Next Article in Journal
Engineered Bone Marrow Stem Cell-Sheets Alleviate Renal Damage in a Rat Chronic Glomerulonephritis Model
Previous Article in Journal
The Relationship between Iron and LRRK2 in a 6-OHDA-Induced Parkinson’s Disease Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metabolism as a New Avenue for Hepatocellular Carcinoma Therapy

1
Interdisciplinary Department of Medicine, School of Medicine, University of Bari, Piazza G. Cesare 11, 70124 Bari, Italy
2
Department of Life, Health and Environmental Sciences, University of L’Aquila, 67100 L’Aquila, Italy
*
Authors to whom correspondence should be addressed.
These authors equally contributed to this work.
Int. J. Mol. Sci. 2023, 24(4), 3710; https://doi.org/10.3390/ijms24043710
Submission received: 14 January 2023 / Revised: 8 February 2023 / Accepted: 10 February 2023 / Published: 13 February 2023
(This article belongs to the Special Issue New Insights into Hepatocellular Carcinoma)

Abstract

:
Hepatocellular carcinoma is today the sixth leading cause of cancer-related death worldwide, despite the decreased incidence of chronic hepatitis infections. This is due to the increased diffusion of metabolic diseases such as the metabolic syndrome, diabetes, obesity, and nonalcoholic steatohepatitis (NASH). The current protein kinase inhibitor therapies in HCC are very aggressive and not curative. From this perspective, a shift in strategy toward metabolic therapies may represent a promising option. Here, we review current knowledge on metabolic dysregulation in HCC and therapeutic approaches targeting metabolic pathways. We also propose a multi-target metabolic approach as a possible new option in HCC pharmacology.

Graphical Abstract

1. Hepatocellular Carcinoma

The liver is a central organ of metabolic homeostasis as it is responsible for the detoxification process of xenobiotics, bile acid production, regulation of lipid metabolism, glycogen storage, amino acid biosynthesis, and nitrogen waste disposal through the urea cycle. In addition, the liver is involved in supporting immune function [1]. The predominant cell type in the liver parenchyma is the hepatocyte, which accounts for approximately 85% of the liver mass. Hepatocytes perform most of the metabolic functions of the liver.
Hepatocellular carcinoma (HCC), the most common primary cancer of the liver that originates in hepatocytes, is the sixth most commonly diagnosed neoplasm and the third leading cause of cancer-related death worldwide. From an epidemiological point of view, the incidence of HCC has decreased in some countries, such as China, Japan, and South Korea, while it has increased in Western countries [2,3]. Despite the reduced incidence of chronic hepatitis infections, recent projections estimate that HCC will rise to the third principal cause of death in Western countries by 2030 [4], and between 2020 and 2040, an increase of 55% is expected in the number of new HCC cases [2,3]. This alarming epidemiology can be explained by the increased incidence of metabolic diseases, such as the metabolic syndrome (MetS), diabetes, obesity, and nonalcoholic fatty liver disease (NAFLD) [5]. A direct link between the onset of NAFLD and the development of HCC has been reported [6]. Current therapeutic interventions for HCC include surgery, transarterial procedures (including TACE and TARE), and tyrosine kinase inhibitors (i.e., sorafenib, regorafenib, and lenvatinib), often in combination with immunotherapy (i.e., pembrolizumab and nivolumab) [5,7]. Despite these interventions, the prognosis of HCC patients is often poor, and pharmacological approaches are invasive and display many side effects. For this reason, a broader understanding of HCC pathogenesis is needed to develop efficacious and less aggressive pharmacological tools. To this end, the study of metabolism can provide concrete opportunities and provide interesting insights into this topic, as has happened in recent years.

2. Cancer and Metabolism

Genetics dominated cancer research for a long time after the pioneering work of Otto Warburg and Albert Szent-Gyorgyi in the 1930s [8,9,10,11,12]. Dysregulation of cellular metabolism has now reemerged as an important driver of cancer [13]. After decades of research primarily focused on tumor genetics, tumor cell metabolism has received renewed attention [13]. Currently, the scientific community is paying great attention to cellular metabolism and metabolic plasticity in tumor cell biology and has investigated the role of specific metabolic pathways in tumorigenesis [14,15]. Glycolysis, lactic acid fermentation, and oxidative phosphorylation (OXPHOS) are key metabolic processes that are now widely studied and considered targets for inhibiting tumor growth and overcoming drug resistance. In general, normal cells rely on the metabolic axis glucose→pyruvate→TCA/OXPHOS (oxidative glycolysis), whereas tumor cells shift towards lactic fermentation, glucose→pyruvate→lactate (fermentative glycolysis). This “fermentative signature” characterizes the Warburg effect, which is aerobic glycolysis or lactic acid fermentation regardless of the presence of hypoxic conditions [16]. Tumor cells primarily rely on the Warburg effect to generate energy, although this process is energetically inefficient (36 versus 4 ATP molecules/mol glucose) [17]. The switch to fermentative glycolysis provides several advantages to tumor cells, such as fast biosynthesis of ATP molecules, an increased supply of biosynthesis intermediates [17], and reduced generation of reactive oxygen species (ROS), which protects cancer cells from apoptosis [18]. Two steps are finely tuned: the conversion of pyruvate into lactate, which is catalyzed by lactate dehydrogenase (LDH), and the activation of OXPHOS.
What triggers the Warburg effect is still being debated. Hypoxia, considered an activating stimulus, has recently been ruled out as a major factor, as some tumors develop the Warburg effect even under normoxic conditions [17]. It has been suggested that the Warburg phenotype is activated by an increased demand for NAD+ that is not matched with an adequate supply of ATP [19]. In general, it is now believed that the very early activation step of the Warburg effect could be an alteration of the metabolic control necessary for tumor cells to acquire a cell-autonomous nutrient uptake and an anabolic profile. Nevertheless, this initial trigger has not been fully elucidated. The mechanisms by which the Warburg effect supports drug resistance are also still under debate. It is known that the glycolytic environment can influence the response of cancer cells to drugs [20]. The relevance of the Warburg effect in tumorigenesis and drug resistance is still under debate [18,20], and the role of OXPHOS in cancer biology remains controversial. For example, some studies have shown that defective OXPHOS promotes tumorigenicity in hepatocellular carcinoma (HCC) [21,22], while others have reported inhibition of OXPHOS as a therapeutic strategy for HCC [23,24]. Similar differences have been described in other types of malignancies [25,26].

3. Hepatocellular Carcinoma and Metabolism

Metabolic dysregulation is emerging as an important risk factor for HCC [6]. As in other tumors, the neoplastic transformation of a normal hepatocyte is characterized by a general metabolic rewiring. Transformed hepatocytes display the Warburg phenotype, which in the liver is characterized by rearrangements in the expression and phosphorylation of key glycolytic enzymes [27]. Indeed, more than fifty years ago, metabolic reprogramming to enhance glycolysis was shown to be central to transforming hepatocytes [28]. Glucose metabolism rewiring can find a mechanistic explanation in the hypoxic milieu in which HCC lies. In fact, in normal hepatic tissue, the median O2 partial pressure is about 30 mmHg, whereas in the HCC intratumor region, it is roughly 6 mmHg [29,30,31]. In line with this, HIF-1α is highly expressed in HCC [32]. In this hypoxic microenvironment, LDH activity is increased [33] and the pyruvate dehydrogenase complex (PDH) is reduced [34], thus obtaining an increase in lactate production and a dampening of the conversion of pyruvate into acetyl-CoA, which contribute to the instauration of a glycolytic phenotype.

3.1. Glucose Metabolism Rewiring

Glucose homeostasis is rearranged in HCC, leading to a change in the expression and activity of transporters and enzymes, which in turn leads to a rewiring of the metabolite flux. The Warburg phenotype in HCC is not characterized by increased gluconeogenesis because fructose 1,6-bisphosphatase 1 (FBP1) and phosphoenolpyruvate carboxykinase 1 and 2 (PEPCK 1 and 2) expression are reduced [35]. In addition, the expression of glucose transporters GLUT1 and GLUT2 is increased in HCC, and while GLUT2 overexpression worsens the prognosis of patients, the inhibition of GLUT1 expression attenuates the malignant behavior of HCC cells [36,37,38]. The conversion of glucose into glucose-6-phosphate (G6Ph) is a highly regulated metabolic reaction catalyzed by the hexokinase (HK) enzyme, which has five isoforms: HK1 to HK4 and the hexokinase domain containing 1 (HKDC1). The expression of HK2, the prevalent isoform in HCC, is associated with reduced survival in HCC patients [39], and, accordingly, HK2 inhibition dampens glycolysis in favor of OXPHOS, relieving the tumor phenotype [40]. HKDC1 expression has also been shown to exacerbate HCC aggressiveness and prognosis [41].
The glycolytic phenotype is strengthened by the modification of the expression and activity of other enzymes. For example, glyceraldehyde-3-phosphate dehydrogenase (GAPDH) [42,43] and pyruvate kinase 2 (PKM2) [44] are upregulated in HCC.
HCC also exhibits an increased dependence on glutamine metabolism [45,46]. Glutamine undergoes different metabolic routes in HCC. For example, glutaminolysis leads to α-ketoglutarate (α-KG) production, which supports the TCA cycle for energy production in the context of low-efficiency OXPHOS. Moreover, α-KG can be converted into citrate, thus fueling the biosynthesis of lipids. However, there is no consensus regarding the function of glutamine in HCC. Some authors proposed a general glutaminolysis fate with a negligible contribution to reductive carboxylation, which leads to lactate production [40]. This notion was not supported by other authors who found inhibition of the pathway leading to α-KG production [35], possibly reflecting the intrinsic heterogeneity of tumor tissue [47]. LDH is a key enzyme in metabolic reprogramming in many tumors and HCC. Five isoforms have been described in humans, and LDHA and LDHB are the most expressed isoforms in the liver. In HCC, serum LDH is a reliable tumor biomarker [48,49]. Lactate is central to metabolic rearrangements in most tumors and HCC, and elevated serum levels have been observed in HCC patients [50]. In HCC, part of lactate is secreted in the extracellular medium by monocarboxylate transporters (MCT), in particular by MCT4, whose expression is correlated with proliferation through Ki-67 [51]. This process is regulated by the transmembrane protein CD147, which elicits MCT1 expression, which in turn increases lactate externalization [52]. MCT1 expression, just like MCT4, promotes HCC growth [52]. Higher plasma levels of lactate have been detected in patients with NASH compared with healthy subjects, supporting the potential role of lactate as an active metabolic mediator of the disease [53].

3.2. Pentose Phosphate Pathway

The Pentose Phosphate Pathway (PPP) plays a central role in the metabolic reprogramming of tumor cells and HCC. The main outcome of the PPP is the production of ribose-5-phosphate, which is the precursor for the synthesis of nucleotides that are needed for tumor cell growth. In addition, NADPH is another product of the PPP that sustains lipid biosynthesis and keeps the tumor cell in an antioxidant environment by maintaining thioredoxin and glutathione in a reduced state. This may help tumor cells protect themselves from apoptotic cell death and develop resistance to drug therapy [54]. Although the levels of ribulose 5-phosphate (Ru5P) and ribose 5-phosphate (R5P) are decreased due to their rapid utilization by cancer cells, the transcription of most enzymes involved in the PPP is elicited in HCC [55]. Among the enzymes that are involved in PPP rewiring, glucose-6-Ph dehydrogenase (G6PDH) should be mentioned. The expression of this enzyme was shown to result in increased metastasis, drug resistance, and decreased survival in HCC. Notably, HBV protein X (HBx) triggers G6PDH expression through a Nrf2-mediated mechanism [56]. Furthermore, in non-oxidative PPP, Nrf-2 upregulates transketolase (TKT) by reducing oxidative stress and increasing purine biosynthesis during hepatocarcinogenesis [57].

