negative bacterial lipopolysaccharide upregulates the expression of LINC00152 and promotes the invasion of CRC cells by introducing histone lactylation on its promoter, reducing the binding efficiency of inhibitory factor YY1 to it (82).
In lung squamous cell carcinoma, solute carrier family 2 member 1 (SLC2A1) upregulation is associated with elevated protein lactylation and SPP1+ macrophage abundance, driving therapy resistance (83). In advanced GBM, monocyte-derived macrophages dominate the TME and secrete IL-10 via PERK-driven histone lactylation, inhibiting T-cell activity. PERK deletion abolishes lactylation, restores T-cell function and delays tumor growth (84). Furthermore, single-cell RNA sequencing has confirmed lactylation-related gene expression across glioma-infiltrating immune cells (monocytes, macrophages, CD8+ T cells), and lactylation scores have been reported to be associated with lipogenesis, DNA repair and mTORC1 signaling (85). HK-3 indirectly affects neuroblastoma progression by recruiting and polarizing M2-like macrophages through the PI3K/AKT/CXCL14 axis. After knocking down the expression of HK3, the levels of histone lysine lactylation in neuroblastoma cell lines SK-N-SH and SK-N-BE (2) are significantly reduced, and the lactate concentration in the supernatant of SK-N-SH/SK-N-BE (2) cell cultures is significantly reduced. These findings indicate that in neuroblastoma HK3 affects lactate secretion in the microenvironment and regulates histone lactylation (86). In the TME of malignant pleural effusion (MPE), the antitumor activity of CD8+ T cells and NKT-like cells is inhibited, and the function of Treg cells is enhanced. The glycolytic pathway and pyruvate metabolism are highly activated in a distinct subpopulation of NKT-like cells expressing FOXP3 in MPE. Notably, a small molecule inhibitor of MCT1, 7ACC2, has been shown to reduce FOXP3 expression and histone lactylation levels in NKT-like cells (87).
In addition to regulating metabolic reprogramming, and epigenetic and immune pathways, lactylation drives multi-faceted treatment resistance in cancer. Previous proteomic profiling identified 1,438 lactylation sites across 772 proteins in HCC tissues, implicating lactylation in amino acid metabolism, ribosomal function and fatty acid metabolism (88). Lactylation of USP14 and ATP-binding cassette (ABC) subfamily F member 1 is associated with drug sensitivity. Notably, patients with high-risk HCC respond better to sorafenib, whereas low-risk patients exhibit elevated Tumor Immune Dysfunction and Exclusion scores and benefit from immunotherapy (89). Lactylation also activates oncogenic pathways (Wnt, MAPK, mTOR, NOTCH) via nuclear receptor 6A1, oxysterol-binding protein 2 and UNC119B, linking Kla to immune evasion and therapy resistance (90). Apicidin treatment reduces lactylation, suppressing HCC migration and proliferation, although clinical validation and mechanistic crosstalk with other PTMs require further study (91).
In GC, copper stress lactylates METTL16 at K229, enhancing copper ionophore (elesclomol) efficacy. Combining elesclomol with the SIRT2 inhibitor AGK2 amplifies copper toxicity, suggesting a therapeutic strategy to overcome chemoresistance (92,93). In CRC, H3K18la upregulation in bevacizumab-resistant tumors, which upregulates autophagy protein RUBCNL/Pacer through BECN1 interaction, enhances autophagosome maturation and survival (94). In diapause-like CRC, SMC4 and phosphoglycerate mutase 1 loss disrupts F-actin assembly and upregulates ABC transporters via histone lactylation, reducing chemotherapy sensitivity (95). Aldolase B-activated PDH kinase 1 drives lactylation of circulating carcinoembryonic antigen, activating CEACAM6 to promote proliferation and chemoresistance in CRC (96).
Hypoxia in the TME amplifies lactate production, global lactylation and immunosuppression (97,98). Hypoxia upregulates SOX9 lactylation, enhancing stemness and invasion in NSCLC (99). AKR1B10 promotes LDHA-driven lactate accumulation, stimulating H4K12la to activate CCNB1 and accelerate DNA replication, conferring pemetrexed resistance in lung cancer brain metastasis (100). This aligns with findings that tumor resistance stems from heterogeneity and immunosuppressive TME interactions (101).
In BCa, H3K18la drives the key transcription factors YBX1 and YY1 associated with cisplatin resistance, thereby promoting cisplatin resistance (102). Sema3A suppresses VEGFA-induced colony formation, proliferation and PD-L1 expression in PCa. Sema3A, along with sema3B/C/E, predicts biochemical recurrence in low- to intermediate-risk PCa post-prostatectomy (103). Prognostic models combining the lactylation-related genes ALDOA, DDX39A, H2AX, KIF2C and RACGAP1 has been shown to predict disease-free survival and treatment response in PCa. These genes are highly expressed in castration-resistant tumors, underscoring the clinical relevance of lactylation (104). In PTEN-deficient metastatic castration-resistant PCa, a PI3K inhibitor (PI3Ki) has been reported to reduce histone lactylation within tumor-associated macrophages (TAMs), resulting in their anticancer phagocytic activation, which is augmented by ADT/aPD-1 treatment and abrogated by feedback activation of the Wnt/β-catenin pathway. Furthermore, co-targeting Wnt/β-catenin signaling with LGK-974 in combination with PI3Ki, has previously demonstrated durable tumor control in 100% of mice via H3K18lac suppression and complete TAM activation (105,106). Gambogic acid disrupts this cycle by recruiting SIRT1 to delactylate chaperone protein CNPY3, inducing lysosomal rupture and pyroptotic cell death (107). Lactylation-related genes influence BCa growth, immune microenvironment remodeling and therapy resistance by modulating lactate transport within the TME (108). In addition, H4K12la contributes to chemoresistance and adverse immune responses in breast cancer (109).
Lactylation modulation represents an emerging frontier for targeted cancer therapy, with implications for treatment response prediction and combination strategies. Yu et al (110) identified 14 lactylation-related genes in ovarian cancer (OC) using The Cancer Genome Atlas database, stratifying patients into low- and high-risk groups. The low-risk profile was associated with metabolic processes (thermogenesis and OXPHOS) and immune responses (neutrophil extracellular traps and IL-17 signaling). High-risk features were associated with cell adhesion (proteoglycans and adhesive plaques) and carcinogenic signaling (Wnt and extracellular matrix-receptor interactions). Notably, lactylation-related genes were shown to be closely related to tumor classification and immunity, and OC based on lactylation-related characteristics was indicated to have a good prognostic performance.
Similarly, in skin cutaneous melanoma (CM), Kaplan-Meier survival analysis revealed that the lactylation low/TME high group had the highest overall survival (OS) rate, whereas the lactylation high/TME low group had the lowest OS rate. In addition, CM immunotherapy was revealed to be more suitable for patients in the lactylation low/TME high group (111). Notably, calmodulin-like protein 5 is a core lactylation-associated gene in CM, which is associated with patient survival and immune infiltration (112).
