Hostname: page-component-7bb8b95d7b-pwrkn Total loading time: 0 Render date: 2024-09-26T16:51:44.014Z Has data issue: false hasContentIssue false

Lactate, histone lactylation and cancer hallmarks

Published online by Cambridge University Press:  09 January 2023

Xinyu Lv
Affiliation:
Wuxi School of Medicine, Jiangnan University, Wuxi 214122, China
Yingying Lv
Affiliation:
Department of Clinical Laboratory, Shanghai Pudong New Area People's Hospital, Shanghai 201299, China
Xiaofeng Dai*
Affiliation:
Wuxi School of Medicine, Jiangnan University, Wuxi 214122, China National Local Joint Engineering Research Center for Precision Surgery & Regenerative Medicine, Shaanxi Provincial Center for Regenerative Medicine and Surgical Engineering, The First Affiliated Hospital of Xi'an Jiaotong University, Xi'an 710061, China
*
Author for correspondence: Xiaofeng Dai, E-mail: xiaofeng.dai@jiangnan.edu.cn
Rights & Permissions [Opens in a new window]

Abstract

Histone lactylation, an indicator of lactate level and glycolysis, has intrinsic connections with cell metabolism that represents a novel epigenetic code affecting the fate of cells including carcinogenesis. Through delineating the relationship between histone lactylation and cancer hallmarks, we propose histone lactylation as a novel epigenetic code priming cells toward the malignant state, and advocate the importance of identifying novel therapeutic strategies or dual-targeting modalities against lactylation toward effective cancer control. This review underpins important yet less-studied area in histone lactylation, and sheds insights on its clinical impact as well as possible therapeutic tools targeting lactylation.

Type
Review
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press

Introduction

Nucleosome, being the basic repeating unit of eukaryotic chromatin, is composed of a histone octamer (H3, H4, H2A and H2B) and 147 bp DNA packaged in the form of a bead on a string (Ref. Reference de la Cruz Munoz-Centeno, Millan-Zambrano and Chavez1). The terminal tail of nucleosome is subjected to various epigenetic modifications such as methylation, acetylation, succinylation phosphorylation, SUMOylation and ubiquitination. While some of these epigenetic marks such as H3K27me3 and H3K9me3 are inheritable, some such as H3K36me3 and acetylation codes have unknown heritability (Refs Reference Mikkelsen2, Reference Bernstein, Meissner and Lander3). Accumulated evidence has suggested the critical and diversified roles of epigenetic codes in cell development and decision making under physiological and pathological conditions including complex disorders such as cancers.

Intensive efforts have been dedicated to profile the epigenome of human cells, and increasing number of novel epigenetic codes have been discovered including lactylation with the development of high-throughput sequencing technologies. Histone lactylation was firstly reported in 2019 by Zhang et al. as an addition of a lactyl (La) group to the lysine amino acid residues in the tails of histone proteins (Ref. Reference Zhang4). Ever since then, histone lactylation has been consecutively reported in a diverse spectrum of organisms including cancers such as ocular melanoma (Ref. Reference Yu5), non-small cell lung cancer (Ref. Reference Jiang6), sarcoma (Ref. Reference Hua7), immunity relevant cells such as macrophage (Ref. Reference Cui8), and plants such as rice grain (Ref. Reference Meng9). Similar to other histone modification codes such as acetylation, methylation, phosphorylation, ubiquitination, lactylation was identified as another epigenomic modifier of gene expression (Ref. Reference Zhang4).

By systematically reviewing the metabolism of lactate and its association with histone lactylation in ‘Lactate and histone lactylation’ section and discussing the functionalities of histone lactylation in fostering cancer hallmarks in ‘Histone lactylation and cancer hallmarks’ section, this review identifies unresolved issues in lactylation for future research and advocates targeting lactylation as an innovative onco-therapeutics, alone or in combination with other treatment strategies, toward enhanced clinical outcome in ‘Discussion’ section.

Lactate and histone lactylation

Lactate, an end product of glycolysis (Fig. 1), had once been misinterpreted as a waste since its discovery in 1780. Accumulating evidence has suggested lactate as a universal metabolic fuel for normal tissues such as skeletal muscles (Ref. Reference Nikooie, Moflehi and Zand10), heart (Ref. Reference Matejovic, Radermacher and Fontaine11), brain (Ref. Reference Dienel12) and malignant cells (Ref. Reference Bonuccelli13) to contribute to the cell fate decision making process such as macrophage polarisation. It is also considered as a metabolic buffer connecting glycolysis and oxidative phosphorylation (Ref. Reference Haas14). In addition, lactate can function as a signalling molecule with a variety of modulatory roles such as immune cell modulation (Ref. Reference Lee15), lipolysis (Ref. Reference Ahmed16), wound healing (Ref. Reference Weinreich17) and cellular homoeostasis maintenance (Refs Reference Jha and Morrison18, Reference Jain19). Lactate is abundantly produced in the tumour milieu and probably the most significant metabolic hallmark of cancer cells, known as the Warburg effect or the glycolytic switch (Refs Reference Vaupel and Multhoff20, Reference Chen21, Reference Kes22) (Fig. 2). Non-malignant lymphocytes or stromal cells such as tumour-associated macrophage (TAM) and cancer-associated fibroblast (CAF) also contribute to lactate accumulation in the tumour microenvironment (TME), called the reverse Warburg effect (Ref. Reference Vlachostergios23). Lactate secreted from hypoxic tumour cells can be up-taken by normoxic tumour cells to allow the diffuse of glucose toward the more hypoxic cells, and such a lactate-based metabolic symbiosis supports the survival of both hypoxic and normoxic cancer cells in the acidic environment (Refs Reference Kes22, Reference Sonveaux24). The lactate concentration is significantly higher in grade III (7.7 ± 2.9 mM) than grade II (5.5 ± 2.4 mM) breast cancer tissues, positively correlated with Nottingham Prognostic Index and negatively associated with lactate dehydrogenase A (LDHA) level (Ref. Reference Cheung25). Isotope tracer measurements show a rapid lactate exchange between the tumour and circulation; most pyruvates produced from tumour cells are converted into lactates and excreted, and most pyruvates fed into the tricarboxylic acid (TCA) cycle in tumour cells come from circulating lactates produced elsewhere (Ref. Reference Hui26). In consistent with this, hyperpolarised [1-13C]pyruvate is converted more rapidly to lactate in tumours than the benign adjacent tissue; and the malignant tissue exhibits elevated levels of hyperpolarised [1-13C]lactate and [1-13C]lactate/[1-13C]pyruvate ratio on external infusion of hyperpolarised pyruvate in prostate cancers (Ref. Reference Nelson27).

Fig. 1. Lactate metabolism. The main biochemical players in glycolysis and the TCA cycle that participate in lactate metabolism.

