Next Article in Journal
Metabolic Changes in Pseudomonas oleovorans Isolated from Contaminated Construction Material Exposed to Varied Biocide Treatments
Previous Article in Journal
Gut Microbiota and Sinusoidal Vasoregulation in MASLD: A Portal Perspective
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Crosstalk between Epigenetics and Metabolic Reprogramming in Metabolic Dysfunction-Associated Steatotic Liver Disease-Induced Hepatocellular Carcinoma: A New Sight

1
College of Basic Medical Science, Heilongjiang University of Chinese Medicine, Harbin 150040, China
2
College of Pharmacy, Heilongjiang University of Chinese Medicine, Harbin 150040, China
3
Key Laboratory of Basic and Application Research of Beiyao, Heilongjiang University of Chinese Medicine, Ministry of Education, Harbin 150040, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Metabolites 2024, 14(6), 325; https://doi.org/10.3390/metabo14060325
Submission received: 30 April 2024 / Revised: 1 June 2024 / Accepted: 5 June 2024 / Published: 8 June 2024
(This article belongs to the Section Endocrinology and Clinical Metabolic Research)

Abstract

:
Epigenetic and metabolic reprogramming alterations are two important features of tumors, and their reversible, spatial, and temporal regulation is a distinctive hallmark of carcinogenesis. Epigenetics, which focuses on gene regulatory mechanisms beyond the DNA sequence, is a new entry point for tumor therapy. Moreover, metabolic reprogramming drives hepatocellular carcinoma (HCC) initiation and progression, highlighting the significance of metabolism in this disease. Exploring the inter-regulatory relationship between tumor metabolic reprogramming and epigenetic modification has become one of the hot directions in current tumor metabolism research. As viral etiologies have given way to metabolic dysfunction-associated steatotic liver disease (MASLD)-induced HCC, it is urgent that complex molecular pathways linking them and hepatocarcinogenesis be explored. However, how aberrant crosstalk between epigenetic modifications and metabolic reprogramming affects MASLD-induced HCC lacks comprehensive understanding. A better understanding of their linkages is necessary and urgent to improve HCC treatment strategies. For this reason, this review examines the interwoven landscape of molecular carcinogenesis in the context of MASLD-induced HCC, focusing on mechanisms regulating aberrant epigenetic alterations and metabolic reprogramming in the development of MASLD-induced HCC and interactions between them while also updating the current advances in metabolism and epigenetic modification-based therapeutic drugs in HCC.

1. Introduction

Globally, HCC exhibits a 5-year survival rate of only 18%, emphasizing the poor prognosis of the disease [1]. With the increase in the prevalence of metabolic syndrome and antiviral drugs and alcohol control increasingly widespread, HCC from nonviral origins has been increasing rapidly, particularly in developed countries [2]. Non-alcoholic fatty liver disease (NAFLD), now better known as metabolic dysfunction-associated steatotic liver disease (MASLD), and its progression to non-alcoholic steatohepatitis (NASH), more recently referred to as metabolic dysfunction-associated steatohepatitis (MASH) [3], is becoming the most dominant risk factor for HCC [4]. Despite this, MASLD-related HCC has received relatively little attention due to MASLD patients being at a higher risk for cardiovascular events. The diagnosis of liver cancer is difficult in the early stages, and the prognosis is poor for patients who receive a late-stage diagnosis [5]. Understanding how MASLD progresses to malignant liver cancer lesions is key to early prevention and reversal of malignant transformation.
Epigenetic regulation is one of the hot topics in studying the pathogenesis and therapeutic intervention of HCC. Epigenetics research involves the changes in heritable gene expression or cellular expression that occur without altering DNA sequences [6]. Both the occurrence and development progress of MASLD and HCC are regulated by epigenetics [7], which is mostly reversible and high in plasticity [8]. It is, therefore, evident that epigenetic mechanisms play an important role in HCC initiation and progression, even in collaboration with other environmental influences.
Metabolic reprogramming refers to the process by which cells change their metabolic pathways and regulate related factors to adapt to specific environments or meet specific energy requirements [9], which has recently become a popular research area. As a result of the increased energetic and biosynthetic demands, cancer cells possess a distinct metabolic pattern that reduces oxidative stress and fulfills energy and biosynthesis needs [10]. Cancer cells have active pathways for nutrient uptake, synthesis, storage, conversion, and ATP production. These pathways include glycolysis, fatty acid (FA) synthesis, and amino acid metabolism [11]. Compared with virus-associated HCC, both MASLD and MASLD-induced HCC are associated with metabolic reprogramming. Consequently, linking the metabolic reprogramming that characterizes cancer with MASLD-dysregulated metabolism might be conducive to the discovery of novel metabolic readouts.
Epigenomes and tumor metabolism interact bidirectionally with the genetic and molecular factors that regulate cancer. On the one hand, alterations in the expression or activity of epigenetic enzymes can have a wide range of direct or indirect effects on cellular metabolism; on the other hand, metabolic reprogramming participates in epigenetic regulation by regulating epigenetic enzyme activity through fluctuations in metabolites and by transmitting information about the metabolic state of the cell to the nucleus [12]. A comprehensive understanding of key enzyme expression patterns and metabolic pathway interactions is crucial for appreciating the occurrence, progression, and treatment resistance of HCC. In certain contexts, metabolic alterations and epigenetic modifications may hinder immunosurveillance and facilitate immune escape, contributing to tumor progression. Since metabolic reprogramming in the immune cells is an important and very hot topic that has been reviewed elsewhere [13], it is only briefly mentioned in this review.
However, we still lack a comprehensive understanding of the tight interplay between epigenetic modifications, as well as metabolic reprogramming in MASLD-induced HCC and how their aberrant crosstalk affects tumorigenesis and progression. In this review, we focus on the mechanisms by which epigenetics and metabolic reprogramming influence and regulate the occurrence of HCC in the context of MASLD and update recent advances in the treatment of HCC by tumor-targeted epigenetic agents and employ targeting metabolic reprogramming. By gaining deeper insights into the interplay of these factors, we can develop more targeted and efficient therapies for HCC.

2. MASLD-Induced HCC

2.1. HCC Tumorigenesis

The intricate relationship between MASLD and HCC is influenced by the interplay of various pathogenic pathways. A theory known as the multi-parallel hit theory offers a more comprehensive explanation for the development of MASLD and its progression to HCC [14], including abnormal lipid metabolism, insulin resistance (IR), oxidative stress, inflammatory response, genetic alterations, and dysbiosis in gut microbiota (Figure 1). While most chronic liver diseases progress to cirrhosis before the onset of HCC, this sequence does not always apply to MASLD-related HCC. This is particularly true in cases of “lean MASLD”, where patients are without cirrhosis or obese. The development of HCC in these patients is believed to be influenced by factors such as endotoxin-related inflammation in the gut, as well as adipocytokines, leptin, and adiponectin [15].
Almost all the possible features of MASLD provide fertile ground for the advancement of hepatocarcinogenesis. This is largely driven by both lipotoxicity and IR, ultimately leading to increased fibrogenesis, inflammation, and abnormal cellular proliferation, as well as decreases in apoptotic cell death, necroptosis, and autophagy. CD36 is considered the primary driver of lipotoxicity associated with FAs in the progression of MASLD [16]. Recent research by Tao et al. demonstrated that the enhanced expression of CD36 could expedite the advancement of HCC by boosting the levels of aldo-keto reductases family 1 member C2 (AKR1C2) and enhancing the uptake of FAs [17].
The level of oxidative DNA damage in the liver is a significant factor in the development of HCC in patients with MASH. Compared with patients with HCC caused by other factors, those with MASH-induced HCC have been found to have significantly higher levels of oxidative DNA damage in their background liver [18]. Significantly, the development of oxidative stress and DNA damage in liver cells, specifically 8-hydroxy-2-deoxyguanosine (8-OHdG, a mutagenic DNA adduct resulting from lipid peroxidation), was identified as a critical factor in the onset of MASH-related HCC [19].
In the initial stages of tumor formation, autophagy inhibits tumorigenesis by eliminating aberrant cells. However, once a tumor is well-established, autophagy facilitates tumor growth and fuels HCC survival, thus contributing to the persistence and progression of cancer [20]. The impairment of autophagy in tumor cells due to metabolic stress leads to the elevated retainment of damaged mitochondria, enhancing oxidative stress and DNA damage [21]. Both the Western diet and fructose have been demonstrated to accelerate mitochondrial depolarization and mitophagy, resulting in mitochondrial dysfunction in the initial phases of MASH, which ultimately contributes to the progression of MASH toward HCC [22]. In MASLD/MASH, cardiolipin peroxidation was observed, and the oxidated cardiolipin represents a mitophagic signal to promote mitochondrial dismissal. Furthermore, cardiolipin frequently decreases during HCC progression, potentially serving as a mechanism to avoid apoptosis [23].
Familial, twin, and epidemiological studies have indicated that MASLD has a strong heritable component. Both common and rare mutations contribute to MASLD pathogenesis and the transition from MASH to HCC [24]. A recently identified MTTP rs745447480 variant, which encodes microsomal triglyceride transfer protein (MTP), causes progressive MASLD with subsequent cirrhosis and HCC in homozygotes [25]. Pinyol et al. identified the molecular specificity of NASH-HCC, with the TERT promoter, CTNNB1, TP53, and ACVR2A as the most commonly mutated genes. In particular, the mutation rate of the potential tumor suppressor gene ACVR2A is higher in NASH-HCC compared with other etiologies [26]. Remarkably, somatic mutations in MASH mice did not did not reveal increased tumorigenesis [27].
Changes in gut microbiota and metabolites in individuals with MASLD can result in the development of liver cancer [28]. The dysregulation of gut microbiota may cause abnormal epigenetic changes, impacting gene expression and playing a role in the pathogenesis of MASLD [29].
A substantial body of research indicates that the activation of the rat sarcoma virus (RAS)/rapidly accelerated fibrosarcoma (RAF)/mitogen-activated protein kinase (MEK)/extracellular signal-regulated kinase (ERK) pathway may contribute to HCC development [30]. The NASH microenvironment may damage hepatocytes, activating the RAS/RAF/MEK/ERK pathway. In this scenario, a hepatocyte exhibiting elevated MYC levels at the point of oncogenic transformation will give rise to an HCC [31]. RAS/RAF/MEK/ERK are the main signaling pathways associated with RAS. RAS family genes are the most prevalent proto-oncogenes in human tumors, and KRAS is the RAS member with the highest mutation rate and is strongly associated with poor prognosis in HCC [32]. For instance, in the CTNNB1 mutation, KRAS activates c-Met to promote HCC, and β-catenin inhibition effectively suppresses HCC progression [33].
MASLD-induced HCC development may also include the following mechanisms. For instance, recent progress in single-cell RNA sequencing has revealed hepatic stellate cells (HSCs) with unique functions [34]. HSCs can be categorized into myHSC and cyHSC subgroups, with the former promoting tumor growth and the latter inhibiting it [35]. Interestingly, not only the overall level of fibrosis in the liver but also the ratio of these different HSC subpopulations may contribute to creating a liver microenvironment that is more conducive to the development of tumors. The mechanisms of MASLD/MASH transition to HCC are briefly shown in Table 1.

2.2. Epigenetic Dysregulation in the Pathogenesis of MASLD-Induced HCC

Epigenetic alterations influenced by variables like age and environment may also influence the advancement of HCC related to MASLD. A growing body of proof has indicated that epigenetic modifications, including DNA methylation, histone modification, RNA modification, and non-coding RNA (ncRNA)-mediated processes, are significantly linked with the progression of MASLD and HCC [36,37].

2.2.1. DNA Methylation

In genetics, methylation is a heritable enzyme-mediated chemical transformation catalyzed by DNA methyltransferases (DNMTs). This process involves transferring a methyl group from S-adenosyl methionine (SAM) to carbon 5 of the cytosine ring. DNA methylation primarily occurs at cytosine bases (C) when followed by guanine (G), thereby referred to as CpG sites. Generally, methylation in the promoter region results in transcriptional repression, whereas methylation in the gene region promotes gene expression—as described in the majority of cancers, including HCC [38]. In cancer, three types of aberrations in DNA methylation occur: hypermethylation of the CpGIs in the promoter regions of tumor suppressor genes, altered expression of DNMTs, and global hypomethylation of genes and repetitive sequences, thereby leading to genomic instability and oncogene activation [39].
Aberrant DNA methylation is a crucial initiating factor in the development of cancer in individuals with MASLD [40]. Tian et al. demonstrated the role of MASH-specific DNA methylation changes in the progression toward MASH-associated multistage HCC. This process induces gene silencing of genes implicated in DNA repair, lipid metabolism, glucose metabolism, and the progression of fibrosis via DNMT [41]. In particular, genes such as epidermal growth factor receptor, estrogen receptor 1, and glycine N-methyltransferase (GNMT) are methylated concurrently in cases of HCC and MASH with advanced fibrosis, supporting the idea of the simultaneous accumulation of CpG island methylator phenotype changes in the pathogenesis of HCC [42].
A growing body of research suggests significant and differential methylation of key genes responsible for lipid homeostasis, insulin signaling, DNA repair, liver tissue remodeling, and fibrosis progression during the progression of MASH to HCC [43]. In lean mice with MASH-HCC, variations in methylation were identified in genes associated with HCC progression and prognosis, as well as genes related to lipid metabolism. On the other hand, in obese mice with MASH-HCC, methylation differences were found in genes that were already known to be linked with HCC [44]. This highlights the importance of understanding the role of methylation in different populations with MASH-related HCC (Table 2), as it could provide valuable insights into potential treatment strategies and prognosis for patients.

2.2.2. Histone Modification

Histone modifications play a key role in epigenetic processes that control transcription, DNA replication, and repair of damage, as well as the segregation of chromosomes during tumor development [45]. Histone proteins undergo various modifications, including acetylation, methylation, phosphorylation, ubiquitination, ribosylation, and SUMOylation. Among these, acetylation has been widely studied. Histone acetylase (HAT) promotes gene transcription by catalyzing histone acetylation, while histone deacetylase (HDAC) induces gene silencing through histone deacetylation. Herranz et al. showed that HAT1 induced in HCC plays important roles in promoting tumorigenesis and poor prognosis and is essential for the development of steatosis in mice [46]. Alongside this, HDAC8 has been identified as a key player in MASH-induced HCC, driving oncogenic pathways and chromatin modifications. Knocking down HDAC8 reverses IR and decreases tumorigenicity associated with MASLD [47]. Other research has demonstrated that the deacetylation of histone H4 lysine 16 leads to the suppression of cell death-related genes, contributing to the initiation of MASH-induced HCC [48]. In addition, FASN acetylation frequently decreases in human HCC and correlates with high levels of HDAC3. The use of HDAC3 inhibitors destabilizes FASN proteins and inhibits the growth of HCC (Table 2) [49].
Table 2. Target genes related to DNA methylations and histone modifications in MASLD-induced HCC.
Table 2. Target genes related to DNA methylations and histone modifications in MASLD-induced HCC.
MechanismExperimental Model/Sample DataTarget GeneReference
DNA hypomethylationNAFLD liver tissue and corresponding HCC tissue from HCC patientsDCAF4L2, CKLF, TRIM4, PRC1,
UBE2C, TUBA1B
[41]
Gene promoter hypermethylationStelic mouse model of non-alcoholic steatohepatitis-derived HCCGNMT, EGFR, ESR1[42]
Differential methylationMouse models of lean and obese NASH-HCCIn lean NASH-HCC (CHCHD2, FSCN1, ZDHHC12, PNPLA6, LDLRAP1);
In obese NASH-HCC (RNF217, GJA8, PTPRE, PSAPL1, LRRC8D)
[44]
Histone acetylationThe transcriptomic data of human liver samples were integrated from publicly available datasetsHAT1[46]
Histone acetylationLO2, HepG2, Bel-7404, and PLC5cells;
Lentiviral-mediated shRNA knockdown in obesity-promoted NASH and HCC mouse models
SREBP-1[47]
Histone acetylationSTAM NASH-related hepatocarcinogenesis mouse modelCell death-related genes[48]
Histone acetylationHEK293T, HCT116, and ZR-75-30 cell lines; Human hepatocellular carcinoma samplesHDAC3, FASN[49]
Abbreviations: DCAF4L2, DDB1 and CUL4 associated factor 4 like 2; CKLF, chemokine-like factor; TRIM4, tripartite motif-containing protein 4; PRC1, polycomb repressive complex 1; UBE2C, ubiquitin-conjugating enzyme E2C; TUBA1B; tubulin alpha 1b; GNMT, glycine N-methyltransferase; EGFR, epidermal growth factor receptor; ESR1, estrogen receptor 1; CHCHD2, coiled-coil-helix-coiled-coil-helix domain containing 2; FSCN1, fascin actin-bundling protein 1; PNPLA6, patatin-like phospholipase domain-containing protein 6; LDLRAP1, low-density lipoprotein receptor adapter protein 1; RNF217, Ring Finger Protein 217; GJA8, gap junction protein alpha 8; PTPRE, protein Tyrosine Phosphatase Receptor Type E; LRRC8D, leucine-rich repeat-containing 8; HAT1, histone Acetyltransferase 1; SREBP-1, sterol regulatory element-binding protein 1; HDAC3, histone deacetylase; FASN, fatty acid synthase.

2.2.3. NcRNAs

Non-coding RNAs (ncRNAs) are a category of RNAs that lack protein-coding potency and can influentially alter diverse cellular processes and participate in pathogenic mechanisms. The class of ncRNAs in question includes microRNAs (miRNAs), long non-coding RNAs (lncRNAs), and circular RNAs (circRNAs). These molecules do not code for proteins but affect gene expression [50]. ncRNA may have significant regulatory functions in the initiation and progression of MASLD-HCC (Table 3).
Accumulating evidence suggests that miRNAs play a significant role in the epigenetic dysregulation of metabolic processes in MASLD and HCC. One of the most notable is miR-122. The reduction of miR-122 has been identified as a direct inducer of MASH-associated HCC [51]. Likewise, research conducted by Rodrigues et al. revealed that miR-21 acts as a crucial instigator of the whole-spectrum MASLD advancement and assumes an equally significant part in the evolution of MASLD to HCC [52]. It is worth mentioning that several miRNAs are differentially expressed in the transitions from MASLD to MASH to HCC, with some being overexpressed in tumors (miR-155, miR-193b, miR-182) while others are downregulated in HCC (miR-20a, miR-200c, miR-483) [53,54,55].
In the context of MASLD-HCC, lncRNAs are likely to impact the susceptibility to liver disease, potentially playing a role in the development of HCC [56]. HULC, the first lncRNA identified to be specifically overexpressed in HCC, has also been found to have increased expression in the liver tissue of MASLD rats [57,58]. Higher levels of lncRNA-PVT1 have been associated with advanced stages of MASLD in patients with HCC, indicating its potential as a diagnostic biomarker for identifying advanced MASLD stages [59]. Research has demonstrated that the activation of adipogenesis requires the lncRNA NEAT1 in a miR-140-dependent manner [60]. NEAT1 is associated with the development of liver fibrosis, MASLD, and HCC while serving as a preventative in the pathogenesis of liver failure by suppressing the inflammatory response [61]. Additionally, research by Wang et al. revealed that LINC01468 is upregulated in liver tissues during MASLD-HCC progression, and silencing this lncRNA can inhibit HCC tumorigenesis through lipid metabolism regulation [62]. These findings highlight the promise of lncRNAs as emerging prognostic markers and therapeutic targets in cancer treatment.
A multitude of conserved binding sites on circRNA function as a “miRNA sponge” by inhibiting miRNA activity through interactions with miRNA AGO proteins [63]. In HCC, certain circRNAs are dysregulated and can impact key processes associated with the progression from MASLD to HCC, including control over lipogenesis, fibrosis, and cellular proliferation [64]. While circMTO1 adversely affects the progression of HCC, CDR1 and circ_0067934 can enhance the invasion and metastasis in HCC [65]. The circRNA_0046366-related rebalancing of lipid homeostasis results in a significant reduction in TG, which in turn leads to amelioration of the hepatocellular steatosis phenotype. The upregulation of circRNA_0046366 abrogates the miR-34a-dependent inhibition of PPARα [66].
Table 3. Relevant dysregulated ncRNAs associated with alterations in MASLD-induced HCC.
Table 3. Relevant dysregulated ncRNAs associated with alterations in MASLD-induced HCC.
ncRNAExperimental ModelExpressionFunctionReference
miR-122NAFLD, NASH, HCC patientsDownregulatedSilence FRAT2 to avoid dysfunction of metabolism causing liver damage and the dysfunction of apoptosis through the dysregulation of TIMP1[51]
miR-21NAFLD-HCC patients, NAFLD-HCC mouse modelsUpregulatedThrough normalizing liver PPARα, miR-21 inhibition and suppression significantly reduced liver damage, inflammation, and fibrogenesis[52]
miR-182C57BL/6J mouse models were long-term HF or LF diet-fedUpregulatedShows early and significant dysregulation in the hepatocarcinogenesis process, and Cyld and Foxo1 as miR-182 target genes[53,54]
miR-483HCC patients, HepG2, SK-Hep1, and Hep3B cells, NAFLD mouse modelsDownregulatedInhibits cell steatosis and fibrogenic signaling[55]
lncRNA HULCHep3B cells, HCC patients, NAFLD rat modelsUpregulatedPromotes HCC growth and metastasis Promotes NAFLD development[57,58]
lncRNA PVT1NAFLD-HCC patientsUpregulatedCirculating could be a useful diagnostic biomarker for discriminating advanced stages[59]
lncRNA NEAT1HepG2, LO2 cellsUpregulatedPromotes adipogenesis, lipogenesis, and lipid absorption[61]
LINC01468THLE2 and the HCC cell lines, NAFLD-HCC patientsUpregulatedLINC01468-mediated lipogenesis promotes HCC progression through CUL4A-linked degradation of SHIP2[62]
circMTO1HCC patients, HCC-bearing male nude mice modelsDownregulatedCircMTO1 inhibits HCC growth by upregulation of p21 via sponging miR-9[65]
circRNA_0046366HepG2 cellsDownregulatedInhibits hepatic steatosis through miR-34a/PPARα[66]
Abbreviations: TIMP1, tissue inhibitor of metalloproteinases-1; PPARα, peroxisome proliferator-activated receptor α; Cyld, cylindromatosis; Foxo1, forkhead box transcription factor O1; CUL4A, cullin 4A; SHIP2, inositol 5′-phosphatase 2.