Lactate accumulation in the TME promotes immune evasion and tumor survival, but lactylation also offers therapeutic opportunities. Preclinical studies have highlighted strategies such as inhibiting lactate production (for example, via LDHA inhibitors) or disrupting histone lactylation [such as with demethylzeylasteral (DML), royal jelly acid (RJA) or fargesin] to reduce tumor energy supply or enhance immunotherapy efficacy (Fig. 2). However, these findings are limited by preclinical risks, as results from homogeneous mouse models may overestimate efficacy in humans, where TME heterogeneity remains unresolved.
Pan et al (113) demonstrated that DML can reduce lactate production by inhibiting glycolytic metabolic pathways, thereby suppressing the lactylation of histone H3 in liver cancer stem cells (LCSCs), especially at the H3K9la and H3K56la sites. Notably, DML inhibited the proliferation and migration of LCSCs, and promoted apoptosis. Furthermore, in a nude mouse transplanted tumor model, it was confirmed that DML exerted an anti-liver cancer effect by regulating histone H3 lactylation. However, Gong et al (114) argued that whether the results in the aforementioned study by Pan et al (113) should be attributed to histone lactylation was still controversial. Gong et al (114) suggested that the molecular weight markers of histones H1 and H4 are very similar to histone H3, and have been found to be involved in post-translational lactylation modifications. Therefore, detecting protein bands of various molecular weights by the naked eye alone may not be sensitive enough to distinguish them. It may be more appropriate to identify lactylated histones and their modification sites through the tandem mass spectrometry (MS/MS) spectra of Kla-containing peptides. In addition, liquid chromatography (LC)-MS/MS and chromatin immunoprecipitation-seq/RNA-seq multi-omics may be used to more accurately explore the regulatory mechanism of lactylation sites on HCC progression (115).
RJA has good in vitro and in vivo antitumor effects. RJA interferes with the production of lactate and inhibits the lactylation of H3K9la and H3K14la sites, affecting glycolytic metabolic pathways and inhibiting the development of HCC (116). The natural product fargesin inhibits cellular lactate production, and the expression of LDHA, LDHB, PKM2 and SLC2A1, and suppresses aerobic glycolysis and H3 histone lactylation in A549 NSCLC cells by targeting PKM2 (117). However, these compounds exhibit off-target effects; for example, DML cross-reacts with acetylation pathways, and efficacy is primarily validated in immunocompromised mouse models lacking human TME heterogeneity.
In a mouse liver cancer cell model (Hepa1-6), LDHA inhibitors have been shown to suppress the immunoregulatory effects of Treg cells in the TME by reducing lactate concentration. Notably, combined treatment with anti-PD-1 and LDHA inhibitors has a stronger antitumor effect than anti-PD-1 alone (118). The anti-epileptic drug, stiripentol enhances the sensitivity of cancer cells to temozolomide by inhibiting lactylation. Lactylation is upregulated in relapsed GBM tissues and temozolomide-resistant cells, mainly since H3K9la confers temozolomide resistance in GBM through Luc7L2-mediated intron-7 retention of MLH1. Stiripentol enhances temozolomide sensitivity in GBM by inhibiting LDH activity and H3K9la-mediated resistance (119). Oxamate promotes immune activation of chimeric antigen receptor (CAR)-T cells infiltrating tumors. In a GBM mouse model, oxamate has been shown to promote the immune activation of tumor-infiltrating CAR-T cells through altering the phenotypes of immune molecules and increasing Treg-cell infiltration (120). However, murine models cannot fully recapitulate human metabolic plasticity. PD-1/PD-L1 is a major inhibitory checkpoint pathway that regulates immune escape in patients with cancer, and its participation and inhibition notably reshape the pattern of tumor clearance. Immune checkpoint inhibition targeting PD-1/PD-L1 is a reliable tumor therapy (121). Notably, lactate treatment in PCa cells can increase the expression of HIF-1α and PD-L1, while restricting the expression of sema3A, whereas silencing the expression of MCT4 reverses this process. Evodiamine blocks lactate-induced angiogenesis by restricting the histone lactylation and expression of HIF-1α in PCa, further enhances Sema3A transcription, inhibits PD-L1 transcription, and induces ferroptosis by decreasing the expression of low GSH peroxidase 4 (122).
Proteomics analysis has shown that H4K12la is elevated in triple-negative breast cancer (TNBC) tissues, and H4K12la has been shown to be positively associated with the proliferative marker Ki-67 and negatively associated with OS rate (123). However, it is still necessary to validate H4K12la as a biomarker for TNBC resistance.
While lactylation is a promising therapeutic target, integrating preclinical findings with clinical practice requires: i) Interdisciplinary collaboration; for example, integrating metabolomics, epigenetics and immunology to address the complexity of the underlying mechanism. ii) Robust clinical validation, with staged trials to evaluate safety, efficacy and related biomarkers. iii) A patient-centered approach to develop guidelines for targeted lactylation therapy based on tumor subtypes, TME background and metabolic vulnerability. By addressing these challenges, lactate regulation could become a cornerstone of precision oncology.
Tumor cells produce lactate through glycolysis, resulting in local acidification and altering the pH value of the TME. The acidified environment affects the cellular components of the TME, and affects tumor growth and spread. Lactate acts as a signaling molecule to regulate tumor growth, invasion, metastasis and therapeutic response. Furthermore, lactylation is involved in critical biological processes such as glycolytic cell function, macrophage polarization, angiogenesis, mitochondrial activity and nervous system regulation. Lactylation also modulates tumor behavior by influencing signal transduction pathways. The interplay between lactylation and the immune cycle is complex, involving metabolism, immune cell function, inflammatory responses and immune regulation (Fig. 3). Therefore, intervening in lactylation may be a new target for tumor therapy.
The structural complexity and functional diversity of histones necessitate further exploration of their lactylation in cancer to elucidate its role and mechanism in tumor development. H3K18la is associated with shifts in cellular states, with quantitative lactylation changes at promoters and enhancers potentially driving these transitions (124). H3K9la levels are upregulated in endothelial cells in response to VEGF stimulation, and hyperlactylation of H3K9 inhibits expression of the lactylation eraser HDAC2, whereas upregulation of HDAC2 decreases H3K9la and suppresses angiogenesis (125,126). Elevated H3K9la levels in LCSCs, HCC and GBM are associated with tumorigenicity, drug resistance and immunosuppression (127). Histone lactylation thus influences gene expression, tumor progression and immune evasion, underscoring the need to determine its underlying mechanisms for developing novel therapies.