3.3. TCA Cycle Rewiring

The TCA cycle produces NADH and FADH2, which fuel OXPHOS activity. The TCA cycle is dysregulated at several points in HCC. Pyruvate dehydrogenase kinase 4 (PDK4), which phosphorylates and inhibits pyruvate dehydrogenase (PDH), is downregulated in HCC and associated with increased lipid biosynthesis and poor prognosis [58]. Succinate dehydrogenase (SDH) [succinate→fumarate] and fumarate hydratase (FH) [fumarate→malate] are deficient in HCC. Moreover, succinate and fumarate-mediated stabilization of HIF-α enhances glycolysis and angiogenesis [59]. Interestingly, SDHB deficiency promotes the malignancy of HCC, triggering the Warburg effect [60].

3.4. OXPHOS Rewiring

In normal cells, OXPHOS utilizes NADH and FADH2 from the TCA cycle for ATP production. Several pieces of evidence associate defective OXPHOS with enhanced HCC development through different mechanisms, for example, mitoribosome defects or by TGFβ-mediated reduction of OXPHOS without the involvement of the glycolytic pathway [21,22]. HCV-mediated reduced expression of OXPHOS protein subunits leads to a Warburg-like phenotype, eventually promoting HCC development [61]. Membrane trafficking-involved Rab3A GTPase is overexpressed and also post-translationally modified by the addition of N-acetylglucosamine (O-GlcNAcylated) in HCC. Usually, unmodified Rab3 inhibits the metastatic process in HCC by boosting OXPHOS, but O-GlcNAcylation impedes this function [62]. Notably, mitochondrial miR-181a-5p is upregulated in HCC, and this results in a downregulation of cytochrome B and cytochrome C oxidase 2 (COX2) with consequent impairment of the electron transport chain. Furthermore, glucose metabolism is reprogrammed by increasing the expression of the glucose transporters GLUT1 and hexokinase 2 (HK2), leading to high glucose uptake and upregulation of LDH activity [63]. Interestingly, the correlation between cell stemness, metabolism, and HCC is provided by the homeobox transcription factor NANOG2, whose expression is increased in alcohol- and obesity/HCV-induced HCC mouse models as well as in human tumor-initiating cells, where NANOG inhibits OXPHOS [64].

3.5. Lipid Metabolism Rewiring

Lipid metabolism is one of the important metabolic tasks accomplished by the liver. Altered lipid metabolism homeostasis is responsible for many pathologic conditions. Dyslipidemias may lead to diabetes, obesity, and liver steatosis, and all these conditions have been correlated with HCC onset [65,66]. De novo lipogenesis (DNL) is very well regulated, and several hindrances have been reported for DNL in HCC. In general, DNL follows this route: pyruvate→acetyl-CoA+oxaloacetate (OAA)→citrate [inside mitochondria]. In conditions of carbohydrate abundance, citrate is exported into the cytoplasm where ATP-citrate lyase (ACLY) converts citrate into acetyl-CoA+OAA. Then, lipid biosynthesis takes place in the cytoplasm, mainly by acetyl-CoA carboxylase (ACC) [acetyl-CoA→malonyl-CoA] and fatty acid synthase (FASN) [malonyl-CoA→palmitoyl-CoA] [67]. DNL works in close association with the PPP to produce the necessary NADPH. DNL deregulations have been reported in HCC-predisposing conditions, such as NAFLD [68]. In addition, increased expression of several DNL enzymes as well as enzymes involved in NADPH biosynthesis, such as G6PD and malic enzyme, has been reported [35]. FASN overexpression is associated with increased cancer growth and survival and a worse prognosis in patients [68,69]. The acetyl-CoA can also derive from acetate, which is imported into cells by MCTs transporters. Acetyl-CoA synthetase enzymes (ACSSs) then convert acetate into acetyl-CoA. ACSS2 is present in the cytosol, while ACSS1 and ACSS3 are mitochondrial isoforms. Notably, the expression of ACSS1 was increased and correlated with enhanced malignant features in HCC [35].
HCC uses DNL and exogenous fatty acids to meet its growth requirements [70], and increased free fatty acid uptake through the fatty acid translocase CD36 correlates with the initiation and progression of HCC [71].
The liver β-oxidizes fatty acids in either peroxisomes or mitochondria, where they are transported by carnitine palmitoyltransferase-I (CPT-I), which condenses carnitine with very long Acyl-CoA [72]. Interestingly, NASH-induced rats show reduced CPT-I activity as well as defective complexes I and II of the mitochondrial electron transport chain (ETC) [73,74]. Furthermore, HCC patients showed defective mitochondrial β-oxidation due to decreased expression of several enzymes involved [35]. Thus, impaired mitochondrial functions can be considered predisposing conditions for HCC development.
Cholesterol homeostasis is another important determinant of HCC development. The key liver enzyme that controls cholesterol biosynthesis is 3-hydroxy-3-methylglutaryl-CoA reductase (HMG-CoA reductase), whose reaction product is mevalonate. Some reports have described an association between HMG-CoA reductase/mevalonate and HCC. In fact, from a clinical point of view, several reports described statins as dampening the probability of developing HCC. Corroborating observations come from in vitro and mouse studies [75,76,77,78,79]. The pro-HCC effects of bile acids, cholic acid, and deoxycholic acid have also been described [80,81]. Phospholipid homeostasis is also involved in HCC pathogenesis. Cirrhotic rats showed altered activity of phospholipid biosynthesis enzymes [82]. Interestingly, HCC patients have high serum and tissue levels of phosphatidylcholine (PC) [83].

3.6. Nucleotide Metabolism Rewiring

Nucleotide metabolism is essential for nucleic acid biosynthesis, and this is particularly true for tumor cells that have an increased proliferative rate compared to healthy cells. In this regard, the PPP provides R5Ph, which is needed along with amino acids for nucleic acid biosynthesis. In HCC, the expression of several genes involved in this biosynthetic pathway is increased. In particular, three enzymes constitute the enzyme complex that catalyzes the initial reaction in pyrimidine biosynthesis: carbamoyl phosphate synthase 2, aspartate transcarbamylase, and dihydroorotase [84,85]. Furthermore, other rate-limiting enzymes, such as thymidylate synthase (TYMS), thymidine kinase 1 (TK1), and deoxythymidylate kinase (DTYMK), are upregulated in HCC and are associated with poor prognosis and cancer stemness [86,87,88]. The expression of most of the enzymes involved in purine biosynthesis is also increased in HCC [55,87,88].

3.7. Protein and Amino Acid Metabolism Rewiring

The liver is also a central hub for protein and amino acid homeostasis, so this imbalance due to pathological conditions reflects changes in protein and amino acid levels. For example, negative correlations between serum albumin levels and tumor diameter, tumor multifocality, portal vein thrombosis, and α-fetoprotein, have been reported [89]. The liver is also responsible for the biosynthesis of non-essential amino acids (NEAAs), which are involved in many biochemical pathways. In many tumors, NEAA production is increased due to the rewiring of glucose metabolism, which increases the demand for oxidizable substrates [90,91,92,93]. For example, in HCC patients, the levels of most NEAAs were significantly higher compared to healthy subjects, while branched-chain amino acids (BCAAs) were decreased compared to aromatic amino acids [94,95]. Interestingly, NEAA actively supports HCC by mediating adaptation to hypoxia [96]. It is worth mentioning here the behavior of three amino acids: glutamine, aspartic acid, and glycine. In HCC, glutamine and aspartate levels are elevated, and glycine levels are decreased [83].

3.8. Urea Cycle Rewiring

In HCC, most urea cycle enzymes are downregulated, along with low concentrations of urea cycle metabolites such as arginine, ornithine, and citrulline, which are common in advanced HCC stages. Two key enzymes in the urea cycle, argininosuccinate synthase 1 (ASS1) and carbamoyl phosphate synthase (CPS), show a hypermethylated state in HCC, which explains their reduced expression.

3.9. HCC Tumor Microenvironment Rewiring

Hepatocytes are the main cell type in the liver parenchyma, while the rest is represented by stromal cells, including endothelial cells, immune cells, Kupffer cells, and hepatic stellate cells. Stromal cells are also recruited by HCC cells in the tumor microenvironment (TME). Two important effector cells are tumor-associated macrophages (TAM) and myeloid-derived suppressor cells (MDSC), which are the main cell types found in the tumor infiltrate [97]. The crosstalk between cancer cell-released metabolites and immune cells sustains the tumor ecosystem. Indeed, upon an inflammatory challenge, macrophages acquire an inflammatory cytokine-producing M1-like phenotype, whereas anti-inflammatory stimuli lead to an immunosuppressive M2-like phenotype [98]. Hence, tumor onset is favored by macrophage-induced chronic inflammation. The immunosuppressive M2-like phenotype may then prevail, supporting growth, angiogenesis, and the metastatic process [99]. These observations are in line with the consolidated view that an inflammatory status promotes tumor onset. For example, tumor-secreted cytokines attract monocyte-derived TAM, which switches to a glycolytic fermentative metabolism and survives in the hypoxic TME [99,100]. Notably, M2-like TAMs may be involved in mediating drug resistance and promoting cancer stemness in HCC, resulting in a poorer prognosis for HCC patients [101,102]. Secreted lactate mediates the crosstalk between TAM and cancer cells. This promotes a VEGF-induced transition to an M2-like phenotype and inhibition of NK cells, by favoring MDSC [103,104]. MDSC-mediated immunosuppression can also be boosted by fatty acids, whose uptake is mediated by CD36 [105].

4. Metabolism-Based Pharmacology Approaches in HCC

Several pharmacological approaches have been and are being developed using the HCC metabolic features described above. Several chemicals are currently being investigated, some in their initial preclinical stages and others in clinical trials.

4.1. Inhibition of the Glycolytic Pathway and Pentose Phosphate Pathway

Several glycolysis inhibitors are currently being investigated as possible therapeutic molecules. The glucose analog 2-deoxy-D-glucose (2-DG) is converted to 2-DG-6-phosphate, which inhibits HK2 activity in a non-competitive manner. This inhibition is associated with reduced HCC growth and invasiveness. Moreover, 2-DG synergizes with the effect of sorafenib in HCC cell lines [106,107]. 3-Bromopyruvic acid (3-BrPA) inhibits several enzymes, including HK2, PGK1, succinate dehydrogenase, GAPDH, and LDH. Also, some studies have identified GAPDH as the principal target of 3-BrPA in HCC cell lines [43,108,109,110]. The plant-derived molecule koniginic acid also inhibits GAPDH by binding to the enzyme’s active site. However, the potential applicability of koniginic acid in HCC has not been well demonstrated [111]. Several metabolites from plants are considered potential inhibitors of glycolysis in HCC. Among them, chrysin, a flavonoid found in Chinese medicinal plants, can reduce the expression of HK2, reduce glucose uptake and lactate production, and ultimately trigger the mitochondrial apoptotic pathway. The anti-HCC effects of chrysin have also been described in a mouse model [112]. Likewise, the plant-stress metabolite methyl jasmonate reduces mitochondrial transmembrane potential by sequestering HK2 and thereby elicits apoptosis in HCC cell lines and mice [113]. The naphthoquinone shikonin from Lithospermum erythrorhizon displays anti-HCC effects and enhances the effect of sorafenib [114]. The plant-derived isoflavone genistein inhibits glycolysis by targeting HIFα, HK2, and GLUT1. In addition, genistein enhances in vitro and in vivo sensitivity to sorafenib [115]. Finally, the anthraquinone emodin, present in different plant species, inhibits glycolysis by reducing HK2, PKM2, and LDHA expression [116]. The thiamine antagonist oxythiamine, which inhibits transketolase (TKT), inhibits HCC cell growth and synergizes with sorafenib by eliciting ROS production in vitro and in vivo [57].