2.2.4. m6A Modification

Since the advent of high-throughput sequencing technology, over 170 distinct post-transcriptional RNA modifications have been noticed [67]. Most RNA modifications are found in transfer RNA and ribosomal RNA, but only a few have been found in mRNA. These include m6A, N1-methyladenosine (m1A), and 5-methylcytosine (m5C) [68]. Of the various modifications to mRNA, N6-methyladenosine (m6A) is by far the most abundant and well-studied. These modifications are dynamically reversible processes jointly regulated by three important types of proteins: writers (METTL3/METTL14/WTAP), erasers (FTO/ALKBH5), and readers (YTHDF1-3/YTHDC1-2) [69]. m6A is the most common and crucial alteration for controlling mRNA stability, splicing, and translation in MASLD and HCC [70]. m6A mRNA demethylation of IL-17RA has emerged as a key event before early tumor formation in HCC [71].
In terms of writers, Pan and colleagues discovered that the m6A methyltransferase METTL3 promotes MASLD-HCC [72]. METTL3 also enhances the expression and stability of LINC00958, which targets miR-3619-5p to upregulate hepatocellular carcinoma-derived growth factors, thereby promoting HCC lipogenesis and progression [73].
The m6A demethylase FTO has been widely studied due to its established relationship with obesity. Significantly increased mRNA and protein levels of FTO were observed in the livers of patients with MASLD, and large amounts of fat accumulated in the livers of the patients [74]. FTO levels are elevated in both HCC tissue and cells. Mechanically, this increase in FTO contributes to the growth of HCC by inducing demethylation of pyruvate kinase M2 (PKM2) mRNA and enhancing protein translation [75]. In contrast, Ma et al. found that FTO was downregulated in liver cancer tissues and inhibited the progression of HCC [76]. This is contrary to the fact that the demethylation regulation of FTO overregulates lipid metabolism in hepatocytes and promotes the development of HCC. Therefore, the m6A-modified “eraser” enzyme system FTO has different views on the regulation of HCC in different studies, and further research is needed.
In NAFLD livers, YTHDC2 inhibits the stability of mRNA for SCD1, FASN, SREBP-1c, and ACC1 and blocks their gene expression, resulting in the accumulation of TGs and the progression of MASLD [77]. YTHDF2, a predictor of poor HCC prognosis, promotes hepatocellular carcinoma stem cell phenotype and metastasis by upregulating octamer-binding transcription factor 4 [78]. However, it has also been demonstrated that YTHDF2 can act as a tumor suppressor. YTHDF2 binds directly to the RNA 3′-UTR of the epidermal growth factor receptor and promotes its degradation, thereby inhibiting HCC cell proliferation and growth [79]. In conclusion, epigenetic modifications may increase or inhibit the expression of specific genes, thereby affecting the progression of cells and maintaining cell stability in many ways. In the occurrence and progression of liver diseases, a variety of abnormal changes in epigenetic modification pathways are followed (Figure 2).

3. Metabolic Shifts and Reprogramming in MASLD-Induced HCC

MASLD-HCC grows slower than virus-induced HCC and acquires the ability to promote fat storage during carcinogenesis, i.e., “metabolic reprogramming” [80]. This section examines recent studies on the metabolic characteristics of HCC induced by MASLD, specifically looking at the changes in glucose, FA, and amino acid metabolism. These three key metabolic alterations have been a focal point in the study of HCC (Figure 3).

3.1. Glucose Metabolism

Glucose metabolism is vital for the development of MASLD-HCC [81]. Glucose participates in glycolysis, the pentose phosphate pathway (PPP), and the tricarboxylic acid (TCA) cycle, serving as the primary energy and nutrient supplier for cells.

3.1.1. Glycolysis

The Warburg effect posits that cancer cells undergo a metabolic shift, prioritizing glycolysis over oxidative phosphorylation for energy production [82]. Despite glycolysis being less efficient than oxidative phosphorylation in terms of ATP production per glucose molecule, cancer cells benefit from the higher rate of ATP production through glycolysis (Figure 3A). This metabolic adaptation supports the rapid growth and proliferation of cancer cells [83]. Similarly, an enhanced glycolytic metabolic phenotype has been observed in the progression of HCC associated with MASH. The dysregulation of glycolysis processes plays a crucial role in driving the progression of MASLD to MASH, cirrhosis, and, ultimately, HCC [84]. This metabolic switching is prominent in patients with diabetic co-morbidity. The mutations stemming from MASH can facilitate IR and metabolic reprogramming, specifically an uptick in glycolysis even when oxygen is present, which may increase the risk of MASH-HCC [85]. Recent research has linked the presence of pathological AKR1B1 to metabolic reprogramming induced by hyperglycemia, which can lead to heightened lactate secretion and the Warburg effect in HCC. Thus, MASLD-associated HCC formation is promoted [86]. The altered activity of glycolytic enzymes is a frequent occurrence in HCC. For instance, the enzymes involved in the process of glycolysis (such as hexokinase 2, HK2, and PKM2) exhibit high levels of expression in HCC [87]. Additionally, the novel hexokinase called hexokinase domain containing 1 (HKDC1) has been identified as playing a role in liver cancer progression, with increased expression in MASLD and HCC [88]. In this study by Khan et al. [88], it was found that knockout of HKDC1 significantly impacts glucose flux, energy metabolism, and mitochondrial function, resulting in decreased ATP production. This, in turn, affects cell-cycle progression and induces ER stress.
KRAS mutations lead to the sustained activation of downstream signaling pathways, which, in turn, result in tumorigenesis, rewire cellular metabolism, and promote alterations in the tumor microenvironment [89]. Mutant KRAS is involved in glucose metabolism in multiple ways [90]. One such alteration is the Warburg effect [91]. The metabolic effects of oncogenic KRAS have been explained by transcriptional upregulation of glucose transporters and glycolytic enzymes [92]. A recent study has reported the direct role of the KRAS4A KRAS isoform in glucose metabolism through the regulation of the glycolytic enzyme HK1. This finding is significant because it further complicates the landscape of KRAS-mediated metabolism [93]. Moreover, the oncogenic KRAS has been shown to enhance mitophagy. The mutant form of KRAS activates a mitophagy receptor, NIX, which ultimately results in reduced mitochondrial function and increased glycolysis, thus promoting cell growth and enhancing redox balance [94].

3.1.2. Pentose Phosphate Pathway

The PPP, a crucial branch of glycolysis, is responsible for producing ribonucleotides and providing NADPH [95]. HCC tissue demonstrates more active PPP when contrasted with neighboring normal tissues [96]. This heightened activity indicates the importance of PPP in meeting the needs of cancer cells for ribonucleotide synthesis and maintaining proper redox balance. As the rate-limiting enzyme in PPP, glucose-6-phosphate dehydrogenase (G6PD) has been the most studied one in HCC. G6PD has been identified as a key factor in promoting the migration and invasion of HCC cells in laboratory settings, primarily through facilitating the process of epithelial-to-mesenchymal transition (EMT) via the signal transducer and activator of the transcription 3 signaling pathway [43]. Metabolic analysis has revealed that the PPP plays a critical role in conferring resistance to regorafenib in HCC [97]. An increase in PPP function enabled cells resistant to regorafenib to better withstand the oxidative stress due to the medication. The interaction between G6PD, PI3K/AKT, NADK, and NADP+ might create a self-regulatory mechanism that controls the resistance to regorafenib in HCC. KRAS also stimulates the hexosamine biosynthesis pathway (HBP) and the non-oxidative arm of the PPP through MAPK-dependent signaling cascades, ultimately facilitating tumor survival [32].

3.1.3. Tricarboxylic Acid Cycle

The TCA cycle is a critical metabolic pathway that controls cellular energy production and aids in the creation of macro-molecules while maintaining the cell’s redox balance. Studies have shown that the TCA cycle is upregulated during the progression of HCC. This change is believed to impact the function of enzymes involved in glutamine metabolism, malate/aspartate, and citrate/pyruvate shuttle, all of which are crucial for the development and advancement of HCC [98]. These disrupted metabolic pathways in the body are associated with impaired mitochondrial adaptation, leading to impairment in antioxidant activity and the disruption of ATP, which is critical in the transition from MASH to HCC [99]. Also, the production of reductive equivalents in the TCA cycle and their subsequent oxidation has a direct impact on the development of steatohepatitis through the induction of ER stress and reactive oxygen species production [100]. During glucose deprivation, cells activate p53 and peroxisome proliferator-activated receptor γ coactivator 1α (PGC1α) in a pathway that depends on AMP-activated protein kinase (AMPK). This activation of the AMPK-p38PGC1α axis ultimately benefits cancer cells by promoting oxidative metabolism.

3.2. Lipid Metabolism

A shared feature of HCC and MASLD is the reprogramming in lipid metabolism. The main characteristic feature of MASLD is the dysregulation of lipid metabolism, which is closely associated with the development of HCC [101] (Figure 3A). Our emphasis will be on how changes in lipid metabolism within cancer cells contribute to the progression of MASLD to HCC since the lipid metabolic rearrangement of immune cells within the tumor microenvironment (TME) has been reviewed in other studies [102].

3.2.1. Fatty Acid Metabolism

The fundamental cause of MASLD is the abnormal accumulation of TGs due to disordered FA metabolism in liver cells. This indicates that the persistent dysregulation of FA metabolism may be the most basic mechanism underlying the progression from MASLD to HCC. There is currently a debate surrounding the mechanisms of FA uptake and utilization in HCC cells, and this continues to persist. One perspective argues that there has been a notable rise in the use of FA transport to meet the energy needs of HCC cells [17]. An alternative viewpoint suggests that HCC cells tend to engage in de novo lipogenesis (DNL) instead of depending on external sources of FAs [103]. This divergence in opinion could be attributed to differences in the functionality of various FA transporters.
In the context of HCC, lipoprotein lipase (LPL) has been demonstrated to boost tumor advancement by increasing the absorption of exogenous lipids. Specifically, both the LPL/FABP4/CPT1 axis and the Zinc-fingers and homeoboxes 2-LPL axis involved in FA metabolism reprogramming promote the transformation of MASLD to malignancy. Disruption of these pathways has been proven to halt HCC advancement effectively [104,105].
Normal cells mainly take up and acquire lipids through exogenous sources, whereas cancer cells are more dependent on DNL to maintain lipid homeostasis to satisfy their own proliferation and growth needs, and this process is accompanied by a high expression of sterol regulatory element-binding protein 1c (SREBP1c), fatty acid synthase (FASN), ATP-citrate lyase (ACLY), and stearoyl-CoA desaturase 1 (SCD1) [106,107]. DNL levels are elevated in patients with MASLD and HCC [108]. The accumulation of lipids in the liver leads to metabolic reprogramming, characterized by a confluence of cellular and metabolic modifications, alongside the accumulation of potentially harmful metabolites [109]. Serum metabolic profiling further revealed the alteration of metabolites associated with MASH-HCC. These altered metabolites are involved in the regulation of lipid metabolism through peroxisome proliferator-activated receptorα (PPARα), FA metabolism, biogenic amine synthesis, suppression of necroptosis, and the mechanistic target of rapamycin (mTOR) signaling pathway [110]. The oncogenic KRAS activates downstream signaling through the AKT, which eventually activates the ACLY enzyme, thus enhancing the conversion of citrate to acetyl-CoA and increasing DNL and sterol biosynthesis [111].
In HCC induced by MASLD, there is a tendency for the fatty acid oxidation (FAO) pathway to be suppressed, which serves to protect HCC cells from lipotoxicity [112]. MASH leads to an uneven distribution of oxygen within the liver lobules due to inflammation and the presence of fibrotic scars, ultimately resulting in a state of hypoxia. Patients with MASLD-induced HCC have been found to exhibit activation of the mTOR pathway, increased lipid accumulation, and upregulated hypoxia-inducible transcription factor (HIF)-2α [113]. Consequently, the activation of HIF-1α combined with the secretion of inflammatory cytokines could drive metabolic reprogramming, promote the growth of new blood vessels, and stimulate cell proliferation, ultimately facilitating the transition from MASH to HCC [114]. Fujinuma and colleagues demonstrated that the forkhead box K1 (FOXK1) protein inhibits the process of hepatic FAO in a manner dependent on mTORC1. Deletion of FOXK1 improves the progression from MASLD to MASH and, eventually, HCC by boosting the breakdown of lipids within liver cells [115]. ACSL3, a member of the long-chain acyl-CoA synthetase family, and ACSL4 play crucial roles in FA activation and are commonly upregulated in HCC [116]. In MASH, the overexpression of ACSL4 leads to enhanced steatosis by inhibiting FAO [117], whereas elevated ACSL4 levels in hepatomas contribute to heightened lipogenesis by indirectly boosting SREBP1 activity [118].
Lipid metabolism plays a crucial role in the communication between HCC cells and the TME. MASH contributes to the development of HCC by creating a proinflammatory microenvironment [119]. One critical aspect is the excessive activation of the classical inflammatory signaling pathway NF-kB. The activation of the NF-kB pathway can lead to the secretion of proinflammatory cytokines such as tumor necrosis factorα, interleukin-1 (IL-1), and IL-6, and transforming growth factor-b, which are linked to liver ailments [120]. These molecular signaling pathways promote the development of MASLD to HCC. Also, shifts in the immunological pattern compromised the hepatic immune system, transforming its anti-tumor function into a carcinogenesis-promoting process. Dysregulation of the gut microbiota through the gut-liver axis leads to immune activation and disrupted bile acid (BA) signaling, both of which have been shown to significantly impact the development and progression of MASLD [121].

3.2.2. Cholesterol Metabolism

In addition to FA metabolism, cholesterol metabolic reprogramming in HCC cells can also promote the development of tumors [122]. Cancer cells maintain the production of cholesterol to support their growth and change the microenvironment, even in the presence of ample sterols. Research shows that disrupted cholesterol metabolism is a significant factor in the development of MASH and HCC [123]. Cholesterol synthesis was shown to promote the growth of HCC, even in the absence of FASN [124], which indicates a crosstalk between DNL and cholesterol synthesis. Liu et al. found that abnormally elevated cholesterol in HCC cells accelerated the process of malignant lesions from MASLD to HCC in mouse liver by inducing the lncRNA SNHG6 to localize in the ER-lysosome interaction region and activating the mTORC1 signaling pathway [125]. Importantly, positive correlations have been found between hypercholesterolemia and reductions in natural killer T (NKT) cells in patients with obesity, MASLD, and MASLD-related HCC. The study conducted by Tang et al. demonstrated that contrary to its role in triggering proinflammatory signaling in the liver, obesity-induced cholesterol accumulation through mTORC1/SREBP2 signaling activation specifically inhibits NKT cell-mediated anti-tumor surveillance within the liver [126]. Recent research by Li et al. demonstrated that signalosome 6 (CSN6) levels are elevated in HCC and act as an activator of hydroxymethylglutaryl-CoA synthase 1 (HMGCS1) in the mevalonate pathway, promoting the development of tumors [127]. The authors also illustrated that inhibiting CSN6 or HMGCS1 can potentially control the growth of liver cancer caused by MASLD. In addition, Hu et al. verified the idea that a feedback loop between cholesterol synthesis and the PPP contributes to the development of HCC. Inhibition of the PPP halted cholesterol formation, consequently hindering HCC in c-Myc mice [128].
It is becoming increasingly clear that the relationship between cholesterol metabolism, BAs, and HCC is significant, particularly in the context of MASH. The synthesis of BAs from cholesterol in hepatocytes acts not only as a digestive process but also as a signaling mechanism that influences the development of HCCV [83]. Trafficking of cholesterol to mitochondria through steroidogenic acute regulatory protein 1 (STARD1) is the rate-limiting step in the alternative pathway of BA generation [129]. Conde et al. discovered that STARD1 plays a significant role in MASH-driven HCC by promoting the generation of BAs in the mitochondrial acidic pathway. The products of this pathway have been found to stimulate hepatocytes, leading to pluripotency, self-renewal, and inflammation within the liver [130].

3.2.3. MUFAs and PUFAs

During the progression of MASLD to HCC, there is a notable rearrangement of the serum lipidome. Research analyzing lipidomic profiles in human samples of MASLD-associated HCC has revealed a decrease in glycerophospholipids containing polyunsaturated fatty acids (PUFAs), like arachidonic acid (C20:4), accompanied by an increase in monounsaturated fatty acids (MUFAs), like oleic acid (C18:1) [131]. Nevertheless, it is unclear whether this dysregulation is a trigger for HCC or a compensatory mechanism that may rescue the phenotype. An increase in saturated fatty acids within cells and a decrease in membrane phospholipid unsaturation induced ER stress-associated cell death, and the activation of ER stress signaling plays an essential role in the onset of MASH-induced HCC. In a noncanonical pathway, MASH disrupts the negative feedback control of ring finger protein 43 (RNF43)/zinc and ring finger 3 (ZNFR3) in the WNT/β-catenin pathway [132]. Mutations in RNF43 and ZNRF3 alter lipid metabolism, specifically affecting unsaturated fatty acids and acyl-CoA biosynthesis in the context of MASH.

3.3. Amino Acid Metabolism

The liver is one of the central organs designed to control amino acid metabolism, which appears strongly enhanced in HCC patients [133]. In particular, HCC is usually accompanied by metabolic alterations in glutamine, branched-chain amino acid (BCAA), urea cycle, and one-carbon metabolism [134] (Figure 3A).

3.3.1. Glutamine Metabolism

Rapidly proliferating cancer cells highly rely on glutamine for their energy needs [135]. The high demand for glutamine in rapidly dividing HCC cells results in a metabolic rewiring that leads to a “glutamine addiction” phenotype. This phenomenon is marked by HCC cells exhibiting an elevated glutamine uptake and subsequently increased glutaminolysis [136]. Normal liver cells mainly produce glutaminase2 (GLS2); however, in the progression of liver cancer, an alteration in metabolism driven by the MYC gene shifts the expression from GLS2 to GLS1, supporting the altered glutamine metabolism that occurs in HCC tumor cells [137]. Alternatively, high levels of GLS1 have been found to correlate positively with late-stage clinicopathological features in cancer and poor prognosis, possibly attributed to its capacity to trigger the pro-proliferative pathways Akt/GSK3β/cyclinD1 and ROS/Wnt/β-catenin [138,139]. Proline biosynthesis via glutamine may also be important for cancer cell growth [140]. Tang et al. demonstrated that the metabolic axis involving glutamine, proline, and hydroxyproline exerts its effects by modulating the activity of the HIF1a in HCC and contributing to resistance to the drug sorafenib [141]. More importantly, these findings indicate that a hypoxic TME exists in HCC. In HCC, the conversion of glutamine into α-ketoglutarate is utilized to sustain glucometabolic intermediates in a glucose-deprived TME [142]. This metabolic reprogramming, particularly evident during metastasis, allows for the continued operation of the TCA cycle through anaplerosis in nutrient-deficient HCC cells. This metabolic adaptability provides a survival advantage to HCC cells [143].

3.3.2. Branched-Chain Amino Acid

BCAAs are a group of necessary amino acids comprising leucine, isoleucine, and valine. According to Takegoshi and colleagues, incorporating BCAAs into the diet of mice can halt the advancement of MASH and deter the emergence of HCC [144]. By contrast, the impairment of BCAA catabolism has been linked to the development and progression of HCC, as well as a decrease in overall survival rates [145]. Tajiri et al. found that levels of leucine were lower in MASH-HCC than in MASH patients. Pathway analysis revealed a notable increase in leucine and isoleucine degradation pathways [146]. In a recent study, Ahmed et al. also observed a reduction in the amino acid leucine in MASH-HCC patients compared with those with MASH [110].

3.3.3. Urea Cycle

The urea cycle, also known as the ornithine cycle, is vital for preventing the buildup of ammonia levels [138]. The ammonia buildup in the body due to reduced urea cycle function could be a considerable mechanism for MASLD to deteriorate into HCC [147]. The urea cycle dysfunction in MASLD is thought to be linked to changes in the regulation of genes responsible for urea cycle enzymes and an elevation in liver cell senescence [148]. The overabundance of fructose acquired through the Western diet causes uric acid production, generating reactive oxygen species (ROS) in both liver cells and fat cells. The persistent suppression of the urea cycle in liver cancer changes the metabolic pathway from producing arginine to synthesizing pyrimidines [85]. In patients with HCC, there is a notable decrease in the expression of genes related to key enzymes of the urea cycle and associated metabolites, such as citrulline, arginine, and ornithine [149].

3.3.4. One-Carbon Metabolism

One-carbon metabolism provides a substrate for methylation and is also a crucial pathway for the formation of nucleotides and reductants. The metabolism of serine, glycine, and methionine is closely interconnected with the production of 1C units. Overall, one-carbon metabolism is a complex network of interconnected pathways. A recent study by Li et al. also showed that dietary folate supplementation in DEN/HFD-induced mouse models promoted tumor development due to folate interactions with methionine and 1C metabolism in the liver [150]. 1C units derived from glycine in HCC cells support the progression of tumors by promoting purine and pyrimidine biosynthesis through the flux of the glycine cleavage system [151]. The downregulation of GNMT and betaine homocysteine methyltransferase, two crucial enzymes in the methionine cycle, has been documented in HCC. The regulation of these enzymes is disrupted in the context of HCC, potentially impacting the one-carbon metabolism [152]. In addition, not only are many enzymes involved in one-carbon metabolism altered in HCC, but they are also affected in the setting of MASH [153].

4. Interactions between Epigenetic Modifications and HCC Cell Metabolism

A bidirectional regulatory mechanism between metabolic remodeling and epigenetics exists in tumors, where many intermediate metabolites can act as substrates or cofactors to regulate chromatin modification and gene expression. On the other hand, epigenetic dysregulation mediates a unique metabolic microenvironment within tumors, which together are involved in tumor progression and treatment resistance [154].