In GC, lactate-driven lactylation of NBS1 promotes homologous recombination (HR)-mediated DNA repair, fostering chemoresistance, while inhibition of lactate production has emerged as a promising strategy (128). Similarly, MRE11 lactylation at K673, mediated by cyclic-AMP response binding protein-binding protein (CBP), enhances HR repair, whereas targeting CBP or LDH sensitizes tumors to chemotherapy in preclinical models (129). KRAS-driven lactylation activates circATXN7, which sequesters NF-κB p65 in the cytoplasm, promoting immune evasion (130). However, the potential for resistance to lactylation-targeted therapies remains a concern, as tumors may exploit alternative DNA repair pathways or upregulate compensatory PTMs, highlighting the need for combination therapies.
While preclinical studies have demonstrated the potential of lactylation inhibition (such as by targeting LDH or CBP) to enhance chemosensitivity and reduce immune evasion, clinical translation faces notable hurdles (131,132). Small molecule inhibitors and antibodies targeting lactylation-related proteins have shown promising prospects in preclinical models, and targeted lactylation is expected to improve the immunosuppressive TME. For example, it has been shown that activating the SIRT3 lactase function can restore the anti-leukemia activity of NK cells (133). In addition, irinotecan has been found to possess new functions of inhibiting lactylation. Their safety is known and they can enter clinical trials more quickly (134). A number of small-molecule inhibitors also exhibit synergistic effects when used in combination with existing chemotherapy drugs or immunotherapies. Furthermore, lactylation levels vary among different tumor types and individuals Therefore, specific lactylation sites may be considered new and effective targets for tumor diagnosis and treatment. Furthermore, developing highly specific inhibitors that can precisely target specific lactylated proteins or sites is currently a challenge, and it is necessary to avoid interfering with other similar post-translational modifications (such as acetylation). How to efficiently deliver drugs to tumor tissues and penetrate into the interior of tumor cells is a common problem faced by clinical application. New delivery systems such as nano-strategies may be the solution. In addition, the lactylation regulatory network is complex, and tumor cells may develop drug resistance through other compensatory pathways; thus, it is necessary to continuously explore the mechanism and develop combined strategies (135,136).
Lactylation detection remains technically challenging. While high-performance LC-MS/MS, LC/MS and anti-Kla immunoblotting are widely used, improved precision is needed to distinguish lactylation from other PTMs, particularly given the overlapping molecular weights of histones H1, H3 and H4 (114). Advanced proteomics analysis and site-specific antibodies could address these limitations. Clinically, interventions must account for tissue-specific oxygen gradients and lactylation heterogeneity. Although lactylation biomarkers could monitor therapeutic response, validation in human cohorts is lacking. Early-phase trials of lactylation inhibitors are needed to assess safety, efficacy and optimal dosing.
As understanding of the role of lactylation in tumor biology deepens, targeting this pathway is expected to address treatment resistance and immunosuppression in some of the most aggressive malignant tumors, especially those driven by glycolytic metabolism and immune rejection. Future success will depend on a careful balance of efficacy and toxicity, the intelligent design of combination regimens, and the development of diagnostic tools to identify patients most likely to benefit from lactylation-directed therapy.
Not applicable.
This work was supported by Research Funds of Center for Xin'an Medicine and Modernization of Traditional Chinese Medicine of institute of Health and Medicine, Hefei Comprehensive National Science Center (grant no. 2023CXMMTCM025), the Anhui Education Department (grant no. gxgwfx2022019), the Anhui University of Chinese Medicine (grant no. 2022BHTNXA05, DT2400001288) and the Anhui Administration of Traditional Chinese Medicine (grant no. 2022CCYB10).
Not applicable.
XW, JC and BW collected and reviewed the literature, and wrote the review. YL, XZ and YS collected and reviewed the literature. CM made substantial contributions to the conception and design of the work. YH is responsible for the design of this work, the final approval of the version to be published, and for all aspects of the work in ensuring that questions related to the accuracy or integrity of any part of the work are appropriately investigated and resolved. 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.
|
Zhu W, Fan C, Hou Y and Zhang Y: Lactylation in tumor microenvironment and immunotherapy resistance: New mechanisms and challenges. Cancer Lett. 627:2178352025. View Article : Google Scholar | |
|
Brooks GA, Curl CC, Leija RG, Osmond AD, Duong JJ and Arevalo JA: Tracing the lactate shuttle to the mitochondrial reticulum. Exp Mol Med. 54:1332–1347. 2022. View Article : Google Scholar : PubMed/NCBI | |
|
Fan H, Yang F, Xiao Z, Luo H, Chen H, Chen Z, Liu Q and Xiao Y: Lactylation: Novel epigenetic regulatory and therapeutic opportunities. Am J Physiol Endocrinol Metab. 324:E330–E338. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Lv X, Lv Y and Dai X: Lactate, histone lactylation and cancer hallmarks. Expert Rev Mol Med. 25:e72023. 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 | |
|
Li S, Dong L and Wang K: Current and future perspectives of lysine lactylation in cancer. Trends Cell Biol. 35:190–193. 2025. View Article : Google Scholar | |
|
Yu X, Yang J, Xu J, Pan H, Wang W, Yu X and Shi S: Histone lactylation: From tumor lactate metabolism to epigenetic regulation. Int J Biol Sci. 20:1833–1854. 2024. View Article : Google Scholar | |
|
Zhang D, Tang Z, Huang H, Zhou G, Cui C, Weng Y, Liu W, Kim S, Lee S, Perez-Neut M, et al: Metabolic regulation of gene expression by histone lactylation. Nature. 574:575–580. 2019. View Article : Google Scholar : PubMed/NCBI | |