4.2. Lipid Metabolism Rewiring Inhibition

To date, little evidence exists to suggest the possibility of lipid metabolism as an effective therapeutic target in HCC. For example, the inhibition of the central lipogenic enzyme acetyl-CoA carboxylase (ACC) has been considered for pharmacological purposes. Indeed, some chemical inhibitors of ACC are now under investigation, such as ND-654, an allosteric inhibitor of ACC1 and ACC2, which also enhances the pharmacological effects of sorafenib [117]. Several fatty acid synthase (FASN) inhibitors, including orlistat, C93, C75, GSK2194069, and GSK837149A, have been evaluated in preclinical testing [118,119]. These compounds, however, show several side effects and significant toxicity. One stearoyl-CoA desaturase 1 (SCD1) inhibitor, A939572, was shown to be efficacious in HCC cell lines and to increase sorafenib sensitivity by modulating ER stress-induced differentiation [120]. The transcription factor Liver X receptor alpha (LXRα) is involved in lipid homeostasis and is under evaluation as an anti-HCC target. TO901317 has been reported to increase LXRα expression and decrease glucose uptake by inhibiting GLUT1 expression [121]. Finally, the squalene synthase (SQS) enzyme in the cholesterol biosynthetic pathway has been considered a druggable target for HCC. For example, YM-53601 was reported to inhibit SQS in vivo and synergize with doxorubicin to inhibit HCC growth [122]. Moreover, an ACLY inhibitor was recently shown to induce a pro-apoptotic effect in in vitro models of HCC [123]. No evidence is available so far about the potential toxicity of these last-mentioned compounds.
Of note, an important aspect to consider in this context is the role of circadian rhythms in the regulation of lipid metabolism and, hence, in controlling the development of predisposing dysmetabolic conditions leading to HCC, such as NAFLD. Indeed, dysregulation of the circadian clock is emerging as an important aspect in the pathogenesis of several diseases [124,125]. In particular, the role of NAFLD in the development of HCC has been recognized [126]. Thus, strategies aimed at targeting the circadian clock machinery can be considered as a possible option in the management of HCC and predisposing dysmetabolic conditions.

4.3. Other Potential Metabolic Targets for HCC

So far, chemicals targeting other metabolic pathways have not provided conclusive indications. For example, it has been proposed to target epigenetic regulation and ubiquitin-dependent degradation to increase fructose-1,6-bisphosphatase 1 (FBP1) expression, which is decreased in HCC [127]. Several molecules are under testing in this respect, such as histone deacetylase inhibitors or lysine-specific histone demethylase 1 inhibitors [128,129,130]. Still, these chemicals show low specificity and a pleiotropic action. Proteasome inhibitors and dexamethasone have also been suggested to regain FBP1 levels, but likewise, the scarce selectivity results in many side effects [131]. Glutaminase 1 (GLS1), which converts glutamine into glutamate, is a key enzyme in tumor cell metabolism because it can promote lipid and nucleotide biosynthesis and reduce ROS generation, thereby protecting tumor cells from apoptosis. Some GLS1 inhibitors have been tested, but they showed toxicity and low selectivity [132,133]. Chemical inhibitors of glutamine uptake are also under evaluation [134,135]. Among them, the plant metabolite berberine shows interesting effects in vitro [136]. Some chemicals targeting metabolism are now under clinical trial evaluation, but the results so far are not conclusive.
In this scenario, it is important to mention that there is growing interest in the potential use of AMP-activated protein kinase (AMPK) activators to improve metabolic abnormalities and better manage neoplastic conditions, including HCC [137]. In fact, some AMPK activators are now being considered as possible drugs for the treatment of metabolic disorders and cancer [138]. In HCC, a recent report suggested that AMPK activation shows therapeutic potential [139]. Still, caution is suggested since contrasting results have been reported for cancer therapy [140]. Interestingly, AMPK activation was shown to prevent thrombotic conditions by phosphorylation and consequent inactivation of ACC [141]. This can represent an important aspect since a prothrombotic imbalance was observed in NASH/cirrhosis, which predicts the development of HCC [142].
An overview of the metabolic targets in HCC, for which chemical inhibitors are available, is shown in Figure 1.

5. Multitarget Metabolic Systems

5.1. Plant-Based Multi-Pathway Approach

As mentioned above, metabolism-based single-target approaches still show several critical points in terms of efficacy and side effects. To this end, a synergistic approach targeting multiple metabolic pathways or processes may offer some advantages to achieve higher efficacy, a lower dose, and lower toxicity. For example, plant extracts, because their composition is a blend of several metabolites, can provide an interesting option to target several metabolic pathways at the same time. Plant extracts have not been explored much in this direction. However, several studies have reported interesting effects of plant extracts on several types of tumors, most likely due to synergistic effects targeting multiple pathways [143,144,145,146,147]. The plant-derived polyphenol molecule resveratrol acts on multiple cellular targets and exhibits a wide range of biological activities, including positive effects on various metabolic pathways [148]. For example, resveratrol reduces the utilization of glucose and amino acids for energy production and the release of lactic acid in HCC cells [149]. Furthermore, resveratrol-derived natural metabolites have multiple effects on glucose metabolism in hepatoma cells [150]. Recently, we investigated the role of Crithmum maritimum, an edible wild plant that grows spontaneously along the Mediterranean and Atlantic coasts. We performed an extensive screen for the chemical composition and biological effects of Crithmum maritimum harvested along the Apulian coast. Notably, we found a significant effect of an ethyl acetate extract of Crithmum maritimum in inhibiting HCC cell growth by exerting synergistic effects on several metabolic pathways involved in promoting HCC, such as glycolysis/Warburg effect, lipid and cholesterol metabolism, and amino acid metabolism. We also identified the composition of the extract, which contains high concentrations of falcarindiol and chlorogenic acid, both of which have anti-cancer properties [151]. Interestingly, Crithmum maritimum modulates the bioenergetic profile of HCC cells by increasing OXPHOS and decreasing LDH activity, and hence lactate production [152]. Moreover, Crithmum maritimum increases the expression of markers typical of healthy hepatocytes [153]. Of note, Crithmum maritimum can also be a valuable option to improve conventional pharmacological treatments with sorafenib [154]. Indeed, we found that Crithmum maritimum can be used in association with sorafenib to decrease the dose and the toxicity. This is an interesting example of how a multi-targeting approach in metabolism can improve conventional pharmacological cytotoxic approaches in HCC. As far as we know, this approach has not been much investigated, and more research is needed to fully uncover its actual potential.

5.2. Targets Exploiting Synergistic Effects on Metabolic Pathways

In addition, and in support of a plant-based multi-pathway approach, selective inhibition of receptors controlling metabolism may offer an additional translational opportunity for HCC pharmacological management. In this respect, we have recently demonstrated that LPA receptor 6 (LPAR6) is directly involved in the control of HCC cell metabolism [155] and that ectopic expression of LPAR6 in HCC cells drives sorafenib resistance by triggering a “metabolic switch”, which increases lactic acid fermentation at the expense of OXPHOS. This sorafenib resistance can be overcome by reducing lactic acid fermentation through inhibition of LPAR6 using our recently developed novel LPAR6 antagonists [156,157,158]. LPAR6 has a pro-tumorigenic role in HCC [159] by controlling the trans-differentiation of peritumoral tissue fibroblasts (PTFs) into carcinoma-associated fibroblasts (CAFs) [160] and its overexpression leads to a worse clinical outcome in HCC patients [159]. In this context, we emphasize the importance of the balance between fermentative glycolysis/OXPHOS in tumor development and cancer drug resistance [161].
Thus, metabolism-based interventions should be considered a valid option to support current therapies in patients with HCC. Current and proposed metabolism-based interventions are shown in Figure 2.

6. Conclusions and Future Perspectives

After several decades in which cancer research has been focused on tumor genetics and molecular biology, a new interest has arisen in the role of metabolism in driving and sustaining carcinogenesis. As discussed above, many biological pathways have been extensively studied, and several enzymes have been identified as potential therapeutic targets. Still, the complex intersections between the different pathways require a more structured, multitarget, process-directed approach aimed at targeting more processes at the same time to minimize possible “escape” strategies of the tumor. The biological mechanisms by which normal cells become transformed and acquire a tumor phenotype are still not fully elucidated. This is an essential step for the basic knowledge of how cells work and how they start to not work properly. From this perspective, a deeper understanding of how cellular adaptation is driven by metabolic processes is central.
HCC is becoming a serious health issue, especially in Western countries, and there is an extreme need for novel and effective therapeutic approaches. Many biochemical pathways are dysregulated in HCC, and a richer knowledge of these fields has the concrete potential to disclose new perspectives for HCC prevention and therapy. In this context, targeting single enzymes involved in the regulation of metabolic pathways can be an option. Still, so far, the results obtained are not conclusive in terms of efficacy and tolerability. This may be due to the intrinsic pleiotropic capacity of tumor cells to bypass a single inhibitory step. For this reason, a synergistic multitarget/multi-process approach can prove useful. In this respect, plant extracts may represent a valuable tool due to their blended composition, which can target many processes at the same time. Coupled with this approach, selective targeting of specific metabolism-controlling receptors may offer a supplemental option, as described above for LPAR6.
In conclusion, alterations in cellular metabolism are important drivers of the tumorigenesis process and the promotion of drug resistance [162,163,164,165]. In this context, the balance between glycolysis and OXPHOS is a central point, and the mechanisms by which this is regulated are still not fully understood. We believe that the complex interplay between tumor cell metabolism and drug resistance is an important aspect, and that the balance between OXPHOS and fermentative glycolysis is critical for determining the cellular behavior during neoplastic transformation and the acquisition of drug resistance by tumor cells. Furthermore, it is worth mentioning that, in our opinion, modulating the balance between OXPHOS and lactic acid fermentation to trigger OXPHOS and inhibit fermentation could be considered a strategy to reduce tumor invasiveness and drug resistance. Indeed, we recently demonstrated that inhibition of OXPHOS in favor of lactic acid fermentation elicits drug resistance in HCC [156]. Based on this observation, approaches to modulate the metabolism of fermenting tumor cells may provide tools to overcome cancer drug resistance. In particular, we propose the activation of OXPHOS and the inhibition of lactic acid fermentation/Warburg phenotype to reduce tumor growth and drug resistance, as well as a pro-differentiation strategy. In line with this principle, this “metabolic approach” could be independent of the genetic background of each tumor, and this could be exploitable for therapeutic purposes in different types of malignancies.