4.1. DNA Methylation and Tumor Metabolism in HCC

The significant association between hepatic DNA methylation and IR in patients with MASLD underscores the potential mechanism by which MASLD may develop and worsen over time [155]. It is noteworthy that genes related to specific metabolic pathways exhibited hypermethylation and decreased expression levels, while numerous genes involved in tissue repair displayed hypomethylation and increased expression levels, indicating a potential dysregulation in lipid metabolism, oxidative stress response, fibrogenesis, and possibly even carcinogenesis [156]. GNMT regulates glucose homeostasis, and GNMT deficiency may have downstream effects on various metabolic pathways. Analyzing metabolism revealed heightened lipid buildup, polyamine creation and degradation, and transsulfuration in GNMT-deficient mice. This suggests that the absence of GNMT results in metabolic reprogramming, shifting carbons away from gluconeogenesis toward pathways utilizing S-adenosylmethionine (SAM) [157].
Deficiency of carbamoylphosphate synthase 1 (CPS1), an enzyme in the urea cycle, leads to an increase in ammonia levels and triggers the activation of FAO. This process generates ATP, which supports the rapid proliferation of HCC cells [158]. Notably, CPS1 was discovered to undergo hypermethylation in HCC, which is associated with a decrease in CPS1 mRNA expression. Hypermethylated transcription-repressed genes related to ureagenesis and amino acid metabolism have been observed in patients with MASLD. The findings from both experimental studies and human research support the idea that methylation of specific gene promoters can impact the functioning of the urea cycle [159].

4.2. Histone Modification and Tumor Metabolism in HCC

Nuclear factor erythroid 2-related factor 2 (Nrf2) is a crucial transcription factor controlling the response to oxidative stress, and it is involved in metabolic reprogramming in various types of cancer. Specifically, in the absence of Nrf2, there is a decrease in acetyl-CoA production, which in turn leads to a reduction in histone acetylation within tumors [160]. Zhao and colleagues identified that ubiquitin protein ligase E3 component N-recognin 7 (UBR7) plays a protective function in the development of HCC by hindering metabolic reprogramming toward aerobic glycolysis [161]. UBR7 is involved in monoubiquitinating histone H2B and acting as a histone chaperone for post-nucleosomal histone H3 [162,163]. Malic enzyme 1 (ME1) is an NADP (+)-dependent enzyme that links glycolysis and the TCA cycle. Using immunoprecipitation and mass spectrum data, Fu and colleagues demonstrated that deacetylation of phosphoglycerate mutase 5 boosts the activity of the ME1, thereby stimulating lipid synthesis and the proliferation of liver cancer cells [164]. Metabolic rewiring of glycoxenogenic enzyme phosphoenolpyruvate carboxykinase 1 (PCK1) disrupts hexosamine synthesis pathway-mediated O-GlcNAcylation and induces cataplerosis in the TCA cycle [165,166]. Gou et al. recently reported that PCK1 stimulates the synthesis of SAM via the serine synthesis pathway [167]. The PCK1-dependent generation of SAM boosts H3K9me3 modification on the promoter of S100A11, leading to the downregulation of the PI3K/AKT signaling pathway and ultimately suppressing the progression of HCC.
The unique structure of the histone variant macroH2A1 may assist cancer cells in cell cycle regulation, as well as DNA repair and transcription [168]. macroH2A1 successfully altered the metabolism of carbohydrates and lipids in HCC cells to promote the transformation into cancer stem cells by increasing lipid accumulation via activation of the LXR pathway [169].
Histone lactylation is a newly discovered histone modification that involves the addition of a lactyl group to the lysine residue of histones [170]. This modification has been found to have a positive association with glycolytic rate and show unique changes over time compared with histone acetylation. Interestingly, histone lactylation may induce gene transcription via E1A-binding protein (p300) and P53, thereby stimulating macrophage transition to the late-phase M2 phenotype.

4.3. ncRNA and Tumor Metabolism in HCC

The NcRNA-mediated reprogramming of FA metabolism is a key process from MASLD to HCC [171]. Numerous ncRNAs can directly control the activity of enzymes that synthesize lipids in liver cancer cells, leading to DNL. Of these, FASN and SCD are the most extensively researched enzymes. LncRNA ARSR has been identified as a promoter of FA accumulation, cell proliferation, and invasion in hepatocytes as a result of stimulating FASN and SCD activity by upregulating Yes-related protein 1 [172]. A newly discovered oncogenic lncRNA, RP11-386G11.10, functions as an endogenous RNA for miR-345-3p, regulating the expression of downstream lipogenic enzymes like FASN, which results in lipid accumulation within HCC cells [173]. Liu and colleagues additionally discovered that lncSNHG6 activates the interaction between mTOR and fas-associated factor family member 2 at ER-lysosome contacts and enhances the recruitment of mTORC1 to lysosomes in a cholesterol-dependent manner, further promoting the transition from MASLD to HCC [125]. Additionally, some lncRNA (Tacc1, lnc027912, Mef2c) are also involved in the reprogramming of lipid metabolism, such as mTOR/AMPK/SREBP 1c regulation in MASLD [174,175]. In addition, ncRNA can regulate the expression of FA transporter-related proteins. As an illustration, the inhibition of fatty acid-binding protein 1 (FABP1) expression by miR-603 contributes to the promotion of FA metabolism and synthesis-related proteins. This, in turn, elevates cellular oxidative stress levels, ultimately facilitating the metastasis of HCC [176].
Several miRNAs were reported to regulate glycolysis. This is, in fact, true for miR-125b, which targets HK2, or miR-34a, which is lactate dehydrogenase (LDHA), an essential enzyme in glycolysis, thereby restraining glycolysis in HCC cells [177,178]. The downregulation of miR-122 in MASH targets PKM2, resulting in enhanced glycolysis [179]. Moreover, factors beyond genetic mutations also play a role in the activation of the RAS/RAF/MEK/ERK pathway in HCC. Epigenetic mechanisms may contribute to tumor progression. Several miRNAs have been identified as potential activators of the RAS/RAF/MEK/ERK pathway in HCC [180]. The tumor-promoting factor SMYD3, targeted by miR-346, is significantly upregulated to trigger MAPK signaling, which facilitates the development of RAS-driven HCC and leads to the methylation of MAP3K2 [181].
In addition, circRNA MAT2B functions as a decoy for miR-338-3p, promoting the expression of PKM2 and increasing aerobic glycolysis, which contributes to the progression of HCC in hypoxic conditions [182]. A novel identified circRNA, circRHBDD1, has been found to enhance aerobic glycolysis and limit the efficacy of anti-programmed death-1 (PD-1) therapy in HCC [183]. LncRNAs are also involved in regulating the expression of transporter proteins and metabolic enzymes in glucose metabolism and regulate aerobic glycolysis through signaling pathways such as LKB1/AMPK, HIF, etc. The activation of RAS signaling pathways and associated lncRNAs are early molecular events in the reprogramming events [184]. lncRNAs can encode some microproteins. The microprotein KRASIM interacts with and co-localizes with the KRAS protein in the cytoplasm of HCC cells. The overexpression of KRASIM reduces the level of KRAS protein, leading to the inhibition of the ERK signaling pathway in HCC cells [185]. Chen et al. found that overexpressed lncRNASNHG6 in HCC cells binds to block proliferation 1 (BOP1) proteins to promote glycolysis and cancer cell proliferation while inhibiting apoptosis [186]. Xu and colleagues discovered that exosomes derived from tumor-associated macrophages (TAMs) have a significant impact on the regulation of glucose metabolism and cell proliferation within HCC cells [187]. Specifically, they found that TAMs release M2 macrophage polarization-associated lncRNA via exosomes to tumor cells to enhance the stability of aldehyde dehydrogenase 1A3, thereby promoting aerobic glycolysis and proliferation in HCC cells.

4.4. m6A Modification and Tumor Metabolism in HCC

In the m6A methylome of mice with MASLD, genes associated with lipid metabolism displayed marked hypermethylation. The differential m6A methylation could potentially affect lipid metabolism-related genes through RNA splicing factors, ultimately impacting the regulation of lipid metabolism [188]. Through protein expression analysis, immunoprecipitation-qPC, and RNA sequencing, Yang et al. demonstrated that the upregulated METTL14 interacts with the mRNA of ACLY and SCD1, causing a shift in their expression profiles [36]. This alteration ultimately exacerbates FA synthesis and lipid accumulation, contributing to the advancement of MASLD and HCC. Additionally, Pan and colleagues showed that the m6A writer METTL3 triggers the m6A-SCAP-cholesterol pathway, resulting in the suppression of anti-tumor CD8+ T cells, subsequently facilitating the development of MASLD-HCC [72]. Recently, METTL5 has been demonstrated to enhance cell proliferation and aggrandize FA metabolism by impacting both the FAO and DNL pathways [189]. These findings highlight a crucial regulatory process in which m6A modification influences lipidomics in MASLD and HCC.
FTO, an important m6A demethylase, is increased in MASLD liver patients, leading to decreased FAO and increased lipid accumulation [190]. Furthermore, FTO activates the SREBP/cell death-inducing DFF45-like effector C signaling pathway in an m6A-dependent manner and increases lipid accumulation in HCC [191].
In addition, the m6A reader YTHDC2 can balance the hepatic lipid metabolism by modifying the mRNA of lipid metabolism-related genes [77]. IGF2BP2, another distinct reader of m6A, could potentially intervene in TAG accumulation by influencing the degradation of CPT1A and PPARa mRNA [192]. Further research has validated that dysregulation of IGF2BP2 is associated with the progression of MASLD to HCC [193].
5-methylcytosine (m5C) is an important mRNA modification as well. Nucleolar protein 2 (NOP2), also known as NSUN1, is an M5C methyltransferase [194]. Zhang et al. identified that NOP2 relies on m5C modification to maintain c-Myc stability, leading to the Warburg effect in HCC cells [195]. C-Myc, an oncoprotein, is responsible for activating nearly all genes related to glycolysis and is crucial for regulating glycolysis in normal oxygen levels [196].

5. Therapeutic Potential of Targeted Epigenetic Modifiers and Metabolic Reprogramming in HCC

5.1. Pre-Clinical Studies

5.1.1. Epigenetic Targets

Epigenetic drugs are a group of compounds that target disturbed epigenetic changes in different disease states. The role of epigenetic drugs in liver disease management is most clearly defined in the field of HCC. These drugs are mainly inhibitors of epigenetic-related enzymes, including DNMT, histone methyltransferase (HMT), histone demethylase (HDM), HAT, HDAC, etc. [7] (Figure 4).
Experimental therapies focusing on DNA methylation-specific mechanisms in MASLD and HCC have been emerging [197]. Targeting aberrant DNA methylation using DNMT inhibitors (DNMTis) is currently being explored. Liu et al. introduced a novel DNA methylation inhibitor, Guadecitabine (SGI-110), that has demonstrated promising anti-proliferative effects on HCC cell lines [198]. Moreover, SGI-110 effectively reversed the silencing of endogenous retroviruses in liver cancer cells, which in turn boosted the immune system’s response to liver cancer and could be utilized to enhance the sensitivity of immune checkpoint inhibitors in organisms [199].
The efficacy of HDAC inhibitors (HDACis) in experimental HCC has been shown through various studies [200]. Illustratively, panobinostat, a non-selective pan-HDACi, has exhibited the competence to trigger apoptosis, reprogram cancer cell metabolism, and mitigate tumor angiogenesis [201]. In HCC, bromodomain and PHD finger containing-1 (BRPF1) induce the activation of oncogenes by stimulating gene promoter H3K14 acetylation, and BRPF1-targeted inhibitor GSK5959 demonstrates a promising capacity to ameliorate tumor progression in murine models of HCC [202]. The integration of oncolytic and epigenetic therapies is a promising strategy for the management of multiple cancers [203]. Telomelysin, a telomerase-specific oncolytic adenovirus, and AR42, an HDACi, have shown anti-cancer properties in pre-clinical studies involving human HCC [204].
The development of specialized drugs for histone demethylation in HCC is equally appealing and encouraging [205]. Epigenetic inhibitors targeting JmjC lysine HDM, including JIB-04, GSK-J4, and SD-70, have been reported to attenuate HCC aggressiveness and viability [206] and, more notably, protect against the progression of experimental MASH-related HCC [207].
In addition, ncRNAs are important epigenetic mediators in the progression of multiple HCCs, which are key targets for effective cancer therapy [208]. Bergamini et al. identified miR-494 as a metabolic driver of HCC cells to a glycolytic phenotype and established the potential of antimiR-494 oligonucleotides used in conjunction with sorafenib and metabolic agents [209]. Significant advancements have been made in the optimization of delivery strategies and chemical modification techniques for ncRNA targets such as exosomes, lipid particles, and polymers in recent years [210]. These approaches targeting ncRNAs are more straightforward than traditional methods that focus on protein-binding inhibitors. For instance, administering adenoviral vectors carrying the tumor suppressor lncRNA PRAL significantly decelerates the progression of HCC in mice [211]. Zuo et al. developed a new PLGA PEG nanoplatform encapsulating si-LINC00958 for treating HCC [73]. This platform displayed accurate tumor-targeting, cellular drug uptake, and controllable release. Parallel to this, the assessment of circulating miRNA profiles is expected to provide a non-invasive tool to evaluate and supervise the severity of HCC [212].
In recent years, pre-clinical studies targeting m6A-specific modifications have been emerging. Zhang et al. found that lipid nanoparticles targeting the m6A reader YTHDF1 could inhibit stemness and augment sensitivity to targeted therapies in HCC cells, thereby enhancing the potency of sorafenib and lenvatinib in vivo [213]. Furthermore, Pan et al. found that nanoparticle siMETTL3 or METTL3 specific inhibitor STM2457 can elicit a profound anti-tumor immune response, and targeting m6A writer METTL3 may induce MASLD-HCC tumor regression, especially when combined with anti-PD-1 therapy [72].
Currently, the application of non-oncology agents for novel cancer treatments is gathering steam [214]. Wang et al. demonstrated that the combination of the non-oncology agent meticrane, which is typically prescribed for essential hypertension, and the epigenetic drug (5AC/DNMT1) inhibited the viability of liver cancer cells more effectively than either drug alone [215].

5.1.2. Targeting Metabolic Reprogramming in HCC

Various research studies have concentrated on inhibiting enzymes in the FA biosynthesis pathway to hinder the growth of HCC cells [107]. Specifically, enzymes such as acetyl-CoA carboxylase (ACC), SCD, and FASN have been targeted through pharmacological means to impede lipid synthesis. Researchers have shown in a pre-clinical study that FASN inhibitors may potentially enhance the effectiveness of HCC therapies [216]. Building on this finding, a more recent study by Shueng and colleagues showed that Orlistat’s ability to inhibit FASN could address resistance to targeted drug therapy by disrupting HCC’s metabolic reprogramming [217]. Additionally, targeting upstream transcription factors could be a promising approach for developing therapeutic interventions for MASH-HCC. Obeticholic acid was found to function as an agonist for the farnesoid X receptor, thereby impacting FA metabolism to restrict the onset and development of MASH-HCC [218].
Saturation of NADH shuttles fuels aerobic glycolysis in cancer cell proliferation, according to a new study by Wang et al. [219]. Consequently, metabolic therapy for HCC entails dual targeting of Warburg effects and oxidative phosphorylation. The PKM2 nuclear translocation appears to be critical for activating aerobic glycolysis in HCC, while PKM1 steers metabolism toward oxidative phosphorylation [220]. By targeting PKM splicing and favoring the expression of the PKM1 isoform, antisense oligonucleotide treatment effectively inhibits HCC dependence on aerobic glycolysis, leading to a reduction in HCC cell proliferation [221]. Furthermore, protein arginine methyltransferase 3 (PRMT3) is a key driver of glycolysis in HCC cells. The PRMT3-specific inhibitor SGC707 prevents PRMT3-mediated LDHA methylation and indirectly suppresses glycolysis, as well as HCC progression [222]. Targeting glucose transporter1 (Glut1) has been investigated for HCC treatment. For example, BAY-876 is a Glut1 antagonist, and a single injection of microcrystalline BAY-876 into HCC tumor tissues led to the inhibition of glucose uptake, proliferation, and EMT of HCC [223]. Li et al. identified that Ilicicolin H can target phosphoglycerate kinase 1, a highly expressed enzyme in HCC cell lines, which inhibits the lactate production and glucose uptake of HCC cells [224]. Several natural compounds have demonstrated significant anti-HCC properties by targeting glycolysis genes or proteins. Oleuropein, for example, can hinder HCC cell glycolysis through glucose-6-phosphate isomerase inhibition, resulting in potent anti-tumor effects in animal models without any adverse side effects [225]. Erianin effectively suppresses pyruvate carboxylase enzyme activity, impairs glycolysis, and induces oxidative stress, resulting in a shortage of needed energy for the growth of HCC cells [226]. Deoxyelephantopin was found to have an impact on key metabolic processes in HepG2 cells, specifically suppressing glycolysis and reducing glucose uptake and lactic acid production, and all of the influences through the PI3K/Akt/mTOR/HIF-1α signaling pathway [227]. Recently, Wu et al. found that HuaChanSu has anti-HCC properties by inhibiting G6PD in PPP [228]. The compound is expected to inhibit PPP flux by suppressing NADPH production and reducing nucleotide levels, providing a potential therapeutic approach for the treatment of HCC.
Metformin is a well-known medication commonly used to manage individuals with type 2 diabetes mellitus [229]. Recent studies have found that metformin not only mitigates the Warburg effect by inhibiting phosphofructokinase-1 [230] but also promotes FAO [231], demonstrating promising anti-tumor effects in pre-clinical studies of HCC. In addition to metformin, other classes of antidiabetic medications have also been considered for their potential benefits in treating HCC. The ability of sodium-glucose cotransporter 2 inhibitor Canagliflozin to inhibit glucose uptake in HCC cells expands its potential applications beyond diabetes management, offering promise for patients with HCC [232]. CP-91149 is a non-insulin-dependent diabetic drug that targets glycogen phosphorylase [233]. Barot et al. found that CP-91149 inhibits glycogenolysis, interfering with glycolysis and the PPP and causing mitochondrial dysfunction in HepG2 cells [234]. Recently, Syamprasad et al. indicated that the aldose reductase inhibitor epalrestat and its analog NAR1-29 could be combined with antidiabetic therapy to combat diabetes-induced MASLD and even HCC [65].
Concerning amino acid metabolism, the development of novel therapeutics against HCC is expected to target the metabolic vulnerability of glutamine-dependent HCC. Although the glutaminase inhibitor CB-839 had limited effect on HCC, it induced the apoptosis of HCC and inhibited HCC xenografts in mice when combined with V-9302, a novel inhibitor of glutamine transporter alanine-serine-cysteine transporter 2 [235].

5.2. Clinical Trials

5.2.1. Epigenetic Drugs

Clinical validation of epigenetic drugs for HCC is currently underway based on the experimental evidence described above. A few clinical trials have demonstrated that epigenetic drugs can improve the sensitivity of tumors to chemotherapy, prompting further investigation of epigenetic inhibitors in combination with conventional chemotherapeutic agents [236].
SGI-110 is a second-generation DNMTi that concluded a phase II clinical trial in 2020 (NCT01752933) in combination with sorafenib and oxaliplatin for HCC. Similarly, another DNMTi, called decitabine, showed a positive clinical response and a safe toxicity profile for advanced HCC patients in phase II clinical trials [237]. Encouragingly, decitabine and an HMT inhibitor targeting G9a (BIX-01294) demonstrated a dual inhibitory effect and showed strong potential for HCC treatment in the clinic [238].
Belistat, an HDACi, successfully prevented further tumor progression in Phase I/II clinical trials for the treatment of patients with unresectable HCC [239]. In addition, the preliminary efficacy of the pan-HDACi resimonstat in conjunction with sorafenib has been demonstrated in the treatment of advanced HCC [240]. Further research and larger-scale clinical trials are warranted to validate these initial findings and determine the optimal dosing and treatment regimens for maximizing the therapeutic benefits of HDAC inhibitors in HCC patients.

5.2.2. Targeting Metabolic Reprogramming in HCC

Currently, treatments targeting FA anabolism are leading the way, particularly those directed against FASN, which are in clinical trials. TVB-2640 is the first FASN inhibitor to undergo clinical studies and has demonstrated positive clinical responses in patients with ovarian and breast cancer [241]. The utilization of a TVB-2640 in a clinical trial for MASH demonstrated effectiveness in the reduction of liver fat and improvement of biochemical biomarkers [242]. A phase II clinical trial of PF-5221304 for the treatment of patients with fibrosing MASH (NCT04321031) concluded with a favorable efficacy and safety profile. In light of the excellent performance of ACC inhibitors in regulating liver metabolism, it has been hypothesized that PF-05221304 may exert a therapeutic effect on MASH-induced HCC or increase the sensitivity of tumor cells to immunotherapy.
In the battle against cancer challenges, researchers have focused on targeting glucose metabolism by primarily using agents that block glycolysis in HCC treatment. It is important to highlight that the phase I clinical trial for the treatment of HCC using oroxylin A has been authorized by the National Medical Products Administration (NMPA; ChiCTR2100051434). Studies revealed that oroxylin A targets transketolase directly to inhibit non-oxidative PPP and trigger p53 signaling, resulting in anti-cancer activity [243]. In 2022, Olutasidenib (FT-2102), a selective inhibitor of mutant isocitrate dehydrogenase 1, was approved by the U.S. Food and Drug Administration to treat relapsed/refractory acute myeloid leukemia. A completed phase Ib/II study (NCT03684811) has shown that Olutasidenib has some therapeutic activity and safety in the treatment of patients with advanced solid tumors, including HCC (NCT03684811). These approaches aim to leverage the synergistic effects of these agents in targeting glucose metabolism in cancer cells, thereby potentially enhancing the therapeutic outcomes for patients. All of these studies prompt deeper consideration of how a combination of targeted therapies for metabolism and dietary interventions might maximize anti-cancer effects, optimize HCC treatment, and improve patient prognosis. Agents targeting deregulated metabolism and epigenetics for the treatment of MASLD-HCC in pre-clinical or in-clinical trials have been summarized in Table 4.