Fig. 2. Simplified illustration on associations between lactate and cancer hallmarks, as well as possible therapeutic strategies targeting lactate and lactylation. Lactates, once generated from glycolysis (with LDHA being the enzyme catalysing the last step) and entered in the tumour microenvironment (TME) including tumour, stromal, various types of immune cells, and blood vessels, increases the TME acidity and thus contributes to ‘onco-therapeutic resistance’. Lactates, generated from peripheral tissues such as tumour associated fibroblasts (TAF), enter cells via MCT1/4 and are reutilised toward enhanced glycolysis via lactate shuttle, contributing to ‘metabolic reprogramming’. Lactate accumulation and acidification of the TME are also relevant to ‘tumour-associated inflammation’ by, e.g., stimulating TAM toward M2-like polarisation that is associated with enhanced production of cytokines and chemokines such as IL6 and CCL5, whereas prolonged inflammation triggers altered profiling of oncogenes and tumour suppressors that promote carcinogenesis. Lactates can also aid tumour cells in ‘immunosurveillance evasion by, e.g., suppressing the antigen presentation ability of dendritic cells and triggering apoptosis of NK cells. Accelerated glycolysis toward excess lactate accumulation can cause or be the result of mutations of tumour driver genes such as p53 and HIF-1α, and thus be associated with ‘genome instability’. Lactate functions through receptors such as MCT1/4 and GPR81, the mutations of which halt ‘cancer growth’. Lactate promotes ‘tumour angiogenesis and metastasis’ via, e.g., stabilizing HIF1α, triggering TAM polarisation toward the M2 state that is pro-angiogenic accompanied with over-representation of Arg1 and Vegf, ameliorating conjugations with the extracellular matrix components and increasing TME acidity to enable subsequent cancer cell migration. Regarding therapeutic opportunities, targeting MCT1/4 abolishes the resistance of cancer cells to MET/EGFR tyrosine kinase inhibitors, and cold atmospheric plasma (CAP), an emerging onco-therapeutic strategy, has demonstrated its efficacy in suppressing LDHA. Bold text in caption signifies the possible therapeutic strategies targeting lactate and lactylation.

Lactagenesis (augmented lactate production) has been proposed to explain the Warburg effect that drives carcinogenesis (Ref. Reference San-Millan and Brooks28), which is featured by five steps, i.e., enhanced glucose uptake, increased glycolytic enzyme expression and activity, decreased mitochondrial function, elevated lactate generation, accumulation and release, as well as upregulated monocarboxylate transporters 1/4 (MCT1/4) for accelerated lactate shuttle (Ref. Reference San-Millan and Brooks28).

Histone lactylation is sensitive to lactate level. Glycolysis inhibitors reducing lactate production decrease lysine lactylation, mitochondrial inhibitors or hypoxia elevating lactate generation can increase lysine lactylation (Ref. Reference Izzo and Wellen29). For instance, lysine lactylation is abolished when LDHA is not functional (Ref. Reference Zhang4); and is enhanced on the increase of cellular lactate level under conditions such as hypoxia, M1 macrophage polarisation, glucose supplementation and treatment with mitochondrial inhibitor rotenon (Ref. Reference Zhang4). Lactate is necessary for histone lactylation that stimulates the expression of genes responsible for switching macrophages from the M1 to M2 phenotype and favours cancer initiation and progression (Ref. Reference Chen30). By adopting four orthogonal methods, Zhang et al. demonstrated histone lysine lactylation as an in vivo protein post-translational modification derived from lactate that represents a new avenue for deciphering the roles of lactate under diversified physiological and pathological conditions including cancers (Ref. Reference Zhang4).

Similar with other epigenetic codes such as methylation and acetylation, lactylation is regulated by writers (i.e., enzymes that transfer the lactyl moieties to the targeted proteins) and erasers (i.e., enzymes that remove the lactyl moieties from the targeted proteins), and functions together with its readers (i.e., proteins that identify lactylation to take on corresponding functionalities). Lactyl-CoA, which has been detected by liquid chromatography mass spectrometry (LC-MS) in mammalian cells and tissues (Ref. Reference Varner31), offers the substrate during the enzymatic lactylation, and lactyl-glutathione is involved in the lactyl moiety transfer of non-enzymatic lactylation. Relatively little has been reported on the writer, eraser and reader of lactylation except for p300 (Refs Reference Zhang4, Reference Cui8), the first lactylation writer so far identified. As different forms of epigenetic codes may share the use of enzymes such as the ‘writing’ function of p300 in lactylation and acetylation (Refs Reference Chen30, Reference Wang32), it is possible that enzymes with writing, erasing and reading roles in other epigenetic marks take on similar functionalities in lactylation.

Histone lactylation and cancer hallmarks

As pointed out by Dr Hanahan and Dr Weinberg in 2011, tumours have gained an additional layer of complexity over the already identified 6 basic hallmarks in 2000 (i.e., sustained proliferation, apoptosis resistance, growth suppressor evasion, replicative immortality, tumour angiogenesis, invasion and metastasis) (Ref. Reference Hanahan and Weinberg33) by recruiting and communicating with a repertoire of ostensibly normal cells that constitute to the TME, which enable 4 other cancer hallmarks, i.e., metabolic reprogramming, tumour-associated inflammation, immunosurveillance evasion and genome instability (Ref. Reference Hanahan and Weinberg34).

Since lactylation is derived from lactate that is one end product of glycolysis, it has intrinsic connections with cell metabolism and TME. We start with 3 enabling cancer characteristics (i.e., metabolic reprogramming, tumour-associated inflammation and immunosurveillance evasion) that have well-documented tight connections with lactylation in the following subsections. Then, we discuss possible connections between histone lactylation and ‘genome instability’, the last enabling cancer hallmark that underlies an area deserving more attention. Lastly, we re-organise the six basic hallmarks into ‘cell growth’, ‘tumour angiogenesis/metastasis’, ‘drug resistance’, and discuss the roles of lactylation in driving these cancer traits.

Histone lactylation and metabolic reprogramming

Reprogrammed metabolism is a well-known hallmark of cancer (Ref. Reference Hanahan and Weinberg34). Lactylation reflects the level of lactate (an important metabolite) that, in turn, drives lactylation. This builds an intrinsic connection between lactylation and cell metabolism.

Histone lactylation functions as a linker between reprogrammed cell metabolism and disordered transcriptome in cancer cells (Ref. Reference Izzo and Wellen29). Altered cell metabolism in malignant cells may affect the level of lactate as represented by an altered histone lactylation landscape (Ref. Reference Izzo and Wellen29), and alterations in the histone lactylation profile of cancer cells may change the transcriptomic profile to adapt to the reprogrammed cell metabolism in the chaotic state. For instance, lactate was shown to modulate cellular metabolism by down-regulating the mRNA levels of glycolytic enzymes hexokinase 1 (HK1) and pyruvate kinase (PKM) and up-regulating that of TCA cycle enzymes succinate dehydrogenase complex flavoprotein subunit A (SDHA) and isocitrate dehydrogenase NAD( + )3 non-catalytic subunit gamma (IDH3G) through promoting histone lactylation in the promoter regions of these genes in non-small cell lung cancer (Ref. Reference Jiang6). Besides, a positive correlation was observed between histone lactylation and Arg1 expression in TAMs isolated from B16F10 melanoma and LLC1 lung tumour cells (Ref. Reference Izzo and Wellen29); and exogenous lactate was shown to enhance the transcription of vascular endothelial growth factor A (Vegfa) during TAM functional polarisation (Ref. Reference Colegio35).

Histone lactylation and tumour-associated inflammation

In addition to functioning as an intermediate metabolite of energy source and biosynthetic pathway, lactate accumulation also occurs during local inflammation and thus is associated with tumour-associated inflammation, another hallmark of cancer (Ref. Reference Hanahan and Weinberg34).