Author Contributions

Conceptualization, D.G., M.M. and A.M.; writing—original draft, D.G., M.M. and A.M.; writing—review & editing, D.G., C.S., M.M. and A.M.; supervision, M.M. and A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Trefts, E.; Gannon, M.; Wasserman, D.H. The liver. Curr. Biol. 2017, 27, R1147–R1151. [Google Scholar] [CrossRef] [PubMed]
  2. Rumgay, H.; Ferlay, J.; de Martel, C.; Georges, D.; Ibrahim, A.S.; Zheng, R.; Wei, W.; Lemmens, V.; Soerjomataram, I. Global, regional and national burden of primary liver cancer by subtype. Eur. J. Cancer 2022, 161, 108–118. [Google Scholar] [CrossRef] [PubMed]
  3. Rumgay, H.; Arnold, M.; Ferlay, J.; Lesi, O.; Cabasag, C.J.; Vignat, J.; Laversanne, M.; McGlynn, K.A.; Soerjomataram, I. Global burden of primary liver cancer in 2020 and predictions to 2040. J. Hepatol. 2022, 77, 1598–1606. [Google Scholar] [CrossRef] [PubMed]
  4. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef]
  5. Forner, A.; Reig, M.; Bruix, J. Hepatocellular carcinoma. Lancet 2018, 391, 1301–1314. [Google Scholar] [CrossRef]
  6. Agosti, P.; Sabba, C.; Mazzocca, A. Emerging metabolic risk factors in hepatocellular carcinoma and their influence on the liver microenvironment. Biochim. Biophys. Acta Mol. Basis Dis. 2018, 1864, 607–617. [Google Scholar] [CrossRef]
  7. Kudo, M. Systemic Therapy for Hepatocellular Carcinoma: Latest Advances. Cancers 2018, 10, 412. [Google Scholar] [CrossRef]
  8. Warburg, O.; Wind, F.; Negelein, E. The Metabolism of Tumors in the Body. J. Gen. Physiol. 1927, 8, 519–530. [Google Scholar] [CrossRef]
  9. Warburg, O. On the origin of cancer cells. Science 1956, 123, 309–314. [Google Scholar] [CrossRef]
  10. Szent-Gyorgyi, A.; Hegyeli, A.; Mc, L.J. Cancer therapy: A possible new approach. Science 1963, 140, 1391–1392. [Google Scholar] [CrossRef]
  11. Szent-Gyorgyi, A. The living state and cancer. Physiol. Chem. Phys. 1980, 12, 99–110. [Google Scholar] [CrossRef] [PubMed]
  12. Szent-Gyorgyi, A. Bioelectronics and cancer. J. Bioenerg. 1973, 4, 533–562. [Google Scholar] [CrossRef]
  13. Fendt, S.M.; Frezza, C.; Erez, A. Targeting Metabolic Plasticity and Flexibility Dynamics for Cancer Therapy. Cancer Discov. 2020, 10, 1797–1807. [Google Scholar] [CrossRef] [PubMed]
  14. Hajaj, E.; Sciacovelli, M.; Frezza, C.; Erez, A. The context-specific roles of urea cycle enzymes in tumorigenesis. Mol. Cell 2021, 81, 3749–3759. [Google Scholar] [CrossRef] [PubMed]
  15. Endicott, M.; Jones, M.; Hull, J. Amino acid metabolism as a therapeutic target in cancer: A review. Amino Acids 2021, 53, 1169–1179. [Google Scholar] [CrossRef] [PubMed]
  16. Vaupel, P.; Multhoff, G. Revisiting the Warburg effect: Historical dogma versus current understanding. J. Physiol. 2021, 599, 1745–1757. [Google Scholar] [CrossRef]
  17. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg effect: The metabolic requirements of cell proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef]
  18. Liberti, M.V.; Locasale, J.W. The Warburg Effect: How Does it Benefit Cancer Cells? Trends Biochem. Sci. 2016, 41, 211–218. [Google Scholar] [CrossRef]
  19. Luengo, A.; Li, Z.; Gui, D.Y.; Sullivan, L.B.; Zagorulya, M.; Do, B.T.; Ferreira, R.; Naamati, A.; Ali, A.; Lewis, C.A.; et al. Increased demand for NAD(+) relative to ATP drives aerobic glycolysis. Mol. Cell 2021, 81, 691–707.e6. [Google Scholar] [CrossRef]
  20. Liu, C.; Jin, Y.; Fan, Z. The Mechanism of Warburg Effect-Induced Chemoresistance in Cancer. Front. Oncol. 2021, 11, 698023. [Google Scholar] [CrossRef]
  21. Soukupova, J.; Malfettone, A.; Hyrossova, P.; Hernandez-Alvarez, M.I.; Penuelas-Haro, I.; Bertran, E.; Junza, A.; Capellades, J.; Giannelli, G.; Yanes, O.; et al. Role of the Transforming Growth Factor-beta in regulating hepatocellular carcinoma oxidative metabolism. Sci. Rep. 2017, 7, 12486. [Google Scholar] [CrossRef] [PubMed]
  22. Lee, Y.K.; Lim, J.J.; Jeoun, U.W.; Min, S.; Lee, E.B.; Kwon, S.M.; Lee, C.; Yoon, G. Lactate-mediated mitoribosomal defects impair mitochondrial oxidative phosphorylation and promote hepatoma cell invasiveness. J. Biol. Chem. 2017, 292, 20208–20217. [Google Scholar] [CrossRef] [PubMed]
  23. Tan, J.L.; Li, F.; Yeo, J.Z.; Yong, K.J.; Bassal, M.A.; Ng, G.H.; Lee, M.Y.; Leong, C.Y.; Tan, H.K.; Wu, C.S.; et al. New High-Throughput Screening Identifies Compounds That Reduce Viability Specifically in Liver Cancer Cells That Express High Levels of SALL4 by Inhibiting Oxidative Phosphorylation. Gastroenterology 2019, 157, 1615–1629. [Google Scholar] [CrossRef] [PubMed]
  24. Lee, S.; Lee, J.S.; Seo, J.; Lee, S.H.; Kang, J.H.; Song, J.; Kim, S.Y. Targeting Mitochondrial Oxidative Phosphorylation Abrogated Irinotecan Resistance in NSCLC. Sci. Rep. 2018, 8, 15707. [Google Scholar] [CrossRef] [PubMed]
  25. Nayak, A.P.; Kapur, A.; Barroilhet, L.; Patankar, M.S. Oxidative Phosphorylation: A Target for Novel Therapeutic Strategies Against Ovarian Cancer. Cancers 2018, 10, 337. [Google Scholar] [CrossRef]
  26. Ashton, T.M.; McKenna, W.G.; Kunz-Schughart, L.A.; Higgins, G.S. Oxidative Phosphorylation as an Emerging Target in Cancer Therapy. Clin. Cancer Res. 2018, 24, 2482–2490. [Google Scholar] [CrossRef]
  27. Xia, H.; Huang, Z.; Xu, Y.; Yam, J.W.P.; Cui, Y. Reprogramming of central carbon metabolism in hepatocellular carcinoma. Biomed. Pharmacother. 2022, 153, 113485. [Google Scholar] [CrossRef]
  28. Lo, C.H.; Farina, F.; Morris, H.P.; Weinhouse, S. Glycolytic regulation in rat liver and hepatomas. Adv. Enzym. Regul. 1968, 6, 453–464. [Google Scholar] [CrossRef]
  29. Vaupel, P.; Mayer, A. Hypoxia in cancer: Significance and impact on clinical outcome. Cancer Metastasis Rev. 2007, 26, 225–239. [Google Scholar] [CrossRef]
  30. Vaupel, P.; Hockel, M.; Mayer, A. Detection and characterization of tumor hypoxia using pO2 histography. Antioxid. Redox Signal. 2007, 9, 1221–1235. [Google Scholar] [CrossRef]
  31. McKeown, S.R. Defining normoxia, physoxia and hypoxia in tumours-implications for treatment response. Br. J. Radiol. 2014, 87, 20130676. [Google Scholar] [CrossRef]
  32. Wada, H.; Nagano, H.; Yamamoto, H.; Yang, Y.; Kondo, M.; Ota, H.; Nakamura, M.; Yoshioka, S.; Kato, H.; Damdinsuren, B.; et al. Expression pattern of angiogenic factors and prognosis after hepatic resection in hepatocellular carcinoma: Importance of angiopoietin-2 and hypoxia-induced factor-1 alpha. Liver Int. 2006, 26, 414–423. [Google Scholar] [CrossRef] [PubMed]
  33. Ma, L.; Li, G.; Zhu, H.; Dong, X.; Zhao, D.; Jiang, X.; Li, J.; Qiao, H.; Ni, S.; Sun, X. 2-Methoxyestradiol synergizes with sorafenib to suppress hepatocellular carcinoma by simultaneously dysregulating hypoxia-inducible factor-1 and -2. Cancer Lett. 2014, 355, 96–105. [Google Scholar] [CrossRef] [PubMed]
  34. Korotchkina, L.G.; Patel, M.S. Site specificity of four pyruvate dehydrogenase kinase isoenzymes toward the three phosphorylation sites of human pyruvate dehydrogenase. J. Biol. Chem. 2001, 276, 37223–37229. [Google Scholar] [CrossRef]
  35. Bjornson, E.; Mukhopadhyay, B.; Asplund, A.; Pristovsek, N.; Cinar, R.; Romeo, S.; Uhlen, M.; Kunos, G.; Nielsen, J.; Mardinoglu, A. Stratification of Hepatocellular Carcinoma Patients Based on Acetate Utilization. Cell Rep. 2015, 13, 2014–2026. [Google Scholar] [CrossRef] [PubMed]
  36. Su, T.S.; Tsai, T.F.; Chi, C.W.; Han, S.H.; Chou, C.K. Elevation of facilitated glucose-transporter messenger RNA in human hepatocellular carcinoma. Hepatology 1990, 11, 118–122. [Google Scholar] [CrossRef] [PubMed]
  37. Kim, Y.H.; Jeong, D.C.; Pak, K.; Han, M.E.; Kim, J.Y.; Liangwen, L.; Kim, H.J.; Kim, T.W.; Kim, T.H.; Hyun, D.W.; et al. SLC2A2 (GLUT2) as a novel prognostic factor for hepatocellular carcinoma. Oncotarget 2017, 8, 68381–68392. [Google Scholar] [CrossRef] [PubMed]
  38. Amann, T.; Hellerbrand, C. GLUT1 as a therapeutic target in hepatocellular carcinoma. Expert Opin. Ther. Targets 2009, 13, 1411–1427. [Google Scholar] [CrossRef]
  39. Gong, L.; Cui, Z.; Chen, P.; Han, H.; Peng, J.; Leng, X. Reduced survival of patients with hepatocellular carcinoma expressing hexokinase II. Med. Oncol. 2012, 29, 909–914. [Google Scholar] [CrossRef]
  40. DeWaal, D.; Nogueira, V.; Terry, A.R.; Patra, K.C.; Jeon, S.M.; Guzman, G.; Au, J.; Long, C.P.; Antoniewicz, M.R.; Hay, N. Hexokinase-2 depletion inhibits glycolysis and induces oxidative phosphorylation in hepatocellular carcinoma and sensitizes to metformin. Nat. Commun. 2018, 9, 446. [Google Scholar] [CrossRef] [Green Version]
  41. Zhang, Z.; Huang, S.; Wang, H.; Wu, J.; Chen, D.; Peng, B.; Zhou, Q. High expression of hexokinase domain containing 1 is associated with poor prognosis and aggressive phenotype in hepatocarcinoma. Biochem. Biophys. Res. Commun. 2016, 474, 673–679. [Google Scholar] [CrossRef]