6. Conclusions and Future Perspectives

In tumor cells, epigenetic modifications and metabolic reprogramming are highly intertwined. The close interaction between the two is similarly manifested in the progression of MASLD to HCC. Ahmed et al. [110] demonstrated significant differences in pathway regulation during MASH progression to HCC, including the downregulation of FA metabolism, biogenic amine synthesis, mTOR signaling, PPAR-α-related lipid metabolism, and amino acid metabolism. Conversely, upregulated signaling pathways in MASH-HCC patients involved DNA repair, BA metabolism, cholesterol metabolism, and methylation pathways. A deep understanding of the relationship between epigenetic modifications and metabolomics, as well as the purposeful treatment of relevant targets, is of great relevance for the development of targeted drugs.
Quantitative systems pharmacology (QSP) model-based drug transformation research is an emerging international cutting-edge drug development paradigm [244]. Several mature QSP models have been developed and put into the study of glucose metabolism and the Warburg effect, changing the shortcomings of previous clinical pharmacological models lacking mechanisms [245]. Various emerging technologies based on genomics and metabolomics have also come to the fore in the study of tumor heterogeneity, cancer clone evolution, and hepatocyte network, which are undoubtedly powerful tools for understanding MASLD-HCC metabolic reprogramming deeply [246,247]. More importantly, it is essential to consider the impact of metabolites on different cell types within the liver cancer microenvironment.
The Barcelona Clinic Liver Cancer (BCLC) system is currently the most widely used staging system for liver cancer worldwide and has been endorsed by many guidelines, including the European Organization for Research and Treatment of Cancer. Its unique advantage lies in its comprehensive consideration of the general condition of the patient, tumor status, liver function, and the preferred treatment method according to different stages [248] (Figure 5). Note that the absence of etiology-specific clinical practice guidelines for HCC currently reflects the complex nature of HCC and the challenges in formulating tailored approaches for different patient populations [249]. Since MASLD is associated with liver cancer risk, proper caloric restriction, weight loss, and diabetes management are important strategies for liver cancer prevention; proper nutrition is equally important in the management of HCC. By ensuring that patients are receiving enough calories and nutrients, healthcare professionals can help improve treatment outcomes and overall quality of life for individuals battling HCC [250]. However, the majority of MASLD-HCC is diagnosed at an advanced stage, making systemic therapy the only option [251,252]. Although immunotherapy has been approved for the treatment of MASLD-related HCC, emerging evidence suggests that this specific subset of patients may not respond well to this treatment option [253]. Emerging fields in the treatment of MASH-HCC include vaccination with peptides or DNA, adoptive transfer of immune cells, and monoclonal antibodies against PD-1, etc. [119]. These strategies are undergoing further investigation to determine their effectiveness in combination with existing standard treatments for metabolic-associated HCC. Interestingly, some drugs and natural products used to treat MASLD may also exhibit anti-HCC effects. For instance, glucagon-like peptide 1 receptor agonist (GLP-1 RA) is a novel medication for diabetes that has demonstrated efficacy in enhancing liver function and diminishing hepatic fat levels in individuals suffering from MASLD [254]. The effectiveness of GLP-1 RA has also been positively demonstrated in HCC mice [255]. However, further research is needed to determine if the metabolic effects of GLP-1 RA are solely responsible for these benefits or if its anti-inflammatory properties also play a role. Compounds such as gastrodin, curcumin, genistein, and silymarin have demonstrated significant efficacy against MASLD and also exhibited anti-HCC activity in vivo and in vitro [256]. Ultimately, bridging the gap between pre-clinical research and clinical application will be crucial in harnessing the full therapeutic potential of these natural compounds for the treatment of HCC.
Given that HCC may manifest in MASLD patients before cirrhosis onset, balancing necessary early diagnosis with unnecessary invasive testing is one of the major challenges in the current management of HCC [257]. State-of-the-art epigenetic techniques have the potential to pinpoint the exact epigenetic modifications associated with MASLD and HCC, improving the accuracy and timeliness of disease diagnosis. Furthermore, researchers can develop innovative methods for diagnosing cancer without invasive procedures by exploiting HCC cells’ dependence on metabolic reprogramming [258]. Previous research has primarily focused on the role of miRNAs in regulating HCC metabolism while overlooking the involvement of lncRNAs and circRNAs in the intricate regulation of multiple genes, among them miRNAs, in HCC. It is imperative to conduct studies that explore the comprehensive biological significance of ncRNAs and their capacity for reprogramming.
HCC treatment research has evolved from a focus on efficacy to an exploration of underlying mechanisms. However, current research still lacks breadth and depth due to limited scale, unclear targets, and insufficient clinical trials. To overcome these deficiencies, it is essential to seamlessly incorporate established research with cutting-edge approaches, such as leveraging artificial intelligence for high-throughput screening, integrating proteomics with network pharmacology, and conducting clinical events-based research. Therefore, a comprehensive understanding of the characteristic alterations of metabolic and epigenetic modifications in the body during tumorigenesis and development and a systematic summary of past findings, current trends, and future research directions in metabolism and epigenetics will help to develop new combined therapeutic strategies and ideas for targeting the characteristic metabolic-epigenetic alterations in tumors.

Author Contributions

Conceptualization, A.L. and R.W.; writing—original draft preparation, A.L., R.W., Y.Z. and P.Z.; writing—review and editing, A.L., R.W. and Y.Z.; figure design, and editing, A.L., R.W. and P.Z.; supervision, J.Y.; project administration, J.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Central Government Supports Local College Reform Projects, grant number 2020YQ05, and the National Natural Science Foundation of China, Heilongjiang Touyan Innovation Team Program, grant number 81603418.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