Macrophages, a heterogeneous cell cohort, play critical roles in regulating immune response and maintaining tissue homoeostasis, whereas its plasticity is modulated at least partly through epigenetic dynamics during inflammation (Ref. Reference Geissmann36). There are two types of macrophages, i.e., the proinflammatory M1 state and the immune regulatory M2 state (Ref. Reference Wynn, Chawla and Pollard37). Transition of macrophages from the M1 to the M2 phenotype is vital for switching healthy cells from the inflammatory state back to immune homoeostasis. B-cell adapter for PI3 K (BCAP) promotes the transition of macrophages from an early inflammatory M1 signature to a late reparative M2 profile by elevating lactate production that is translated into enhanced histone lactylation and expression of reparative macrophage genes including forkhead box O1 (Foxo1) and glycogen synthase kinase-3β (Gsk3β) (Ref. Reference Irizarry-Caro38). Lactate promotes homoeostatic macrophage polarisation by transcriptionally modifying the expression of mitochondrial antiviral-signalling protein and thus inhibiting pro-inflammatory interferon-mediated signalling (Ref. Reference Ivashkiv39). Lactate-derived histone lysine lactylation (including H3K4, H3K18, H4K5, H4 K) induces the expression of homoeostatic genes such as arginase 1 (Arg1), which is highly expressed and secreted by the immunosuppressive myeloid-derived suppressor cells (MDSC) and TAMs during the transition of macrophages from the M1 to the M2-like phenotype (Ref. Reference Zhang4). Other view considers lysine lactylation as a consequence rather than a cause of macrophage activation co-occurring incidentally with Arg1-dependent metabolic rewiring under inflammation (Ref. Reference Dichtl40).

While high lactate and low pH in inflamed tissues under hypoxia is a condition beneficial to pathogen clearance by confining T cells to the inflammatory site, it is harmful during tumour-associated inflammation via suppressing the cytolytic function of CD8+ T cells or inducing the Th17 phenotype of CD4+ T cells (Ref. Reference Siska41). TAM exhibits the M1 proinflammatory phenotype with the anti-tumour activity at the tumour initiation stage, and skews to the M2 phenotype during cancer progression (Ref. Reference Wynn, Chawla and Pollard37). Lactate, under hypoxia, mediates the immunosuppressive effects of efferocytosis by inducing the expression of anti-inflammatory genes (Ref. Reference Ivashkiv39).

Histone lactylation and immunosurveillance evasion

Cancers are featured by the ability of evading immunosurveillance (Ref. Reference Hanahan and Weinberg34). Lactate in the TME has been shown to aid tumour cells in escaping immune surveillance by remodelling T cells and macrophages into the immunosuppressive phenotypes such as tumour-promoting Tregs and M2-like TAMs (Ref. Reference Dichtl40). Lactate adversely affects the recruitment of CTLs into the TME via suppressing their proliferation, function and movement (Ref. Reference Brand42), and induces the apoptosis of natural killer (NK) cells (Ref. Reference Harmon43). Lactate also inhibits cytokine production, and thereby reduces the cytotoxic effect (Ref. Reference Hua7). Besides, tumour-derived lactate helps malignant cells evade immunosurveillance by suppressing the antigen presentation ability of dendritic cells (Ref. Reference Gottfried44), and promotes the development of MDSC that suppresses the innate and adaptive immunities (Ref. Reference Husain45).

Histone lactylation and genome instability

Little evidence has been reported so far on the direct association between histone lactylation and genome instability. Gate keepers such as p53, once perturbed, may result in accelerated genome mutation that ultimately leads to cancer initiation and the evolve of all other cancer hallmarks.

Lactagenesis has been shown to be orchestrated by genetic mutations, e.g., over-represented expression of genes encoding hypoxia inducible factor 1 subunit alpha (Hif-1α) or cellular MYC (c-Myc) is associated with decreased mitochondrial function and elevated LDHA level (Ref. Reference Chen30). Low p53 and high LDHA expression are associated with poor breast cancer overall survival, with demonstrated regulatory role of p53 on LDHA being reported (Ref. Reference Zhou46). In addition, the modulatory functionality of p53 on other critical factors involved in lactylation and lactate production such as MCT1 has been documented (Ref. Reference Boidot47). Thus, it is possible that lactylation is the consequence but not the cause of genome instability. However, we cannot exclude the possibility that lactylation contributes to genome instability by perturbing the transcriptome of cancer driver genes, given the crosstalk of lactylation with other epigenetic coding such as acetylation on histones. There also exists the possibility that lactylation occurs on DNA/RNA sequences besides proteins, similar to what we have witnessed on the discovery of mRNA acetylation (Ref. Reference Liu48).

Histone lactylation and cancer growth

Lactate functions through monocarboxylic acid transporters such as MCT1/4 and G protein-coupled receptors such as GPR81 (also known as hydroxycarboxylic acid receptor 1) (Refs Reference Moussaieff49, Reference Hadzic50). In particular, MCT4 is primarily expressed in highly glycolytic cells such as white muscle fibres to facilitate lactate export in response to hypoxia (Ref. Reference Luo51), and MCT1 is predominantly present in red muscle fibres to consume secreted lactate for further oxidation (Ref. Reference Garcia52) (Fig. 2); GPR81 signalling is adopted by bone marrow-derived inflammatory neutrophils for lactate release (Ref. Reference Khatib-Massalha53).

Most hallmarks of cancers are relevant to the mechanism of malignant cells toward uncontrolled growth or resistance to death. Aberrant lactate production can foster cells with this hallmark. For example, tumour-produced lactate is eliminated by deleting MCT1 in lung cancer cells (Ref. Reference Faubert54), the growth of leukaemia cells is arrested by inhibiting MCT1 (Refs Reference Pivovarova and MacGregor55, Reference Saulle56) or MCT4 (Ref. Reference Saulle56), and the proliferation of invasive bladder cancer cells is arrested through selective inhibition of MCT4 (Ref. Reference Todenhofer57). Cancer-produced lactate activates GPR81 (Ref. Reference Xie58), and GPR81 deletion halts breast cancer growth both in vitro and in vivo (Ref. Reference Brown59), suggesting the promotive role of lactate on breast cancer proliferation.

Histone lactylation and tumour angiogenesis/metastasis

Most cancer-associated death events are caused by tumour metastasis where tumour angiogenesis prepares the network of blood vessels supplying tumours with a supportive microenvironment toward local or distant metastasis, both of which are basic cancer hallmarks (Ref. Reference Hanahan and Weinberg34).

Lactate (especially tumour-derived lactate) is an angiogenesis promoter that participates in angiogenesis via stabilizing HIF1α (Refs Reference Haaga and Haaga60, Reference Vallee, Guillevin and Vallee61, Reference Lu, Forbes and Verma62), and triggers TAM polarisation toward the M2 state that is pro-angiogenic accompanied with over-represented expression of Arg1 and Vegf (Ref. Reference Colegio35). Lactate can induce the expression of another pro-angiogenic factor, interleukin-8, which sustains new blood vessels maturation during tumour angiogenesis (Ref. Reference Vegran63).

Lactate, in the TME, was reported to be capable of ameliorating conjugations with the extracellular matrix components to enable subsequent cancer cell migration by adjusting the binding of integrins on tumour cells (Ref. Reference Webb64). Decreased extracellular pH as a result of lactagenesis further facilitates the motility and invasiveness of cancer cells (Ref. Reference Busco65). Increased lactate levels exhibit a positive correlation with amplified metastatic potential in various human primary carcinomas (Ref. Reference Walenta and Mueller-Klieser66). For example, lactate triggers Tgfβ2 expression in glioma cells that activates matrix metalloproteinase-2 (Ref. Reference Baumann67), and elevates Klhdc8a expression that enhances the proliferation, migration and invasion of high-grade glioma cells (Ref. Reference Zhu68). A positive correlation has been established between cold atmospheric plasma (CAP)-induced lactate addiction and activated epithelial-mesenchymal transition in prostate cancer cells (Ref. Reference Ippolito69). Supplementing exogenous lactate to cancer cells enhances the motility of different tumour cells (Ref. Reference Goetze70), and promotes the migration of endothelial cells.