|
Irizarry-Caro RA, McDaniel MM, Overcast GR, Jain VG, Troutman TD and Pasare C: TLR signaling adapter BCAP regulates inflammatory to reparatory macrophage transition by promoting histone lactylation. Proc Natl Acad Sci USA. 117:30628–30638. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Wang J, Yang P, Yu T, Gao M, Liu D, Zhang J, Lu C, Chen X, Zhang X and Liu Y: Lactylation of PKM2 suppresses inflammatory metabolic adaptation in pro-inflammatory macrophages. Int J Biol Sci. 18:6210–6225. 2022. View Article : Google Scholar | |
|
Hua Q, Mi B, Xu F, Wen J, Zhao L, Liu J and Huang G: Hypoxia-induced lncRNA-AC020978 promotes proliferation and glycolytic metabolism of non-small cell lung cancer by regulating PKM2/HIF-1α axis. Theranostics. 10:4762–4778. 2020. 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 | |
|
Wu Z, Wu H, Dai Y, Wang Z, Han H, Shen Y, Zhang R and Wang X: A pan-cancer multi-omics analysis of lactylation genes associated with tumor microenvironment and cancer development. Heliyon. 10:e274652024. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang X, Li Y and Chen Y: Development of a comprehensive gene signature linking hypoxia, glycolysis, lactylation, and metabolomic insights in gastric cancer through the integration of bulk and single-cell RNA-Seq data. Biomedicines. 11:29482023. View Article : Google Scholar : PubMed/NCBI | |
|
Huang H, Chen K, Zhu Y, Hu Z, Wang Y, Chen J, Li Y, Li D and Wei P: A multi-dimensional approach to unravel the intricacies of lactylation related signature for prognostic and therapeutic insight in colorectal cancer. J Transl Med. 22:2112024. 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 | |
|
Zeng Y, Yu T, Lou Z, Chen L, Pan L and Ruan B: Emerging function of main RNA methylation modifications in the immune microenvironment of digestive system tumors. Pathol Res Pract. 256:1552682024. View Article : Google Scholar | |
|
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. 5:61–79. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Yang H, Zou X, Yang S, Zhang A, Li N and Ma Z: Identification of lactylation related model to predict prognostic, tumor infiltrating immunocytes and response of immunotherapy in gastric cancer. Front Immunol. 14:11499892023. View Article : Google Scholar | |
|
Mao Y, Zhang J, Zhou Q, He X, Zheng Z, Wei Y, Zhou K, Lin Y, Yu H, Zhang H, et al: Hypoxia induces mitochondrial protein lactylation to limit oxidative phosphorylation. Cell Res. 34:13–30. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Zong Z, Xie F, Wang S, Wu X, Zhang Z, Yang B and Zhou F: Alanyl-tRNA synthetase, AARS1, is a lactate sensor and lactyltransferase that lactylates p53 and contributes to tumorigenesis. Cell. 187:2375–2392.e33. 2024. View Article : Google Scholar | |
|
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. 134:e1745872024. View Article : Google Scholar | |
|
Zhou J, Xu W, Wu Y, Wang M, Zhang N, Wang L, Feng Y, Zhang T, Wang L and Mao A: GPR37 promotes colorectal cancer liver metastases by enhancing the glycolysis and histone lactylation via Hippo pathway. Oncogene. 42:3319–3330. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Miao Z, Zhao X and Liu X: Hypoxia induced β-catenin lactylation promotes the cell proliferation and stemness of colorectal cancer through the wnt signaling pathway. Exp Cell Res. 422:1134392023. View Article : Google Scholar : PubMed/NCBI | |
|
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. 81:651–666. 2024. View Article : Google Scholar | |
|
Song F, Hou C, Huang Y, Liang J, Cai H, Tian G, Jiang Y, Wang Z and Hou J: Lactylome analyses suggest systematic lysine-lactylated substrates in oral squamous cell carcinoma under normoxia and hypoxia. Cell Signal. 120:1112282024. View Article : Google Scholar | |
|
Wang M, He T, Meng D, Lv W, Ye J, Cheng L and Hu J: BZW2 modulates lung adenocarcinoma progression through glycolysis-mediated IDH3G lactylation modification. J Proteome Res. 22:3854–3865. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Xie B, Lin J, Chen X, Zhou X, Zhang Y, Fan M, Xiang J, He N, Hu Z and Wang F: CircXRN2 suppresses tumor progression driven by histone lactylation through activating the Hippo pathway in human bladder cancer. Mol Cancer. 22:1512023. View Article : Google Scholar : PubMed/NCBI | |
|
Wang R, Xu F, Yang Z, Cao J, Hu L and She Y: The mechanism of PFK-1 in the occurrence and development of bladder cancer by regulating ZEB1 lactylation. BMC Urol. 24:592024. View Article : Google Scholar | |
|
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. 31:779–791. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Pandkar MR, Sinha S, Samaiya A and Shukla S: Oncometabolite lactate enhances breast cancer progression by orchestrating histone lactylation-dependent c-Myc expression. Transl Oncol. 37:1017582023. View Article : Google Scholar | |
|
Hou X, Ouyang J, Tang L, Wu P, Deng X, Yan Q, Shi L, Fan S, Fan C, Guo C, et al: KCNK1 promotes proliferation and metastasis of breast cancer cells by activating lactate dehydrogenase A (LDHA) and up-regulating H3K18 lactylation. PLoS Biol. 22:e30026662024. View Article : Google Scholar | |
|
He Y, Ji Z, Gong Y, Fan L, Xu P, Chen X, Miao J, Zhang K, Zhang W, Ma P, et al: Numb/Parkin-directed mitochondrial fitness governs cancer cell fate via metabolic regulation of histone lactylation. Cell Rep. 42:1120332023. View Article : Google Scholar : PubMed/NCBI | |
|
Wang X, Ying T, Yuan J, Wang Y, Su X, Chen S, Zhao Y, Zhao Y, Sheng J, Teng L, et al: BRAFV600E restructures cellular lactylation to promote anaplastic thyroid cancer proliferation. Endocr Relat Cancer. 30:e2203442023. View Article : Google Scholar : PubMed/NCBI | |
|
Marcucci F and Rumio C: Tumor cell glycolysis-at the crossroad of epithelial-mesenchymal transition and autophagy. Cells. 11:10412022. View Article : Google Scholar : PubMed/NCBI | |
|