Figure 1, Figure 2, Figure 3, Figure 4 and Figure 5 were created with BioRender.com, accessed on 19 April 2024.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer statistics. CA Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef]
  2. Yang, J.; Pan, G.; Guan, L.; Liu, Z.; Wu, Y.; Liu, Z.; Lu, W.; Li, S.; Xu, H.; Ouyang, G. The burden of primary liver cancer caused by specific etiologies from 1990 to 2019 at the global, regional, and national levels. Cancer Med. 2022, 11, 1357–1370. [Google Scholar] [CrossRef]
  3. Rinella, M.E.; Lazarus, J.V.; Ratziu, V.; Francque, S.M.; Sanyal, A.J.; Kanwal, F.; Romero, D.; Abdelmalek, M.F.; Anstee, Q.M.; Arab, J.P.; et al. A multisociety Delphi consensus statement on new fatty liver disease nomenclature. Hepatology 2023, 78, 1966–1986. [Google Scholar] [CrossRef]
  4. Cho, Y.; Kim, B.H.; Park, J.W. Preventive strategy for nonalcoholic fatty liver disease-related hepatocellular carcinoma. Clin. Mol. Hepatol. 2023, 29, S220–S227. [Google Scholar] [CrossRef]
  5. Cucarull, B.; Tutusaus, A.; Rider, P.; Hernaez-Alsina, T.; Cuno, C.; Garcia, D.F.P.; Colell, A.; Mari, M.; Morales, A. Hepatocellular carcinoma: Molecular pathogenesis and therapeutic advances. Cancers 2022, 14, 621. [Google Scholar] [CrossRef]
  6. Shi, Y.; Qi, W. Histone modifications in NAFLD: Mechanisms and potential therapy. Int. J. Mol. Sci. 2023, 24, 14653. [Google Scholar] [CrossRef]
  7. Chen, L.; Huang, W.; Wang, L.; Zhang, Z.; Zhang, F.; Zheng, S.; Kong, D. The effects of epigenetic modification on the occurrence and progression of liver diseases and the involved mechanism. Expert Rev. Gastroenterol. Hepatol. 2020, 14, 259–270. [Google Scholar] [CrossRef]
  8. Inoue, Y.; Suzuki, Y.; Kunishima, Y.; Washio, T.; Morishita, S.; Takeda, H. High-fat diet in early life triggers both reversible and persistent epigenetic changes in the medaka fish (Oryzias latipes). BMC Genom. 2023, 24, 472. [Google Scholar] [CrossRef]
  9. Martinez-Reyes, I.; Chandel, N.S. Cancer metabolism: Looking forward. Nat. Rev. Cancer 2021, 21, 669–680. [Google Scholar] [CrossRef]
  10. Finley, L. What is cancer metabolism? Cell 2023, 186, 1670–1688. [Google Scholar] [CrossRef]
  11. Khalil, A.A.; Elfert, A.; Ghanem, S.; Helal, M.; Abdelsattar, S.; Elgedawy, G.; Obada, M.; Abdel-Samiee, M.; El-Said, H. The role of metabolomics in hepatocellular carcinoma. Egypt. Liver J. 2021, 11, 41. [Google Scholar] [CrossRef]
  12. Sun, L.; Zhang, H.; Gao, P. Metabolic reprogramming and epigenetic modifications on the path to cancer. Protein Cell 2022, 13, 877–919. [Google Scholar] [CrossRef]
  13. Oura, K.; Morishita, A.; Hamaya, S.; Fujita, K.; Masaki, T. The roles of epigenetic regulation and the tumor microenvironment in the mechanism of resistance to systemic therapy in hepatocellular carcinoma. Int. J. Mol. Sci. 2023, 24, 2805. [Google Scholar] [CrossRef]
  14. Polyzos, S.A.; Chrysavgis, L.; Vachliotis, I.D.; Chartampilas, E.; Cholongitas, E. Nonalcoholic fatty liver disease and hepatocellular carcinoma: Insights in epidemiology, pathogenesis, imaging, prevention and therapy. Semin. Cancer Biol. 2023, 93, 20–35. [Google Scholar] [CrossRef]
  15. Ito, T.; Nguyen, M.H. Perspectives on the underlying etiology of HCC and its effects on treatment outcomes. J. Hepatocell. Carcinoma 2023, 10, 413–428. [Google Scholar] [CrossRef]
  16. Rada, P.; Gonzalez-Rodriguez, A.; Garcia-Monzon, C.; Valverde, A.M. Understanding lipotoxicity in NAFLD pathogenesis: Is CD36 a key driver? Cell Death Dis. 2020, 11, 802. [Google Scholar] [CrossRef]
  17. Tao, L.; Ding, X.; Yan, L.; Xu, G.; Zhang, P.; Ji, A.; Zhang, L. CD36 accelerates the progression of hepatocellular carcinoma by promoting fas absorption. Med. Oncol. 2022, 39, 202. [Google Scholar] [CrossRef]
  18. Tanaka, S.; Miyanishi, K.; Kobune, M.; Kawano, Y.; Hoki, T.; Kubo, T.; Hayashi, T.; Sato, T.; Sato, Y.; Takimoto, R.; et al. Increased hepatic oxidative DNA damage in patients with nonalcoholic steatohepatitis who develop hepatocellular carcinoma. J. Gastroenterol. 2013, 48, 1249–1258. [Google Scholar] [CrossRef]
  19. Kakehashi, A.; Suzuki, S.; Ishii, N.; Okuno, T.; Kuwae, Y.; Fujioka, M.; Gi, M.; Stefanov, V.; Wanibuchi, H. Accumulation of 8-hydroxydeoxyguanosine, l-arginine and glucose metabolites by liver tumor cells are the important characteristic features of metabolic syndrome and non-alcoholic steatohepatitis-associated hepatocarcinogenesis. Int. J. Mol. Sci. 2020, 21, 7746. [Google Scholar] [CrossRef]
  20. Ramai, D.; Facciorusso, A.; Vigandt, E.; Schaf, B.; Saadedeen, W.; Chauhan, A.; di Nunzio, S.; Shah, A.; Giacomelli, L.; Sacco, R. Progressive liver fibrosis in non-alcoholic fatty liver disease. Cells 2021, 10, 3401. [Google Scholar] [CrossRef]
  21. Mathew, R.; Karp, C.M.; Beaudoin, B.; Vuong, N.; Chen, G.; Chen, H.Y.; Bray, K.; Reddy, A.; Bhanot, G.; Gelinas, C.; et al. Autophagy suppresses tumorigenesis through elimination of p62. Cell 2009, 137, 1062–1075. [Google Scholar] [CrossRef]
  22. Krishnasamy, Y.; Gooz, M.; Li, L.; Lemasters, J.J.; Zhong, Z. Role of mitochondrial depolarization and disrupted mitochondrial homeostasis in non-alcoholic steatohepatitis and fibrosis in mice. Int. J. Physiol. Pathophysiol. Pharmacol. 2019, 11, 190–204. [Google Scholar]
  23. Longo, M.; Paolini, E.; Meroni, M.; Dongiovanni, P. Remodeling of mitochondrial plasticity: The key switch from NAFLD/NASH to HCC. Int. J. Mol. Sci. 2021, 22, 4173. [Google Scholar] [CrossRef]
  24. Meroni, M.; Longo, M.; Tria, G.; Dongiovanni, P. Genetics is of the essence to face NAFLD. Biomedicines 2021, 9, 1359. [Google Scholar] [CrossRef]
  25. Grove, J.I.; Lo, P.; Shrine, N.; Barwell, J.; Wain, L.V.; Tobin, M.D.; Salter, A.M.; Borkar, A.N.; Cuevas-Ocana, S.; Bennett, N.; et al. Identification and characterisation of a rare mttp variant underlying hereditary non-alcoholic fatty liver disease. JHEP Rep. 2023, 5, 100764. [Google Scholar] [CrossRef]
  26. Pinyol, R.; Torrecilla, S.; Wang, H.; Montironi, C.; Pique-Gili, M.; Torres-Martin, M.; Wei-Qiang, L.; Willoughby, C.E.; Ramadori, P.; Andreu-Oller, C.; et al. Molecular characterisation of hepatocellular carcinoma in patients with non-alcoholic steatohepatitis. J. Hepatol. 2021, 75, 865–878. [Google Scholar] [CrossRef]
  27. Wang, Z.; Zhu, S.; Jia, Y.; Wang, Y.; Kubota, N.; Fujiwara, N.; Gordillo, R.; Lewis, C.; Zhu, M.; Sharma, T.; et al. Positive selection of somatically mutated clones identifies adaptive pathways in metabolic liver disease. Cell 2023, 186, 1968–1984. [Google Scholar] [CrossRef]
  28. Zhang, X.; Coker, O.O.; Chu, E.S.; Fu, K.; Lau, H.; Wang, Y.X.; Chan, A.; Wei, H.; Yang, X.; Sung, J.; et al. Dietary cholesterol drives fatty liver-associated liver cancer by modulating gut microbiota and metabolites. Gut 2021, 70, 761–774. [Google Scholar] [CrossRef]
  29. Tang, R.; Liu, R.; Zha, H.; Cheng, Y.; Ling, Z.; Li, L. Gut Microbiota Induced Epigenetic Modifications in the Non-alcoholic Fatty Liver Disease Pathogenesis. Eng. Life Sci. 2023, 24, e2300016. [Google Scholar] [CrossRef]
  30. Yang, S.; Liu, G. Targeting the ras/raf/mek/erk pathway in hepatocellular carcinoma. Oncol. Lett. 2017, 13, 1041–1047. [Google Scholar] [CrossRef]
  31. D’Artista, L.; Moschopoulou, A.A.; Barozzi, I.; Craig, A.J.; Seehawer, M.; Herrmann, L.; Minnich, M.; Kang, T.W.; Rist, E.; Henning, M.; et al. MYC determines lineage commitment in KRAS-driven primary liver cancer development. J. Hepatol. 2023, 79, 141–149. [Google Scholar] [CrossRef]
  32. Mukhopadhyay, S.; Vander, H.M.; Mccormick, F. The metabolic landscape of RAS-driven cancers from biology to therapy. Nat. Cancer 2021, 2, 271–283. [Google Scholar] [CrossRef]
  33. Tao, J.; Zhang, R.; Singh, S.; Poddar, M.; Xu, E.; Oertel, M.; Chen, X.; Ganesh, S.; Abrams, M.; Monga, S.P. Targeting beta-catenin in hepatocellular cancers induced by coexpression of mutant beta-catenin and k-ras in mice. Hepatology 2017, 65, 1581–1599. [Google Scholar] [CrossRef]
  34. Wang, S.; Li, K.; Pickholz, E.; Dobie, R.; Matchett, K.P.; Henderson, N.C.; Carrico, C.; Driver, I.; Borch, J.M.; Chen, L.; et al. An autocrine signaling circuit in hepatic stellate cells underlies advanced fibrosis in nonalcoholic steatohepatitis. Sci. Transl. Med. 2023, 15, d3949. [Google Scholar] [CrossRef]
  35. Filliol, A.; Saito, Y.; Nair, A.; Dapito, D.H.; Yu, L.X.; Ravichandra, A.; Bhattacharjee, S.; Affo, S.; Fujiwara, N.; Su, H.; et al. Opposing roles of hepatic stellate cell subpopulations in hepatocarcinogenesis. Nature 2022, 610, 356–365. [Google Scholar] [CrossRef]
  36. Yang, Y.; Cai, J.; Yang, X.; Wang, K.; Sun, K.; Yang, Z.; Zhang, L.; Yang, L.; Gu, C.; Huang, X.; et al. Dysregulated m6A modification promotes lipogenesis and development of non-alcoholic fatty liver disease and hepatocellular carcinoma. Mol. Ther. 2022, 30, 2342–2353. [Google Scholar] [CrossRef]
  37. Song, J.; Hao, J.; Lu, Y.; Ding, X.; Li, M.; Xin, Y. Increased m6A Modification of BDNF MRNA via FTO Promotes Neuronal Apoptosis Following Aluminum-Induced Oxidative Stress. Environ. Pollut. 2024, 349, 123848. [Google Scholar] [CrossRef]
  38. Pirola, C.J.; Sookoian, S. Epigenetics factors in nonalcoholic fatty liver disease. Expert Rev. Gastroenterol. Hepatol. 2022, 16, 521–536. [Google Scholar] [CrossRef]
  39. Braghini, M.R.; Lo, R.O.; Romito, I.; Fernandez-Barrena, M.G.; Barbaro, B.; Pomella, S.; Rota, R.; Vinciguerra, M.; Avila, M.A.; Alisi, A. Epigenetic remodelling in human hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 2022, 41, 107. [Google Scholar] [CrossRef]
  40. Tian, Y.; Wong, V.W.; Chan, H.L.; Cheng, A.S. Epigenetic regulation of hepatocellular carcinoma in non-alcoholic fatty liver disease. Semin. Cancer Biol. 2013, 23, 471–482. [Google Scholar] [CrossRef]
  41. Tian, Y.; Arai, E.; Makiuchi, S.; Tsuda, N.; Kuramoto, J.; Ohara, K.; Takahashi, Y.; Ito, N.; Ojima, H.; Hiraoka, N.; et al. Aberrant DNA methylation results in altered gene expression in non-alcoholic steatohepatitis-related hepatocellular carcinomas. J. Cancer Res. Clin. Oncol. 2020, 146, 2461–2477. [Google Scholar] [CrossRef]
  42. Borowa-Mazgaj, B.; de Conti, A.; Tryndyak, V.; Steward, C.R.; Jimenez, L.; Melnyk, S.; Seneshaw, M.; Mirshahi, F.; Rusyn, I.; Beland, F.A.; et al. Gene expression and DNA methylation alterations in the glycine n-methyltransferase gene in diet-induced nonalcoholic fatty liver disease-associated carcinogenesis. Toxicol. Sci. 2019, 170, 273–282. [Google Scholar] [CrossRef]
  43. Gutierrez-Cuevas, J.; Lucano-Landeros, S.; Lopez-Cifuentes, D.; Santos, A.; Armendariz-Borunda, J. Epidemiologic, genetic, pathogenic, metabolic, epigenetic aspects involved in NASH-HCC: Current therapeutic strategies. Cancers 2022, 15, 23. [Google Scholar] [CrossRef]
  44. Hymel, E.; Fisher, K.W.; Farazi, P.A. Differential methylation patterns in lean and obese non-alcoholic steatohepatitis-associated hepatocellular carcinoma. BMC Cancer 2022, 22, 1276. [Google Scholar] [CrossRef]
  45. Ghiraldini, F.G.; Filipescu, D.; Bernstein, E. Solid tumours hijack the histone variant network. Nat. Rev. Cancer 2021, 21, 257–275. [Google Scholar] [CrossRef]
  46. Herranz, J.M.; Lopez-Pascual, A.; Claveria-Cabello, A.; Uriarte, I.; Latasa, M.U.; Irigaray-Miramon, A.; Adan-Villaescusa, E.; Castello-Uribe, B.; Sangro, B.; Arechederra, M.; et al. Comprehensive analysis of epigenetic and epitranscriptomic genes’ expression in human NAFLD. J. Physiol. Biochem. 2023, 79, 901–924. [Google Scholar] [CrossRef]
  47. Tian, Y.; Wong, V.W.; Wong, G.L.; Yang, W.; Sun, H.; Shen, J.; Tong, J.H.; Go, M.Y.; Cheung, Y.S.; Lai, P.B.; et al. Histone deacetylase HDAC8 promotes insulin resistance and beta-catenin activation in NAFLD-associated hepatocellular carcinoma. Cancer Res. 2015, 75, 4803–4816. [Google Scholar] [CrossRef]
  48. de Conti, A.; Dreval, K.; Tryndyak, V.; Orisakwe, O.E.; Ross, S.A.; Beland, F.A.; Pogribny, I.P. Inhibition of the cell death pathway in nonalcoholic steatohepatitis (NASH)-related hepatocarcinogenesis is associated with histone H4 lysine 16 deacetylation. Mol. Cancer Res. 2017, 15, 1163–1172. [Google Scholar] [CrossRef]
  49. Lin, H.P.; Cheng, Z.L.; He, R.Y.; Song, L.; Tian, M.X.; Zhou, L.S.; Groh, B.S.; Liu, W.R.; Ji, M.B.; Ding, C.; et al. Destabilization of fatty acid synthase by acetylation inhibits de novo lipogenesis and tumor cell growth. Cancer Res. 2016, 76, 6924–6936. [Google Scholar] [CrossRef]
  50. Esteller, M. Non-coding RNAs in human disease. Nat. Rev. Genet. 2011, 12, 861–874. [Google Scholar] [CrossRef]
  51. Li, C.W.; Chiu, Y.K.; Chen, B.S. Investigating pathogenic and hepatocarcinogenic mechanisms from normal liver to HCC by constructing genetic and epigenetic networks via big genetic and epigenetic data mining and genome-wide NGS data identification. Dis. Markers 2018, 2018, 8635329. [Google Scholar] [CrossRef]
  52. Rodrigues, P.M.; Afonso, M.B.; Simao, A.L.; Islam, T.; Gaspar, M.M.; O’Rourke, C.J.; Lewinska, M.; Andersen, J.B.; Arretxe, E.; Alonso, C.; et al. miR-21-5p promotes NASH-related hepatocarcinogenesis. Liver Int. 2023, 43, 2256–2274. [Google Scholar] [CrossRef]
  53. Tessitore, A.; Cicciarelli, G.; Del Vecchio, F.; Gaggiano, A.; Verzella, D.; Fischietti, M.; Mastroiaco, V.; Vetuschi, A.; Sferra, R.; Barnabei, R.; et al. MicroRNA Expression Analysis in High Fat Diet-Induced NAFLD-NASH-HCC Progression: Study on C57BL/6J Mice. BMC Cancer 2016, 16, 3. [Google Scholar] [CrossRef]
  54. Compagnoni, C.; Capelli, R.; Zelli, V.; Corrente, A.; Vecchiotti, D.; Flati, I.; Di Vito, N.M.; Angelucci, A.; Alesse, E.; Zazzeroni, F.; et al. MIR-182-5p is upregulated in hepatic tissues from a diet-induced NAFLD/NASH/HCC C57BL/6J mouse model and modulates Cyld and Foxo1 expression. Int. J. Mol. Sci. 2023, 24, 9239. [Google Scholar] [CrossRef]
  55. Niture, S.; Gadi, S.; Qi, Q.; Gyamfi, M.A.; Varghese, R.S.; Rios-Colon, L.; Chimeh, U.; Vandana; Ressom, H.W.; Kumar, D. MicroRNA-483-5p inhibits hepatocellular carcinoma cell proliferation, cell steatosis, and fibrosis by targeting PPARalpha and TIMP2. Cancers 2023, 15, 1715. [Google Scholar] [CrossRef]
  56. Romeo, M.; Dallio, M.; Scognamiglio, F.; Ventriglia, L.; Cipullo, M.; Coppola, A.; Tammaro, C.; Scafuro, G.; Iodice, P.; Federico, A. Role of non-coding RNAs in hepatocellular carcinoma progression: From classic to novel clinicopathogenetic implications. Cancers 2023, 15, 5178. [Google Scholar] [CrossRef]
  57. Panzitt, K.; Tschernatsch, M.M.; Guelly, C.; Moustafa, T.; Stradner, M.; Strohmaier, H.M.; Buck, C.R.; Denk, H.; Schroeder, R.; Trauner, M.; et al. Characterization of HULC, a novel gene with striking up-regulation in hepatocellular carcinoma, as noncoding rna. Gastroenterology 2007, 132, 330–342. [Google Scholar] [CrossRef]
  58. Shen, X.; Guo, H.; Xu, J.; Wang, J. Inhibition of lncRNA HULC improves hepatic fibrosis and hepatocyte apoptosis by inhibiting the MAPK signaling pathway in rats with nonalcoholic fatty liver disease. J. Cell. Physiol. 2019, 234, 18169–18179. [Google Scholar] [CrossRef]
  59. Rashad, N.M.; Khalil, U.; Ahmed, S.M.; Hussien, M.; Sami, M.M.; Wadea, F.M. The expression level of long non-coding RNA PVT1 as a diagnostic marker for advanced stages in patients with nonalcoholic fatty liver disease. Egypt. J. Hosp. Med. 2022, 87, 1825–1834. [Google Scholar] [CrossRef]
  60. Gernapudi, R.; Wolfson, B.; Zhang, Y.; Yao, Y.; Yang, P.; Asahara, H.; Zhou, Q. MicroRNA 140 promotes expression of long noncoding RNA neat1 in adipogenesis. Mol. Cell. Biol. 2016, 36, 30–38. [Google Scholar] [CrossRef]
  61. Bu, F.T.; Wang, A.; Zhu, Y.; You, H.M.; Zhang, Y.F.; Meng, X.M.; Huang, C.; Li, J. LncRNA neat1: Shedding light on mechanisms and opportunities in liver diseases. Liver Int. 2020, 40, 2612–2626. [Google Scholar] [CrossRef]
  62. Wang, H.; Wang, Y.; Lai, S.; Zhao, L.; Liu, W.; Liu, S.; Chen, H.; Wang, J.; Du, G.; Tang, B. LINC01468 drives NAFLD-HCC progression through CUL4A-linked degradation of SHIP2. Cell Death Discov. 2022, 8, 449. [Google Scholar] [CrossRef]
  63. Huang, R.; Duan, X.; Fan, J.; Li, G.; Wang, B. Role of noncoding RNA in development of nonalcoholic fatty liver disease. BioMed Res. Int. 2019, 2019, 8690592. [Google Scholar] [CrossRef]
  64. Zhang, Y.; Wang, Y. Circular RNAs in hepatocellular carcinoma: Emerging functions to clinical significances. Front. Oncol. 2021, 11, 667428. [Google Scholar] [CrossRef]
  65. Han, D.; Li, J.; Wang, H.; Su, X.; Hou, J.; Gu, Y.; Qian, C.; Lin, Y.; Liu, X.; Huang, M.; et al. Circular RNA circMTO1 acts as the sponge of microRNA-9 to suppress hepatocellular carcinoma progression. Hepatology 2017, 66, 1151–1164. [Google Scholar] [CrossRef]
  66. Guo, X.Y.; Sun, F.; Chen, J.N.; Wang, Y.Q.; Pan, Q.; Fan, J.G. CircRNA_0046366 inhibits hepatocellular steatosis by normalization of PPAR signaling. World J. Gastroenterol. 2018, 24, 323–337. [Google Scholar] [CrossRef]
  67. Garbo, S.; Zwergel, C.; Battistelli, C. m6A RNA methylation and beyond—The epigenetic machinery and potential treatment options. Drug Discov. Today 2021, 26, 2559–2574. [Google Scholar] [CrossRef]
  68. Kumari, R.; Ranjan, P.; Suleiman, Z.G.; Goswami, S.K.; Li, J.; Prasad, R.; Verma, S.K. MRNA modifications in cardiovascular biology and disease: With a focus on m6A modification. Cardiovasc. Res. 2022, 118, 1680–1692. [Google Scholar] [CrossRef]
  69. Ushijima, T.; Clark, S.J.; Tan, P. Mapping genomic and epigenomic evolution in cancer ecosystems. Science 2021, 373, 1474–1479. [Google Scholar] [CrossRef]
  70. Chen, Y.; Zhu, Z.; Zhang, L.; Wang, J.; Ren, H. Roles of N6-methyladenosine epitranscriptome in non-alcoholic fatty liver disease and hepatocellular carcinoma. Smart Med. 2023, 2, e20230008. [Google Scholar] [CrossRef]
  71. Gan, X.; Dai, Z.; Ge, C.; Yin, H.; Wang, Y.; Tan, J.; Sun, S.; Zhou, W.; Yuan, S.; Yang, F. Fto promotes liver inflammation by suppressing m6A mRNA methylation of il-17ra. Front. Oncol. 2022, 12, 989353. [Google Scholar] [CrossRef]
  72. Pan, Y.; Chen, H.; Zhang, X.; Liu, W.; Ding, Y.; Huang, D.; Zhai, J.; Wei, W.; Wen, J.; Chen, D.; et al. Mettl3 drives NAFLD-related hepatocellular carcinoma and is a therapeutic target for boosting immunotherapy. Cell Rep. Med. 2023, 4, 101144. [Google Scholar] [CrossRef]
  73. Zuo, X.; Chen, Z.; Gao, W.; Zhang, Y.; Wang, J.; Wang, J.; Cao, M.; Cai, J.; Wu, J.; Wang, X. m6A-mediated upregulation of linc00958 increases lipogenesis and acts as a nanotherapeutic target in hepatocellular carcinoma. J. Hematol. Oncol. 2020, 13, 5. [Google Scholar] [CrossRef]
  74. Guo, J.; Ren, W.; Li, A.; Ding, Y.; Guo, W.; Su, D.; Hu, C.; Xu, K.; Chen, H.; Xu, X.; et al. Fat mass and obesity-associated gene enhances oxidative stress and lipogenesis in nonalcoholic fatty liver disease. Dig. Dis. Sci. 2013, 58, 1004–1009. [Google Scholar] [CrossRef]
  75. Li, J.; Zhu, L.; Shi, Y.; Liu, J.; Lin, L.; Chen, X. m6A demethylase fto promotes hepatocellular carcinoma tumorigenesis via mediating pkm2 demethylation. Am. J. Transl. Res. 2019, 11, 6084–6092. [Google Scholar] [PubMed]
  76. Ma, J.Z.; Yang, F.; Zhou, C.C.; Liu, F.; Yuan, J.H.; Wang, F.; Wang, T.T.; Xu, Q.G.; Zhou, W.P.; Sun, S.H. Mettl14 suppresses the metastatic potential of hepatocellular carcinoma by modulating N6-methyladenosine-dependent primary microRNA processing. Hepatology 2017, 65, 529–543. [Google Scholar] [CrossRef]
  77. Zhou, B.; Liu, C.; Xu, L.; Yuan, Y.; Zhao, J.; Zhao, W.; Chen, Y.; Qiu, J.; Meng, M.; Zheng, Y.; et al. N6-methyladenosine reader protein yt521-b homology domain-containing 2 suppresses liver steatosis by regulation of mRNA stability of lipogenic genes. Hepatology 2021, 73, 91–103. [Google Scholar] [CrossRef]
  78. Zhang, C.; Huang, S.; Zhuang, H.; Ruan, S.; Zhou, Z.; Huang, K.; Ji, F.; Ma, Z.; Hou, B.; He, X. Ythdf2 promotes the liver cancer stem cell phenotype and cancer metastasis by regulating oct4 expression via m6A RNA methylation. Oncogene 2020, 39, 4507–4518. [Google Scholar] [CrossRef]