Histone lactylation and onco-therapy resistance

Malignant cells may develop chemo- or radio- resistance against onco-therapeutics which is typically associated with acquired cancer stemness during the course of treatment (Refs Reference Yue71, Reference Lopez-Menendez72). Targeting glycolysis together with existing therapeutics has been proposed to overcome therapeutic resistance such as in the treatment of melanoma (Ref. Reference Cascone73).

Lactate substantially contributes to TME acidification and cancer cell drug resistance, as many drugs are weak bases that can be easily impaired by the acid milieu (Ref. Reference Taylor74). Specifically, charged molecules cannot freely penetrate through cell membrane; thus, the acidic TME hampers the cellular uptake of weak base drugs such as anthracyclines, anthracenediones, campothecins, Vinca alkaloids (Refs Reference Adar75, Reference De Milito and Fais76, Reference Ellegaard77, Reference Jansen78, Reference Mahoney79, Reference Wojtkowiak80), and complex drugs such as cisplatin (Ref. Reference Federici81). Lactate contributes to the establishment of resistance to epithelial growth factor receptor (EGFR) tyrosine kinase inhibitors in cancer cells (Ref. Reference Apicella82). G-protein coupled receptor 1 (GPR1), a receptor of lactate, is associated with lactate-induced chemoresistance in hepatic cancer cells (Ref. Reference Soni83). Besides, immunotherapies, among other approaches, may lack therapeutic efficacies despite their recognised immense potential in killing cancer cells given the immunosuppressive role of lactate and the acidic TME it fosters.

Lactate also triggers radio-resistance in many types of cancers (Ref. Reference Tang84) due to its demonstrated anti-oxidant properties (Ref. Reference Taddei85). For instance, lactate concentration is positively correlated with the resistance of human head and neck squamous cell carcinoma (HNSCC) to fractioned irradiation (Ref. Reference Sattler86), and lactate dehydrogenase 5 over-representation is associated with the radio-resistance of HNSCC (Ref. Reference Koukourakis87), colorectal (Ref. Reference Koukourakis88) and prostate (Ref. Reference Koukourakis89) cancers. Additionally, high lactate concentration abrogates the sensitivity of cancer cells to oxidative stress toward the evolvement of resistance to hydrogen peroxide, high-dose ascorbate and photodynamic therapy (Ref. Reference Koncosova90).

Discussion

Non-histone lactylation

Similar to other epigenomic codes such as acetylation, lactylation may also occur in non-histone proteins. A global lysine lactylome analysis was conducted in Botrytis cinerea (a fungal pathogen) by LC-MS/MS, where 273 lysine lactylation were identified from 166 proteins. Among these proteins, 36% are distributed in the nucleus, 27% in the mitochondria, and 25% in the cytoplasm (Ref. Reference Gao, Zhang and Liang91). Several proteins with critical functionalities in cancers can be lactylated such as MAPK lactylation at K60. These evidences are suggestive of the prevalence and wide-spread roles of lysine lactylation in cells at both the healthy and abnormal states, and non-histone lactylation that has attracted relatively less attention may represent a future direction deserving intensive investigations.

Crosstalk between lactylation and other epigenomic events

Lactate associates lactylation with other epigenetic codes such as methylation given its pivotal roles in epigenomic reprogramming (Ref. Reference Bhagat92). For instance, tumour-derived lactate promotes α-ketoglutarate (α-KG) production that activates the demethylase TET, resulting in decreased cytosine methylation and enhanced hydroxymethylation during the differentiation of mesenchymal stem cells (MSC) to CAFs. TET proteins are dioxygenases for DNA hydroxymethylation (Ref. Reference Zhang93). 2-Hydroglutarate (2HG) exists in two enantiomers, i.e., R-2HG and S-2HG, both of which inhibit α-KG-dependent dioxygenases including TETs (Ref. Reference Sulkowski94). While R-2HG is an oncometabolite generated from α-KG, S-2HG is generated by LDH or malate dehydrogenase under hypoxia (Ref. Reference Nadtochiy95). These create a negative feedback loop involving LDH, lactate, α-KG, 2HG and TET that collectively orchestrate the crosstalk between lactylation and methylation.

In addition, lactate has been shown to be an endogenous histone deacetylase (HDAC) inhibitor toward enhanced expression of genes associated with HDAC proteins (Refs Reference Moussaieff49, Reference Genders96). Of particular relevance, is the similarity and coordination between lactylation and acetylation. Both types of epigenetic codes prefer lysine and share the use of some enzymes, e.g., p300 as the writer (Refs Reference Chen30, Reference Wang32). Interestingly, p300 is highly enriched in the promoter regions of pluripotency genes such as Oct4, Sall4, c-Myc during reprogramming, suggesting the coordinated roles of lactylation and acetylation as driven by fluctuations between lactate and acetyl-CoA during cell decision making (Refs Reference Chisolm and Weinmann97, Reference Dai98). Besides, histone lactylation can affect RNA modifications and contribute to tumorigenesis. For example, histone lactylation facilitates the expression of genes encoding YTH N6-methyladenosine RNA binding protein 2 (Ythdf2) that recognises m6A-modified period circadian regulator 1 (Per1) and p53 mRNAs (two tumour suppressors) for degradation in ocular melanoma (Ref. Reference Yu5). It is possible that enhanced lactate production in cancer cells as a result of aberrant metabolic reprogramming leads to a higher concentration of lactyl-CoA than acetyl-CoA that drives the epigenetic modification at a certain histone lysine site toward lactylation rather than acetylation; and this will lead to a higher level of histone lactylation than acetylation in cancer cells and accelerated carcinogenesis. Alternatively, co-enzymes of lactylation or acetylation may exist to discriminate the use of lactyl- and acetyl-CoA for the epigenetic modification of a certain site and determine the levels of lactylation and acetylation. Yet, these are all hypotheses that require experimental validations.

Onco-therapeutic opportunities via targeting lactylation

Given accumulated evidence on the positive association of lactylation with carcinogenesis, lactate production or lactate transporters such as LDHA, MCT1/4 and GPR1 have been proposed as novel onco-therapeutic targets, alone or in combination with other anti-cancer strategies (Refs Reference Siska41, Reference Taddei85, Reference Feng99, Reference Feichtinger and Lang100) (Table 1). For instance, LDHA has been proposed as an oncotarget alone or through creating synergies with redox-sensitive onco-therapies in a p53/NAD(H)-dependent manner (Ref. Reference Allison101). Also, targeting the lactate axis can abolish the resistance of cancer cells to EGFR tyrosine kinase inhibitors in vivo (Ref. Reference Apicella82). Curcumin (Ref. Reference Soni83) and LRH7-G5 (Ref. Reference Huang102), via targeting GPR1, can restore tumour cells' sensitivity to chemotherapies. AZD3956, a drug that targets MCT1, is currently under the clinical trial (NCT01791595).

Table 1. Onco-therapeutic approaches targeting the lactate axis

NA represents ‘not available’.

On the other hand, the pyruvate dehydrogenase complex (PDC), catalysing the conversion of pyruvate to acetyl-CoA and playing key roles in histone acetylation (Ref. Reference de Boer and Houten103), is associated with elevated histone lactylation once inhibited (Ref. Reference Zhang4). Pyruvate dehydrogenase kinases (PDKs) are specific kinases of PDCs suppressing their activities. PDK inhibitors such as dichloroacetate (DCA) or diisopropylamine dichloroacetate (DATA) reduce lactate production and histone typrosine lactylation levels that leads to enhanced radio-sensitivity of oesophageal squamous cell carcinoma cells, non-small cell lung cancer cells, glioblastoma cells and breast cancer cells (Refs Reference Dong104, Reference Bonnet105). While the efficacy of DCA as an onco-therapeutic strategy has already been examined in a phase II clinical trial (Ref. Reference Powell106), a superior efficacy was reported for DADA using a breast cancer in vivo model (Ref. Reference Su107).