  42. Liu, S.; Sun, Y.; Jiang, M.; Li, Y.; Tian, Y.; Xue, W.; Ding, N.; Sun, Y.; Cheng, C.; Li, J.; et al. Glyceraldehyde-3-phosphate dehydrogenase promotes liver tumorigenesis by modulating phosphoglycerate dehydrogenase. Hepatology 2017, 66, 631–645. [Google Scholar] [CrossRef] [PubMed]
  43. Ganapathy-Kanniappan, S.; Kunjithapatham, R.; Geschwind, J.F. Glyceraldehyde-3-phosphate dehydrogenase: A promising target for molecular therapy in hepatocellular carcinoma. Oncotarget 2012, 3, 940–953. [Google Scholar] [CrossRef] [PubMed]
  44. Wong, C.C.; Au, S.L.; Tse, A.P.; Xu, I.M.; Lai, R.K.; Chiu, D.K.; Wei, L.L.; Fan, D.N.; Tsang, F.H.; Lo, R.C.; et al. Switching of pyruvate kinase isoform L to M2 promotes metabolic reprogramming in hepatocarcinogenesis. PLoS ONE 2014, 9, e115036. [Google Scholar] [CrossRef]
  45. Cluntun, A.A.; Lukey, M.J.; Cerione, R.A.; Locasale, J.W. Glutamine Metabolism in Cancer: Understanding the Heterogeneity. Trends Cancer 2017, 3, 169–180. [Google Scholar] [CrossRef]
  46. Bode, B.P.; Souba, W.W. Glutamine transport and human hepatocellular transformation. JPEN J. Parenter. Enter. Nutr. 1999, 23, S33–S37. [Google Scholar] [CrossRef] [PubMed]
  47. Yuneva, M.O.; Fan, T.W.; Allen, T.D.; Higashi, R.M.; Ferraris, D.V.; Tsukamoto, T.; Mates, J.M.; Alonso, F.J.; Wang, C.; Seo, Y.; et al. The metabolic profile of tumors depends on both the responsible genetic lesion and tissue type. Cell Metab. 2012, 15, 157–170. [Google Scholar] [CrossRef]
  48. Faloppi, L.; Scartozzi, M.; Bianconi, M.; Svegliati Baroni, G.; Toniutto, P.; Giampieri, R.; Del Prete, M.; De Minicis, S.; Bitetto, D.; Loretelli, C.; et al. The role of LDH serum levels in predicting global outcome in HCC patients treated with sorafenib: Implications for clinical management. BMC Cancer 2014, 14, 110. [Google Scholar] [CrossRef]
  49. Faloppi, L.; Bianconi, M.; Memeo, R.; Casadei Gardini, A.; Giampieri, R.; Bittoni, A.; Andrikou, K.; Del Prete, M.; Cascinu, S.; Scartozzi, M. Lactate Dehydrogenase in Hepatocellular Carcinoma: Something Old, Something New. BioMed Res. Int. 2016, 2016, 7196280. [Google Scholar] [CrossRef]
  50. Chen, Y.; Zhou, J.; Li, J.; Feng, J.; Chen, Z.; Wang, X. Plasma metabolomic analysis of human hepatocellular carcinoma: Diagnostic and therapeutic study. Oncotarget 2016, 7, 47332–47342. [Google Scholar] [CrossRef] [Green Version]
  51. Gao, H.J.; Zhao, M.C.; Zhang, Y.J.; Zhou, D.S.; Xu, L.; Li, G.B.; Chen, M.S.; Liu, J. Monocarboxylate transporter 4 predicts poor prognosis in hepatocellular carcinoma and is associated with cell proliferation and migration. J. Cancer Res. Clin. Oncol. 2015, 141, 1151–1162. [Google Scholar] [CrossRef] [PubMed]
  52. Huang, Q.; Li, J.; Xing, J.; Li, W.; Li, H.; Ke, X.; Zhang, J.; Ren, T.; Shang, Y.; Yang, H.; et al. CD147 promotes reprogramming of glucose metabolism and cell proliferation in HCC cells by inhibiting the p53-dependent signaling pathway. J. Hepatol. 2014, 61, 859–866. [Google Scholar] [CrossRef] [PubMed]
  53. Kalhan, S.C.; Guo, L.; Edmison, J.; Dasarathy, S.; McCullough, A.J.; Hanson, R.W.; Milburn, M. Plasma metabolomic profile in nonalcoholic fatty liver disease. Metabolism 2011, 60, 404–413. [Google Scholar] [CrossRef] [PubMed]
  54. Patra, K.C.; Hay, N. The pentose phosphate pathway and cancer. Trends Biochem. Sci. 2014, 39, 347–354. [Google Scholar] [CrossRef]
  55. Nwosu, Z.C.; Megger, D.A.; Hammad, S.; Sitek, B.; Roessler, S.; Ebert, M.P.; Meyer, C.; Dooley, S. Identification of the Consistently Altered Metabolic Targets in Human Hepatocellular Carcinoma. Cell. Mol. Gastroenterol. Hepatol. 2017, 4, 303–323.e301. [Google Scholar] [CrossRef]
  56. Liu, B.; Fang, M.; He, Z.; Cui, D.; Jia, S.; Lin, X.; Xu, X.; Zhou, T.; Liu, W. Hepatitis B virus stimulates G6PD expression through HBx-mediated Nrf2 activation. Cell Death Dis. 2015, 6, e1980. [Google Scholar] [CrossRef]
  57. Xu, I.M.; Lai, R.K.; Lin, S.H.; Tse, A.P.; Chiu, D.K.; Koh, H.Y.; Law, C.T.; Wong, C.M.; Cai, Z.; Wong, C.C.; et al. Transketolase counteracts oxidative stress to drive cancer development. Proc. Natl. Acad. Sci. USA 2016, 113, E725–E734. [Google Scholar] [CrossRef]
  58. Yang, C.; Wang, S.; Ruan, H.; Li, B.; Cheng, Z.; He, J.; Zuo, Q.; Yu, C.; Wang, H.; Lv, Y.; et al. Downregulation of PDK4 Increases Lipogenesis and Associates with Poor Prognosis in Hepatocellular Carcinoma. J. Cancer 2019, 10, 918–926. [Google Scholar] [CrossRef]
  59. King, A.; Selak, M.A.; Gottlieb, E. Succinate dehydrogenase and fumarate hydratase: Linking mitochondrial dysfunction and cancer. Oncogene 2006, 25, 4675–4682. [Google Scholar] [CrossRef]
  60. Tseng, P.L.; Wu, W.H.; Hu, T.H.; Chen, C.W.; Cheng, H.C.; Li, C.F.; Tsai, W.H.; Tsai, H.J.; Hsieh, M.C.; Chuang, J.H.; et al. Decreased succinate dehydrogenase B in human hepatocellular carcinoma accelerates tumor malignancy by inducing the Warburg effect. Sci. Rep. 2018, 8, 3081. [Google Scholar] [CrossRef] [Green Version]
  61. Gerresheim, G.K.; Roeb, E.; Michel, A.M.; Niepmann, M. Hepatitis C Virus Downregulates Core Subunits of Oxidative Phosphorylation, Reminiscent of the Warburg Effect in Cancer Cells. Cells 2019, 8, 1410. [Google Scholar] [CrossRef] [PubMed]
  62. Wu, W.; Zheng, X.; Wang, J.; Yang, T.; Dai, W.; Song, S.; Fang, L.; Wang, Y.; Gu, J. O-GlcNAcylation on Rab3A attenuates its effects on mitochondrial oxidative phosphorylation and metastasis in hepatocellular carcinoma. Cell Death Dis. 2018, 9, 970. [Google Scholar] [CrossRef] [PubMed]
  63. Zhuang, X.; Chen, Y.; Wu, Z.; Xu, Q.; Chen, M.; Shao, M.; Cao, X.; Zhou, Y.; Xie, M.; Shi, Y.; et al. Mitochondrial miR-181a-5p promotes glucose metabolism reprogramming in liver cancer by regulating the electron transport chain. Carcinogenesis 2020, 41, 972–983. [Google Scholar] [CrossRef] [PubMed]
  64. Chen, C.L.; Uthaya Kumar, D.B.; Punj, V.; Xu, J.; Sher, L.; Tahara, S.M.; Hess, S.; Machida, K. NANOG Metabolically Reprograms Tumor-Initiating Stem-like Cells through Tumorigenic Changes in Oxidative Phosphorylation and Fatty Acid Metabolism. Cell Metab. 2016, 23, 206–219. [Google Scholar] [CrossRef] [PubMed]
  65. Li, X.; Wang, X.; Gao, P. Diabetes Mellitus and Risk of Hepatocellular Carcinoma. BioMed Res. Int. 2017, 2017, 5202684. [Google Scholar] [CrossRef]
  66. Gan, L.; Liu, Z.; Sun, C. Obesity linking to hepatocellular carcinoma: A global view. Biochim. Biophys. Acta Rev. Cancer 2018, 1869, 97–102. [Google Scholar] [CrossRef]
  67. Batchuluun, B.; Pinkosky, S.L.; Steinberg, G.R. Lipogenesis inhibitors: Therapeutic opportunities and challenges. Nat. Rev. Drug Discov. 2022, 21, 283–305. [Google Scholar] [CrossRef]
  68. Donnelly, K.L.; Smith, C.I.; Schwarzenberg, S.J.; Jessurun, J.; Boldt, M.D.; Parks, E.J. Sources of fatty acids stored in liver and secreted via lipoproteins in patients with nonalcoholic fatty liver disease. J. Clin. Investig. 2005, 115, 1343–1351. [Google Scholar] [CrossRef]
  69. Mounier, C.; Bouraoui, L.; Rassart, E. Lipogenesis in cancer progression (review). Int. J. Oncol. 2014, 45, 485–492. [Google Scholar] [CrossRef]
  70. Cao, D.; Song, X.; Che, L.; Li, X.; Pilo, M.G.; Vidili, G.; Porcu, A.; Solinas, A.; Cigliano, A.; Pes, G.M.; et al. Both de novo synthetized and exogenous fatty acids support the growth of hepatocellular carcinoma cells. Liver Int. 2017, 37, 80–89. [Google Scholar] [CrossRef] [Green Version]
  71. Nath, A.; Li, I.; Roberts, L.R.; Chan, C. Elevated free fatty acid uptake via CD36 promotes epithelial-mesenchymal transition in hepatocellular carcinoma. Sci. Rep. 2015, 5, 14752. [Google Scholar] [CrossRef] [PubMed]
  72. Kerner, J.; Hoppel, C. Fatty acid import into mitochondria. Biochim. Biophys. Acta 2000, 1486, 1–17. [Google Scholar] [CrossRef] [PubMed]
  73. Serviddio, G.; Giudetti, A.M.; Bellanti, F.; Priore, P.; Rollo, T.; Tamborra, R.; Siculella, L.; Vendemiale, G.; Altomare, E.; Gnoni, G.V. Oxidation of hepatic carnitine palmitoyl transferase-I (CPT-I) impairs fatty acid beta-oxidation in rats fed a methionine-choline deficient diet. PLoS ONE 2011, 6, e24084. [Google Scholar] [CrossRef] [PubMed]
  74. Serviddio, G.; Bellanti, F.; Tamborra, R.; Rollo, T.; Romano, A.D.; Giudetti, A.M.; Capitanio, N.; Petrella, A.; Vendemiale, G.; Altomare, E. Alterations of hepatic ATP homeostasis and respiratory chain during development of non-alcoholic steatohepatitis in a rodent model. Eur. J. Clin. Investig. 2008, 38, 245–252. [Google Scholar] [CrossRef]
  75. Zhou, Y.Y.; Zhu, G.Q.; Wang, Y.; Zheng, J.N.; Ruan, L.Y.; Cheng, Z.; Hu, B.; Fu, S.W.; Zheng, M.H. Systematic review with network meta-analysis: Statins and risk of hepatocellular carcinoma. Oncotarget 2016, 7, 21753–21762. [Google Scholar] [CrossRef]