Jia M, Yue X, Sun W, Zhou Q, Chang C, Gong W, Feng J, Li X, Zhan R, Mo K, et al: ULK1-mediated metabolic reprogramming regulates Vps34 lipid kinase activity by its lactylation. Sci Adv. 9:eadg49932023. View Article : Google Scholar : PubMed/NCBI | |
|
Sun W, Jia M, Feng Y and Cheng X: Lactate is a bridge linking glycolysis and autophagy through lactylation. Autophagy. 19:3240–3241. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Ashton TM, McKenna WG, Kunz-Schughart LA and Higgins GS: Oxidative phosphorylation as an emerging target in cancer therapy. Clin Cancer Res. 24:2482–2490. 2018. View Article : Google Scholar : PubMed/NCBI | |
|
Sancho P, Barneda D and Heeschen C: Hallmarks of cancer stem cell metabolism. Br J Cancer. 114:1305–1312. 2016. View Article : Google Scholar : PubMed/NCBI | |
|
Payen VL, Mina E, Van Hée VF, Porporato PE and Sonveaux P: Monocarboxylate transporters in cancer. Mol Metab. 33:48–66. 2020. View Article : Google Scholar : PubMed/NCBI | |
|
Cheng Z, Huang H, Li M and Chen Y: Proteomic analysis identifies PFKP lactylation in SW480 colon cancer cells. iScience. 27:1086452023. View Article : Google Scholar | |
|
Huang H, Wang S, Xia H, Zhao X, Chen K, Jin G, Zhou S, Lu Z, Chen T, Yu H, et al: Lactate enhances NMNAT1 lactylation to sustain nuclear NAD+ salvage pathway and promote survival of pancreatic adenocarcinoma cells under glucose-deprived conditions. Cancer Lett. 588:2168062024. View Article : Google Scholar | |
|
Peng T, Sun F, Yang JC, Cai MH, Huai MX, Pan JX, Zhang FY and Xu LM: Novel lactylation-related signature to predict prognosis for pancreatic adenocarcinoma. World J Gastroenterol. 30:2575–2602. 2024. View Article : Google Scholar | |
|
Jiang J, Huang D, Jiang Y, Hou J, Tian M, Li J, Sun L, Zhang Y, Zhang T, Li Z, et al: Lactate modulates cellular metabolism through histone lactylation-mediated gene expression in non-small cell lung cancer. Front Oncol. 11:6475592021. View Article : Google Scholar | |
|
Daw CC, Ramachandran K, Enslow BT, Maity S, Bursic B, Novello MJ, Rubannelsonkumar CS, Mashal AH, Ravichandran J, Bakewell TM, et al: Lactate elicits ER-mitochondrial Mg2+ dynamics to integrate cellular metabolism. Cell. 183:474–489.e17. 2020. View Article : Google Scholar | |
|
Zhang R, Li L and Yu J: Lactate-induced IGF1R protein lactylation promotes proliferation and metabolic reprogramming of lung cancer cells. Open Life Sci. 19:202208742024. View Article : Google Scholar | |
|
Seo SH, Hwang SY, Hwang S, Han S, Park H, Lee YS, Rho SB and Kwon Y: Hypoxia-induced ELF3 promotes tumor angiogenesis through IGF1/IGF1R. EMBO Rep. 23:e529772022. View Article : Google Scholar : PubMed/NCBI | |
|
Longhitano L, Vicario N, Tibullo D, Giallongo C, Broggi G, Caltabiano R, Barbagallo GMV, Altieri R, Baghini M, Di Rosa M, et al: Lactate induces the expressions of MCT1 and HCAR1 to promote tumor growth and progression in glioblastoma. Front Oncol. 12:8717982022. View Article : Google Scholar | |
|
Longhitano L, Giallongo S, Orlando L, Broggi G, Longo A, Russo A, Caltabiano R, Giallongo C, Barbagallo I, Di Rosa M, et al: Lactate rewrites the metabolic reprogramming of uveal melanoma cells and induces quiescence phenotype. Int J Mol Sci. 24:242022. View Article : Google Scholar | |
|
Yang L, Wang X, Liu J, Liu X, Li S, Zheng F, Dong Q, Xu S, Xiong J and Fu B: Prognostic and tumor microenvironmental feature of clear cell renal cell carcinoma revealed by m6A and lactylation modification-related genes. Front Immunol. 14:12250232023. View Article : Google Scholar | |
|
Qu J, Li P and Sun Z: Histone lactylation regulates cancer progression by reshaping the tumor microenvironment. Front Immunol. 14:12843442023. View Article : Google Scholar | |
|
Li F, Si W, Xia L, Yin D, Wei T, Tao M, Cui X, Yang J, Hong T and Wei R: Positive feedback regulation between glycolysis and histone lactylation drives oncogenesis in pancreatic ductal adenocarcinoma. Mol Cancer. 23:902024. View Article : Google Scholar : PubMed/NCBI | |
|
Chen M, Cen K, Song Y, Zhang X, Liou YC, Liu P, Huang J, Ruan J, He J, Ye W, et al: NUSAP1-LDHA-glycolysis-lactate feedforward loop promotes Warburg effect and metastasis in pancreatic ductal adenocarcinoma. Cancer Lett. 567:2162852023. View Article : Google Scholar | |
|
Qiao Z, Li Y, Li S, Liu S and Cheng Y: Hypoxia-induced SHMT2 protein lactylation facilitates glycolysis and stemness of esophageal cancer cells. Mol Cell Biochem. 479:3063–3076. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Wei X, Chen Y, Jiang X, Peng M, Liu Y, Mo Y, Ren D, Hua Y, Yu B, Zhou Y, et al: Mechanisms of vasculogenic mimicry in hypoxic tumor microenvironments. Mol Cancer. 20:72021. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang M, Zhao Y, Liu X, Ruan X, Wang P, Liu L, Wang D, Dong W, Yang C and Xue Y: Pseudogene MAPK6P4-encoded functional peptide promotes glioblastoma vasculogenic mimicry development. Commun Biol. 6:10592023. View Article : Google Scholar | |
|
Liu R, Zou Z, Chen L, Feng Y, Ye J, Deng Y, Zhu X, Zhang Y, Lin J, Cai S, et al: FKBP10 promotes clear cell renal cell carcinoma progression and regulates sensitivity to the HIF2α blockade by facilitating LDHA phosphorylation. Cell Death Dis. 15:642024. View Article : Google Scholar : PubMed/NCBI | |
|
Luo Y, Yang Z, Yu Y and Zhang P: HIF1α lactylation enhances KIAA1199 transcription to promote angiogenesis and vasculogenic mimicry in prostate cancer. Int J Biol Macromol. 222:2225–2243. 2022. View Article : Google Scholar | |
|
Meng Q, Sun H, Zhang Y, Yang X, Hao S, Liu B, Zhou H, Xu ZX and Wang Y: Lactylation stabilizes DCBLD1 activating the pentose phosphate pathway to promote cervical cancer progression. J Exp Clin Cancer Res. 43:362024. View Article : Google Scholar : PubMed/NCBI | |
|
Meng Q, Zhang Y, Sun H, Yang X, Hao S, Liu B, Zhou H, Wang Y and Xu ZX: Human papillomavirus-16 E6 activates the pentose phosphate pathway to promote cervical cancer cell proliferation by inhibiting G6PD lactylation. Redox Biol. 71:1031082024. View Article : Google Scholar | |
|
Wei S, Zhang J, Zhao R, Shi R, An L, Yu Z, Zhang Q, Zhang J, Yao Y, Li H and Wang H: Histone lactylation promotes malignant progression by facilitating USP39 expression to target PI3K/AKT/HIF-1α signal pathway in endometrial carcinoma. Cell Death Discov. 10:1212024. View Article : Google Scholar : PubMed/NCBI | |