  79. Zhong, L.; Liao, D.; Zhang, M.; Zeng, C.; Li, X.; Zhang, R.; Ma, H.; Kang, T. YTHDF2 suppresses cell proliferation and growth via destabilizing the EGFR mRNA in hepatocellular carcinoma. Cancer Lett. 2019, 442, 252–261. [Google Scholar] [CrossRef]
  80. Benhammou, J.N.; Lin, J.; Aby, E.S.; Markovic, D.; Raman, S.S.; Lu, D.S.; Tong, M.J. Nonalcoholic fatty liver disease-related hepatocellular carcinoma growth rates and their clinical outcomes. Hepatoma Res. 2021, 7, 70. [Google Scholar] [CrossRef]
  81. Wang, S.; Li, Y.; Liu, N.; Shen, W.; Xu, W.; Yao, P. Identification of glucose metabolism-related genes in the progression from nonalcoholic fatty liver disease to hepatocellular carcinoma. Genet. Res. 2022, 2022, 8566342. [Google Scholar] [CrossRef]
  82. Ni, X.; Lu, C.P.; Xu, G.Q.; Ma, J.J. Transcriptional regulation and post-translational modifications in the glycolytic pathway for targeted cancer therapy. Acta Pharmacol. Sin. 2024. [Google Scholar] [CrossRef]
  83. 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 2021, 20, 28. [Google Scholar] [CrossRef]
  84. Go, Y.; Jeong, J.Y.; Jeoung, N.H.; Jeon, J.H.; Park, B.Y.; Kang, H.J.; Ha, C.M.; Choi, Y.K.; Lee, S.J.; Ham, H.J.; et al. Inhibition of pyruvate dehydrogenase kinase 2 protects against hepatic steatosis through modulation of tricarboxylic acid cycle anaplerosis and ketogenesis. Diabetes 2016, 65, 2876–2887. [Google Scholar] [CrossRef]
  85. Dewdney, B.; Roberts, A.; Qiao, L.; George, J.; Hebbard, L. A sweet connection? Fructose’s role in hepatocellular carcinoma. Biomolecules 2020, 10, 496. [Google Scholar] [CrossRef]
  86. Syamprasad, N.P.; Jain, S.; Rajdev, B.; Panda, S.R.; Kumar, G.J.; Shaik, K.M.; Shantanu, P.A.; Challa, V.S.; Jorvekar, S.B.; Borkar, R.M.; et al. Akr1b1 drives hyperglycemia-induced metabolic reprogramming in masld-associated hepatocellular carcinoma. JHEP Rep. 2024, 6, 100974. [Google Scholar] [CrossRef]
  87. Zhang, H.; Zhao, L.; Ren, P.; Sun, X. LncRNA mbnl1-as1 knockdown increases the sensitivity of hepatocellular carcinoma to tripterine by regulating mir-708-5p-mediated glycolysis. Biotechnol. Genet. Eng. Rev. 2023, 1–18. [Google Scholar] [CrossRef]
  88. Khan, M.W.; Terry, A.R.; Priyadarshini, M.; Ilievski, V.; Farooq, Z.; Guzman, G.; Cordoba-Chacon, J.; Ben-Sahra, I.; Wicksteed, B.; Layden, B.T. The hexokinase “hkdc1” interaction with the mitochondria is essential for liver cancer progression. Cell Death Dis. 2022, 13, 660. [Google Scholar] [CrossRef]
  89. Pylayeva-Gupta, Y.; Grabocka, E.; Bar-Sagi, D. Ras oncogenes: Weaving a tumorigenic web. Nat. Rev. Cancer 2011, 11, 761–774. [Google Scholar] [CrossRef]
  90. Kim, D.; Min, D.; Kim, J.; Kim, M.J.; Seo, Y.; Jung, B.H.; Kwon, S.H.; Ro, H.; Lee, S.; Sa, J.K.; et al. Nutlin-3a induces kras mutant/p53 wild type lung cancer specific methuosis-like cell death that is dependent on gfpt2. J. Exp. Clin. Cancer Res. 2023, 42, 338. [Google Scholar] [CrossRef]
  91. Ying, H.; Kimmelman, A.C.; Lyssiotis, C.A.; Hua, S.; Chu, G.C.; Fletcher-Sananikone, E.; Locasale, J.W.; Son, J.; Zhang, H.; Coloff, J.L.; et al. Oncogenic kras maintains pancreatic tumors through regulation of anabolic glucose metabolism. Cell 2012, 149, 656–670. [Google Scholar] [CrossRef]
  92. Kerr, E.M.; Gaude, E.; Turrell, F.K.; Frezza, C.; Martins, C.P. Mutant kras copy number defines metabolic reprogramming and therapeutic susceptibilities. Nature 2016, 531, 110–113. [Google Scholar] [CrossRef]
  93. Amendola, C.R.; Mahaffey, J.P.; Parker, S.J.; Ahearn, I.M.; Chen, W.C.; Zhou, M.; Court, H.; Shi, J.; Mendoza, S.L.; Morten, M.J.; et al. Kras4a directly regulates hexokinase 1. Nature 2019, 576, 482–486. [Google Scholar] [CrossRef]
  94. Humpton, T.J.; Alagesan, B.; Denicola, G.M.; Lu, D.; Yordanov, G.N.; Leonhardt, C.S.; Yao, M.A.; Alagesan, P.; Zaatari, M.N.; Park, Y.; et al. Oncogenic kras induces nix-mediated mitophagy to promote pancreatic cancer. Cancer Discov. 2019, 9, 1268–1287. [Google Scholar] [CrossRef]
  95. Mathew, M.; Nguyen, N.T.; Bhutia, Y.D.; Sivaprakasam, S.; Ganapathy, V. Metabolic signature of warburg effect in cancer: An effective and obligatory interplay between nutrient transporters and catabolic/anabolic pathways to promote tumor growth. Cancers 2024, 16, 504. [Google Scholar] [CrossRef]
  96. Ye, M.; Li, X.; Chen, L.; Mo, S.; Liu, J.; Huang, T.; Luo, F.; Zhang, J. A high-throughput sequencing data-based classifier reveals the metabolic heterogeneity of hepatocellular carcinoma. Cancers 2023, 15, 592. [Google Scholar] [CrossRef]
  97. Yang, H.; Chen, D.; Wu, Y.; Zhou, H.; Diao, W.; Liu, G.; Li, Q. A feedback loop of ppp and pi3k/akt signal pathway drives regorafenib-resistance in HCC. Cancer Metab. 2023, 11, 27. [Google Scholar] [CrossRef]
  98. Lulli, M.; Del, C.L.; Mello, T.; Sukowati, C.; Madiai, S.; Gragnani, L.; Forte, P.; Fanizzi, F.P.; Mazzocca, A.; Rombouts, K.; et al. DNA damage response protein CHK2 regulates metabolism in liver cancer. Cancer Res. 2021, 81, 2861–2873. [Google Scholar] [CrossRef]
  99. Leveille, M.; Estall, J.L. Mitochondrial dysfunction in the transition from NASH to HCC. Metabolites 2019, 9, 233. [Google Scholar] [CrossRef]
  100. Bhat, N.; Mani, A. Dysregulation of lipid and glucose metabolism in nonalcoholic fatty liver disease. Nutrients 2023, 15, 2323. [Google Scholar] [CrossRef]
  101. Paul, B.; Lewinska, M.; Andersen, J.B. Lipid alterations in chronic liver disease and liver cancer. JHEP Rep. 2022, 4, 100479. [Google Scholar] [CrossRef]
  102. Li, X.; Ramadori, P.; Pfister, D.; Seehawer, M.; Zender, L.; Heikenwalder, M. The immunological and metabolic landscape in primary and metastatic liver cancer. Nat. Rev. Cancer 2021, 21, 541–557. [Google Scholar] [CrossRef]
  103. Wang, M.D.; Wang, N.Y.; Zhang, H.L.; Sun, L.Y.; Xu, Q.R.; Liang, L.; Li, C.; Huang, D.S.; Zhu, H.; Yang, T. Fatty acid transport protein-5 (fatp5) deficiency enhances hepatocellular carcinoma progression and metastasis by reprogramming cellular energy metabolism and regulating the ampk-mtor signaling pathway. Oncogenesis 2021, 10, 74. [Google Scholar] [CrossRef]
  104. Wu, Z.; Ma, H.; Wang, L.; Song, X.; Zhang, J.; Liu, W.; Ge, Y.; Sun, Y.; Yu, X.; Wang, Z.; et al. Tumor suppressor zhx2 inhibits NAFLD-HCC progression via blocking lpl-mediated lipid uptake. Cell Death Differ. 2020, 27, 1693–1708. [Google Scholar] [CrossRef]
  105. Yang, H.; Deng, Q.; Ni, T.; Liu, Y.; Lu, L.; Dai, H.; Wang, H.; Yang, W. Targeted inhibition of lpl/fabp4/cpt1 fatty acid metabolic axis can effectively prevent the progression of nonalcoholic steatohepatitis to liver cancer. Int. J. Biol. Sci. 2021, 17, 4207–4222. [Google Scholar] [CrossRef]
  106. Ning, Z.; Guo, X.; Liu, X.; Lu, C.; Wang, A.; Wang, X.; Wang, W.; Chen, H.; Qin, W.; Liu, X.; et al. Usp22 regulates lipidome accumulation by stabilizing PPARgamma in hepatocellular carcinoma. Nat. Commun. 2022, 13, 2187. [Google Scholar] [CrossRef]
  107. Alannan, M.; Fayyad-Kazan, H.; Trezeguet, V.; Merched, A. Targeting lipid metabolism in liver cancer. Biochemistry 2020, 59, 3951–3964. [Google Scholar] [CrossRef]
  108. Roumans, K.; Lindeboom, L.; Veeraiah, P.; Remie, C.; Phielix, E.; Havekes, B.; Bruls, Y.; Brouwers, M.; Stahlman, M.; Alssema, M.; et al. Hepatic saturated fatty acid fraction is associated with de novo lipogenesis and hepatic insulin resistance. Nat. Commun. 2020, 11, 1891. [Google Scholar] [CrossRef]
  109. Nakagawa, H.; Hayata, Y.; Kawamura, S.; Yamada, T.; Fujiwara, N.; Koike, K. Lipid metabolic reprogramming in hepatocellular carcinoma. Cancers 2018, 10, 447. [Google Scholar] [CrossRef]
  110. Ahmed, E.A.; El-Derany, M.O.; Anwar, A.M.; Saied, E.M.; Magdeldin, S. Metabolomics and lipidomics screening reveal reprogrammed signaling pathways toward cancer development in non-alcoholic steatohepatitis. Int. J. Mol. Sci. 2022, 24, 210. [Google Scholar] [CrossRef]
  111. Carrer, A.; Trefely, S.; Zhao, S.; Campbell, S.L.; Norgard, R.J.; Schultz, K.C.; Sidoli, S.; Parris, J.; Affronti, H.C.; Sivanand, S.; et al. Acetyl-coa metabolism supports multistep pancreatic tumorigenesis. Cancer Discov. 2019, 9, 416–435. [Google Scholar] [CrossRef]
  112. Lin, M.; Lv, D.; Zheng, Y.; Wu, M.; Xu, C.; Zhang, Q.; Wu, L. Downregulation of cpt2 promotes tumorigenesis and chemoresistance to cisplatin in hepatocellular carcinoma. Oncotargets Ther. 2018, 11, 3101–3110. [Google Scholar] [CrossRef]
  113. Chen, J.; Chen, J.; Huang, J.; Li, Z.; Gong, Y.; Zou, B.; Liu, X.; Ding, L.; Li, P.; Zhu, Z.; et al. Hif-2alpha upregulation mediated by hypoxia promotes NAFLD-HCC progression by activating lipid synthesis via the pi3k-akt-mtor pathway. Aging 2019, 11, 10839–10860. [Google Scholar] [CrossRef]
  114. Yamada, K.; Tanaka, T.; Kai, K.; Matsufuji, S.; Ito, K.; Kitajima, Y.; Manabe, T.; Noshiro, H. Suppression of NASH-related HCC by farnesyltransferase inhibitor through inhibition of inflammation and hypoxia-inducible factor-1alpha expression. Int. J. Mol. Sci. 2023, 24, 11546. [Google Scholar] [CrossRef]
  115. Fujinuma, S.; Nakatsumi, H.; Shimizu, H.; Sugiyama, S.; Harada, A.; Goya, T.; Tanaka, M.; Kohjima, M.; Takahashi, M.; Izumi, Y.; et al. Foxk1 promotes nonalcoholic fatty liver disease by mediating mtorc1-dependent inhibition of hepatic fatty acid oxidation. Cell Rep. 2023, 42, 112530. [Google Scholar] [CrossRef]
  116. Liu, J.; Waugh, M.G. The Regulation and Functions of ACSL3 and ACSL4 in the Liver and Hepatocellular Carcinoma. Liver Cancer Int. 2022, 4, 2–41. [Google Scholar] [CrossRef]
  117. Duan, J.; Wang, Z.; Duan, R.; Yang, C.; Zhao, R.; Feng, Q.; Qin, Y.; Jiang, J.; Gu, S.; Lv, K.; et al. Therapeutic targeting of hepatic acsl4 ameliorates NASH in mice. Hepatology 2022, 75, 140–153. [Google Scholar] [CrossRef]
  118. Chen, J.; Ding, C.; Chen, Y.; Hu, W.; Yu, C.; Peng, C.; Feng, X.; Cheng, Q.; Wu, W.; Lu, Y.; et al. Acsl4 reprograms fatty acid metabolism in hepatocellular carcinoma via c-myc/srebp1 pathway. Cancer Lett. 2021, 502, 154–165. [Google Scholar] [CrossRef]
  119. Dongiovanni, P.; Meroni, M.; Longo, M.; Fargion, S.; Fracanzani, A.L. Genetics, immunity and nutrition boost the switching from NASH to HCC. Biomedicines 2021, 9, 1524. [Google Scholar] [CrossRef]
  120. Li, H.; Hu, P.; Zou, Y.; Yuan, L.; Xu, Y.; Zhang, X.; Luo, X.; Zhang, Z. Tanshinone iia and hepatocellular carcinoma: A potential therapeutic drug. Front. Oncol. 2023, 13, 1071415. [Google Scholar] [CrossRef]
  121. Song, Q.; Zhang, X. The role of gut-liver axis in gut microbiome dysbiosis associated NAFLD and NAFLD-HCC. Biomedicines 2022, 10, 524. [Google Scholar] [CrossRef]
  122. Peng, L.; Yan, Q.; Chen, Z.; Hu, Y.; Sun, Y.; Miao, Y.; Wu, Y.; Yao, Y.; Tao, L.; Chen, F.; et al. Research progress on the role of cholesterol in hepatocellular carcinoma. Eur. J. Pharmacol. 2023, 938, 175410. [Google Scholar] [CrossRef]
  123. Kim, J.Y.; He, F.; Karin, M. From liver fat to cancer: Perils of the western diet. Cancers 2021, 13, 1095. [Google Scholar] [CrossRef]
  124. Che, L.; Chi, W.; Qiao, Y.; Zhang, J.; Song, X.; Liu, Y.; Li, L.; Jia, J.; Pilo, M.G.; Wang, J.; et al. Cholesterol biosynthesis supports the growth of hepatocarcinoma lesions depleted of fatty acid synthase in mice and humans. Gut 2020, 69, 177–186. [Google Scholar] [CrossRef]
  125. Liu, F.; Tian, T.; Zhang, Z.; Xie, S.; Yang, J.; Zhu, L.; Wang, W.; Shi, C.; Sang, L.; Guo, K.; et al. Long non-coding RNA snhg6 couples cholesterol sensing with mtorc1 activation in hepatocellular carcinoma. Nat. Metab. 2022, 4, 1022–1040. [Google Scholar] [CrossRef]
  126. Tang, W.; Zhou, J.; Yang, W.; Feng, Y.; Wu, H.; Mok, M.; Zhang, L.; Liang, Z.; Liu, X.; Xiong, Z.; et al. Aberrant cholesterol metabolic signaling impairs antitumor immunosurveillance through natural killer t cell dysfunction in obese liver. Cell. Mol. Immunol. 2022, 19, 834–847. [Google Scholar] [CrossRef]
  127. Li, K.; Zhang, J.; Lyu, H.; Yang, J.; Wei, W.; Wang, Y.; Luo, H.; Zhang, Y.; Jiang, X.; Yi, H.; et al. CSN6-SPOP-HMGCS1 axis promotes hepatocellular carcinoma progression via YAP1 activation. Adv. Sci. 2024, 11, e2306827. [Google Scholar] [CrossRef]
  128. Hu, J.; Liu, N.; Song, D.; Steer, C.J.; Zheng, G.; Song, G. A positive feedback between cholesterol synthesis and the pentose phosphate pathway rather than glycolysis promotes hepatocellular carcinoma. Oncogene 2023, 42, 2892–2904. [Google Scholar] [CrossRef]
  129. Galano, M.; Venugopal, S.; Papadopoulos, V. Role of star and scp2/scpx in the transport of cholesterol and other lipids. Int. J. Mol. Sci. 2022, 23, 12115. [Google Scholar] [CrossRef]
  130. Conde, D.L.R.L.; Garcia-Ruiz, C.; Vallejo, C.; Baulies, A.; Nunez, S.; Monte, M.J.; Marin, J.; Baila-Rueda, L.; Cenarro, A.; Civeira, F.; et al. Stard1 promotes NASH-driven HCC by sustaining the generation of bile acids through the alternative mitochondrial pathway. J. Hepatol. 2021, 74, 1429–1441. [Google Scholar] [CrossRef]
  131. Tan, S.; Israeli, E.; Ericksen, R.E.; Chow, P.; Han, W. The altered lipidome of hepatocellular carcinoma. Semin. Cancer Biol. 2022, 86, 445–456. [Google Scholar] [CrossRef]
  132. Belenguer, G.; Mastrogiovanni, G.; Pacini, C.; Hall, Z.; Dowbaj, A.M.; Arnes-Benito, R.; Sljukic, A.; Prior, N.; Kakava, S.; Bradshaw, C.R.; et al. Rnf43/znrf3 loss predisposes to hepatocellular-carcinoma by impairing liver regeneration and altering the liver lipid metabolic ground-state. Nat. Commun. 2022, 13, 334. [Google Scholar] [CrossRef]
  133. Zhao, Y.; Zhang, J.; Wang, S.; Jiang, Q.; Xu, K. Identification and validation of a nine-gene amino acid metabolism-related risk signature in HCC. Front. Cell. Dev. Biol. 2021, 9, 731790. [Google Scholar] [CrossRef]
  134. Foglia, B.; Beltra, M.; Sutti, S.; Cannito, S. Metabolic reprogramming of HCC: A new microenvironment for immune responses. Int. J. Mol. Sci. 2023, 24, 7463. [Google Scholar] [CrossRef]
  135. Yin, H.; Liu, Y.; Dong, Q.; Wang, H.; Yan, Y.; Wang, X.; Wan, X.; Yuan, G.; Pan, Y. The mechanism of extracellular cypb promotes glioblastoma adaptation to glutamine deprivation microenvironment. Cancer Lett. 2024, 216862. [Google Scholar] [CrossRef]
  136. Khan, S.U.; Khan, M.U. The role of amino acid metabolic reprogramming in tumor development and immunotherapy. Biochem. Mol. Biol. 2022, 7, 6–12. [Google Scholar] [CrossRef]
  137. Zhu, W.W.; Lu, M.; Wang, X.Y.; Zhou, X.; Gao, C.; Qin, L.X. The fuel and engine: The roles of reprogrammed metabolism in metastasis of primary liver cancer. Genes Dis. 2020, 7, 299–307. [Google Scholar] [CrossRef]
  138. Du, D.; Liu, C.; Qin, M.; Zhang, X.; Xi, T.; Yuan, S.; Hao, H.; Xiong, J. Metabolic dysregulation and emerging therapeutical targets for hepatocellular carcinoma. Acta Pharm. Sin. B 2022, 12, 558–580. [Google Scholar] [CrossRef]
  139. Wang, Z.; Liu, F.; Fan, N.; Zhou, C.; Li, D.; Macvicar, T.; Dong, Q.; Bruns, C.J.; Zhao, Y. Targeting glutaminolysis: New perspectives to understand cancer development and novel strategies for potential target therapies. Front. Oncol. 2020, 10, 589508. [Google Scholar] [CrossRef]
  140. Linder, S.J.; Bernasocchi, T.; Martinez-Pastor, B.; Sullivan, K.D.; Galbraith, M.D.; Lewis, C.A.; Ferrer, C.M.; Boon, R.; Silveira, G.G.; Cho, H.M.; et al. Inhibition of the proline metabolism rate-limiting enzyme p5cs allows proliferation of glutamine-restricted cancer cells. Nat. Metab. 2023, 5, 2131–2147. [Google Scholar] [CrossRef]
  141. Tang, L.; Zeng, J.; Geng, P.; Fang, C.; Wang, Y.; Sun, M.; Wang, C.; Wang, J.; Yin, P.; Hu, C.; et al. Global metabolic profiling identifies a pivotal role of proline and hydroxyproline metabolism in supporting hypoxic response in hepatocellular carcinoma. Clin. Cancer Res. 2018, 24, 474–485. [Google Scholar] [CrossRef]
  142. Marsico, M.; Santarsiero, A.; Pappalardo, I.; Convertini, P.; Chiummiento, L.; Sardone, A.; Di Noia, M.A.; Infantino, V.; Todisco, S. Mitochondria-mediated apoptosis of HCC cells triggered by knockdown of glutamate dehydrogenase 1: Perspective for its inhibition through quercetin and permethylated anigopreissin A. Biomedicines 2021, 9, 1664. [Google Scholar] [CrossRef]
  143. Li, Y.; Li, B.; Xu, Y.; Qian, L.; Xu, T.; Meng, G.; Li, H.; Wang, Y.; Zhang, L.; Jiang, X.; et al. Got2 silencing promotes reprogramming of glutamine metabolism and sensitizes hepatocellular carcinoma to glutaminase inhibitors. Cancer Res. 2022, 82, 3223–3235. [Google Scholar] [CrossRef]
  144. Takegoshi, K.; Honda, M.; Okada, H.; Takabatake, R.; Matsuzawa-Nagata, N.; Campbell, J.S.; Nishikawa, M.; Shimakami, T.; Shirasaki, T.; Sakai, Y.; et al. Branched-chain amino acids prevent hepatic fibrosis and development of hepatocellular carcinoma in a non-alcoholic steatohepatitis mouse model. Oncotarget 2017, 8, 18191–18205. [Google Scholar] [CrossRef]
  145. Ericksen, R.E.; Lim, S.L.; Mcdonnell, E.; Shuen, W.H.; Vadiveloo, M.; White, P.J.; Ding, Z.; Kwok, R.; Lee, P.; Radda, G.K.; et al. Loss of bcaa catabolism during carcinogenesis enhances mtorc1 activity and promotes tumor development and progression. Cell Metab. 2019, 29, 1151–1165. [Google Scholar] [CrossRef]
  146. Tajiri, K.; Shimizu, Y. Branched-chain amino acids in liver diseases. World J. Gastroenterol. 2013, 19, 7620–7629. [Google Scholar] [CrossRef]
  147. Pichon, C.; Nachit, M.; Gillard, J.; Vande, V.G.; Lanthier, N.; Leclercq, I.A. Impact of l-ornithine l-aspartate on non-alcoholic steatohepatitis-associated hyperammonemia and muscle alterations. Front. Nutr. 2022, 9, 1051157. [Google Scholar] [CrossRef]
  148. Thomsen, K.L.; Eriksen, P.L.; Kerbert, A.J.; De Chiara, F.; Jalan, R.; Vilstrup, H. Role of ammonia in NAFLD: An unusual suspect. JHEP Rep. 2023, 5, 100780. [Google Scholar] [CrossRef]
  149. Tenen, D.G.; Chai, L.; Tan, J.L. Metabolic alterations and vulnerabilities in hepatocellular carcinoma. Gastroenterol. Rep. 2021, 9, 1–13. [Google Scholar] [CrossRef]
  150. Li, J.T.; Yang, H.; Lei, M.Z.; Zhu, W.P.; Su, Y.; Li, K.Y.; Zhu, W.Y.; Wang, J.; Zhang, L.; Qu, J.; et al. Dietary folate drives methionine metabolism to promote cancer development by stabilizing MAT IIA. Signal Transduct. Target. Ther. 2022, 7, 192. [Google Scholar] [CrossRef]
  151. Mukha, D.; Fokra, M.; Feldman, A.; Sarvin, B.; Sarvin, N.; Nevo-Dinur, K.; Besser, E.; Hallo, E.; Aizenshtein, E.; Schug, Z.T.; et al. Glycine decarboxylase maintains mitochondrial protein lipoylation to support tumor growth. Cell Metab. 2022, 34, 775–782. [Google Scholar] [CrossRef]
  152. Simile, M.M.; Cigliano, A.; Paliogiannis, P.; Daino, L.; Manetti, R.; Feo, C.F.; Calvisi, D.F.; Feo, F.; Pascale, R.M. Nuclear localization dictates hepatocarcinogenesis suppression by glycine n-methyltransferase. Transl. Oncol. 2022, 15, 101239. [Google Scholar] [CrossRef]
  153. Schwartz, L.M.; Persson, E.C.; Weinstein, S.J.; Graubard, B.I.; Freedman, N.D.; Mannisto, S.; Albanes, D.; Mcglynn, K.A. Alcohol consumption, one-carbon metabolites, liver cancer and liver disease mortality. PLoS ONE 2013, 8, e78156. [Google Scholar] [CrossRef]
  154. Zheng, Q.; Maksimovic, I.; Upad, A.; David, Y. Non-enzymatic covalent modifications: A new link between metabolism and epigenetics. Protein Cell 2020, 11, 401–416. [Google Scholar] [CrossRef]
  155. Sookoian, S.; Rosselli, M.S.; Gemma, C.; Burgueno, A.L.; Fernandez, G.T.; Castano, G.O.; Pirola, C.J. Epigenetic regulation of insulin resistance in nonalcoholic fatty liver disease: Impact of liver methylation of the peroxisome proliferator-activated receptor gamma coactivator 1alpha promoter. Hepatology 2010, 52, 1992–2000. [Google Scholar] [CrossRef]
  156. Murphy, S.K.; Yang, H.; Moylan, C.A.; Pang, H.; Dellinger, A.; Abdelmalek, M.F.; Garrett, M.E.; Ashley-Koch, A.; Suzuki, A.; Tillmann, H.L.; et al. Relationship between methylome and transcriptome in patients with nonalcoholic fatty liver disease. Gastroenterology 2013, 145, 1076–1087. [Google Scholar] [CrossRef]
  157. Hughey, C.C.; Trefts, E.; Bracy, D.P.; James, F.D.; Donahue, E.P.; Wasserman, D.H. Glycine n-methyltransferase deletion in mice diverts carbon flux from gluconeogenesis to pathways that utilize excess methionine cycle intermediates. J. Biol. Chem. 2018, 293, 11944–11954. [Google Scholar] [CrossRef]