Many targeted therapies and immunotherapies fail due to the evolved cancer cell resistance, which triggers the development of duel-targeting strategies such as the combined use of duvelisib and rituximab in the treatment of chronic lymphocytic leukaemia (Ref. Reference Davids108) and the aforementioned combinatorial strategies involving targets of the lactate axis. Despite its great promise, dual-targeting still relies on limited receptor-mediated signalling and does not represent the ultimate option for cancer cure. The call for emerging onco-therapeutics from a novel perspective is thus imperative and timely. CAP, being an emerging onco-therapeutic approach against cancer cells with multi-modality nature, does not rely on any single receptor or pathway to take on actions (Ref. Reference Dai109). Among the many evidences demonstrating its selectivity against cancer cells, CAP was shown capable of suppressing LDHA and creating synergies with other drugs toward enhanced anti-cancer efficacy. Thus, onco-therapeutic strategies taking advantages of the glycolysis switch as represented by lactylation and acetylation with the aid of CAP may shift the paradigm of anti-cancer investigations into an innovative era that possibly leads to the ultimate cure of cancer.

Conclusion

Studies on lactylation and its clinical impact are in its infancy. Comprehensively delineating the landscape of how lactylation coordinates with other epigenetic codes toward reprogrammed cell metabolism and rewired fates is urgently needed toward effective methodological design against cancers. These include investigations on non-histone lactylation, novel functionalities they represent (and in particular during cancer initiation and progression), as well as specific writers, erasers and readers that may involve.

Unlike the other layers of epigenetic coding marks that largely play dual roles in cancer initiation and progression, all evidence on lactylation so far reported have associated it with the oncogenic role. Yet, strategies targeting lactylation and their clinical translation are still at the infant stage. Thus, novel onco-therapeutic approaches taking advantages of lactylation are urgently called for to resolve tumours or rewire drug resistant malignant cells toward the sensitive state, which represent an encouraging trend in the future.

Authors’ contribution

X. Dai conceptualised the idea, drafted the manuscript, and finalised the figures. X. Lv contributed in literature searching and figure preparation. X. Dai provided the financial support.

Financial support

This study was funded by the National Natural Science Foundation of China (Grant No. 81972789), Fundamental Research Funds for the Central Universities (Grant No. JUSRP22011), Technology Development Funding of Wuxi (Grant No. WX18IVJN017).

Conflict of interest

The authors declare no competing interest.

Footnotes

*

These authors contribute equally to this work.

References

de la Cruz Munoz-Centeno, M, Millan-Zambrano, G and Chavez, S (2012) A matter of packaging: influence of nucleosome positioning on heterologous gene expression. Methods in Molecular Biology 824, 5164.CrossRefGoogle ScholarPubMed
Mikkelsen, TS et al. (2007) Genome-wide maps of chromatin state in pluripotent and lineage-committed cells. Nature 448, 553560.CrossRefGoogle Scholar
Bernstein, BE, Meissner, A and Lander, ES (2007) The mammalian epigenome. Cell 128, 669681.CrossRefGoogle ScholarPubMed
Zhang, D et al. (2019) Metabolic regulation of gene expression by histone lactylation. Nature 574, 575580.CrossRefGoogle ScholarPubMed
Yu, J et al. (2021) Histone lactylation drives oncogenesis by facilitating m(6)A reader protein YTHDF2 expression in ocular melanoma. Genome Biology 22, 85.CrossRefGoogle ScholarPubMed
Jiang, J et al. (2021) Lactate modulates cellular metabolism through histone lactylation-mediated gene expression in non-small cell lung cancer. Frontiers in Oncology 11, 647559.CrossRefGoogle ScholarPubMed
Hua, G et al. (2014) Targeting glucose metabolism in chondrosarcoma cells enhances the sensitivity to doxorubicin through the inhibition of lactate dehydrogenase-A. Oncology Reports 31, 27272734.CrossRefGoogle ScholarPubMed
Cui, H et al. (2021) Lung myofibroblasts promote macrophage profibrotic activity through lactate-induced histone lactylation. American Journal of Respiratory Cell and Molecular Biology 64, 115125.CrossRefGoogle ScholarPubMed
Meng, X et al. (2021) Comprehensive analysis of lysine lactylation in rice (Oryza sativa) grains. Journal of Agricultural and Food Chemistry 69, 82878297.CrossRefGoogle ScholarPubMed
Nikooie, R, Moflehi, D and Zand, S (2021) Lactate regulates autophagy through ROS-mediated activation of ERK1/2/m-TOR/p-70S6K pathway in skeletal muscle. Journal of Cell Communication and Signaling 15, 107123.CrossRefGoogle ScholarPubMed
Matejovic, M, Radermacher, P and Fontaine, E (2007) Lactate in shock: a high-octane fuel for the heart? Intensive Care Medicine 33, 406408.CrossRefGoogle Scholar
Dienel, GA (2014) Lactate shuttling and lactate use as fuel after traumatic brain injury: metabolic considerations. Journal of Cerebral Blood Flow & Metabolism 34, 17361748.CrossRefGoogle ScholarPubMed
Bonuccelli, G et al. (2010) Ketones and lactate “fuel” tumor growth and metastasis: evidence that epithelial cancer cells use oxidative mitochondrial metabolism. Cell Cycle 9, 35063514.CrossRefGoogle ScholarPubMed
Haas, R et al. (2016) Intermediates of metabolism: from bystanders to signalling molecules. Trends in Biochemical Sciences 41, 460471.CrossRefGoogle ScholarPubMed
Lee, TY (2021) Lactate: a multifunctional signaling molecule. Yeungnam University Journal of Medicine 38, 183193.CrossRefGoogle ScholarPubMed
Ahmed, K et al. (2010) An autocrine lactate loop mediates insulin-dependent inhibition of lipolysis through GPR81. Cell Metabolism 11, 311319.CrossRefGoogle ScholarPubMed
Weinreich, J et al. (2011) Rapamycin-induced impaired wound healing is associated with compromised tissue lactate accumulation and extracellular matrix remodeling. European Surgical Research 47, 3944.CrossRefGoogle ScholarPubMed
Jha, MK and Morrison, BM (2020) Lactate transporters mediate Glia-neuron metabolic crosstalk in homeostasis and disease. Frontiers in Cellular Neuroscience 14, 589582.CrossRefGoogle ScholarPubMed
Jain, M et al. (2020) A D-lactate dehydrogenase from rice is involved in conferring tolerance to multiple abiotic stresses by maintaining cellular homeostasis. Scientific Reports 10, 12835.CrossRefGoogle Scholar
Vaupel, P and Multhoff, G (2021) The Warburg effect: historical dogma versus current rationale. Advances in Experimental Medicine and Biology 1269, 169177.CrossRefGoogle ScholarPubMed
Chen, J et al. (2020) Warburg effect is a cancer immune evasion mechanism against macrophage immunosurveillance. Frontiers in Immunology 11, 621757.CrossRefGoogle ScholarPubMed
Kes, MMG et al. (2020) Oncometabolites lactate and succinate drive pro-angiogenic macrophage response in tumors. Biochimica et Biophysica Acta, Reviews on Cancer 1874, 188427.CrossRefGoogle ScholarPubMed
Vlachostergios, PJ et al. (2015) Elevated lactic acid is a negative prognostic factor in metastatic lung cancer. Cancer Biomarkers: Section A of Disease Markers 15, 725734.CrossRefGoogle ScholarPubMed
Sonveaux, P et al. (2012) Targeting the lactate transporter MCT1 in endothelial cells inhibits lactate-induced HIF-1 activation and tumor angiogenesis. PLoS ONE 7, e33418.CrossRefGoogle ScholarPubMed
Cheung, SM et al. (2020) Lactate concentration in breast cancer using advanced magnetic resonance spectroscopy. British Journal of Cancer 123, 261267.CrossRefGoogle ScholarPubMed
Hui, S et al. (2017) Glucose feeds the TCA cycle via circulating lactate. Nature 551, 115118.CrossRefGoogle ScholarPubMed
Nelson, SJ et al. (2013) Metabolic imaging of patients with prostate cancer using hyperpolarized [1-(1)(3)C]pyruvate. Science Translational Medicine 5, 198–108.CrossRefGoogle Scholar
San-Millan, I and Brooks, GA (2017) Reexamining cancer metabolism: lactate production for carcinogenesis could be the purpose and explanation of the Warburg effect. Carcinogenesis 38, 119133.Google ScholarPubMed
Izzo, LT and Wellen, KE (2019) Histone lactylation links metabolism and gene regulation. Nature 574, 492493.CrossRefGoogle ScholarPubMed
Chen, AN et al. (2021) Lactylation, a novel metabolic reprogramming code: current Status and prospects. Frontiers in Immunology 12, 688910.CrossRefGoogle ScholarPubMed
Varner, EL et al. (2020) Quantification of lactoyl-CoA (lactyl-CoA) by liquid chromatography mass spectrometry in mammalian cells and tissues. Open Biology 10, 200187.CrossRefGoogle ScholarPubMed
Wang, T et al. (2021) Acetylation of lactate dehydrogenase B drives NAFLD progression by impairing lactate clearance. Journal of Hepatology 74, 10381052.CrossRefGoogle ScholarPubMed
Hanahan, D and Weinberg, RA (2000) The hallmarks of cancer. Cell 100, 5770.CrossRefGoogle ScholarPubMed
Hanahan, D and Weinberg, RA (2011) Hallmarks of cancer: the next generation. Cell 144, 646674.CrossRefGoogle ScholarPubMed
Colegio, OR et al. (2014) Functional polarization of tumour-associated macrophages by tumour-derived lactic acid. Nature 513, 559563.CrossRefGoogle ScholarPubMed
Geissmann, F et al. (2010) Development of monocytes, macrophages, and dendritic cells. Science 327, 656661.CrossRefGoogle ScholarPubMed
Wynn, TA, Chawla, A and Pollard, JW (2013) Macrophage biology in development, homeostasis and disease. Nature 496, 445455.CrossRefGoogle ScholarPubMed
Irizarry-Caro, RA et al. (2020) TLR signaling adapter BCAP regulates inflammatory to reparatory macrophage transition by promoting histone lactylation. Proceedings of the National Academy of Sciences of the USA 117, 3062830638.CrossRefGoogle ScholarPubMed
Ivashkiv, LB (2020) The hypoxia-lactate axis tempers inflammation. Nature Reviews Immunology 20, 8586.CrossRefGoogle ScholarPubMed
Dichtl, S et al. (2021) Lactate and IL6 define separable paths of inflammatory metabolic adaptation. Science Advances 7, 3505.CrossRefGoogle ScholarPubMed
Siska, PJ et al. (2020) The immunological Warburg effect: can a metabolic-tumor-stroma score (MeTS) guide cancer immunotherapy? Immunological Reviews 295, 187202.CrossRefGoogle ScholarPubMed
Brand, A et al. (2016) LDHA-Associated lactic acid production blunts tumor immunosurveillance by T and NK cells. Cell Metabolism 24, 657671.CrossRefGoogle Scholar
Harmon, C et al. (2019) Lactate-mediated acidification of tumor microenvironment induces apoptosis of liver-resident NK cells in colorectal liver metastasis. Cancer Immunology Research 7, 335346.CrossRefGoogle ScholarPubMed
Gottfried, E et al. (2006) Tumor-derived lactic acid modulates dendritic cell activation and antigen expression. Blood 107, 20132021.CrossRefGoogle ScholarPubMed
Husain, Z et al. (2013) Tumor-derived lactate modifies antitumor immune response: effect on myeloid-derived suppressor cells and NK cells. Journal of Immunology 191, 14861495.CrossRefGoogle ScholarPubMed
Zhou, Y et al. (2019) P53/lactate dehydrogenase A axis negatively regulates aerobic glycolysis and tumor progression in breast cancer expressing wild-type p53. Cancer Science 110, 939949.CrossRefGoogle ScholarPubMed
Boidot, R et al. (2012) Regulation of monocarboxylate transporter MCT1 expression by p53 mediates inward and outward lactate fluxes in tumors. Cancer Research 72, 939948.CrossRefGoogle ScholarPubMed
Liu, CL et al. (2009) Lactate inhibits lipolysis in fat cells through activation of an orphan G-protein-coupled receptor, GPR81. Journal of Biological Chemistry 284, 28112822.CrossRefGoogle ScholarPubMed
Moussaieff, A et al. (2015) Glycolysis-mediated changes in acetyl-CoA and histone acetylation control the early differentiation of embryonic stem cells. Cell Metabolism 21, 392402.CrossRefGoogle ScholarPubMed
Hadzic, A et al. (2020) The lactate receptor HCA1 Is present in the choroid plexus, the Tela Choroidea, and the Neuroepithelial lining of the dorsal part of the third ventricle. International Journal of Molecular Sciences 21, 6451.CrossRefGoogle ScholarPubMed
Luo, F et al. (2017) Enhanced glycolysis, regulated by HIF-1alpha via MCT-4, promotes inflammation in arsenite-induced carcinogenesis. Carcinogenesis 38, 615626.CrossRefGoogle ScholarPubMed
Garcia, CK et al. (1994) Molecular characterization of a membrane transporter for lactate, pyruvate, and other monocarboxylates: implications for the Cori cycle. Cell 76, 865873.CrossRefGoogle ScholarPubMed
Khatib-Massalha, E et al. (2020) Lactate released by inflammatory bone marrow neutrophils induces their mobilization via endothelial GPR81 signaling. Nature Communications 11, 3547.CrossRefGoogle ScholarPubMed
Faubert, B et al. (2017) Lactate metabolism in human lung tumors. Cell 171, 358371. e9.CrossRefGoogle ScholarPubMed
Pivovarova, AI and MacGregor, GG (2018) Glucose-dependent growth arrest of leukemia cells by MCT1 inhibition: feeding Warburg's sweet tooth and blocking acid export as an anticancer strategy. Biomedicine & Pharmacotherapy 98, 173179.CrossRefGoogle ScholarPubMed
Saulle, E et al. (2020) Targeting lactate metabolism by inhibiting MCT1 or MCT4 impairs leukemic cell proliferation, induces two different related death-pathways and increases chemotherapeutic sensitivity of acute myeloid leukemia cells. Frontiers in Oncology 10, 621458.CrossRefGoogle ScholarPubMed
Todenhofer, T et al. (2018) Selective inhibition of the lactate transporter MCT4 reduces growth of invasive bladder cancer. Molecular Cancer Therapeutics 17, 27462755.CrossRefGoogle ScholarPubMed
Xie, Q et al. (2020) A lactate-induced snail/STAT3 pathway drives GPR81 expression in lung cancer cells. Biochimica et Biophysica Acta, Molecular Basis of Disease 1866, 165576.CrossRefGoogle ScholarPubMed
Brown, TP et al. (2020) The lactate receptor GPR81 promotes breast cancer growth via a paracrine mechanism involving antigen-presenting cells in the tumor microenvironment. Oncogene 39, 32923304.CrossRefGoogle Scholar
Haaga, JR and Haaga, R (2013) Acidic lactate sequentially induced lymphogenesis, phlebogenesis, and arteriogenesis (ALPHA) hypothesis: lactate-triggered glycolytic vasculogenesis that occurs in normoxia or hypoxia and complements the traditional concept of hypoxia-based vasculogenesis. Surgery 154, 632637.CrossRefGoogle ScholarPubMed
Vallee, A, Guillevin, R and Vallee, JN (2018) Vasculogenesis and angiogenesis initiation under normoxic conditions through Wnt/beta-catenin pathway in gliomas. Reviews in the Neurosciences 29, 7191.CrossRefGoogle ScholarPubMed
Lu, H, Forbes, RA and Verma, A (2002) Hypoxia-inducible factor 1 activation by aerobic glycolysis implicates the Warburg effect in carcinogenesis. Journal of Biological Chemistry 277, 2311123115.CrossRefGoogle ScholarPubMed
Vegran, F et al. (2011) Lactate influx through the endothelial cell monocarboxylate transporter MCT1 supports an NF-kappaB/IL-8 pathway that drives tumor angiogenesis. Cancer Research 71, 25502560.CrossRefGoogle ScholarPubMed
Webb, BA et al. (2011) Dysregulated pH: a perfect storm for cancer progression. Nature Reviews Cancer 11, 671677.CrossRefGoogle Scholar
Busco, G et al. (2010) NHE1 Promotes invadopodial ECM proteolysis through acidification of the peri-invadopodial space. FASEB Journal 24, 39033915.CrossRefGoogle ScholarPubMed
Walenta, S and Mueller-Klieser, WF (2004) Lactate: mirror and motor of tumor malignancy. Seminars in Radiation Oncology 14, 267274.CrossRefGoogle ScholarPubMed
Baumann, F et al. (2009) Lactate promotes glioma migration by TGF-beta2-dependent regulation of matrix metalloproteinase-2. Neuro-Oncology 11, 368380.CrossRefGoogle ScholarPubMed
Zhu, X et al. (2020) Lactate induced up-regulation of KLHDC8A (Kelch domain-containing 8A) contributes to the proliferation, migration and apoptosis of human glioma cells. Journal of Cellular and Molecular Medicine 24, 1169111702.CrossRefGoogle Scholar
Ippolito, L et al. (2019) Cancer-associated fibroblasts promote prostate cancer malignancy via metabolic rewiring and mitochondrial transfer. Oncogene 38, 53395355.CrossRefGoogle ScholarPubMed
Goetze, K et al. (2011) Lactate enhances motility of tumor cells and inhibits monocyte migration and cytokine release. International Journal of Oncology 39, 453463.Google ScholarPubMed
Yue, J et al. (2021) LncRNAs link cancer stemness to therapy resistance. American Journal of Cancer Research 11, 10511068.Google ScholarPubMed
Lopez-Menendez, C et al. (2021) E2A modulates stemness, metastasis, and therapeutic resistance of breast cancer. Cancer Research 81, 45294544.CrossRefGoogle ScholarPubMed
Cascone, T et al. (2018) Increased tumor glycolysis characterizes immune resistance to adoptive T cell therapy. Cell Metabolism 27, 977987. e4.CrossRefGoogle ScholarPubMed
Taylor, S et al. (2015) Microenvironment acidity as a major determinant of tumor chemoresistance: proton pump inhibitors (PPIs) as a novel therapeutic approach. Drug Resistance Updates: Reviews and Commentaries in Antimicrobial and Anticancer Chemotherapy 23, 6978.CrossRefGoogle Scholar
Adar, Y et al. (2012) Imidazoacridinone-dependent lysosomal photodestruction: a pharmacological Trojan horse approach to eradicate multidrug-resistant cancers. Cell Death & Disease 3, e293.CrossRefGoogle ScholarPubMed
De Milito, A and Fais, S (2005) Tumor acidity, chemoresistance and proton pump inhibitors. Future Oncology 1, 779786.CrossRefGoogle ScholarPubMed
Ellegaard, AM et al. (2013) Sunitinib and SU11652 inhibit acid sphingomyelinase, destabilize lysosomes, and inhibit multidrug resistance. Molecular Cancer Therapeutics 12, 20182030.CrossRefGoogle ScholarPubMed
Jansen, G et al. (1999) Multiple mechanisms of resistance to polyglutamatable and lipophilic antifolates in mammalian cells: role of increased folylpolyglutamylation, expanded folate pools, and intralysosomal drug sequestration. Molecular Pharmacology 55, 761769.Google ScholarPubMed
Mahoney, BP et al. (2003) Tumor acidity, ion trapping and chemotherapeutics. I. Acid pH affects the distribution of chemotherapeutic agents in vitro. Biochemical Pharmacology 66, 12071218.CrossRefGoogle ScholarPubMed
Wojtkowiak, JW et al. (2011) Drug resistance and cellular adaptation to tumor acidic pH microenvironment. Molecular Pharmaceutics 8, 20322038.CrossRefGoogle ScholarPubMed
Federici, C et al. (2014) Exosome release and low pH belong to a framework of resistance of human melanoma cells to cisplatin. PLoS ONE 9, e88193.CrossRefGoogle Scholar
Apicella, M et al. (2018) Increased lactate secretion by cancer cells sustains non-cell-autonomous adaptive resistance to MET and EGFR targeted therapies. Cell Metabolism 28, 848865. e6.CrossRefGoogle ScholarPubMed
Soni, VK et al. (2020) Curcumin circumvent lactate-induced chemoresistance in hepatic cancer cells through modulation of hydroxycarboxylic acid receptor-1. International Journal of Biochemistry & Cell Biology 123, 105752.CrossRefGoogle ScholarPubMed
Tang, L et al. (2018) Role of metabolism in cancer cell radioresistance and radiosensitization methods. Journal of Experimental & Clinical Cancer Research: CR 37, 87.CrossRefGoogle ScholarPubMed
Taddei, ML et al. (2020) Lactate in sarcoma microenvironment: much more than just a waste product. Cells 9, 102109.CrossRefGoogle ScholarPubMed
Sattler, UG et al. (2010) Glycolytic metabolism and tumour response to fractionated irradiation. Radiotherapy & Oncology 94, 102109.CrossRefGoogle ScholarPubMed
Koukourakis, MI et al. (2009) Lactate dehydrogenase 5 expression in squamous cell head and neck cancer relates to prognosis following radical or postoperative radiotherapy. Oncology 77, 285292.CrossRefGoogle ScholarPubMed
Koukourakis, MI et al. (2011) Prognostic and predictive role of lactate dehydrogenase 5 expression in colorectal cancer patients treated with PTK787/ZK 222584 (vatalanib) antiangiogenic therapy. Clinical Cancer Research 17, 48924900.CrossRefGoogle ScholarPubMed
Koukourakis, MI et al. (2014) Lactate dehydrogenase 5 isoenzyme overexpression defines resistance of prostate cancer to radiotherapy. British Journal of Cancer 110, 22172223.CrossRefGoogle ScholarPubMed
Koncosova, M et al. (2021) Inhibition of mitochondrial metabolism leads to selective eradication of cells adapted to acidic microenvironment. International Journal of Molecular Sciences 22, 594743.CrossRefGoogle ScholarPubMed
Gao, M, Zhang, N and Liang, W (2020) Systematic analysis of lysine lactylation in the plant fungal pathogen Botrytis cinerea. Frontiers in Microbiology 11, 594743.CrossRefGoogle ScholarPubMed
Bhagat, TD et al. (2019) Lactate-mediated epigenetic reprogramming regulates formation of human pancreatic cancer-associated fibroblasts. Elife 8, 5091.CrossRefGoogle ScholarPubMed
Zhang, M et al. (2021) Ten-eleven translocation 1 mediated-DNA hydroxymethylation is required for myelination and remyelination in the mouse brain. Nature Communications 12, 5091.CrossRefGoogle ScholarPubMed
Sulkowski, PL et al. (2017) 2-Hydroxyglutarate produced by neomorphic IDH mutations suppresses homologous recombination and induces PARP inhibitor sensitivity. Science Translational Medicine 9, 2018820197.CrossRefGoogle ScholarPubMed
Nadtochiy, SM et al. (2016) Acidic pH is a metabolic switch for 2-hydroxyglutarate generation and signaling. Journal of Biological Chemistry 291, 2018820197.CrossRefGoogle ScholarPubMed
Genders, AJ et al. (2019) A physiological drop in pH decreases mitochondrial respiration, and HDAC and Akt signaling, in L6 myocytes. American Journal of Physiology. Cell Physiology 316, C404C414.CrossRefGoogle ScholarPubMed
Chisolm, DA and Weinmann, AS (2018) Connections between metabolism and epigenetics in programming cellular differentiation. Annual Review of Immunology 36, 221246.CrossRefGoogle ScholarPubMed
Dai, X et al. (2021) Histone lactylation: epigenetic mark of glycolytic switch. Trends in Genetics 38, 124127.CrossRefGoogle ScholarPubMed
Feng, J et al. (2017) Tumor cell-derived lactate induces TAZ-dependent upregulation of PD-L1 through GPR81 in human lung cancer cells. Oncogene 36, 58295839.CrossRefGoogle ScholarPubMed
Feichtinger, RG and Lang, R (2019) Targeting L-lactate metabolism to overcome resistance to immune therapy of melanoma and other tumor entities. Journal of Oncology 2019, 2084195.CrossRefGoogle ScholarPubMed
Allison, SJ et al. (2014) Identification of LDH-A as a therapeutic target for cancer cell killing via (i) p53/NAD(H)-dependent and (ii) p53-independent pathways. Oncogenesis 3, e102.CrossRefGoogle Scholar
Huang, C et al. (2020) A screened GPR1 peptide exerts antitumor effects on triple-negative breast cancer. Molecular Therapy Oncolytics 18, 602612.CrossRefGoogle ScholarPubMed
de Boer, VC and Houten, SM (2014) A mitochondrial expatriate: nuclear pyruvate dehydrogenase. Cell 158, 910.CrossRefGoogle ScholarPubMed
Dong, G et al. (2016) Diisopropylamine dichloroacetate enhances radiosensitization in esophageal squamous cell carcinoma by increasing mitochondria-derived reactive oxygen species levels. Oncotarget 7, 6817068178.CrossRefGoogle ScholarPubMed
Bonnet, S et al. (2007) A mitochondria-K + channel axis is suppressed in cancer and its normalization promotes apoptosis and inhibits cancer growth. Cancer Cell 11, 3751.CrossRefGoogle ScholarPubMed
Powell, SF et al. (2022) Phase II study of dichloroacetate, an inhibitor of pyruvate dehydrogenase, in combination with chemoradiotherapy for unresected, locally advanced head and neck squamous cell carcinoma. Investigational New Drugs 40, 622633.CrossRefGoogle ScholarPubMed
Su, L et al. (2016) Superior anti-tumor efficacy of diisopropylamine dichloroacetate compared with dichloroacetate in a subcutaneous transplantation breast tumor model. Oncotarget 7, 6572165731.CrossRefGoogle Scholar
Davids, MS et al. (2021) A phase 1b/2 study of duvelisib in combination with FCR (DFCR) for frontline therapy for younger CLL patients. Leukemia 35, 10641072.CrossRefGoogle ScholarPubMed
Dai, X et al. (2018) The emerging role of gas plasma in oncotherapy. Trends in Biotechnology 36, 11831198.CrossRefGoogle ScholarPubMed
Pajak, B et al. (2019) 2-Deoxy-d-glucose and its analogs: from diagnostic to therapeutic agents. International Journal of Molecular Sciences 21(1), 234.CrossRefGoogle ScholarPubMed
Li, W et al. (2017) Benserazide, a dopadecarboxylase inhibitor, suppresses tumor growth by targeting hexokinase 2. Journal of Experimental & Clinical Cancer Research 36(1), 58.CrossRefGoogle ScholarPubMed
Zhang, Q et al. (2014) Hexokinase II inhibitor, 3-BrPA induced autophagy by stimulating ROS formation in human breast cancer cells. Genes and Cancer 5(3-4), 100112.CrossRefGoogle ScholarPubMed
Kim, DJ et al. (2019) Tristetraprolin-mediated hexokinase 2 expression regulation contributes to glycolysis in cancer cells. Molecular Biology of the Cell 30(5), 542553.CrossRefGoogle ScholarPubMed
An, MX et al. (2017) BAG3 directly stabilizes Hexokinase 2 mRNA and promotes aerobic glycolysis in pancreatic cancer cells. Journal of Cell Biology 216(12), 40914105.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1. Lactate metabolism. The main biochemical players in glycolysis and the TCA cycle that participate in lactate metabolism.