  76. Wada, A.; Fukui, K.; Sawai, Y.; Imanaka, K.; Kiso, S.; Tamura, S.; Shimomura, I.; Hayashi, N. Pamidronate induced anti-proliferative, apoptotic, and anti-migratory effects in hepatocellular carcinoma. J. Hepatol. 2006, 44, 142–150. [Google Scholar] [CrossRef]
  77. Singh, S.; Singh, P.P.; Singh, A.G.; Murad, M.H.; Sanchez, W. Statins are associated with a reduced risk of hepatocellular cancer: A systematic review and meta-analysis. Gastroenterology 2013, 144, 323–332. [Google Scholar] [CrossRef]
  78. Ridruejo, E.; Romero-Caimi, G.; Obregon, M.J.; Kleiman de Pisarev, D.; Alvarez, L. Potential Molecular Targets of Statins in the Prevention of Hepatocarcinogenesis. Ann. Hepatol. 2018, 17, 490–500. [Google Scholar] [CrossRef]
  79. Honda, Y.; Aikata, H.; Honda, F.; Nakano, N.; Nakamura, Y.; Hatooka, M.; Morio, K.; Kobayashi, T.; Fukuhara, T.; Nagaoki, Y.; et al. Clinical outcome and prognostic factors in hepatocellular carcinoma patients with bone metastases medicated with zoledronic acid. Hepatol. Res. 2017, 47, 1053–1060. [Google Scholar] [CrossRef]
  80. Yamada, S.; Takashina, Y.; Watanabe, M.; Nagamine, R.; Saito, Y.; Kamada, N.; Saito, H. Bile acid metabolism regulated by the gut microbiota promotes non-alcoholic steatohepatitis-associated hepatocellular carcinoma in mice. Oncotarget 2018, 9, 9925–9939. [Google Scholar] [CrossRef] [Green Version]
  81. Lozano, E.; Sanchez-Vicente, L.; Monte, M.J.; Herraez, E.; Briz, O.; Banales, J.M.; Marin, J.J.; Macias, R.I. Cocarcinogenic effects of intrahepatic bile acid accumulation in cholangiocarcinoma development. Mol. Cancer Res. 2014, 12, 91–100. [Google Scholar] [CrossRef] [PubMed]
  82. Stanca, E.; Serviddio, G.; Bellanti, F.; Vendemiale, G.; Siculella, L.; Giudetti, A.M. Down-regulation of LPCAT expression increases platelet-activating factor level in cirrhotic rat liver: Potential antiinflammatory effect of silybin. Biochim. Biophys. Acta 2013, 1832, 2019–2026. [Google Scholar] [CrossRef] [PubMed]
  83. Huang, Q.; Tan, Y.X.; Yin, P.Y.; Ye, G.Z.; Gao, P.; Lu, X.; Wang, H.Y.; Xu, G.W. Metabolic characterization of hepatocellular carcinoma using nontargeted tissue metabolomics. Cancer Res. 2013, 73, 4992–5002. [Google Scholar] [CrossRef] [PubMed]
  84. Liu, H.; Dong, H.; Robertson, K.; Liu, C. DNA methylation suppresses expression of the urea cycle enzyme carbamoyl phosphate synthetase 1 (CPS1) in human hepatocellular carcinoma. Am. J. Pathol. 2011, 178, 652–661. [Google Scholar] [CrossRef]
  85. Cancer Genome Atlas Research Network; Wheeler, D.A.; Roberts, L.R. Comprehensive and Integrative Genomic Characterization of Hepatocellular Carcinoma. Cell 2017, 169, 1327–1341.e23. [Google Scholar] [CrossRef]
  86. Yeh, H.W.; Lee, S.S.; Chang, C.Y.; Hu, C.M.; Jou, Y.S. Pyrimidine metabolic rate limiting enzymes in poorly-differentiated hepatocellular carcinoma are signature genes of cancer stemness and associated with poor prognosis. Oncotarget 2017, 8, 77734–77751. [Google Scholar] [CrossRef]
  87. Uhlen, M.; Zhang, C.; Lee, S.; Sjostedt, E.; Fagerberg, L.; Bidkhori, G.; Benfeitas, R.; Arif, M.; Liu, Z.; Edfors, F.; et al. A pathology atlas of the human cancer transcriptome. Science 2017, 357, eaan2507. [Google Scholar] [CrossRef]
  88. Peng, X.; Chen, Z.; Farshidfar, F.; Xu, X.; Lorenzi, P.L.; Wang, Y.; Cheng, F.; Tan, L.; Mojumdar, K.; Du, D.; et al. Molecular Characterization and Clinical Relevance of Metabolic Expression Subtypes in Human Cancers. Cell Rep. 2018, 23, 255–269. [Google Scholar] [CrossRef]
  89. Carr, B.I.; Guerra, V. Serum albumin levels in relation to tumor parameters in hepatocellular carcinoma patients. Int. J. Biol. Mrk. 2017, 32, e391–e396. [Google Scholar] [CrossRef]
  90. Possemato, R.; Marks, K.M.; Shaul, Y.D.; Pacold, M.E.; Kim, D.; Birsoy, K.; Sethumadhavan, S.; Woo, H.K.; Jang, H.G.; Jha, A.K.; et al. Functional genomics reveal that the serine synthesis pathway is essential in breast cancer. Nature 2011, 476, 346–350. [Google Scholar] [CrossRef] [Green Version]
  91. Locasale, J.W.; Grassian, A.R.; Melman, T.; Lyssiotis, C.A.; Mattaini, K.R.; Bass, A.J.; Heffron, G.; Metallo, C.M.; Muranen, T.; Sharfi, H.; et al. Phosphoglycerate dehydrogenase diverts glycolytic flux and contributes to oncogenesis. Nat. Genet. 2011, 43, 869–874. [Google Scholar] [CrossRef] [PubMed]
  92. Jain, M.; Nilsson, R.; Sharma, S.; Madhusudhan, N.; Kitami, T.; Souza, A.L.; Kafri, R.; Kirschner, M.W.; Clish, C.B.; Mootha, V.K. Metabolite profiling identifies a key role for glycine in rapid cancer cell proliferation. Science 2012, 336, 1040–1044. [Google Scholar] [CrossRef] [PubMed]
  93. Chaneton, B.; Hillmann, P.; Zheng, L.; Martin, A.C.L.; Maddocks, O.D.K.; Chokkathukalam, A.; Coyle, J.E.; Jankevics, A.; Holding, F.P.; Vousden, K.H.; et al. Serine is a natural ligand and allosteric activator of pyruvate kinase M2. Nature 2012, 491, 458–462. [Google Scholar] [CrossRef]
  94. Stepien, M.; Duarte-Salles, T.; Fedirko, V.; Floegel, A.; Barupal, D.K.; Rinaldi, S.; Achaintre, D.; Assi, N.; Tjonneland, A.; Overvad, K.; et al. Alteration of amino acid and biogenic amine metabolism in hepatobiliary cancers: Findings from a prospective cohort study. Int. J. Cancer 2016, 138, 348–360. [Google Scholar] [CrossRef] [PubMed]
  95. Gao, R.; Cheng, J.; Fan, C.; Shi, X.; Cao, Y.; Sun, B.; Ding, H.; Hu, C.; Dong, F.; Yan, X. Serum Metabolomics to Identify the Liver Disease-Specific Biomarkers for the Progression of Hepatitis to Hepatocellular Carcinoma. Sci. Rep. 2015, 5, 18175. [Google Scholar] [CrossRef] [PubMed]
  96. 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] [PubMed]
  97. Qian, B.Z.; Pollard, J.W. Macrophage diversity enhances tumor progression and metastasis. Cell 2010, 141, 39–51. [Google Scholar] [CrossRef] [PubMed]
  98. Biswas, S.K.; Mantovani, A. Macrophage plasticity and interaction with lymphocyte subsets: Cancer as a paradigm. Nat. Immunol. 2010, 11, 889–896. [Google Scholar] [CrossRef] [PubMed]
  99. Biswas, S.K.; Chittezhath, M.; Shalova, I.N.; Lim, J.Y. Macrophage polarization and plasticity in health and disease. Immunol. Res. 2012, 53, 11–24. [Google Scholar] [CrossRef] [PubMed]
  100. Biswas, S.K.; Sica, A.; Lewis, C.E. Plasticity of macrophage function during tumor progression: Regulation by distinct molecular mechanisms. J. Immunol. 2008, 180, 2011–2017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Wan, S.; Zhao, E.; Kryczek, I.; Vatan, L.; Sadovskaya, A.; Ludema, G.; Simeone, D.M.; Zou, W.; Welling, T.H. Tumor-associated macrophages produce interleukin 6 and signal via STAT3 to promote expansion of human hepatocellular carcinoma stem cells. Gastroenterology 2014, 147, 1393–1404. [Google Scholar] [CrossRef] [PubMed]
  102. Budhu, A.; Forgues, M.; Ye, Q.H.; Jia, H.L.; He, P.; Zanetti, K.A.; Kammula, U.S.; Chen, Y.; Qin, L.X.; Tang, Z.Y.; et al. Prediction of venous metastases, recurrence, and prognosis in hepatocellular carcinoma based on a unique immune response signature of the liver microenvironment. Cancer Cell 2006, 10, 99–111. [Google Scholar] [CrossRef] [PubMed]
  103. Husain, Z.; Huang, Y.; Seth, P.; Sukhatme, V.P. Tumor-derived lactate modifies antitumor immune response: Effect on myeloid-derived suppressor cells and NK cells. J. Immunol. 2013, 191, 1486–1495. [Google Scholar] [CrossRef] [PubMed]
  104. Colegio, O.R.; Chu, N.Q.; Szabo, A.L.; Chu, T.; Rhebergen, A.M.; Jairam, V.; Cyrus, N.; Brokowski, C.E.; Eisenbarth, S.C.; Phillips, G.M.; et al. Functional polarization of tumour-associated macrophages by tumour-derived lactic acid. Nature 2014, 513, 559–563. [Google Scholar] [CrossRef]
  105. Al-Khami, A.A.; Zheng, L.; Del Valle, L.; Hossain, F.; Wyczechowska, D.; Zabaleta, J.; Sanchez, M.D.; Dean, M.J.; Rodriguez, P.C.; Ochoa, A.C. Exogenous lipid uptake induces metabolic and functional reprogramming of tumor-associated myeloid-derived suppressor cells. Oncoimmunology 2017, 6, e1344804. [Google Scholar] [CrossRef]
  106. Wang, L.; Yang, Q.; Peng, S.; Liu, X. The combination of the glycolysis inhibitor 2-DG and sorafenib can be effective against sorafenib-tolerant persister cancer cells. OncoTargets Ther. 2019, 12, 5359–5373. [Google Scholar] [CrossRef]
  107. Tomizawa, M.; Shinozaki, F.; Motoyoshi, Y.; Sugiyama, T.; Yamamoto, S.; Ishige, N. 2-Deoxyglucose and sorafenib synergistically suppress the proliferation and motility of hepatocellular carcinoma cells. Oncol. Lett. 2017, 13, 800–804. [Google Scholar] [CrossRef]
  108. Sun, X.; Sun, G.; Huang, Y.; Hao, Y.; Tang, X.; Zhang, N.; Zhao, L.; Zhong, R.; Peng, Y. 3-Bromopyruvate regulates the status of glycolysis and BCNU sensitivity in human hepatocellular carcinoma cells. Biochem. Pharmacol. 2020, 177, 113988. [Google Scholar] [CrossRef]
  109. Pereira da Silva, A.P.; El-Bacha, T.; Kyaw, N.; dos Santos, R.S.; da-Silva, W.S.; Almeida, F.C.; Da Poian, A.T.; Galina, A. Inhibition of energy-producing pathways of HepG2 cells by 3-bromopyruvate. Biochem. J. 2009, 417, 717–726. [Google Scholar] [CrossRef]
  110. Ganapathy-Kanniappan, S.; Geschwind, J.F.; Kunjithapatham, R.; Buijs, M.; Vossen, J.A.; Tchernyshyov, I.; Cole, R.N.; Syed, L.H.; Rao, P.P.; Ota, S.; et al. Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) is pyruvylated during 3-bromopyruvate mediated cancer cell death. Anticancer Res. 2009, 29, 4909–4918. [Google Scholar]
  111. Liberti, M.V.; Dai, Z.; Wardell, S.E.; Baccile, J.A.; Liu, X.; Gao, X.; Baldi, R.; Mehrmohamadi, M.; Johnson, M.O.; Madhukar, N.S.; et al. A Predictive Model for Selective Targeting of the Warburg Effect through GAPDH Inhibition with a Natural Product. Cell Metab. 2017, 26, 648–659.e8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Xu, D.; Jin, J.; Yu, H.; Zhao, Z.; Ma, D.; Zhang, C.; Jiang, H. Chrysin inhibited tumor glycolysis and induced apoptosis in hepatocellular carcinoma by targeting hexokinase-2. J. Exp. Clin. Cancer Res. 2017, 36, 44. [Google Scholar] [CrossRef] [PubMed]
  113. Li, J.; Chen, K.; Wang, F.; Dai, W.; Li, S.; Feng, J.; Wu, L.; Liu, T.; Xu, S.; Xia, Y.; et al. Methyl jasmonate leads to necrosis and apoptosis in hepatocellular carcinoma cells via inhibition of glycolysis and represses tumor growth in mice. Oncotarget 2017, 8, 45965–45980. [Google Scholar] [CrossRef]
  114. Liu, T.; Li, S.; Wu, L.; Yu, Q.; Li, J.; Feng, J.; Zhang, J.; Chen, J.; Zhou, Y.; Ji, J.; et al. Experimental Study of Hepatocellular Carcinoma Treatment by Shikonin through Regulating PKM2. J Hepatocell. Carcinoma 2020, 7, 19–31. [Google Scholar] [CrossRef] [PubMed]
  115. Li, S.; Li, J.; Dai, W.; Zhang, Q.; Feng, J.; Wu, L.; Liu, T.; Yu, Q.; Xu, S.; Wang, W.; et al. Genistein suppresses aerobic glycolysis and induces hepatocellular carcinoma cell death. Br. J. Cancer 2017, 117, 1518–1528. [Google Scholar] [CrossRef] [PubMed]
  116. Xing, Y.X.; Li, M.H.; Tao, L.; Ruan, L.Y.; Hong, W.; Chen, C.; Zhao, W.L.; Xu, H.; Chen, J.F.; Wang, J.S. Anti-Cancer Effects of Emodin on HepG2 Cells as Revealed by (1)H NMR Based Metabolic Profiling. J. Proteome Res. 2018, 17, 1943–1952. [Google Scholar] [CrossRef]
  117. Wei, L.; Harriman, G.; Ghoshal, S.; Moaven, O.; Greenwood, J.; Bhat, S.; Westlin, W.F.; Harwood, H.J.; Kapeller, R.; Tanabe, K.K.; et al. Combination therapy with a liver selective acetyl-CoA carboxylase inhibitor ND-654 and sorafenib improves efficacy in the treatment of cirrhotic rats with hepatocellular carcinoma. Cancer Res. 2016, 76, 3781. [Google Scholar] [CrossRef]
  118. Zhou, W.; Simpson, P.J.; McFadden, J.M.; Townsend, C.A.; Medghalchi, S.M.; Vadlamudi, A.; Pinn, M.L.; Ronnett, G.V.; Kuhajda, F.P. Fatty acid synthase inhibition triggers apoptosis during S phase in human cancer cells. Cancer Res. 2003, 63, 7330–7337. [Google Scholar]
  119. Flavin, R.; Peluso, S.; Nguyen, P.L.; Loda, M. Fatty acid synthase as a potential therapeutic target in cancer. Future Oncol. 2010, 6, 551–562. [Google Scholar] [CrossRef]
  120. Ma, M.K.F.; Lau, E.Y.T.; Leung, D.H.W.; Lo, J.; Ho, N.P.Y.; Cheng, L.K.W.; Ma, S.; Lin, C.H.; Copland, J.A.; Ding, J.; et al. Stearoyl-CoA desaturase regulates sorafenib resistance via modulation of ER stress-induced differentiation. J. Hepatol. 2017, 67, 979–990. [Google Scholar] [CrossRef]
  121. Xiong, T.; Li, Z.; Huang, X.; Lu, K.; Xie, W.; Zhou, Z.; Tu, J. TO901317 inhibits the development of hepatocellular carcinoma by LXRalpha/Glut1 decreasing glycometabolism. Am. J. Physiol. Gastrointest. Liver Physiol. 2019, 316, G598–G607. [Google Scholar] [CrossRef] [PubMed]
  122. Montero, J.; Morales, A.; Llacuna, L.; Lluis, J.M.; Terrones, O.; Basanez, G.; Antonsson, B.; Prieto, J.; Garcia-Ruiz, C.; Colell, A.; et al. Mitochondrial cholesterol contributes to chemotherapy resistance in hepatocellular carcinoma. Cancer Res. 2008, 68, 5246–5256. [Google Scholar] [CrossRef] [Green Version]
  123. Zheng, Y.H.; Zhou, Q.Q.; Zhao, C.; Li, J.J.; Yu, Z.P.; Zhu, Q.D. ATP citrate lyase inhibitor triggers endoplasmic reticulum stress to induce hepatocellular carcinoma cell apoptosis via p-eIF2 alpha/ATF4/CHOP axis. J. Cell. Mol. Med. 2021, 25, 1468–1479. [Google Scholar] [CrossRef] [PubMed]
  124. Gnocchi, D.; Pedrelli, M.; Hurt-Camejo, E.; Parini, P. Lipids around the Clock: Focus on Circadian Rhythms and Lipid Metabolism. Biology 2015, 4, 104–132. [Google Scholar] [CrossRef] [PubMed]
  125. Gnocchi, D.; Bruscalupi, G. Circadian Rhythms and Hormonal Homeostasis: Pathophysiological Implications. Biology 2017, 6, 10. [Google Scholar] [CrossRef]
  126. Gnocchi, D.; Custodero, C.; Sabba, C.; Mazzocca, A. Circadian rhythms: A possible new player in non-alcoholic fatty liver disease pathophysiology. J. Mol. Med. 2019, 97, 741–759. [Google Scholar] [CrossRef]
  127. Jin, X.; Pan, Y.; Wang, L.; Zhang, L.; Ravichandran, R.; Potts, P.R.; Jiang, J.; Wu, H.; Huang, H. MAGE-TRIM28 complex promotes the Warburg effect and hepatocellular carcinoma progression by targeting FBP1 for degradation. Oncogenesis 2017, 6, e312. [Google Scholar] [CrossRef]
  128. Yang, J.; Jin, X.; Yan, Y.; Shao, Y.; Pan, Y.; Roberts, L.R.; Zhang, J.; Huang, H.; Jiang, J. Inhibiting histone deacetylases suppresses glucose metabolism and hepatocellular carcinoma growth by restoring FBP1 expression. Sci. Rep. 2017, 7, 43864. [Google Scholar] [CrossRef]
  129. Schmidt, D.M.Z.; McCafferty, D.G. trans-2-phenylcyclopropylamine is a mechanism-based inactivator of the histone demethylase LSD1. Biochemistry 2007, 46, 4408–4416. [Google Scholar] [CrossRef]
  130. Pan, D.N.; Mao, C.X.; Wang, Y.X. Suppression of Gluconeogenic Gene Expression by LSD1-Mediated Histone Demethylation. PLoS ONE 2013, 8, e66294. [Google Scholar] [CrossRef]
  131. Ma, R.H.; Zhang, W.G.; Tang, K.; Zhang, H.F.; Zhang, Y.; Li, D.P.; Li, Y.; Xu, P.W.; Luo, S.Q.; Cai, W.Q.; et al. Switch of glycolysis to gluconeogenesis by dexamethasone for treatment of hepatocarcinoma. Nat. Commun. 2013, 4, 2508. [Google Scholar] [CrossRef] [PubMed]
  132. Shukla, K.; Ferraris, D.V.; Thomas, A.G.; Stathis, M.; Duvall, B.; Delahanty, G.; Alt, J.; Rais, R.; Rojas, C.; Gao, P.; et al. Design, Synthesis, and Pharmacological Evaluation of Bis-2-(5-phenylacetamido-1,2,4-thiadiazol-2-yl)ethyl Sulfide 3 (BPTES) Analogs as Glutaminase Inhibitors. J. Med. Chem. 2012, 55, 10551–10563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Chen, Z.; Li, D.; Xu, N.; Fang, J.Z.; Yu, Y.; Hou, W.; Ruan, H.Q.; Zhu, P.P.; Ma, R.C.; Lu, S.Y.; et al. Novel 1,3,4-Selenadiazole-Containing Kidney-Type Glutaminase Inhibitors Showed Improved Cellular Uptake and Antitumor Activity. J. Med. Chem. 2019, 62, 589–603. [Google Scholar] [CrossRef]
  134. Schulte, M.L.; Fu, A.; Zhao, P.; Li, J.; Geng, L.; Smith, S.T.; Kondo, J.; Coffey, R.J.; Johnson, M.O.; Rathmell, J.C.; et al. Pharmacological blockade of ASCT2-dependent glutamine transport leads to antitumor efficacy in preclinical models. Nat. Med. 2018, 24, 194–202. [Google Scholar] [CrossRef] [PubMed]
  135. Jin, H.J.; Wang, S.Y.; Zaal, E.A.; Wang, C.; Wu, H.Q.; Bosma, A.; Jochems, F.; Isima, N.; Jin, G.Z.; 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] [PubMed]
  136. Zhang, P.C.; Wang, Q.C.; Lin, Z.B.; Yang, P.J.; Dou, K.F.; Zhang, R.H. Berberine Inhibits Growth of Liver Cancer Cells by Suppressing Glutamine Uptake. OncoTargets Ther. 2019, 12, 1751–1763. [Google Scholar] [CrossRef] [PubMed]
  137. Steinberg, G.R.; Hardie, D.G. New insights into activation and function of the AMPK. Nat. Rev. Mol. Cell Biol. 2022. [Google Scholar] [CrossRef]
  138. Goodman, M.; Liu, Z.; Zhu, P.; Li, J. AMPK Activators as a Drug for Diabetes, Cancer and Cardiovascular Disease. Pharm. Regul. Aff. 2014, 3, 118. [Google Scholar] [CrossRef]
  139. Tseng, H.I.; Zeng, Y.S.; Lin, Y.C.J.; Huang, J.W.; Lin, C.L.; Lee, M.H.; Yang, F.W.; Fang, T.P.; Mar, A.C.; Su, J.C. A novel AMPK activator shows therapeutic potential in hepatocellular carcinoma by suppressing HIF1 alpha-mediated aerobic glycolysis. Mol. Oncol. 2022, 16, 2274–2294. [Google Scholar] [CrossRef]
  140. Vara-Ciruelos, D.; Russell, F.M.; Hardie, D.G. The strange case of AMPK and cancer: Dr Jekyll or Mr Hyde? (dagger). Open Biol. 2019, 9, 190099. [Google Scholar] [CrossRef]
  141. Lepropre, S.; Kautbally, S.; Octave, M.; Ginion, A.; Onselaer, M.B.; Steinberg, G.R.; Kemp, B.E.; Hego, A.; Wera, O.; Brouns, S.; et al. AMPK-ACC signaling modulates platelet phospholipids and potentiates thrombus formation. Blood 2018, 132, 1180–1192. [Google Scholar] [CrossRef]
  142. Ciavarella, A.; Gnocchi, D.; Custodero, C.; Lenato, G.M.; Fiore, G.; Sabba, C.; Mazzocca, A. Translational insight into prothrombotic state and hypercoagulation in nonalcoholic fatty liver disease. Thromb. Res. 2021, 198, 139–150. [Google Scholar] [CrossRef]
  143. Kwan, Y.P.; Saito, T.; Ibrahim, D.; Al-Hassan, F.M.; Ein Oon, C.; Chen, Y.; Jothy, S.L.; Kanwar, J.R.; Sasidharan, S. Evaluation of the cytotoxicity, cell-cycle arrest, and apoptotic induction by Euphorbia hirta in MCF-7 breast cancer cells. Pharm. Biol. 2016, 54, 1223–1236. [Google Scholar] [CrossRef]
  144. Karna, P.; Chagani, S.; Gundala, S.R.; Rida, P.C.; Asif, G.; Sharma, V.; Gupta, M.V.; Aneja, R. Benefits of whole ginger extract in prostate cancer. Br. J. Nutr. 2012, 107, 473–484. [Google Scholar] [CrossRef]
  145. Jo, K.J.; Cha, M.R.; Lee, M.R.; Yoon, M.Y.; Park, H.R. Methanolic extracts of Uncaria rhynchophylla induce cytotoxicity and apoptosis in HT-29 human colon carcinoma cells. Plant Foods Hum. Nutr. 2008, 63, 77–82. [Google Scholar] [CrossRef]
  146. Gill, M.S.A.; Saleem, H.; Ahemad, N. Plant Extracts and their Secondary Metabolites as Modulators of Kinases. Curr. Top. Med. Chem. 2020, 20, 1093–1104. [Google Scholar] [CrossRef]
  147. Encalada, M.A.; Hoyos, K.M.; Rehecho, S.; Berasategi, I.; de Ciriano, M.G.; Ansorena, D.; Astiasaran, I.; Navarro-Blasco, I.; Cavero, R.Y.; Calvo, M.I. Anti-proliferative effect of Melissa officinalis on human colon cancer cell line. Plant Foods Hum. Nutr. 2011, 66, 328–334. [Google Scholar] [CrossRef]
  148. Pirola, L.; Frojdo, S. Resveratrol: One molecule, many targets. IUBMB Life 2008, 60, 323–332. [Google Scholar] [CrossRef]
  149. Massimi, M.; Tomassini, A.; Sciubba, F.; Sobolev, A.P.; Devirgiliis, L.C.; Miccheli, A. Effects of resveratrol on HepG2 cells as revealed by H-1-NMR based metabolic profiling. Biochim. Biophys. Acta (BBA)-Gen. Subj. 2012, 1820, 1–8. [Google Scholar] [CrossRef]
  150. Teng, W.D.; Yin, W.J.; Zhao, L.; Ma, C.W.; Huang, J.Q.; Ren, F.Z. Resveratrol metabolites ameliorate insulin resistance in HepG2 hepatocytes by modulating IRS-1/AMPK. RSC Adv. 2018, 8, 36034–36042. [Google Scholar] [CrossRef]
  151. Gnocchi, D.; Del Coco, L.; Girelli, C.R.; Castellaneta, F.; Cesari, G.; Sabba, C.; Fanizzi, F.P.; Mazzocca, A. 1H-NMR metabolomics reveals a multitarget action of Crithmum maritimum ethyl acetate extract in inhibiting hepatocellular carcinoma cell growth. Sci. Rep. 2021, 11, 1259. [Google Scholar] [CrossRef]
  152. Gnocchi, D.; Sabba, C.; Mazzocca, A. The Edible Plant Crithmum maritimum Shows Nutraceutical Properties by Targeting Energy Metabolism in Hepatic Cancer. Plant Foods Hum. Nutr. 2022, 77, 481–483. [Google Scholar] [CrossRef]
  153. Gnocchi, D.; Sabba, C.; Mazzocca, A. Crithmum maritimum Improves Sorafenib Sensitivity by Decreasing Lactic Acid Fermentation and Inducing a Pro-Hepatocyte Marker Profile in Hepatocellular Carcinoma. Plant Foods Hum. Nutr. 2022. [Google Scholar] [CrossRef]
  154. Gnocchi, D.; Castellaneta, F.; Cesari, G.; Fiore, G.; Sabba, C.; Mazzocca, A. Treatment of liver cancer cells with ethyl acetate extract of Crithmum maritimum permits reducing sorafenib dose and toxicity maintaining its efficacy. J. Pharm. Pharm. 2021, 73, 1369–1376. [Google Scholar] [CrossRef]
  155. Lippolis, R.; Gnocchi, D.; Santacroce, L.; Siciliano, R.A.; Mazzeo, M.F.; Scacco, S.; Sabba, C.; Mazzocca, A. A distinctive protein signature induced by lysophosphatidic acid receptor 6 (LPAR6) expression in hepatocellular carcinoma cells. Biochem. Biophys. Res. Commun. 2020, 526, 1150–1156. [Google Scholar] [CrossRef]
  156. Gnocchi, D.; Kurzyk, A.; Mintrone, A.; Lentini, G.; Sabba, C.; Mazzocca, A. Inhibition of LPAR6 overcomes sorafenib resistance by switching glycolysis into oxidative phosphorylation in hepatocellular carcinoma. Biochimie 2022, 202, 180–189. [Google Scholar] [CrossRef]
  157. Gnocchi, D.; Kapoor, S.; Nitti, P.; Cavalluzzi, M.M.; Lentini, G.; Denora, N.; Sabba, C.; Mazzocca, A. Novel lysophosphatidic acid receptor 6 antagonists inhibit hepatocellular carcinoma growth through affecting mitochondrial function. J. Mol. Med. 2020, 98, 179–191. [Google Scholar] [CrossRef]
  158. Gnocchi, D.; Cavalluzzi, M.M.; Mangiatordi, G.F.; Rizzi, R.; Tortorella, C.; Spennacchio, M.; Lentini, G.; Altomare, A.; Sabba, C.; Mazzocca, A. Xanthenylacetic Acid Derivatives Effectively Target Lysophosphatidic Acid Receptor 6 to Inhibit Hepatocellular Carcinoma Cell Growth. ChemMedChem 2021, 16, 2121–2129. [Google Scholar] [CrossRef]
  159. Mazzocca, A.; Dituri, F.; De Santis, F.; Filannino, A.; Lopane, C.; Betz, R.C.; Li, Y.Y.; Mukaida, N.; Winter, P.; Tortorella, C.; et al. Lysophosphatidic acid receptor LPAR6 supports the tumorigenicity of hepatocellular carcinoma. Cancer Res. 2015, 75, 532–543. [Google Scholar] [CrossRef]
  160. Mazzocca, A.; Dituri, F.; Lupo, L.; Quaranta, M.; Antonaci, S.; Giannelli, G. Tumor-secreted lysophostatidic acid accelerates hepatocellular carcinoma progression by promoting differentiation of peritumoral fibroblasts in myofibroblasts. Hepatology 2011, 54, 920–930. [Google Scholar] [CrossRef]
  161. Gnocchi, D.; Sabba, C.; Mazzocca, A. Lactic acid fermentation: A maladaptive mechanism and an evolutionary throwback boosting cancer drug resistance. Biochimie 2023. [Google Scholar] [CrossRef] [PubMed]
  162. Rodriguez-Hernandez, M.A.; de la Cruz-Ojeda, P.; Lopez-Grueso, M.J.; Navarro-Villaran, E.; Requejo-Aguilar, R.; Castejon-Vega, B.; Negrete, M.; Gallego, P.; Vega-Ochoa, A.; Victor, V.M.; et al. Integrated molecular signaling involving mitochondrial dysfunction and alteration of cell metabolism induced by tyrosine kinase inhibitors in cancer. Redox Biol. 2020, 36, 101510. [Google Scholar] [CrossRef] [PubMed]
  163. Rodriguez-Hernandez, M.A.; de la Cruz-Ojeda, P.; Gallego, P.; Navarro-Villaran, E.; Stankova, P.; Del Campo, J.A.; Kucera, O.; Elkalaf, M.; Maseko, T.E.; Cervinkova, Z.; et al. Dose-dependent regulation of mitochondrial function and cell death pathway by sorafenib in liver cancer cells. Biochem. Pharm. 2020, 176, 113902. [Google Scholar] [CrossRef] [PubMed]
  164. Rodriguez-Hernandez, M.A.; Chapresto-Garzon, R.; Cadenas, M.; Navarro-Villaran, E.; Negrete, M.; Gomez-Bravo, M.A.; Victor, V.M.; Padillo, F.J.; Muntane, J. Differential effectiveness of tyrosine kinase inhibitors in 2D/3D culture according to cell differentiation, p53 status and mitochondrial respiration in liver cancer cells. Cell Death Dis. 2020, 11, 339. [Google Scholar] [CrossRef]
  165. Nwosu, Z.C.; Pioronska, W.; Battello, N.; Zimmer, A.D.; Dewidar, B.; Han, M.; Pereira, S.; Blagojevic, B.; Castven, D.; Charlestin, V.; et al. Severe metabolic alterations in liver cancer lead to ERK pathway activation and drug resistance. EBioMedicine 2020, 54, 102699. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Metabolic enzymes that can be targeted in HCC. The metabolic pathways are depicted in white boxes, while metabolic targets are reported in light brown boxes. See the text for details. Illustrations have been realized with Biorender.
Figure 1. Metabolic enzymes that can be targeted in HCC. The metabolic pathways are depicted in white boxes, while metabolic targets are reported in light brown boxes. See the text for details. Illustrations have been realized with Biorender.
Ijms 24 03710 g001
Figure 2. Current therapies and metabolism-based interventions for HCC. New metabolism-based therapeutics may provide promising tools for the treatment of HCC. See the text for details. Illustrations have been realized with Biorender.
Figure 2. Current therapies and metabolism-based interventions for HCC. New metabolism-based therapeutics may provide promising tools for the treatment of HCC. See the text for details. Illustrations have been realized with Biorender.
Ijms 24 03710 g002
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

Gnocchi, D.; Sabbà, C.; Massimi, M.; Mazzocca, A. Metabolism as a New Avenue for Hepatocellular Carcinoma Therapy. Int. J. Mol. Sci. 2023, 24, 3710. https://doi.org/10.3390/ijms24043710

AMA Style

Gnocchi D, Sabbà C, Massimi M, Mazzocca A. Metabolism as a New Avenue for Hepatocellular Carcinoma Therapy. International Journal of Molecular Sciences. 2023; 24(4):3710. https://doi.org/10.3390/ijms24043710

Chicago/Turabian Style

Gnocchi, Davide, Carlo Sabbà, Mara Massimi, and Antonio Mazzocca. 2023. "Metabolism as a New Avenue for Hepatocellular Carcinoma Therapy" International Journal of Molecular Sciences 24, no. 4: 3710. https://doi.org/10.3390/ijms24043710

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