  158. Wu, T.; Luo, G.; Lian, Q.; Sui, C.; Tang, J.; Zhu, Y.; Zheng, B.; Li, Z.; Zhang, Y.; Zhang, Y.; et al. Discovery of a carbamoyl phosphate synthetase 1-deficient HCC subtype with therapeutic potential through integrative genomic and experimental analysis. Hepatology 2021, 74, 3249–3268. [Google Scholar] [CrossRef]
  159. Sodum, N.; Kumar, G.; Bojja, S.L.; Kumar, N.; Rao, C.M. Epigenetics in NAFLD/NASH: Targets and therapy. Pharmacol. Res. 2021, 167, 105484. [Google Scholar] [CrossRef]
  160. Xi, C.; Pang, J.; Barrett, A.; Horuzsko, A.; Ande, S.; Mivechi, N.F.; Zhu, X. Nrf2 drives hepatocellular carcinoma progression through acetyl-coa-mediated metabolic and epigenetic regulatory networks. Mol. Cancer Res. 2023, 21, 1079–1092. [Google Scholar] [CrossRef]
  161. Zhao, L.; Kang, M.; Liu, X.; Wang, Z.; Wang, Y.; Chen, H.; Liu, W.; Liu, S.; Li, B.; Li, C.; et al. Ubr7 inhibits HCC tumorigenesis by targeting keap1/nrf2/bach1/hk2 and glycolysis. J. Exp. Clin. Cancer Res. 2022, 41, 330. [Google Scholar] [CrossRef]
  162. Dasgupta, A.; Mondal, P.; Dalui, S.; Das, C.; Roy, S. Molecular characterization of substrate-induced ubiquitin transfer by ubr7-phd finger, a newly identified histone h2bk120 ubiquitin ligase. FEBS J. 2022, 289, 1842–1857. [Google Scholar] [CrossRef]
  163. Hogan, A.K.; Sathyan, K.M.; Willis, A.B.; Khurana, S.; Srivastava, S.; Zasadzinska, E.; Lee, A.S.; Bailey, A.O.; Gaynes, M.N.; Huang, J.; et al. UBR7 acts as a histone chaperone for post-nucleosomal histone H3. EMBO J. 2021, 40, e108307. [Google Scholar] [CrossRef]
  164. Fu, G.; Li, S.T.; Jiang, Z.; Mao, Q.; Xiong, N.; Li, X.; Hao, Y.; Zhang, H. PGAM5 deacetylation mediated by SIRT2 facilitates lipid metabolism and liver cancer proliferation. Acta Biochim. Biophys. Sin. 2023, 55, 1370–1379. [Google Scholar] [CrossRef]
  165. Liu, M.X.; Jin, L.; Sun, S.J.; Liu, P.; Feng, X.; Cheng, Z.L.; Liu, W.R.; Guan, K.L.; Shi, Y.H.; Yuan, H.X.; et al. Metabolic reprogramming by PCK1 promotes TCA cataplerosis, oxidative stress and apoptosis in liver cancer cells and suppresses hepatocellular carcinoma. Oncogene 2018, 37, 1637–1653. [Google Scholar] [CrossRef]
  166. Xiang, J.; Chen, C.; Liu, R.; Gou, D.; Chang, L.; Deng, H.; Gao, Q.; Zhang, W.; Tuo, L.; Pan, X.; et al. Gluconeogenic enzyme PCK1 deficiency promotes CHK2 O-glcnacylation and hepatocellular carcinoma growth upon glucose deprivation. J. Clin. Investig. 2021, 131, e144703. [Google Scholar] [CrossRef]
  167. Gou, D.; Liu, R.; Shan, X.; Deng, H.; Chen, C.; Xiang, J.; Liu, Y.; Gao, Q.; Li, Z.; Huang, A.; et al. Gluconeogenic enzyme PCK1 supports s-adenosylmethionine biosynthesis and promotes H3K9me3 modification to suppress hepatocellular carcinoma progression. J. Clin. Investig. 2023, 133, e161713. [Google Scholar] [CrossRef]
  168. Zhang, L.; Ye, B.; Xu, Z.; Li, X.; DM, C.; Shao, Z. Genome-wide identification of mammalian cell-cycle invariant and mitotic-specific macroh2a1 domains. Biosci. Trends 2023, 17, 393–400. [Google Scholar] [CrossRef]
  169. Rivas, S.I.; Romito, I.; Maugeri, A.; Lo, R.O.; Giallongo, S.; Mazzoccoli, G.; Oben, J.A.; Li, V.G.; Mazza, T.; Alisi, A.; et al. A lipidomic signature complements stemness features acquisition in liver cancer cells. Int. J. Mol. Sci. 2020, 21, 8452. [Google Scholar] [CrossRef]
  170. Zhang, D.; Tang, Z.; Huang, H.; Zhou, G.; Cui, C.; Weng, Y.; Liu, W.; Kim, S.; Lee, S.; Perez-Neut, M.; et al. Metabolic regulation of gene expression by histone lactylation. Nature 2019, 574, 575–580. [Google Scholar] [CrossRef]
  171. Xu, K.; Xia, P.; Chen, X.; Ma, W.; Yuan, Y. Ncrna-mediated fatty acid metabolism reprogramming in HCC. Trends Endocrinol. Metab. 2023, 34, 278–291. [Google Scholar] [CrossRef]
  172. Chi, Y.; Gong, Z.; Xin, H.; Wang, Z.; Liu, Z. Long noncoding RNA lncARSR promotes nonalcoholic fatty liver disease and hepatocellular carcinoma by promoting YAP1 and activating the IRS2/AKT pathway. J. Transl. Med. 2020, 18, 126. [Google Scholar] [CrossRef]
  173. Xu, K.; Xia, P.; Gongye, X.; Zhang, X.; Ma, S.; Chen, Z.; Zhang, H.; Liu, J.; Liu, Y.; Guo, Y.; et al. A novel lncRNA rp11-386g11.10 reprograms lipid metabolism to promote hepatocellular carcinoma progression. Mol. Metab. 2022, 63, 101540. [Google Scholar] [CrossRef]
  174. Sui, J.; Pan, D.; Yu, J.; Wang, Y.; Sun, G.; Xia, H. Identification and evaluation of hub long noncoding RNAs and mRNAs in high fat diet induced liver steatosis. Nutrients 2023, 15, 948. [Google Scholar] [CrossRef]
  175. Chu, K.; Zhao, N.; Hu, X.; Feng, R.; Zhang, L.; Wang, G.; Li, W.; Liu, L. LncNONMMUG027912 alleviates lipid accumulation through AMPKα/mTOR/SREBP1C axis in nonalcoholic fatty liver. Biochem. Biophys. Res. Commun. 2022, 618, 8–14. [Google Scholar] [CrossRef]
  176. Lin, Y.X.; Wu, X.B.; Zheng, C.W.; Zhang, Q.L.; Zhang, G.Q.; Chen, K.; Zhan, Q.; An, F.M. Mechanistic investigation on the regulation of fabp1 by the il-6/mir-603 signaling in the pathogenesis of hepatocellular carcinoma. BioMed Res. Int. 2021, 2021, 8579658. [Google Scholar] [CrossRef]
  177. Zhang, H.F.; Wang, Y.C.; Han, Y.D. MicroRNA-34a inhibits liver cancer cell growth by reprogramming glucose metabolism. Mol. Med. Rep. 2018, 17, 4483–4489. [Google Scholar] [CrossRef]
  178. Jiang, J.X.; Gao, S.; Pan, Y.Z.; Yu, C.; Sun, C.Y. Overexpression of microRNA-125b sensitizes human hepatocellular carcinoma cells to 5-fluorouracil through inhibition of glycolysis by targeting hexokinase ii. Mol. Med. Rep. 2014, 10, 995–1002. [Google Scholar] [CrossRef]
  179. Inomata, Y.; Oh, J.W.; Taniguchi, K.; Sugito, N.; Kawaguchi, N.; Hirokawa, F.; Lee, S.W.; Akao, Y.; Takai, S.; Kim, K.P.; et al. Downregulation of mir-122-5p activates glycolysis via pkm2 in kupffer cells of rat and mouse models of non-alcoholic steatohepatitis. Int. J. Mol. Sci. 2022, 23, 5230. [Google Scholar] [CrossRef]
  180. Masliah-Planchon, J.; Garinet, S.; Pasmant, E. RAS-MAPK pathway epigenetic activation in cancer: Mirnas in action. Oncotarget 2016, 7, 38892–38907. [Google Scholar] [CrossRef]
  181. Zhu, W.; Qian, J.; Ma, L.; Ma, P.; Yang, F.; Shu, Y. MiR-346 suppresses cell proliferation through smyd3 dependent approach in hepatocellular carcinoma. Oncotarget 2017, 8, 65218–65229. [Google Scholar] [CrossRef]
  182. Li, Q.; Pan, X.; Zhu, D.; Deng, Z.; Jiang, R.; Wang, X. Circular RNA MAT2B promotes glycolysis and malignancy of hepatocellular carcinoma through the miR-338-3p/PKM2 axis under hypoxic stress. Hepatology 2019, 70, 1298–1316. [Google Scholar] [CrossRef]
  183. Cai, J.; Chen, Z.; Zhang, Y.; Wang, J.; Zhang, Z.; Wu, J.; Mao, J.; Zuo, X. CircRHBDD1 augments metabolic rewiring and restricts immunotherapy efficacy via m6A modification in hepatocellular carcinoma. Mol. Ther. Oncolytics 2022, 24, 755–771. [Google Scholar] [CrossRef]
  184. Kim, D.H.; Marinov, G.K.; Pepke, S.; Singer, Z.S.; He, P.; Williams, B.; Schroth, G.P.; Elowitz, M.B.; Wold, B.J. Single-cell transcriptome analysis reveals dynamic changes in lncRNA expression during reprogramming. Cell Stem Cell 2015, 16, 88–101. [Google Scholar] [CrossRef]
  185. Xu, W.; Deng, B.; Lin, P.; Liu, C.; Li, B.; Huang, Q.; Zhou, H.; Yang, J.; Qu, L. Ribosome profiling analysis identified a KRAS-interacting microprotein that represses oncogenic signaling in hepatocellular carcinoma cells. Sci. China Life Sci. 2020, 63, 529–542. [Google Scholar] [CrossRef]
  186. Chen, K.; Wang, X.; Wei, B.; Sun, R.; Wu, C.; Yang, H.J. LncRNA SNHG6 promotes glycolysis reprogramming in hepatocellular carcinoma by stabilizing the BOP1 protein. Anim. Cells Syst. 2022, 26, 369–379. [Google Scholar] [CrossRef]
  187. Xu, M.; Zhou, C.; Weng, J.; Chen, Z.; Zhou, Q.; Gao, J.; Shi, G.; Ke, A.; Ren, N.; Sun, H.; et al. Tumor associated macrophages-derived exosomes facilitate hepatocellular carcinoma malignance by transferring lncMMPA to tumor cells and activating glycolysis pathway. J. Exp. Clin. Cancer Res. 2022, 41, 253. [Google Scholar] [CrossRef]
  188. Luo, Z.; Zhang, Z.; Tai, L.; Zhang, L.; Sun, Z.; Zhou, L. Comprehensive analysis of differences of N6-methyladenosine RNA methylomes between high-fat-fed and normal mouse livers. Epigenomics 2019, 11, 1267–1282. [Google Scholar] [CrossRef]
  189. Peng, H.; Chen, B.; Wei, W.; Guo, S.; Han, H.; Yang, C.; Ma, J.; Wang, L.; Peng, S.; Kuang, M.; et al. N6-methyladenosine (m6A) in 18S rRNA promotes fatty acid metabolism and oncogenic transformation. Nat. Metab. 2022, 4, 1041–1054. [Google Scholar] [CrossRef]
  190. Mizuno, T.M. Fat mass and obesity associated (FTO) gene and hepatic glucose and lipid metabolism. Nutrients 2018, 10, 1600. [Google Scholar] [CrossRef]
  191. Chen, A.; Chen, X.; Cheng, S.; Shu, L.; Yan, M.; Yao, L.; Wang, B.; Huang, S.; Zhou, L.; Yang, Z.; et al. Fto promotes srebp1c maturation and enhances cidec transcription during lipid accumulation in hepg2 cells. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2018, 1863, 538–548. [Google Scholar] [CrossRef]
  192. Regue, L.; Minichiello, L.; Avruch, J.; Dai, N. Liver-specific deletion of igf2 mRNA binding protein-2/imp2 reduces hepatic fatty acid oxidation and increases hepatic triglyceride accumulation. J. Biol. Chem. 2019, 294, 11944–11951. [Google Scholar] [CrossRef]
  193. Simon, Y.; Kessler, S.M.; Bohle, R.M.; Haybaeck, J.; Kiemer, A.K. The insulin-like growth factor 2 (igf2) mrna-binding protein p62/igf2bp2-2 as a promoter of NAFLD and HCC? Gut 2014, 63, 861–863. [Google Scholar] [CrossRef]
  194. Liao, H.; Gaur, A.; Mcconie, H.; Shekar, A.; Wang, K.; Chang, J.T.; Breton, G.; Denicourt, C. Human nop2/nsun1 regulates ribosome biogenesis through non-catalytic complex formation with box c/d snornps. Nucleic Acids Res. 2022, 50, 10695–10716. [Google Scholar] [CrossRef]
  195. Zhang, H.; Zhai, X.; Liu, Y.; Xia, Z.; Xia, T.; Du, G.; Zhou, H.; Franziska, S.D.; Bazhin, A.V.; Li, Z.; et al. Nop2-mediated m5c modification of c-myc in an eif3a-dependent manner to reprogram glucose metabolism and promote hepatocellular carcinoma progression. Research 2023, 6, 184. [Google Scholar] [CrossRef]
  196. Wang, X.Y.; Wei, Y.; Hu, B.; Liao, Y.; Wang, X.; Wan, W.H.; Huang, C.X.; Mahabati, M.; Liu, Z.Y.; Qu, J.R.; et al. c-Myc-driven glycolysis polarizes functional regulatory B cells that trigger pathogenic inflammatory responses. Signal Transduct. Target. Ther. 2022, 7, 105. [Google Scholar] [CrossRef]
  197. Gimeno-Valiente, F.; Lopez-Rodas, G.; Castillo, J.; Franco, L. Alternative splicing, epigenetic modifications and cancer: A dangerous triangle, or a hopeful one? Cancers 2022, 14, 560. [Google Scholar] [CrossRef]
  198. Liu, M.; Zhang, L.; Li, H.; Hinoue, T.; Zhou, W.; Ohtani, H.; El-Khoueiry, A.; Daniels, J.; O’Connell, C.; Dorff, T.B.; et al. Integrative epigenetic analysis reveals therapeutic targets to the DNA methyltransferase inhibitor guadecitabine (sgi-110) in hepatocellular carcinoma. Hepatology 2018, 68, 1412–1428. [Google Scholar] [CrossRef]
  199. Topper, M.J.; Vaz, M.; Marrone, K.A.; Brahmer, J.R.; Baylin, S.B. The emerging role of epigenetic therapeutics in immuno-oncology. Nat. Rev. Clin. Oncol. 2020, 17, 75–90. [Google Scholar] [CrossRef]
  200. Hai, R.; Yang, D.; Zheng, F.; Wang, W.; Han, X.; Bode, A.M.; Luo, X. The emerging roles of hdacs and their therapeutic implications in cancer. Eur. J. Pharmacol. 2022, 931, 175216. [Google Scholar] [CrossRef]
  201. Toh, T.B.; Lim, J.J.; Chow, E.K. Epigenetics of hepatocellular carcinoma. Clin. Transl. Med. 2019, 8, 13. [Google Scholar] [CrossRef]
  202. Cheng, C.L.; Tsang, F.H.; Wei, L.; Chen, M.; Chin, D.W.; Shen, J.; Law, C.T.; Lee, D.; Wong, C.C.; Ng, I.O.; et al. Bromodomain-containing protein BRPF1 is a therapeutic target for liver cancer. Commun. Biol. 2021, 4, 888. [Google Scholar] [CrossRef]
  203. Shirbhate, E.; Veerasamy, R.; Boddu, S.; Tiwari, A.K.; Rajak, H. Histone deacetylase inhibitor-based oncolytic virotherapy: A promising strategy for cancer treatment. Drug Discov. Today 2022, 27, 1689–1697. [Google Scholar] [CrossRef]
  204. Lin, Z.Z.; Hu, M.C.; Hsu, C.; Wu, Y.M.; Lu, Y.S.; Ho, J.A.; Yeh, S.H.; Chen, P.J.; Cheng, A.L. Synergistic efficacy of telomerase-specific oncolytic adenoviral therapy and histone deacetylase inhibition in human hepatocellular carcinoma. Cancer Lett. 2023, 556, 216063. [Google Scholar] [CrossRef]
  205. Liu, Y.; Chen, M. Histone demethylation profiles in nonalcoholic fatty liver disease and prognostic values in hepatocellular carcinoma: A bioinformatic analysis. Curr. Issues Mol. Biol. 2023, 45, 3640–3657. [Google Scholar] [CrossRef]
  206. Bayo, J.; Fiore, E.J.; Dominguez, L.M.; Real, A.; Malvicini, M.; Rizzo, M.; Atorrasagasti, C.; Garcia, M.G.; Argemi, J.; Martinez, E.D.; et al. A comprehensive study of epigenetic alterations in hepatocellular carcinoma identifies potential therapeutic targets. J. Hepatol. 2019, 71, 78–90. [Google Scholar] [CrossRef]
  207. Juhling, F.; Hamdane, N.; Crouchet, E.; Li, S.; El, S.H.; Mukherji, A.; Fujiwara, N.; Oudot, M.A.; Thumann, C.; Saviano, A.; et al. Targeting clinical epigenetic reprogramming for chemoprevention of metabolic and viral hepatocellular carcinoma. Gut 2021, 70, 157–169. [Google Scholar] [CrossRef]
  208. Dai, E.; Zhu, Z.; Wahed, S.; Qu, Z.; Storkus, W.J.; Guo, Z.S. Epigenetic modulation of antitumor immunity for improved cancer immunotherapy. Mol. Cancer 2021, 20, 171. [Google Scholar] [CrossRef]
  209. Bergamini, C.; Leoni, I.; Rizzardi, N.; Melli, M.; Galvani, G.; Coada, C.A.; Giovannini, C.; Monti, E.; Liparulo, I.; Valenti, F.; et al. MiR-494 induces metabolic changes through g6pc targeting and modulates sorafenib response in hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 2023, 42, 145. [Google Scholar] [CrossRef]
  210. Winkle, M.; El-Daly, S.M.; Fabbri, M.; Calin, G.A. Noncoding RNA therapeutics—Challenges and potential solutions. Nat. Rev. Drug Discov. 2021, 20, 629–651. [Google Scholar] [CrossRef]
  211. Zhou, C.C.; Yang, F.; Yuan, S.X.; Ma, J.Z.; Liu, F.; Yuan, J.H.; Bi, F.R.; Lin, K.Y.; Yin, J.H.; Cao, G.W.; et al. Systemic genome screening identifies the outcome associated focal loss of long noncoding RNA PRAL in hepatocellular carcinoma. Hepatology 2016, 63, 850–863. [Google Scholar] [CrossRef]
  212. Aggeletopoulou, I.; Kalafateli, M.; Tsounis, E.P.; Triantos, C. Epigenetic regulation in lean nonalcoholic fatty liver disease. Int. J. Mol. Sci. 2023, 24, 12864. [Google Scholar] [CrossRef]
  213. Zhang, X.; Su, T.; Wu, Y.; Cai, Y.; Wang, L.; Liang, C.; Zhou, L.; Wang, S.; Li, X.X.; Peng, S.; et al. N6-methyladenosine reader YTHDF1 promotes stemness and therapeutic resistance in hepatocellular carcinoma by enhancing notch1 expression. Cancer Res. 2024, 84, 827–840. [Google Scholar] [CrossRef]
  214. Zhang, Z.; Zhou, L.; Xie, N.; Nice, E.C.; Zhang, T.; Cui, Y.; Huang, C. Overcoming cancer therapeutic bottleneck by drug repurposing. Signal Transduct. Target. Ther. 2020, 5, 113. [Google Scholar] [CrossRef]
  215. Wang, Y.; Sharma, A.; Ge, F.; Chen, P.; Yang, Y.; Liu, H.; Liu, H.; Zhao, C.; Mittal, L.; Asthana, S.; et al. Non-oncology drug (meticrane) shows anti-cancer ability in synergy with epigenetic inhibitors and appears to be involved passively in targeting cancer cells. Front. Oncol. 2023, 13, 1157366. [Google Scholar] [CrossRef]
  216. Wang, H.; Zhou, Y.; Xu, H.; Wang, X.; Zhang, Y.; Shang, R.; O’Farrell, M.; Roessler, S.; Sticht, C.; Stahl, A.; et al. Therapeutic efficacy of fasn inhibition in preclinical models of HCC. Hepatology 2022, 76, 951–966. [Google Scholar] [CrossRef]
  217. Shueng, P.W.; Chan, H.W.; Lin, W.C.; Kuo, D.Y.; Chuang, H.Y. Orlistat resensitizes sorafenib-resistance in hepatocellular carcinoma cells through modulating metabolism. Int. J. Mol. Sci. 2022, 23, 6501. [Google Scholar] [CrossRef]
  218. Attia, Y.M.; Tawfiq, R.A.; Gibriel, A.A.; Ali, A.A.; Kassem, D.H.; Hammam, O.A.; Elmazar, M.M. Activation of fxr modulates socs3/jak2/stat3 signaling axis in a NASH-dependent hepatocellular carcinoma animal model. Biochem. Pharmacol. 2021, 186, 114497. [Google Scholar] [CrossRef]
  219. Wang, Y.; Stancliffe, E.; Fowle-Grider, R.; Wang, R.; Wang, C.; Schwaiger-Haber, M.; Shriver, L.P.; Patti, G.J. Saturation of the mitochondrial NADH shuttles drives aerobic glycolysis in proliferating cells. Mol. Cell 2022, 82, 3270–3283. [Google Scholar] [CrossRef]
  220. Jiao, F.; Dai, W.; Mao, Y.; Wu, L.; Li, J.; Chen, K.; Yu, Q.; Kong, R.; Li, S.; Zhang, J.; et al. Simvastatin re-sensitizes hepatocellular carcinoma cells to sorafenib by inhibiting HIF-1α/PPAR-γ/PKM2-mediated glycolysis. J. Exp. Clin. Cancer Res. 2020, 39, 24. [Google Scholar] [CrossRef]
  221. Ma, W.K.; Voss, D.M.; Scharner, J.; Costa, A.; Lin, K.T.; Jeon, H.Y.; Wilkinson, J.E.; Jackson, M.; Rigo, F.; Bennett, C.F.; et al. ASO-based PKM splice-switching therapy inhibits hepatocellular carcinoma growth. Cancer Res. 2022, 82, 900–915. [Google Scholar] [CrossRef]
  222. Lei, Y.; Han, P.; Chen, Y.; Wang, H.; Wang, S.; Wang, M.; Liu, J.; Yan, W.; Tian, D.; Liu, M. Protein arginine methyltransferase 3 promotes glycolysis and hepatocellular carcinoma growth by enhancing arginine methylation of lactate dehydrogenase A. Clin. Transl. Med. 2022, 12, e686. [Google Scholar] [CrossRef]
  223. Chen, H.; Wu, Q.; Liu, P.; Cao, T.; Deng, M.; Liu, Y.; Huang, J.; Hu, Y.; Fu, N.; Zhou, K.; et al. Mechanism, clinical significance, and treatment strategy of warburg effect in hepatocellular carcinoma. J. Nanomater. 2021, 2021, 5164100. [Google Scholar] [CrossRef]
  224. Li, M.; Zhang, A.; Qi, X.; Yu, R.; Li, J. A novel inhibitor of pgk1 suppresses the aerobic glycolysis and proliferation of hepatocellular carcinoma. Biomed. Pharmacother. 2023, 158, 114115. [Google Scholar] [CrossRef]
  225. Hong, Z.; Lu, Y.; Liu, B.; Ran, C.; Lei, X.; Wang, M.; Wu, S.; Yang, Y.; Wu, H. Glycolysis, a new mechanism of oleuropein against liver tumor. Phytomedicine 2023, 114, 154770. [Google Scholar] [CrossRef]
  226. Sheng, Y.; Chen, Y.; Zeng, Z.; Wu, W.; Wang, J.; Ma, Y.; Lin, Y.; Zhang, J.; Huang, Y.; Li, W.; et al. Identification of pyruvate carboxylase as the cellular target of natural bibenzyls with potent anticancer activity against hepatocellular carcinoma via metabolic reprogramming. J. Med. Chem. 2022, 65, 460–484. [Google Scholar] [CrossRef]
  227. Wu, H.; Zhou, S.; Wang, C.; Zhou, H.; Li, F.; Chen, X.; Zhang, L.; Qin, Y. Effects of deoxydibilin on proliferation, migration and glycolysis of hepatocellular carcinoma HepG2 cells. Chin. Tradit. Pat Med. 2023, 45, 208–212. (In Chinese) [Google Scholar]
  228. Wu, Q.; Ge, X.L.; Geng, Z.K.; Wu, H.; Yang, J.Y.; Cao, S.R.; Yang, A.L. Huachansu suppresses the growth of hepatocellular carcinoma cells by interfering with pentose phosphate pathway through down-regulation of g6pd enzyme activity and expression. Heliyon 2024, 10, e25144. [Google Scholar] [CrossRef]
  229. Bhalla, K.; Hwang, B.J.; Dewi, R.E.; Twaddel, W.; Goloubeva, O.; Wong, K.; Saxena, N.K.; Biswal, S.; Girnun, G.D. Metformin prevents liver tumorigenesis by inhibiting pathways driving hepatic lipogenesis. Cancer Prev. Res. 2012, 5, 544–552. [Google Scholar] [CrossRef]
  230. Hu, L.; Zeng, Z.; Xia, Q.; Liu, Z.; Feng, X.; Chen, J.; Huang, M.; Chen, L.; Fang, Z.; Liu, Q.; et al. Metformin attenuates hepatoma cell proliferation by decreasing glycolytic flux through the hif-1alpha/pfkfb3/pfk1 pathway. Life Sci. 2019, 239, 116966. [Google Scholar] [CrossRef]
  231. He, L. Metformin and systemic metabolism. Trends Pharmacol. Sci. 2020, 41, 868–881. [Google Scholar] [CrossRef]
  232. Hendryx, M.; Dong, Y.; Ndeke, J.M.; Luo, J. Sodium-glucose cotransporter 2 (sglt2) inhibitor initiation and hepatocellular carcinoma prognosis. PLoS ONE 2022, 17, e274519. [Google Scholar] [CrossRef]
  233. Martin, W.H.; Hoover, D.J.; Armento, S.J.; Stock, I.A.; Mcpherson, R.K.; Danley, D.E.; Stevenson, R.W.; Barrett, E.J.; Treadway, J.L. Discovery of a human liver glycogen phosphorylase inhibitor that lowers blood glucose in vivo. Proc. Natl. Acad. Sci. USA 1998, 95, 1776–1781. [Google Scholar] [CrossRef]