Figure 1

Fig. 2. Simplified illustration on associations between lactate and cancer hallmarks, as well as possible therapeutic strategies targeting lactate and lactylation. Lactates, once generated from glycolysis (with LDHA being the enzyme catalysing the last step) and entered in the tumour microenvironment (TME) including tumour, stromal, various types of immune cells, and blood vessels, increases the TME acidity and thus contributes to ‘onco-therapeutic resistance’. Lactates, generated from peripheral tissues such as tumour associated fibroblasts (TAF), enter cells via MCT1/4 and are reutilised toward enhanced glycolysis via lactate shuttle, contributing to ‘metabolic reprogramming’. Lactate accumulation and acidification of the TME are also relevant to ‘tumour-associated inflammation’ by, e.g., stimulating TAM toward M2-like polarisation that is associated with enhanced production of cytokines and chemokines such as IL6 and CCL5, whereas prolonged inflammation triggers altered profiling of oncogenes and tumour suppressors that promote carcinogenesis. Lactates can also aid tumour cells in ‘immunosurveillance evasion by, e.g., suppressing the antigen presentation ability of dendritic cells and triggering apoptosis of NK cells. Accelerated glycolysis toward excess lactate accumulation can cause or be the result of mutations of tumour driver genes such as p53 and HIF-1α, and thus be associated with ‘genome instability’. Lactate functions through receptors such as MCT1/4 and GPR81, the mutations of which halt ‘cancer growth’. Lactate promotes ‘tumour angiogenesis and metastasis’ via, e.g., stabilizing HIF1α, triggering TAM polarisation toward the M2 state that is pro-angiogenic accompanied with over-representation of Arg1 and Vegf, ameliorating conjugations with the extracellular matrix components and increasing TME acidity to enable subsequent cancer cell migration. Regarding therapeutic opportunities, targeting MCT1/4 abolishes the resistance of cancer cells to MET/EGFR tyrosine kinase inhibitors, and cold atmospheric plasma (CAP), an emerging onco-therapeutic strategy, has demonstrated its efficacy in suppressing LDHA. Bold text in caption signifies the possible therapeutic strategies targeting lactate and lactylation.

Figure 2

Table 1. Onco-therapeutic approaches targeting the lactate axis