|
Zhao Y, Jiang J, Zhou P, Deng K, Liu Z, Yang M, Yang X, Li J, Li R and Xia J: H3K18 lactylation-mediated VCAM1 expression promotes gastric cancer progression and metastasis via AKT-mTOR-CXCL1 axis. Biochem Pharmacol. 222:1161202024. View Article : Google Scholar | |
|
Li XM, Yang Y, Jiang FQ, Hu G, Wan S, Yan WY, He XS, Xiao F, Yang XM, Guo X, et al: Histone lactylation inhibits RARγ expression in macrophages to promote colorectal tumorigenesis through activation of TRAF6-IL-6-STAT3 signaling. Cell Rep. 43:1136882024. View Article : Google Scholar : PubMed/NCBI | |
|
Liu M, Gu L, Zhang Y, Li Y, Zhang L, Xin Y, Wang Y and Xu ZX: LKB1 inhibits telomerase activity resulting in cellular senescence through histone lactylation in lung adenocarcinoma. Cancer Lett. 595:2170252024. View Article : Google Scholar | |
|
Zheng P, Mao Z, Luo M, Zhou L, Wang L, Liu H, Liu W and Wei S: Comprehensive bioinformatics analysis of the solute carrier family and preliminary exploration of SLC25A29 in lung adenocarcinoma. Cancer Cell Int. 23:2222023. View Article : Google Scholar | |
|
Li L, Li Z, Meng X, Wang X, Song D, Liu Y, Xu T, Qin J, Sun N, Tian K, et al: Histone lactylation-derived LINC01127 promotes the self-renewal of glioblastoma stem cells via the cis-regulating the MAP4K4 to activate JNK pathway. Cancer Lett. 579:2164672023. View Article : Google Scholar | |
|
Yang J, Luo L, Zhao C, Li X, Wang Z, Zeng Z, Yang X, Zheng X, Jie H, Kang L, et al: A positive feedback loop between inactive VHL-triggered histone lactylation and PDGFRβ signaling drives clear cell renal cell carcinoma progression. Int J Biol Sci. 18:3470–3483. 2022. View Article : Google Scholar | |
|
Jin J, Bai L, Wang D, Ding W, Cao Z, Yan P, Li Y, Xi L, Wang Y, Zheng X, et al: SIRT3-dependent delactylation of cyclin E2 prevents hepatocellular carcinoma growth. EMBO Rep. 24:e560522023. View Article : Google Scholar : PubMed/NCBI | |
|
Liao J, Chen Z, Chang R, Yuan T, Li G, Zhu C, Wen J, Wei Y, Huang Z, Ding Z, et al: CENPA functions as a transcriptional regulator to promote hepatocellular carcinoma progression via cooperating with YY1. Int J Biol Sci. 19:5218–5232. 2023. View Article : Google Scholar | |
|
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 USA. 121:e23141281212024. View Article : Google Scholar : PubMed/NCBI | |
|
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 m6A modification promotes immunosuppression of tumor-infiltrating myeloid cells. Mol Cell. 82:1660–1677.e10. 2022. View Article : Google Scholar | |
|
Yu J, Chai P, Xie M, Ge S, Ruan J, Fan X and Jia R: Histone lactylation drives oncogenesis by facilitating m6A reader protein YTHDF2 expression in ocular melanoma. Genome Biol. 22:852021. View Article : Google Scholar | |
|
Gu X, Zhuang A, Yu J, Yang L, Ge S, Ruan J, Jia R, Fan X and Chai P: Histone lactylation-boosted ALKBH3 potentiates tumor progression and diminished promyelocytic leukemia protein nuclear condensates by m1A demethylation of SP100A. Nucleic Acids Res. 52:2273–2289. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Chen B, Deng Y, Hong Y, Fan L, Zhai X, Hu H, Yin S, Chen Q, Xie X, Ren X, et al: Metabolic recoding of NSUN2-mediated m5c modification promotes the progression of colorectal cancer via the NSUN2/YBX1/m5C-ENO1 positive feedback loop. Adv Sci (Weinh). 11:e23098402024. View Article : Google Scholar : PubMed/NCBI | |
|
Takata T, Nakamura A, Yasuda H, Miyake H, Sogame Y, Sawai Y, Hayakawa M, Mochizuki K, Nakao R, Ogata T, et al: Pathophysiological implications of protein lactylation in pancreatic epithelial tumors. Acta Histochem Cytochem. 57:57–66. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Cai D, Yuan X, Cai DQ, Li A, Yang S, Yang W, Duan J, Zhuo W, Min J, Peng L and Wei J: Integrative analysis of lactylation-related genes and establishment of a novel prognostic signature for hepatocellular carcinoma. J Cancer Res Clin Oncol. 149:11517–11530. 2023. View Article : Google Scholar | |
|
Yang H, Yang S, He J, Li W, Zhang A, Li N, Zhou G and Sun B: Glucose transporter 3 (GLUT3) promotes lactylation modifications by regulating lactate dehydrogenase A (LDHA) in gastric cancer. Cancer Cell Int. 23:3032023. View Article : Google Scholar | |
|
Yang D, Yin J, Shan L, Yi X, Zhang W and Ding Y: Identification of lysine-lactylated substrates in gastric cancer cells. iScience. 25:1046302022. View Article : Google Scholar | |
|
Hu X, Ouyang W, Chen H, Liu Z, Lai Z and Yao H: Claudin-9 (CLDN9) promotes gastric cancer progression by enhancing the glycolysis pathway and facilitating PD-L1 lactylation to suppress CD8+ T cell anti-tumor immunity. Cancer Pathog Ther. 3:253–266. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Huang ZW, Zhang XN, Zhang L, Liu LL, Zhang JW, Sun YX, Xu JQ, Liu Q and Long ZJ: STAT5 promotes PD-L1 expression by facilitating histone lactylation to drive immunosuppression in acute myeloid leukemia. Signal Transduct Target Ther. 8:3912023. View Article : Google Scholar : PubMed/NCBI | |
|
Gu J, Xu X, Li X, Yue L, Zhu X, Chen Q, Gao J, Takashi M, Zhao W, Zhao B, et al: Tumor-resident microbiota contributes to colorectal cancer liver metastasis by lactylation and immune modulation. Oncogene. 43:2389–2404. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Wang J, Liu Z, Xu Y, Wang Y, Wang F, Zhang Q, Ni C, Zhen Y, Xu R, Liu Q, et al: Enterobacterial LPS-inducible LINC00152 is regulated by histone lactylation and promotes cancer cells invasion and migration. Front Cell Infect Microbiol. 12:9138152022. View Article : Google Scholar | |
|
Hao B, Dong H, Xiong R, Song C, Xu C, Li N and Geng Q: Identification of SLC2A1 as a predictive biomarker for survival and response to immunotherapy in lung squamous cell carcinoma. Comput Biol Med. 171:1081832024. View Article : Google Scholar : PubMed/NCBI | |
|
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. 57:1105–1123.e8. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Lu X, Zhou Z, Qiu P and Xin T: Integrated single-cell and bulk RNA-sequencing data reveal molecular subtypes based on lactylation-related genes and prognosis and therapeutic response in glioma. Heliyon. 10:e307262024. View Article : Google Scholar : PubMed/NCBI | |
|
Wu X, Mi T, Jin L, Ren C, Wang J, Zhang Z, Liu J, Wang Z, Guo P and He D: Dual roles of HK3 in regulating the network between tumor cells and tumor-associated macrophages in neuroblastoma. Cancer Immunol Immunother. 73:1222024. View Article : Google Scholar : PubMed/NCBI | |
|
Wang ZH, Zhang P, Peng WB, Ye LL, Xiang X, Wei XS, Niu YR, Zhang SY, Xue QQ, Wang HL and Zhou Q: Altered phenotypic and metabolic characteristics of FOXP3+CD3+CD56+ natural killer T (NKT)-like cells in human malignant pleural effusion. Oncoimmunology. 12:21605582022. View Article : Google Scholar : PubMed/NCBI | |
|
Hong H, Chen X, Wang H, Gu X, Yuan Y and Zhang Z: Global profiling of protein lysine lactylation and potential target modified protein analysis in hepatocellular carcinoma. Proteomics. 23:e22004322023. View Article : Google Scholar : PubMed/NCBI | |
|
Cheng Z, Huang H, Li M, Liang X, Tan Y and Chen Y: Lactylation-related gene signature effectively predicts prognosis and treatment responsiveness in hepatocellular carcinoma. Pharmaceuticals (Basel). 16:6442023. View Article : Google Scholar : PubMed/NCBI | |
|
Wu Q, Li X, Long M, Xie X and Liu Q: Integrated analysis of histone lysine lactylation (Kla)-specific genes suggests that NR6A1, OSBP2 and UNC119B are novel therapeutic targets for hepatocellular carcinoma. Sci Rep. 13:186422023. View Article : Google Scholar : PubMed/NCBI | |
|
Lai JP, Sandhu DS, Moser CD, Cazanave SC, Oseini AM, Shire AM, Shridhar V, Sanderson SO and Roberts LR: Additive effect of apicidin and doxorubicin in sulfatase 1 expressing hepatocellular carcinoma in vitro and in vivo. J Hepatol. 50:1112–1121. 2009. View Article : Google Scholar | |
|
Wang Y, Pei P, Yang K, Guo L and Li Y: Copper in colorectal cancer: From copper-related mechanisms to clinical cancer therapies. Clin Transl Med. 14:e17242024. View Article : Google Scholar : PubMed/NCBI | |
|
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 m6A-modification on FDX1 mRNA in gastric cancer. Nat Commun. 14:65232023. View Article : Google Scholar : PubMed/NCBI | |
|
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 : PubMed/NCBI | |
|
Sun X, He L, Liu H, Thorne RF, Zeng T, Liu L, Zhang B, He M, Huang Y, Li M, et al: The diapause-like colorectal cancer cells induced by SMC4 attenuation are characterized by low proliferation and chemotherapy insensitivity. Cell Metab. 35:1563–1579.e8. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Chu YD, Cheng LC, Lim SN, Lai MW, Yeh CT and Lin WR: Aldolase B-driven lactagenesis and CEACAM6 activation promote cell renewal and chemoresistance in colorectal cancer through the Warburg effect. Cell Death Dis. 14:6602023. View Article : Google Scholar : PubMed/NCBI | |
|
Brown TP and Ganapathy V: Lactate/GPR81 signaling and proton motive force in cancer: Role in angiogenesis, immune escape, nutrition, and Warburg phenomenon. Pharmacol Ther. 206:1074512020. View Article : Google Scholar : PubMed/NCBI | |
|
Xia L, Oyang L, Lin J, Tan S, Han Y, Wu N, Yi P, Tang L, Pan Q, Rao S, et al: The cancer metabolic reprogramming and immune response. Mol Cancer. 20:282021. View Article : Google Scholar : PubMed/NCBI | |
|
Yan F, Teng Y, Li X, Zhong Y, Li C, Yan F and He X: Hypoxia promotes non-small cell lung cancer cell stemness, migration, and invasion via promoting glycolysis by lactylation of SOX9. Cancer Biol Ther. 25:23041612024. View Article : Google Scholar : PubMed/NCBI | |
|
Duan W, Liu W, Xia S, Zhou Y, Tang M, Xu M, Lin M, Li X and Wang Q: Warburg effect enhanced by AKR1B10 promotes acquired resistance to pemetrexed in lung cancer-derived brain metastasis. J Transl Med. 21:5472023. View Article : Google Scholar : PubMed/NCBI | |
|
Zheng H, Peng X, Yang S, Li X, Huang M, Wei S, Zhang S, He G, Liu J, Fan Q, et al: Targeting tumor-associated macrophages in hepatocellular carcinoma: Biology, strategy, and immunotherapy. Cell Death Discov. 9:652023. View Article : Google Scholar : PubMed/NCBI | |
|
Li F, Zhang H, Huang Y, Li D, Zheng Z, Xie K, Cao C, Wang Q, Zhao X, Huang Z, et al: Single-cell transcriptome analysis reveals the association between histone lactylation and cisplatin resistance in bladder cancer. Drug Resist Updat. 73:1010592024. View Article : Google Scholar | |
|
Li K, Chen MK, Li LY, Lu MH, Shao ChK, Su ZL, He D, Pang J and Gao X: The predictive value of semaphorins 3 expression in biopsies for biochemical recurrence of patients with low- and intermediate-risk prostate cancer. Neoplasma. 60:683–689. 2013. View Article : Google Scholar | |
|
Pan J, Zhang J, Lin J, Cai Y and Zhao Z: Constructing lactylation-related genes prognostic model to effectively predict the disease-free survival and treatment responsiveness in prostate cancer based on machine learning. Front Genet. 15:13431402024. View Article : Google Scholar | |
|
Chaudagar K, Hieromnimon HM, Khurana R, Labadie B, Hirz T, Mei S, Hasan R, Shafran J, Kelley A, Apostolov E, et al: Reversal of lactate and PD-1-mediated macrophage immunosuppression controls growth of PTEN/p53-deficient prostate cancer. Clin Cancer Res. 29:1952–1968. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Chaudagar K, Hieromnimon HM, Kelley A, Labadie B, Shafran J, Rameshbabu S, Drovetsky C, Bynoe K, Solanki A, Markiewicz E, et al: Suppression of tumor cell lactate-generating signaling pathways eradicates murine PTEN/p53-deficient aggressive-variant prostate cancer via macrophage phagocytosis. Clin Cancer Res. 29:4930–4940. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Zhang XW, Li L, Liao M, Liu D, Rehman A, Liu Y, Liu ZP, Tu PF and Zeng KW: Thermal proteome profiling strategy identifies CNPY3 as a cellular target of gambogic acid for inducing prostate cancer pyroptosis. J Med Chem. 67:10005–10011. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Jiao Y, Ji F, Hou L, Lv Y and Zhang J: Lactylation-related gene signature for prognostic prediction and immune infiltration analysis in breast cancer. Heliyon. 10:e247772024. View Article : Google Scholar : PubMed/NCBI | |
|
Deng J and Liao X: Lysine lactylation (Kla) might be a novel therapeutic target for breast cancer. BMC Med Genomics. 16:2832023. View Article : Google Scholar : PubMed/NCBI | |
|
Yu L, Jing C, Zhuang S, Ji L and Jiang L: A novel lactylation-related gene signature for effectively distinguishing and predicting the prognosis of ovarian cancer. Transl Cancer Res. 13:2497–2508. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
Zhu Y and Song B, Yang Z, Peng Y, Cui Z, Chen L and Song B: Integrative lactylation and tumor microenvironment signature as prognostic and therapeutic biomarkers in skin cutaneous melanoma. J Cancer Res Clin Oncol. 149:17897–17919. 2023. View Article : Google Scholar | |
|
Feng H, Chen W and Zhang C: Identification of lactylation gene CALML5 and its correlated lncRNAs in cutaneous melanoma by machine learning. Medicine (Baltimore). 102:e359992023. View Article : Google Scholar : PubMed/NCBI | |
|
Pan L, Feng F, Wu J, Fan S, Han J, Wang S, Yang L, Liu W, Wang C and Xu K: Demethylzeylasteral targets lactate by inhibiting histone lactylation to suppress the tumorigenicity of liver cancer stem cells. Pharmacol Res. 181:1062702022. View Article : Google Scholar : PubMed/NCBI | |
|
Gong H, Xu HM, Ma YH and Zhang DK: Demethylzeylasteral targets lactate to suppress the tumorigenicity of liver cancer stem cells: It is attributed to histone lactylation? Pharmacol Res. 194:1068692023. View Article : Google Scholar : PubMed/NCBI | |
|
Wang C, Feng F, Pan L and Xu K: Reply to the letter titled: Demethylzeylasteral targets lactate to suppress the tumorigenicity of liver cancer stem cells: Is it attributed to histone lactylation? Pharmacol Res. 194:1068682023. View Article : Google Scholar : 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. 118:1549402023. View Article : Google Scholar : PubMed/NCBI | |
|
Guo Z, Tang Y, Wang S, Huang Y, Chi Q, Xu K and Xue L: Natural product fargesin interferes with H3 histone lactylation via targeting PKM2 to inhibit non-small cell lung cancer tumorigenesis. Biofactors. 50:592–607. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
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. 39:1109862022. View Article : Google Scholar : PubMed/NCBI | |
|
Yue Q, Wang Z, Shen Y, Lan Y, Zhong X, Luo X, Yang T, Zhang M, Zuo B, Zeng T, et al: Histone H3K9 lactylation confers temozolomide resistance in glioblastoma via LUC7L2-Mediated MLH1 intron retention. Adv Sci (Weinh). 11:e23092902024. View Article : Google Scholar : PubMed/NCBI | |
|
Sun T, Liu B, Li Y, Wu J, Cao Y, Yang S, Tan H, Cai L, Zhang S, Qi X, et al: Oxamate enhances the efficacy of CAR-T therapy against glioblastoma via suppressing ectonucleotidases and CCR8 lactylation. J Exp Clin Cancer Res. 42:2532023. View Article : Google Scholar : PubMed/NCBI | |
|
Hu X, Wang J, Chu M, Liu Y, Wang ZW and Zhu X: Emerging role of ubiquitination in the regulation of PD-1/PD-L1 in cancer immunotherapy. Mol Ther. 29:908–919. 2021. 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 | |
|
Cui Z, Li Y, Lin Y, Zheng C, Luo L, Hu D, Chen Y, Xiao Z and Sun Y: Lactylproteome analysis indicates histone H4K12 lactylation as a novel biomarker in triple-negative breast cancer. Front Endocrinol (Lausanne). 15:13286792024. View Article : Google Scholar : PubMed/NCBI | |
|
Galle E, Wong CW, Ghosh A, Desgeorges T, Melrose K, Hinte LC, Castellano-Castillo D, Engl M, de Sousa JA, Ruiz-Ojeda FJ, et al: H3K18 lactylation marks tissue-specific active enhancers. Genome Biol. 23:2072022. View Article : Google Scholar | |
|
Fan W, Zeng S, Wang X, Wang G, Liao D, Li R, He S, Li W, Huang J, Li X, et al: A feedback loop driven by H3K9 lactylation and HDAC2 in endothelial cells regulates VEGF-induced angiogenesis. Genome Biol. 25:1652024. View Article : Google Scholar | |
|
Dai W, Wu G, Liu K, Chen Q, Tao J, Liu H and Shen M: Lactate promotes myogenesis via activating H3K9 lactylation-dependent up-regulation of Neu2 expression. J Cachexia Sarcopenia Muscle. 14:2851–2865. 2023. View Article : Google Scholar : PubMed/NCBI | |
|
Li X, Chen M, Chen X, He X, Li X, Wei H, Tan Y, Min J, Azam T, Xue M, et al: TRAP1 drives smooth muscle cell senescence and promotes atherosclerosis via HDAC3-primed histone H4 lysine 12 lactylation. Eur Heart J. 45:4219–4235. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
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. 631:663–669. 2024. View Article : Google Scholar : PubMed/NCBI | |
|
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. 187:294–311.e21. 2024. View Article : Google Scholar | |
|
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. 15:4992024. View Article : Google Scholar : PubMed/NCBI | |
|
Deng D, Luo Y, Hong Y, Ren X, Zu X and Feng J: Lactylation: A new direction for tumor-targeted therapy. Biochim Biophys Acta Rev Cancer. 1880:1893992025. View Article : Google Scholar : PubMed/NCBI | |
|
Hou X, Hong Z, Zen H, Zhang C, Zhang P, Ma D and Han Z: Lactylation in cancer biology: Unlocking new avenues for research and therapy. Cancer Commun (Lond). Aug 19–2025.(Epub ahead of print). View Article : Google Scholar | |
|
Jin J, Yan P, Wang D, Bai L, Liang H, Zhu X, Zhu H, Ding C, Wei H and Wang Y: Targeting lactylation reinforces NK cell cytotoxicity within the tumor microenvironment. Nat Immunol. 26:1099–1112. 2025. View Article : Google Scholar | |
|
Li X, Zhang C, Mei Y, Zhong W, Fan W, Liu L, Feng Z, Bai X, Liu C, Xiao M, et al: Irinotecan alleviates chemoresistance to anthracyclines through the inhibition of AARS1-mediated BLM lactylation and homologous recombination repair. Signal Transduct Target Ther. 10:2142025. 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 | |
|
Chen X, Yuan Y, Zhou F, Li L, Pu J, Zeng Y and Jiang X: Lactylation: From homeostasis to pathological implications and therapeutic strategies. MedComm (2020). 6:e702262025. View Article : Google Scholar : PubMed/NCBI |