  234. Barot, S.; Stephenson, O.J.; Priya, V.H.; Yadav, A.; Bhutkar, S.; Trombetta, L.D.; Dukhande, V.V. Metabolic alterations and mitochondrial dysfunction underlie hepatocellular carcinoma cell death induced by a glycogen metabolic inhibitor. Biochem. Pharmacol. 2022, 203, 115201. [Google Scholar] [CrossRef]
  235. Jin, H.; Wang, S.; Zaal, E.A.; Wang, C.; Wu, H.; Bosma, A.; Jochems, F.; Isima, N.; Jin, G.; Lieftink, C.; et al. A powerful drug combination strategy targeting glutamine addiction for the treatment of human liver cancer. Elife 2020, 9, e56749. [Google Scholar] [CrossRef]
  236. Italiano, A.; Soria, J.C.; Toulmonde, M.; Michot, J.M.; Lucchesi, C.; Varga, A.; Coindre, J.M.; Blakemore, S.J.; Clawson, A.; Suttle, B.; et al. Tazemetostat, an ezh2 inhibitor, in relapsed or refractory b-cell non-hodgkin lymphoma and advanced solid tumours: A first-in-human, open-label, phase 1 study. Lancet Oncol. 2018, 19, 649–659. [Google Scholar] [CrossRef]
  237. Mei, Q.; Chen, M.; Lu, X.; Li, X.; Duan, F.; Wang, M.; Luo, G.; Han, W. An open-label, single-arm, phase I/II study of lower-dose decitabine based therapy in patients with advanced hepatocellular carcinoma. Oncotarget 2015, 6, 16698–16711. [Google Scholar] [CrossRef]
  238. Barcena-Varela, M.; Caruso, S.; Llerena, S.; Alvarez-Sola, G.; Uriarte, I.; Latasa, M.U.; Urtasun, R.; Rebouissou, S.; Alvarez, L.; Jimenez, M.; et al. Dual targeting of histone methyltransferase g9a and dna-methyltransferase 1 for the treatment of experimental hepatocellular carcinoma. Hepatology 2019, 69, 587–603. [Google Scholar] [CrossRef]
  239. Yeo, W.; Chung, H.C.; Chan, S.L.; Wang, L.Z.; Lim, R.; Picus, J.; Boyer, M.; Mo, F.K.; Koh, J.; Rha, S.Y.; et al. Epigenetic therapy using belinostat for patients with unresectable hepatocellular carcinoma: A multicenter phase i/ii study with biomarker and pharmacokinetic analysis of tumors from patients in the mayo phase ii consortium and the cancer therapeutics research group. J. Clin. Oncol. 2012, 30, 3361–3367. [Google Scholar] [CrossRef]
  240. Tak, W.Y.; Ryoo, B.Y.; Lim, H.Y.; Kim, D.Y.; Okusaka, T.; Ikeda, M.; Hidaka, H.; Yeon, J.E.; Mizukoshi, E.; Morimoto, M.; et al. Phase I/II study of first-line combination therapy with sorafenib plus resminostat, an oral hdac inhibitor, versus sorafenib monotherapy for advanced hepatocellular carcinoma in east asian patients. Investig. New Drugs 2018, 36, 1072–1084. [Google Scholar] [CrossRef]
  241. Falchook, G.; Infante, J.; Arkenau, H.T.; Patel, M.R.; Dean, E.; Borazanci, E.; Brenner, A.; Cook, N.; Lopez, J.; Pant, S.; et al. First-in-human study of the safety, pharmacokinetics, and pharmacodynamics of first-in-class fatty acid synthase inhibitor tvb-2640 alone and with a taxane in advanced tumors. Eclinicalmedicine 2021, 34, 100797. [Google Scholar] [CrossRef]
  242. Loomba, R.; Mohseni, R.; Lucas, K.J.; Gutierrez, J.A.; Perry, R.G.; Trotter, J.F.; Rahimi, R.S.; Harrison, S.A.; Ajmera, V.; Wayne, J.D.; et al. Tvb-2640 (fasn inhibitor) for the treatment of nonalcoholic steatohepatitis: Fascinate-1, a randomized, placebo-controlled phase 2a trial. Gastroenterology 2021, 161, 1475–1486. [Google Scholar] [CrossRef]
  243. Jia, D.; Liu, C.; Zhu, Z.; Cao, Y.; Wen, W.; Hong, Z.; Liu, Y.; Liu, E.; Chen, L.; Chen, C.; et al. Novel transketolase inhibitor oroxylin a suppresses the non-oxidative pentose phosphate pathway and hepatocellular carcinoma tumour growth in mice and patient-derived organoids. Clin. Transl. Med. 2022, 12, e1095. [Google Scholar] [CrossRef]
  244. Sove, R.J.; Verma, B.K.; Wang, H.; Ho, W.J.; Yarchoan, M.; Popel, A.S. Virtual clinical trials of anti-PD-1 and anti-CTLA-4 immunotherapy in advanced hepatocellular carcinoma using a quantitative systems pharmacology model. J. Immunother. Cancer 2022, 10, e005414. [Google Scholar] [CrossRef]
  245. Kapuy, O.; Makk-Merczel, K.; Szarka, A. Therapeutic approach of kras mutant tumours by the combination of pharmacologic ascorbate and chloroquine. Biomolecules 2021, 11, 652. [Google Scholar] [CrossRef]
  246. Lu, Y.; Yang, A.; Quan, C.; Pan, Y.; Zhang, H.; Li, Y.; Gao, C.; Lu, H.; Wang, X.; Cao, P.; et al. A single-cell atlas of the multicellular ecosystem of primary and metastatic hepatocellular carcinoma. Nat. Commun. 2022, 13, 4594. [Google Scholar] [CrossRef]
  247. Guilliams, M.; Bonnardel, J.; Haest, B.; Vanderborght, B.; Wagner, C.; Remmerie, A.; Bujko, A.; Martens, L.; Thone, T.; Browaeys, R.; et al. Spatial proteogenomics reveals distinct and evolutionarily conserved hepatic macrophage niches. Cell 2022, 185, 379–396. [Google Scholar] [CrossRef]
  248. Vogel, A.; Meyer, T.; Sapisochin, G.; Salem, R.; Saborowski, A. Hepatocellular carcinoma. Lancet 2022, 400, 1345–1362. [Google Scholar] [CrossRef]
  249. Foerster, F.; Gairing, S.J.; Muller, L.; Galle, P.R. Nafld-driven HCC: Safety and efficacy of current and emerging treatment options. J. Hepatol. 2022, 76, 446–457. [Google Scholar] [CrossRef]
  250. Anstee, Q.M.; Reeves, H.L.; Kotsiliti, E.; Govaere, O.; Heikenwalder, M. From NASH to HCC: Current concepts and future challenges. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 411–428. [Google Scholar] [CrossRef]
  251. Papadakos, S.P.; Machairas, N.; Stergiou, I.E.; Arvanitakis, K.; Germanidis, G.; Frampton, A.E.; Theocharis, S. Unveiling the Yin-Yang balance of M1 and M2 macrophages in hepatocellular carcinoma: Role of exosomes in tumor microenvironment and immune modulation. Cells 2023, 12, 2036. [Google Scholar] [CrossRef]
  252. Zhang, H.; Zhang, W.; Jiang, L.; Chen, Y. Recent advances in systemic therapy for hepatocellular carcinoma. Biomark. Res. 2022, 10, 3. [Google Scholar] [CrossRef]
  253. Chan, L.L.; Chan, S.L. Novel perspectives in immune checkpoint inhibitors and the management of non-alcoholic steatohepatitis-related hepatocellular carcinoma. Cancers 2022, 14, 1526. [Google Scholar] [CrossRef]
  254. Lee, H.A.; Kim, H.Y. Therapeutic mechanisms and clinical effects of glucagon-like peptide 1 receptor agonists in nonalcoholic fatty liver disease. Int. J. Mol. Sci. 2023, 24, 9324. [Google Scholar] [CrossRef]
  255. Arvanitakis, K.; Koufakis, T.; Kotsa, K.; Germanidis, G. How far beyond diabetes can the benefits of glucagon-like peptide-1 receptor agonists go? A review of the evidence on their effects on hepatocellular carcinoma. Cancers 2022, 14, 4651. [Google Scholar] [CrossRef]
  256. Zhang, C.Y.; Liu, S.; Yang, M. Antioxidant and anti-inflammatory agents in chronic liver diseases: Molecular mechanisms and therapy. World J. Hepatol. 2023, 15, 180–200. [Google Scholar] [CrossRef]
  257. Marrero, J.A.; Kulik, L.M.; Sirlin, C.B.; Zhu, A.X.; Finn, R.S.; Abecassis, M.M.; Roberts, L.R.; Heimbach, J.K. Diagnosis, staging, and management of hepatocellular carcinoma: 2018 practice guidance by the american association for the study of liver diseases. Hepatology 2018, 68, 723–750. [Google Scholar] [CrossRef]
  258. Miallot, R.; Galland, F.; Millet, V.; Blay, J.Y.; Naquet, P. Metabolic landscapes in sarcomas. J. Hematol. Oncol. 2021, 14, 114. [Google Scholar] [CrossRef]
Figure 1. Pathogenesis spectrum of MASLD–HCC. Abbreviations: MASLD, metabolic dysfunction-associated steatotic liver disease; MASH, metabolic dysfunction-associated steatohepatitis; HCC, hepatocellular carcinoma.
Figure 1. Pathogenesis spectrum of MASLD–HCC. Abbreviations: MASLD, metabolic dysfunction-associated steatotic liver disease; MASH, metabolic dysfunction-associated steatohepatitis; HCC, hepatocellular carcinoma.
Metabolites 14 00325 g001
Figure 2. Several types of epigenetic modifications in MASLD-induced HCC. Abbreviations: DNMTs, DNA methyltransferases.
Figure 2. Several types of epigenetic modifications in MASLD-induced HCC. Abbreviations: DNMTs, DNA methyltransferases.
Metabolites 14 00325 g002
Figure 3. Three types of metabolic reprogramming (A) and dysregulation of metabolic pathways in MASLD-HCC (B). As depicted in the literature review, oncogenic pathways are triggered by alterations in various metabolic pathways at the onset of neoplasia. These modifications at the cellular and metabolic levels provide cancer cells with advantages to support their rapid growth and proliferation in response to the hostile tumor environment. Altered regulation of FA and cholesterol metabolism in HCC related to MASLD. Fatty acid β-oxidation is suppressed for adaptation to a lipid-rich environment. Acetyl-CoA is converted into FAs through lipogenesis. HCC cells demonstrate heightened glucose absorption and rely on aerobic glycolysis as their primary energy source. Rather than being oxidized in the mitochondria, pyruvate is predominantly converted into lactate. The increased glucose uptake also meets the substrate demands of the PPP. This is important for nucleotide biosynthesis and nucleic acid replication. Dysregulation of genes and metabolic intermediates involved in amino acid and glutamine metabolism occurs in HCC. This dysregulation can be triggered by aberrantly activated oncogenes and the loss of tumor suppressors, such as ncRNAs. Abbreviation: FFA, free fatty acids; FATPs, fatty acid transport proteins; MUFAs, monounsaturated fatty acids; PUFAs, polyunsaturated fatty acids; SCD, stearoyl-CoA desaturase; FAs, fatty acid; FASN, fatty acid synthase; Glu, glucose; MCTs, monocarboxylate transporters; GLUT1, glucose transporter 1; PPP, pentose phosphate pathway; LDHA, lactate dehydrogenase; Gln, glutamine; ASCT2, alanine-serine-cysteine transporter 2; BCAAs, branched-chain amino acids; SLC7A5, solute carrier family 7 member 5; GLS1, glutaminase 1; OAA, oxaloacetate; TCA cycle, tricarboxylic acid cycle; α-KG, α-ketoglutarate; GDH, glutamate dehydrogenase; TA, transglutaminase; DNL, De novo lipogenesis; PPP, pentose phosphate pathway.
Figure 3. Three types of metabolic reprogramming (A) and dysregulation of metabolic pathways in MASLD-HCC (B). As depicted in the literature review, oncogenic pathways are triggered by alterations in various metabolic pathways at the onset of neoplasia. These modifications at the cellular and metabolic levels provide cancer cells with advantages to support their rapid growth and proliferation in response to the hostile tumor environment. Altered regulation of FA and cholesterol metabolism in HCC related to MASLD. Fatty acid β-oxidation is suppressed for adaptation to a lipid-rich environment. Acetyl-CoA is converted into FAs through lipogenesis. HCC cells demonstrate heightened glucose absorption and rely on aerobic glycolysis as their primary energy source. Rather than being oxidized in the mitochondria, pyruvate is predominantly converted into lactate. The increased glucose uptake also meets the substrate demands of the PPP. This is important for nucleotide biosynthesis and nucleic acid replication. Dysregulation of genes and metabolic intermediates involved in amino acid and glutamine metabolism occurs in HCC. This dysregulation can be triggered by aberrantly activated oncogenes and the loss of tumor suppressors, such as ncRNAs. Abbreviation: FFA, free fatty acids; FATPs, fatty acid transport proteins; MUFAs, monounsaturated fatty acids; PUFAs, polyunsaturated fatty acids; SCD, stearoyl-CoA desaturase; FAs, fatty acid; FASN, fatty acid synthase; Glu, glucose; MCTs, monocarboxylate transporters; GLUT1, glucose transporter 1; PPP, pentose phosphate pathway; LDHA, lactate dehydrogenase; Gln, glutamine; ASCT2, alanine-serine-cysteine transporter 2; BCAAs, branched-chain amino acids; SLC7A5, solute carrier family 7 member 5; GLS1, glutaminase 1; OAA, oxaloacetate; TCA cycle, tricarboxylic acid cycle; α-KG, α-ketoglutarate; GDH, glutamate dehydrogenase; TA, transglutaminase; DNL, De novo lipogenesis; PPP, pentose phosphate pathway.
Metabolites 14 00325 g003
Figure 4. By targeting specific epigenetic modifications, researchers can develop a variety of small molecule inhibitors. Abbreviation: HAT, histone acetyltransferase; Ac, acetylation; Me, methylation; HMT, histone methyltransferase; HDM, histone demethylase; HDAC, histone deacetylase; SAM, s-adenosylmethionine; SAH, s-adenosine homocysteine; DNMT, DNA methyltransferase.
Figure 4. By targeting specific epigenetic modifications, researchers can develop a variety of small molecule inhibitors. Abbreviation: HAT, histone acetyltransferase; Ac, acetylation; Me, methylation; HMT, histone methyltransferase; HDM, histone demethylase; HDAC, histone deacetylase; SAM, s-adenosylmethionine; SAH, s-adenosine homocysteine; DNMT, DNA methyltransferase.
Metabolites 14 00325 g004
Figure 5. Barcelona Clinic Liver Cancer (BCLC) staging and classification.
Figure 5. Barcelona Clinic Liver Cancer (BCLC) staging and classification.
Metabolites 14 00325 g005
Table 1. Pathogenic mechanisms of MASLD-HCC progression.
Table 1. Pathogenic mechanisms of MASLD-HCC progression.
Involved Pathogenic MechanismsOutcomesReferences
LipotoxicityThe molecular mechanism by which CD36 accelerated the progression of HCC was to promote the expression of AKR1C2 and thus enhance FA intake[17]
Oxidative DNA damageOxidative stress and 8-OHdG formation in the DNA in mice liver cells are the important characteristics of MASH-associated hepatocarcinogenesis[19]
MitophagyHepatic mitochondrial depolarization occurs early in mice fed a Western diet, followed by increased mitophagic burden, suppressed mitochondrial biogenesis and dynamics, and mitochondrial depletion, which ultimately contributes to the progression of MASH toward HCC[22]
Genetic factorA rare genetic variant in the gene MTTP has been identified as responsible for the development of progressive MASLD in a four-generation family with no typical disease risk factors[25]
Genetic factorMutational profiling of MASH-HCC tumors revealed TERT promoter (56%), CTNNB1 (28%), TP53 (18%), and ACVR2A (10%) as the most frequently mutated genes. ACVR2A mutation rates were higher in MASH-HCC than in other HCC aetiologies (10% vs. 3%, p < 0.05)[26]
Genetic factorSomatic mutations in MASH mice did not reveal increased tumorigenesis[27]
Dysbiosis in gut microbiotaDietary cholesterol drives NAFLD-HCC formation by inducing the alteration of gut microbiota and metabolites in mice[28]
Other mechanismsThe activation of the RAS/RAF/MEK/ERK pathway may contribute to HCC development; KRAS activation downstream of c-Met is sufficient to induce clinically relevant HCC in cooperation with mutant β-catenin.[30,33]
Other mechanismsThe dynamic shift in HSC subpopulations and their mediators during MASLD is associated with a switch from HCC protection to HCC promotion[35]
Abbreviation: AKR1C2, aldo-keto reductases family 1 member C2; FAs, fatty acids; 8-OHdG, 8-hydroxy-2′-deoxyguanosine; MTTP, microsomal triglyceride transfer protein; TERT, telomerase reverse tranase; CTNNB1, beta-catenin; ACVR2A, activin receptor type 2A; RAS, rat sarcoma virus; RAF, rapidly accelerated fibrosarcoma; MEK, mitogen-activated protein kinase; ERK, extracellular signal-regulated kinase; HSC, hepatic stellate cell.
Table 4. List of some potential anti-cancer drugs targeting epigenetics and metabolism.
Table 4. List of some potential anti-cancer drugs targeting epigenetics and metabolism.
AgentsDevelopmental StageFunctions in HCCTargetReference/Clinical Trial Number
Epigenetic drugs
Guadecitabine (SGI-110)Pre-clinicalAnti-proliferative effects on HCC cell lines; enhances the sensitivity of immune checkpoint inhibitors in organismsDNMT[198,199]
PanobinostatPre-clinicalTriggers apoptosis, reprograms cancer cell metabolism, and mitigates tumor angiogenesisHDAC[201]
GSK5959Pre-clinicalSuppresses HCC cell growthBRPF1[202]
AR42 + TelomelysinPre-clinicalAR42 reduced Telomelysin-induced phospho-Akt activation and enhanced Telomelysin-induced apoptosisHDAC[204]
JIB-04, GSK-J4, SD-70Pre-clinicalAttenuates HCC aggressiveness and viabilityJmjC lysine HDM[206,207]
AntimiR-494 oligonucleotides + SorafenibPre-clinicalInhibition of HCC cells to a glycolytic phenotypemiR-494[209]
lncRNA-PRALPre-clinicalInhibits HCC growth and induces apoptosisp53[211]
PLGA-based nanoplatform encapsulating LINC00958 siRNAPre-clinicalInhibits HCC lipogenesis and progressionLINC00958[73]
STM2457Pre-clinicalElicits a profound anti-tumor immune responseMETTL3[72]
5AC + Meticrane; CUDC-101 + Meticrane; ACY1215 + MeticranePre-clinicalInhibited the viability of liver cancer cellsDNMT1; HDACs; HDAC6[215]
Guadecitabine (SGI-110) + Sorafenib + OxaliplatinPhase II-DNMTNCT01752933
DecitabinePhase II-DNMT[237]
BIX-01294 + Decitabine--HMT G9a + DNMT[238]
BelistatPhase I/II-HDAC[239]
Resimonstat + SorafenibPhase I/II-HDAC[240]
Agents targeted metabolism
TVB3664Pre-clinicalAmeliorates the fatty liver phenotype in the aged mice and AKT-induced hepatic steatosisFASN[216]
OrlistatPre-clinicalDisrupts HCC’s metabolic reprogrammingFASN[217]
Obeticholic acidPre-clinicalAttenuates the development and progression of NASH-dependent HCC, possibly by interfering with SOCS3/Jak2/STAT3 pathwayFXR[218]
SGC707Pre-clinicalPrevents PRMT3-mediated LDHA methylation and indirectly suppresses glycolysis, as well as HCC progressionPRMT3[222]
BAY-876Pre-clinicalInhibits glucose uptake, proliferation, and EMT of HCCGlut1[223]
Ilicicolin HPre-clinicalInhibits the lactate production and glucose uptake of HCC cellsPGK1[224]
OleuropeinPre-clinicalInhibits HCC cell glycolysisG6PI[225]
ErianinPre-clinicalImpairs glycolysis and induces oxidative stressPC[226]
DeoxyelephantopinPre-clinicalInhibits glycolysis and reduces glucose uptake and lactic acid productionPI3K/Akt/mTOR/HIF-1α pathway[227]
HuaChanSuPre-clinicalInhibits PPP fluxG6PD[228]
MetforminPre-clinicalMitigates the Warburg effect and promotes FAOPFK1[230,231]
CanagliflozinPre-clinicalInhibits glucose uptake in HCC cellsSGLT2[232]
CP-91149Pre-clinicalInhibits glycogenolysis, interfering with glycolysis and the PPPPG[234]
Epalrestat/NAR1-29Pre-clinicalCombats diabetes-induced MASLD and even HCCAR[65]
CB-839 + V-9302Pre-clinicalTargets the metabolic vulnerability of glutamine-dependent HCCGLS and ASCT2[235]
TVB-2640Phase IIReduces liver fat and improves biochemical biomarkersFASN[242]
PF-5221304Phase II-ACCNCT04321031
Oroxylin APhase IInhibits non-oxidative PPP and triggers p53 signalingTKChiCTR2100051434
Olutasidenib (FT-2102)Phase Ib/II-IDH1NCT03684811
Abbreviations: DNMT, DNA methyltransferase; HDAC, histone deacetylase; BRPF1, bromodomain and PHD finger containing 1; HDM, histone demethylase; METTL3, methyltransferase 3; HMT, histone methyltransferase; FXR, farnesoid X receptor; PRMT3, protein arginine methyltransferase 3; LDHA, lactate dehydrogenase A; EMT, epithelial-to-mesenchymal transition; PGK1; phosphoglycerate kinase 1; SGLT2, sodium-glucose transport protein 2; PG, protein-glutaminase; AR, aldose reductase; GLS, glutaminase-1; ASCT2, alanine-serine-cysteine transporter 2; ACC, acetyl CoA carboxylase; TK, transketolase; IDH1, isocitrate dehydrogenase 1.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, A.; Wang, R.; Zhao, Y.; Zhao, P.; Yang, J. Crosstalk between Epigenetics and Metabolic Reprogramming in Metabolic Dysfunction-Associated Steatotic Liver Disease-Induced Hepatocellular Carcinoma: A New Sight. Metabolites 2024, 14, 325. https://doi.org/10.3390/metabo14060325

AMA Style

Li A, Wang R, Zhao Y, Zhao P, Yang J. Crosstalk between Epigenetics and Metabolic Reprogramming in Metabolic Dysfunction-Associated Steatotic Liver Disease-Induced Hepatocellular Carcinoma: A New Sight. Metabolites. 2024; 14(6):325. https://doi.org/10.3390/metabo14060325

Chicago/Turabian Style

Li, Anqi, Rui Wang, Yuqiang Zhao, Peiran Zhao, and Jing Yang. 2024. "Crosstalk between Epigenetics and Metabolic Reprogramming in Metabolic Dysfunction-Associated Steatotic Liver Disease-Induced Hepatocellular Carcinoma: A New Sight" Metabolites 14, no. 6: 325. https://doi.org/10.3390/metabo14060325

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop