IMR Press / FBL / Volume 29 / Issue 4 / DOI: 10.31083/j.fbl2904143
Open Access Review
The Metabolism of Coenzyme A and Its Derivatives Plays a Crucial Role in Diseases
Show Less
1 Hubei Key Laboratory of Renal Disease Occurrence and Intervention, Medical School, Hubei Polytechnic University, 435003 Huangshi, Hubei, China
2 Department of Molecular Mechanisms Research and Treatment of Skin Tumors, Huangshi Central Hospital, 435000 Huangshi, Hubei, China
*Correspondence: xiejm922@163.com; xiejumin@hbpu.edu.cn (Jumin Xie); maohui1970@163.com (Hui Mao)
Front. Biosci. (Landmark Ed) 2024, 29(4), 143; https://doi.org/10.31083/j.fbl2904143
Submitted: 22 November 2023 | Revised: 12 January 2024 | Accepted: 6 February 2024 | Published: 9 April 2024
Copyright: © 2024 The Author(s). Published by IMR Press.
This is an open access article under the CC BY 4.0 license.
Abstract

Coenzyme A (CoA) functions as a crucial carrier of acyl groups within cells, playing a fundamental role in regulating acyl transfer reactions and participating in cellular metabolic processes. As the principal substrate and cofactor engaged in diverse metabolic reactions, CoA and its derivatives exert central influence over various physiological processes, primarily modulating lipid and ketone metabolism, as well as protein modification. This paper presents a comprehensive review of the molecular mechanisms by which CoA influences the onset and progression of cancer, cardiovascular disease (CVD), neurodegenerative disorders, and other illnesses. The main focal points include the following. (1) In cancer, enzymes such as acetyl-CoA synthetase 2, ATP citrate lyase, and acetyl-CoA carboxylase regulate lipid synthesis and energy metabolism by modulating acetyl-CoA levels. (2) In CVD, the effects of enzymes such as stearoyl-CoA desaturase-1, 3-hydroxy-3-methylglutaryl-CoA (HMGC) synthase 2, and HMGC reductase on the formation and advancement of these diseases are elucidated by their regulation of CoA metabolism across multiple organs. (3) In neurodegenerative disorders, the significance of CoA in maintaining cholesterol homeostasis in the brain and its implications on the development of such disorders are thoroughly discussed. The metabolic processes involving CoA and its derivatives span all physiological aspects within cells, playing a critical role in the onset and progression of various diseases. Elucidating the role of CoA in these conditions yields important insights that can serve as valuable references and guidance for disease diagnosis, treatment, and drug development.

Keywords
cancer
cardiovascular disease
coenzyme A
lipid metabolism
neurodegenerative disorders
1. Introduction

Coenzyme A (CoA) plays a pivotal role in cellular processes and is synthesized from L-Cysteine, ATP, and vitamin B5 (also known as pantothenic acid) [1]. Its primary function is to facilitate the formation of acyl-CoA compounds through thioester binding with various acyl groups including acetyl, propionyl, succinyl, and 3-hydroxy-3-methylglutaryl (HMG) [2]. These acyl-CoA derivatives serve as essential participants in diverse metabolic pathways within the human body. The metabolism of CoA and its thioester derivatives is intricately linked to a spectrum of diseases [2].

Acetyl-CoA serves as a pivotal intermediary in carbon metabolism, playing a dual role by supporting the tricarboxylic acid (TCA) cycle for ATP generation and facilitating the synthesis of fatty acids (FAs). Moreover, it serves as a crucial substrate for lysine acetylation reactions within cells, which are indispensable for cell growth and survival [3]. Enzymes such as ATP citrate lyase (ACLY), pyruvate dehydrogenase complex (PDC), and acyl-CoA short chain synthetase (ACSS) are responsible for the conversion of citrate, pyruvate, and acetate into acetyl-CoA, with these processes occurring in the cytoplasm and mitochondria [4].

ACLY, which is localized in both the cytoplasm and nucleus, is responsible for converting citrate into acetyl-CoA in the cytoplasm [5]. Another significant contributor to acetyl-CoA in both the nucleus and cytoplasm is ACSS2, which synthesizes acetyl-CoA from acetate [6]. In mammalian cells, ACSS2 is implicated in cytoplasmic lipid synthesis, whereas ACSS1 facilitates ATP production in the mitochondria [7]. ACLY plays a critical role in glucose-dependent fatty acids (FA) synthesis and histone acetylation. The upregulation of ACSS2 and conversion of acetate are the primary compensatory mechanisms in response to ACLY defects. Interestingly, despite the elevated acetyl-CoA levels and increased ACSS2 expression, defects in ACLY lead to reduced acetylation levels of total histone proteins, impaired proliferation, and altered gene expression patterns [8].

The rate-limiting enzyme governing long-chain FA synthesis and mitochondrial FA oxidation (FAO) is acetyl-CoA carboxylase (ACC) [9]. ACC1, predominantly localized in the cytoplasm, generates malonyl-CoA, a key substrate for FA biosynthesis, and acts as an allosteric inhibitor of carnitine palmitoyl-transferase, a key mediator in transporting FAs into the mitochondria for β-oxidation [10]. Additionally, ACC2, localized in the mitochondria and prominently expressed in oxidative tissues such as skeletal muscle and heart, plays a crucial role in regulating FA β-oxidation [9]. Significantly, all three enzymes are found in the nucleus, underscoring the central importance of acetyl-CoA in cellular metabolism, signaling, and epigenetics. This molecular hub is essential for fundamental biological processes, including energy production, macromolecular biosynthesis, and protein modification [5, 11].

Serving as a carrier molecule, CoA actively engages in transferring acyl groups across various metabolic pathways, playing pivotal roles in processes such as the TCA cycle, cholesterol synthesis, and protein modification [12, 13, 14]. The crucial involvement of CoA in cellular metabolism tightly correlates with the onset and progression of cancer [15]. Understanding the intricate interplay between CoA and cancer is paramount, potentially paving the way for novel research avenues, personalized diagnostic tools, and therapeutic approaches in cancer treatment [16, 17, 18]. Accumulating evidence indicates that disturbances in CoA metabolism contribute to the genesis and exacerbation of cardiovascular diseases (CVDs), impacting vital pathways such as the TCA cycle, lipid metabolism, and ketone body utilization [19, 20, 21]. Moreover, neuronal genetic defects in CoA metabolism detrimentally affect cellular energy generation and oxidative stress, leading to neuronal dysfunction and neurodegeneration [22, 23]. Diseases such as Alzheimer’s disease, Parkinson’s disease, and Huntington’s disease, which frequently show disruptions in cholesterol metabolism, energy pathways, and mitochondrial function, are intricately interconnected with CoA-dependent processes [24, 25, 26, 27].

CoA and its derivatives play pivotal roles in multiple diseases. Further exploration of their functions could pave the way for innovative therapeutic interventions and strategies.

2. Functions of CoA and Its Derivatives in Diseases
2.1 Acetyl-CoA and Cancer

Dysregulation of metabolic processes involving diverse substances contributes to the initiation and progression of cancer [28]. While normal cells predominantly rely on mitochondrial oxidative phosphorylation for energy production, cancer cells exhibit a preference for aerobic glycolysis, a phenomenon crucial in cancer cell proliferation, commonly known as the “Warburg effect” [29, 30].

The metabolism of acetyl-CoA metabolism is governed by three enzymes, ACSS2, ACLY, and ACC, all of which exert an impact on lipid metabolism [20, 31]. These enzymes are significantly overexpressed in various cancers. ACSS2 and ACLY enhance acetyl-CoA production, while ACC improves the utilization rate of acetyl-CoA, thereby promoting cell lipid metabolism, the TCA cycle, and histone acetylation. These processes collectively facilitate tumor growth (Fig. 1) [14].

Fig. 1.

Coenzyme A (CoA) and cancer. Condition A delineates acetyl-CoA metabolism in cancer, while condition B describes other CoA metabolism types in cancer. The solid arrow denotes the promotion process, whereas the blunt arrow signifies inhibition. The metabolism of acetyl-CoA in cancer is intricately regulated by acyl-CoA short chain synthetase (ACSS), ATP citrate lyase (ACLY), and acetyl-CoA carboxylase (ACC). These enzymes exert a profound influence on the metabolism of fatty acids (FAs), pyruvate, and other lipids, as well as the tricarboxylic acid (TCA) cycle and histone acetylation—critical processes for cancer cell proliferation and survival. Under conditions of hypoxia and nutrient restriction, cancer cells release hypoxia inducible factor (HIF), a pivotal regulator of acetyl-CoA metabolism. Hypoxia inducible factor acts to foster the growth of cancer cells by modulating acetyl-CoA utilization. Additionally, Hedgehog acyltransferase (Hhat), succinyl-CoA, and HMG-CoA play crucial roles in regulating other CoA metabolism pathways in cancer cells. Notably, HMG-CoA serves as a precursor for mevalonate synthesis through 3-hydroxy-3-methylglutaryl (HMG) reductase. The mevalonate pathway catalytically influences multiple cancer pathways, underscoring the significance of CoA metabolism in cancer development and progression. ROS, Reactive Oxygen Species.

2.1.1 ACSS2

ACSS2 is integral to a diverse array of cellular physiological processes. Initially, it catalyzes the synthesis of acetyl-CoA. Then it exerts regulatory control over signal transduction pathways. Lastly, within the nucleus, ACSS2 participates in the recycling of acetate generated by lysine deacetylase (Kdac) activity, thereby contributing to the energy supply essential for cellular metabolism [32, 33]. Comparative metabolomics and lipidomics studies have revealed that cancer cells utilize acetate as a nutritional source in an ACSS2-dependent manner. Under conditions of low oxygen and lipid depletion, ACSS2 exhibits the capacity to catalyze the reverse reaction from acetyl-CoA to acetate in tumor cells. This suggests that ACSS2 functions as a bidirectional enzyme in cancer cells, potentially playing a buffering role in the metabolism of acetyl-CoA/acetate in cancer [34, 35, 36]. Numerous studies have identified increased expression levels of ACSS2 across various cancer types, contributing to the maintenance of histone acetylation modifications [37, 38]. Conversely, research has demonstrated that diminishing ACSS2-mediated autophagy may promote breast cancer growth [39]. In colorectal cancer, thymine–guanine-interacting factor inhibits ACSS2, resulting in reduced acetyl-CoA production. However, cancer cells exhibit continued growth through the reprogramming of glucose transporter 1 (GLUT1) [40, 41].

In obesity-associated myeloma, elevated ACSS2 expression shows a positive correlation with increased body mass index. ACSS2 promotes the development of obesity-induced myeloma by stabilizing interferon regulatory factor 4. Given its substantial role in cell metabolism and regulation of gene expression, targeting the inhibition of ACSS2 expression may represent a promising strategy for cancer treatment [42].

2.1.2 ACLY

ACLY, a multifunctional enzyme, not only participates in the synthesis of acetyl-CoA but also plays a pivotal role in cell signal transduction by modulating histone acetylation [43]. Functioning as a target for serine/threonine-protein kinase (AKT), ACLY contributes to the conversion of citric acid to acetyl-CoA, establishing a crucial link between glycolysis and lipid metabolism processes [44]. The AKT/ACLY signaling pathway governs histone acetylation in cancer cells, and inhibiting ACLY results in citric acid accumulation, suppression of phosphofructokinase 1 (PFK1) and PFK2 activities, thereby reducing glycolysis and inducing apoptosis in cancer cells [45, 46]. In the context of cancer-related gene regulation, nutritional restriction, and AKT metabolic recombination, the upregulation of AKT1 leads to increased phosphorylation and activation of ACLY, consequently augmenting or sustaining acetyl-CoA production [47]. Sirtuin 6 (SIRT6), on the other hand, diminishes nuclear ACLY activity and acetyl-CoA levels, impeding the histone acetylation process and inhibiting tumor cell invasion [48]. However, within the apoptosis-inducing family, caspase-10 cleaves ACLY, diminishing its activity and reducing nuclear acetyl-CoA and lipid levels, and histone acetylation, thereby preventing cancer onset [49]. The overexpression of ACLY corresponds to increased and progressive lipid synthesis in various cancers [50, 51, 52].

In colorectal cancer, cancer-derived exosomal HSPC111 phosphorylates ACLY, promoting cancer-associated fibroblast-mediated lipid metabolism and elevating acetyl-CoA levels, thereby enhancing liver cancer cell metastasis [53]. In pancreatic cancer, the acetyl-CoA generated by ACLY is utilized for histone acetylation and mevalonate metabolism, fostering the development of pancreatic ductal adenocarcinoma (PDAC) [16]. Studies have demonstrated that the upregulation of ACLY is associated with unfavorable outcomes in gastric cancer, highlighting the feasibility of targeting ACLY for treatment. ACSS2 is believed to function cooperatively with ACLY, suggesting that simultaneous inhibition of ACLY and ACSS2 could potentially hinder the progression and treatment of gastric cancer [54]. Given its multifaceted roles in cell metabolism, signaling, and gene regulation, downregulating ACLY expression offers a promising strategy to mitigate cancer risk across various cell types [55].

2.1.3 ACC1

ACC1 enzymatically converts acetyl-CoA to malonyl-CoA, contributing to FA synthesis and regulating FAO, thereby orchestrating glucose and fat metabolism [56, 57]. ACC1 demonstrates elevated expression across numerous cancer types [58]. In breast cancer, phosphorylation of ACC1 leads to decreased lipid synthesis and elevated acetyl-CoA levels. This triggers increased histone H3 acetylation, activating SMAD2-mediated acetylation, and promoting the epithelial-mesenchymal transition (EMT), ultimately promoting cancer cell metastasis [59]. In prostate cancer, valproic acid inhibits ACC1 expression through the CCAAT enhancer-binding protein alpha and sterol regulatory element-binding protein 1 (SREBP-1) pathways, effectively curbing lipid accumulation and inducing apoptosis [60, 61]. In liver cancer, inhibition of ACC1 or the use of ND-654 inhibitors effectively suppresses de novo lipogenesis (DNL), thereby impeding cancer cell growth [62]. Additionally, in pancreatic cancer, eicosapentaenoic acid reduces ACC1 expression and ACC1-mediated DNL, leading to the inhibition of cell proliferation and survival [63]. These significant findings strongly underscore the crucial role of ACC1 in maintaining cellular energy balance and overall metabolic regulation. Targeting ACC1 expression levels can effectively reduce the proliferation of various cancer cell types.

2.1.4 Histone Acetylation

Histone acetylation constitutes an intricate biochemical process with a profound influence on cellular function and gene expression regulation. First, it initiates structural alterations in chromatin, facilitating a more open configuration. This modification enhances DNA’s affinity for transcription factors, thereby elevating gene transcriptional activity and increasing gene expression [64]. Second, histone acetylation plays a pivotal role in DNA repair and replication processes, contributing to the preservation of genomic integrity. This is achieved by recruiting specific proteins that collaborate with others, ensuring their effective functioning across various stages of the cell cycle [65].

A positive correlation has been established between the levels of acetyl-CoA and histone modification [66]. Acetyl-CoA synthesis is enhanced in the nucleus of cancer cells, where ACSS2 and ACLY facilitate the conversion of acetate and citric acid into acetyl-CoA, thereby promoting the process of histone acetylation [67]. The PDC elevates the levels of acetyl-CoA in the nucleus, resulting in increased acetylation of histones H2B, H3, and H4 [68]. In prostate cancer, the mitochondrial PDC increases citric acid levels for fat synthesis, while the nucleolar PDC enhances acetyl-CoA levels, thus promoting the acetylation of histone H3K9. This leads to increased expression of ACLY and lipid synthesis, maintaining the progression of prostate cancer [69].

Histone hyperacetylation emerges as a contributing factor to DNA damage, demonstrating its potential as a catalyst for liver cancer initiation. Within the liver cancer milieu, steatosis induces chromatin relaxation, enhances the DNA damage marker γH2AX production, and disrupts the production and conversion of acetyl-CoA [70, 71]. Amidst alterations in acetyl-CoA levels and histone acetylation dynamics, the downregulation of acyl-CoA thioesterase 12 precipitates increased acetyl-CoA levels. Simultaneously, the presence of free FAs (FFAs) elevates the acetylation levels of H3K9, H4K8, and H4K16. These molecular events collectively contribute to the upregulation of twist family basic helix–loop–helix transcription factor 2 expression, ultimately promoting the EMT, cancer cell metastasis, and lipid accumulation [72, 73].

Moreover, in PDAC, escalated acetyl-CoA levels drive increased histone acetylation, thereby promoting cancer development [74]. In acute myeloid leukemia, AMP-activated protein kinase (AMPK) plays a pivotal role in stabilizing acetyl-CoA levels, ensuring the maintenance of histone acetylation, facilitating cell proliferation, and recruiting bromodomain containing 4 onto chromatin [75]. In a rat model of colon cancer, the intricate interplay between dietary fat and cellulose elevates the acetylation level of histone H3, induces the expression of genes related to FAO, and promotes tumor development [76]. These multifaceted functions underscore the wide-reaching and pivotal role of histone acetylation in the realms of cell biology and molecular biology.

2.1.5 Fatty Acids

FAs serve diverse physiological functions. First, they act as crucial energy storage molecules within living organisms, adopting the form of triacylglycerol in adipose tissue. This reservoir is subsequently catabolized to provide energy as needed [77]. Second, FAs actively participate in intracellular signal transduction and the synthesis of second messengers, significantly contributing to the regulation of cellular metabolism [78].

FAs play a pivotal role in tumor cells, contributing to cell membrane biosynthesis and serving as a source of metabolic energy under stress conditions [79]. FAs are exclusively utilized through their combination with CoA to form acyl-CoA through an activation step [54]. Subsequently, FAs undergo β-oxidation, generating acetyl-CoA, which enters the TCA cycle and undergoes a condensation reaction with oxaloacetic acid, resulting in the formation of citric acid [80]. Within the mitochondrial matrix, acyl-CoA is enzymatically converted to acetyl-CoA through the process of FAO [81].

In breast cancer, the upregulation of FAO leads to elevated acetyl-CoA levels, consequently promoting the acetylation of histone H3 and H3K27ac. This facilitates the expression of genes associated with the EMT, a critical process in cancer progression [82]. In glioblastoma (GBM), the downregulation of diacylglycerol O-acyltransferase 1 expression results in reduced lipid storage, enhanced FAO of excess FAs, and subsequent accumulation of acetyl-CoA. This cascade of events influences the level of reactive oxygen species and induces cell apoptosis, ultimately inhibiting GBM growth [83].

2.1.6 Pyruvate

Pyruvate intricately participates in various physiological processes in living organisms. First, it functions as an energy storage molecule by participating in TCA cycle within the mitochondria, thereby supplying energy to the cell. Second, pyruvate serves as an intermediate product in the TCA cycle, which is capable of further metabolism to generate essential organic compounds including FAs and cholesterol. These dual roles underscore the pivotal contribution of pyruvate in sustaining cellular energy metabolism and facilitating organic synthesis [84].

The PDC is located in the mitochondria and nuclei, facilitating the connection between glucose metabolism and the TCA cycle for the synthesis of acetyl-CoA using pyruvate as a substrate [85]. Acetyl-CoA holds significant importance in cancer metabolism [86, 87]. The PDC activity is regulated by pyruvate dehydrogenase kinase (PDK) and acetyl-CoA acetyltransferase 1 (ACAT1). Phosphorylation of PDK or acetylation of ACAT1 leads to PDC inhibition, thereby enhancing the Warburg effect and promoting cancer growth [29]. Consequently, only a limited amount of pyruvate enters the TCA cycle, resulting in reduced production of acetyl-CoA within the mitochondria. Moreover, the upregulation of PDK is closely associated with a poor prognosis in cancer [88, 89, 90]. Therefore, pyruvate assumes a distinctive and crucial role in linking aerobic and anaerobic metabolism, serving as a central determinant for sustaining cancer growth [91].

2.1.7 Hypoxia and Nutrient Restriction with Acetyl-CoA

Hypoxia plays a crucial role in influencing acetyl-CoA metabolism in cancer, contributing to the survival and proliferation of specific cancer cells while activating hypoxia inducible factor (HIF) [92]. The upregulation of HIF-1α escalates the expression of GLUT1, facilitating increased glucose uptake and promoting its metabolic conversion to pyruvate, followed by the synthesis of acetyl-CoA [93, 94]. In conditions of hypoxia and nutrient restriction, there is elevation in mitochondrial PDK expression, leading to inactivation of the PDC. Consequently, the synthesis of acetyl-CoA by the PDC within the nucleus is of particular importance [95]. In response to hypoxia and low serum stimulation, ACSS2 expression is induced due to nutrient deficiency. This mechanism involves the translocation of ACSS2 from the cytoplasm to the nucleus, resulting in an elevated level of acetyl-CoA within the nucleus. This promotes lipid synthesis, acetylation of autophagy genes, histones, and transcription factor HIF-2α, enhancing erythropoiesis [96, 97, 98]. Highly expressed ACSS1/2 in cancer cells utilize acetate to increase acetyl-CoA production while simultaneously regulating the level of FA synthetase (FAS) to acetylate H3K9, H3K27, and H3K56, thereby enhancing the survival of hypoxic tumors [99]. During periods of starvation or exercise, the liver generates acetyl-CoA through extensive FAO, surpassing citric acid synthesis. This excess acetyl-CoA enters the TCA cycle, elevates ATP levels, and is utilized for ketone body production [100, 101, 102]. Currently, aerobic and energy metabolism are considered promising strategies for cancer prevention [103].

2.2 Other Types of CoA and Cancer

Acyl-CoA constitutes a category of metabolic intermediates characterized by acyl groups forming thioester bonds with CoA. These molecules play a pivotal role in facilitating acyl transfer within various metabolic processes [104].

Hedgehog acetyltransferase (Hhat) has the capability to attach non-palmitate FAs to a recombinant protein known as sonic hedgehog (Shh), a crucial gene within the Hhat signaling pathway. It has been suggested that heterogenous fatty acylation may impact Shh signaling during development and in cancer cells [105]. Aberrant re-expression and signaling of Shh have been linked to cancer. In addition to palmitic acid, the N-terminal segment of Shh features binding sites for myristate acid, palmitoleic acid, stearic acid, oleic acid, and arachidonic acid [106, 107, 108].

Succinylation plays a significant role in cancer and inflammation. In instances where SIRT5 is absent or succinic dehydrogenase malfunctions, the succinylation process within the mitochondria is disrupted, resulting in succinic acid accumulation. This, in turn, triggers an overabundance of histone succinylation, disrupting the metabolic equilibrium and elevating the cancer risk [109]. The enzyme alpha-ketoglutarate dehydrogenase (α-KGDH) within the nucleus forms an association with lysine acetyltransferase 2A (KAT2A), which binds to succinyl-CoA, ultimately succinylating lysine 79 on histone H3. Inhibiting the nuclear entry of α-KGDH or suppressing KAT2A expression can diminish the expression of target genes, impeding cancer growth [110]. Elevated levels of succinyl-CoA have been observed in high-grade squamous intraepithelial lesions, with succinate-CoA ligase ADP-forming subunit beta (SUCLA2) identified as a driver of anoikis resistance and metastasis. SUCLA2 decreases succinyl-CoA concentration around glutaminase, thereby lowering succinyl-CoA modification on glutaminase [111, 112]. The carcinogenic impact of HMG-CoA synthase 1 (HMGCS1) is evident across various human cancers, with research highlighting its increased expression in tumors such as digestive tract, breast, and uterine cancers [113].

Mevalonic acid (MVA) is synthesized from HMG-CoA through the action of HMG-CoA reductase (HMGCR). The MVA pathway demonstrates increased expression in leukemia, lymphoma, multiple myeloma, breast cancer, liver cancer, pancreatic cancer, esophageal cancer, and prostate cancer [114]. Consequently, inhibiting MVA levels holds promise for potentially impeding cancer growth. Statins, acting as inhibitors of HMGCR, intervene in the MVA pathway, inhibiting the synthesis of cholesterol, geranyl pyrophosphate, and farnesyl pyrophosphate. Early clinical studies have underscored the therapeutic potential of statins in treating acute myeloid leukemia [115]. In PDAC cells, there is activation of ketone body metabolism, with β-hydroxybutyrate (β-HB) serving as an intracellular or systemic fuel, thereby promoting the growth and progression of PDAC. Disruption of HMG CoA lyase (HMGCL) significantly retards the growth of PDAC, resulting in β-HB stimulating its metastasis and dissemination to the liver [116]. Future applications may involve exploring other CoA metabolic pathways for cancer treatment by preventing the proliferation of cancerous cells (Fig. 1).

2.3 CoA and Cardiovascular Disease

CoA and its thioester derivatives play a significant role in CVDs, exerting regulatory control over processes such as FAO, ketogenic reactions, and cholesterol synthesis [117]. Globally, CVD stands out as the foremost and deadliest health concern [118].

The key enzyme, HMG-CoA holds paramount significance in the regulation of CVDs. HMGCS2 and HMGCR, in conjunction with acetyl-CoA, stearoyl-CoA desaturase (SCD1), and fatty acyl-CoA synthetase, establish critical connections to CVDs across various dimensions. The expression levels of these enzymes can differ across different tissues of the human body [119]. The relationship between CVD and CoA metabolism can be delineated across three principal aspects: the liver, kidney, and the heart and cardiovascular system.

2.3.1 Liver, Kidney and CoA Metabolism

The liver and kidneys are vital organs involved in human metabolism and detoxification, exerting an influence on cardiovascular health through their metabolic and regulatory functions [120]. Impairment of liver function can lead to cholesterol accumulation within the liver, contributing to cardiovascular ailments like hypercholesterolemia and atherosclerosis. Anomalous renal function or kidney disorders, including chronic kidney disease and kidney failure, exhibit a close association with CVD [121]. Moreover, kidney ailments have the potential to instigate or exacerbate hypertension, thereby heightening the susceptibility to atherosclerosis and CVDs [122].

The synthesis of ketone bodies intricately intersects with the regulatory mechanisms of CVDs, mediated by CoA [123]. HMGCS2, a rate-limiting enzyme located in the mitochondria, orchestrates the synthesis of ketone bodies by facilitating the production of HMG-CoA [124]. This enzyme is predominantly expressed in the mitochondrial matrix of the liver, actively participating in ketogenesis, and supplying lipid-derived energy during fasting [125, 126]. In mammals, β-hydroxybutyrate (β-HB) stands out as the principal ketone body, synthesized primarily within the liver mitochondria and utilized by extrahepatic organs, notably the heart, thereby influencing cardiac energy metabolism [127]. The synthesis pathway of ketone bodies entails the entry of FFAs into the mitochondria, β-oxidation to form acetoacetyl-CoA, subsequent formation of HMG-CoA through the catalysis of acetyl-CoA by HMGCS2, and the ultimate conversion to b-hydroxybutyric acid (b-HB) through the collaborative actions of HMGCL and NADH [128]. A study conducted by Song et al. [129] revealed that elevated levels of β-HB can induce arrhythmias, contributing to the development of arrhythmic cardiomyopathy (AC). Conversely, Al-zaid et al. [130] discovered cardioprotective effects associated with β-HB. Consequently, fluctuations in β-HB levels present both advantages and disadvantages in the context of CVDs. Investigating the specific mechanisms underlying these effects represent a promising avenue for targeting CVDs in the future (Fig. 2).

Fig. 2.

Synthesis of Low-density lipoprotein cholesterol (LDL-C), β-hydroxybutyrate (β-HB), and fatty acids (FA). LDL-C serves as a crucial marker for atherosclerotic cardiovascular disease (ASCVD), and its synthesis pathway involves acetyl-CoA, HMG-CoA, and mevalonic acid (MVA), ultimately leading to the formation of LDL-C. β-HB is predominantly synthesized in the mitochondrial matrix of the liver, subsequently entering the bloodstream for transport to various tissues and organs. Pyruvate enters the mitochondria and undergoes conversion to acetyl-CoA, which, under the influence of acetyl-CoA carboxylase (ACC), is further transformed into malonyl-CoA. Notably, malonyl-CoA exerts inhibitory effects on FA oxidation (FAO) while concurrently promoting the synthesis of fatty acid synthase (FAS).

Chronic kidney disease precipitates elevated levels of triglycerides (TGs) in the body, thereby fostering dyslipidemia and CVDs [131]. Within pyruvate metabolism, ACC acts as a catalyst for converting acetyl-CoA, producing malonyl-CoA, inhibiting FAO, and supplying substrates for FA biosynthesis [132]. Both synthetic saturated FAs and monounsaturated FAs (MUFAs) become integrated into diverse lipid species, including TGs and phospholipids [133]. By diminishing the levels of synthetic substrates—acetyl-CoA and malonyl-CoA—ACC inhibitors delay the accumulation of lipids through DNL. Inhibiting ACC holds promise for reducing liver malonyl-CoA levels, mitigating lipid toxicity and bolstering FAO [134].

Stearoyl-CoA desaturase (SCD1) catalyzes the production of MUFAs, exerting influence as an energy metabolism regulator [135]. Dyslipidemia commonly observed in chronic kidney disease patients manifests as irregular levels of high-density lipoprotein cholesterol (HDL-C) and TGs [136]. SCD1 localizes alongside diacylglycerol acyltransferase 2 in the endoplasmic reticulum, tightly regulating triacylglycerol (TAG) synthesis in the liver [137]. Overexpression of SCD1 correlates with hypertriglyceridemia and heightened susceptibility to cardiovascular complications [138]. Focused exploration into the modulation of FA metabolism governed by ACC and SCD1 could reveal a deeper comprehension of the onset and progression of dyslipidemia in chronic kidney disease and cardiovascular disorders.

2.3.2 Heart and CoA Metabolism

Cardiac dysfunction is intricately linked to modifications in cardiac substrate and energy metabolism [139]. The heart, being an organ characterized by high energy demands, heavily depends on CoA and its derivatives for the metabolic processing of FAs [140]. Concurrently, serving as the central hub of the cardiovascular system, the heart is vulnerable to various CVDs, encompassing coronary heart disease, cardiomyopathy, and heart failure [141].

Within the context of FAO, CoA plays a crucial role in converting FAs into energy. Genetic defects in CoA biosynthesis disrupt proper FA metabolism, resulting in the accumulation of related metabolites at the site of the enzyme defect. Furthermore, this genetic aberration diminishes the substrate pool available for energy production. The accumulation of metabolites, coupled with ensuing energy deficiency, can adversely impact the functionality of heart cells, culminating in the development of CVDs including cardiomyopathy and arrhythmia [142]. Deficiencies in mitochondrial trifunctional protein (MTP/TFP), stemming from mutations in the hydroxyacyl-CoA dehydrogenase trifunctional multienzyme complex subunit alpha (HADHA) or HADHB genes responsible for critical proteins in CoA biosynthesis, can impair FAO. This impairment, in turn, affects the energy metabolism and function of heart cells. However, the specific mechanistic intricacies remain elusive, and to date, no established treatments exist for MTP/TFP deficiency [143].

Medium chain acyl-CoA dehydrogenase (MCAD) serves as a critical regulator of FAO in the heart. Genetic defects in CoA biosynthesis may lead to abnormal expression of MCAD, impacting the energy metabolism and function of the heart. Patients with heart failure exhibit disruptions in cardiac energy metabolism, prompting consideration of addressing metabolic defects as a potential treatment strategy. Consequently, utilizing the gene delivery method targeting MCAD may help restore heart energy metabolism, thereby safeguarding the heart from damage associated with pathological remodeling [144]. However, further research is needed to determine how genetic defects affect CoA biosynthesis and contribute to the development of CVD.

In the heart, elevated levels of non-histone acetylation enhance FAO in conditions such as heart failure [145]. FA β-oxidation refers to the degradation process of long-chain FAs into acetyl-CoA within the mitochondria. This intricate process relies on the activation of long-chain FAs outside the mitochondria by acyl-CoA synthetase and carnitine palmitoyl transferase 1 (CPT1), followed by the transport of acyl-CoA into the mitochondria [146]. An important regulatory point in FAO is the inhibition of CPT1 by malonyl-CoA, subsequently reducing the uptake of FAs by the mitochondria [147]. Enzymes involved in FA β-oxidation, including fatty acyl CoA synthetase, CPT1, long chain acyl CoA dehydrogenase (LCAD), and β-hydroxyl-CoA dehydrogenase (β-HAD), serve as critical targets of mitochondrial acetylation [148]. The acyl-CoA thioesterase (ACOT) family is actively involved in FA activation and the degradation of fat-derived acyl-CoA, generating FFAs and CoA [149]. Acyl-CoA thioesterase 1 (ACOT1) has demonstrated the ability to protect heart function during diabetes and sepsis by modulating long-chain acyl-CoA (LCA) and FA levels in the cytoplasm [150, 151].

SIRT3 plays a crucial role as a mitochondrial deacetylase, and its activation holds the potential to restore mitochondrial metabolic homeostasis [152]. Acetyl-CoA acetylates mitochondrial proteins involved in FA beta oxidation, such as LCAD and β-HAD, and SIRT3 regulates their activity by deacetylating these proteins [147]. The consumption of a high-fat diet can reduce the expression of SIRT3, leading to increased lysine acetylation of cardiac mitochondrial proteins and accelerated FAO [153, 154]. Consequently, the dysregulation of SIRT3-mediated acetylation of lysine residues on mitochondrial protein may contribute to cardiometabolic inflexibility in individuals with obesity or diabetes.

2.3.3 Cardiovascular System and CoA Metabolism

Serum cholesterol emerges as a crucial focal point for CoA in the regulation of atherosclerotic CVD (ASCVD) [155, 156]. Acyl-CoA constitutes a significant element of CoA derivatives. Elevated plasma levels of acyl-CoA binding protein (ACBP) are linked to increased susceptibility to CVDs, displaying correlations with cardiovascular risk factors such as total free cholesterol and TGs. Conversely, ACBP levels exhibit a negative correlation with the protective HDL levels [157]. HDL is widely recognized as a protective factor against CVD and serves as an indicator for anti-atherosclerosis effects [158]. Globally, ASCVD stands out as the most prevalent and leading cause of mortality among CVDs [159]. Abnormal blood lipid levels, particularly elevated levels of low-density lipoprotein cholesterol (LDL-C), represent a significant risk factor for CVD and are considered a central aspect of ASCVD pathogenesis (Fig. 2). CoA itself does not directly reduce cholesterol levels. The majority of LDL-C is derived from the enzymatic activity of HMGCR, which catalyzes the conversion of HMG-CoA into mevalonate, a crucial precursor in cholesterol synthesis [160]. HMGCR serves as a typical target for statin medications, which competitively bind to the active sites of HMGCR. This inhibition results in reduced serum cholesterol levels, slowing down disease progression [161].

In addition to HMGCR inhibitors, flavonoids have demonstrated the ability to lower LDL-C levels [162]. They promote the oxidative metabolism of FAs and cholesterol by activating CoA and increasing the expression of fatty acyl-CoA oxidase. In the presence of CoA or mercaptan donors, hemoglobin converts ingested flavonoids into conjugates [163].

At lower doses, statins promote angiogenesis, whereas higher doses inhibit protein isopentenylation as well as cell and blood vessel development [161]. Xu et al. [164] discovered that alpha-ketoisovaleric acid and propionyl-CoA, metabolites of valine, strongly activate platelets, leading to propionylation of propionin-3-propionylation and subsequent thrombosis.

2.4 CoA and Neurodegenerative Diseases

Neurodegenerative diseases encompass a spectrum of disorders leading to dysfunction within the nervous system, including well-known conditions such as Alzheimer’s disease, Parkinson’s disease, Huntington’s disease, and amyotrophic lateral sclerosis. An overarching characteristic shared by these diseases is the abnormal aggregation of proteins in the cytoplasm or nucleus. The genetic foundations of neurodegenerative diseases are diverse, involving both inherited disorders resulting from rare genetic mutations and the influences of common risk genes [165].

Among these conditions are pantothenate kinase-related neurodegeneration (PKAN), CoA synthase protein-associated neurodegeneration (CoPAN), protein lipoylation, and pyruvate dehydrogenase E2, four hereditary neurodegenerative diseases. Each of these disorders is linked to deficiencies in distinct metabolic processes within CoA biosynthesis [166]. Two autosomal recessive neurodegenerative diseases stem from mutations in genes encoding enzymes in the CoA pathway. PKAN and CoPAN, representing the initial and final enzymes of the CoA biosynthesis pathway related to neurodegenerative diseases, are part of the heterogeneous group of neurodegenerative diseases (neurodegeneration with brain iron accumulation) characterized by iron deposition in the brain [167].

Genetic variations lead to dysfunction of the pantothenate kinase 2 (PANK2) gene, resulting in reduced activity of PANK2. This alteration indirectly impacts the initial step of CoA biosynthetic pathways, where pantothenic acid is converted to 4-phosphopantothenic acid. The aberrant function of PANK2 causes a decline in the phosphorylation of pantothenic acid to 4-phosphopantothenic acid, leading to the diminished concentration of CoA in vulnerable tissues, including the brain [168]. In fibroblasts of CoPAN patients, there is a notable reduction in CoA synthetase (COASY) expression, accompanied by lower levels of acetyl-CoA and total CoA. COASY, the gene encoding the CoA synthetase, undergoes mutation, resulting in decreased CoA levels. This, in turn, influences intracellular acetyl-CoA levels, leading to an abnormal autophagy process [169].

Studies indicate that disruptions in cholesterol metabolism may be linked to the onset and progression of neurodegenerative diseases [170]. Intracellular cholesterol levels are tightly regulated through four processes: de novo synthesis, uptake, export, and storage [171]. Cholesterol metabolism in the brain operates relatively independently and primarily relies on the synthesis carried out by glial cells and neurons themselves [172]. Cholesterol synthesis takes place primarily in the endoplasmic reticulum and involves the conversion of acetyl-CoA into HMG-CoA through the activity of the irreversible rate-limiting enzyme HMG-CoA synthetase. This step is followed by the action of HMGCR, which converts HMG-CoA into mevalonate, squalene, and lanosterol, constituting the mevalonate pathway. In brain cholesterol synthesis, lanosterol is involved through two pathways: the Bloch pathway in astrocytes and the Kandutsch-Russel pathway in neurons (Fig. 3). The synthesized cholesterol is stored as lipid droplets and can be transported to plasma lipoproteins for the production of other substances [173, 174, 175]. Notably, adult neurons are less efficient at cholesterol synthesis compared to astrocytes [176, 177]. An imbalance in brain cholesterol levels can contribute to the development of neurodegenerative diseases, such as Alzheimer’s disease, Parkinson’s disease, and Huntington’s disease [178, 179].

Fig. 3.

Schematic illustration of CoA metabolism in neurodegenerative diseases. Cholesterol synthesis in neurons predominantly occurs through the Kandutsch-Russell pathway, while in astrocytes, it primarily takes place via the Bloch pathway. Previously, acetyl-CoA was sequentially converted to HMG-CoA, which was then transformed into mevalonate by the action of HMG-CoA reductase (HMGCR), subsequently supplying lanosterol for cholesterol synthesis. In adult neurons, the main source of cholesterol is provided by astrocytes, which supply it through the LDLR-related protein 1/LDLR in the form of apolipoproteins. Upon internalization, these apolipoproteins undergo conversion to free cholesterol within the endosome/lysosome compartment. In cases of cholesterol excess, intracellular esterification takes place, leading to the storage of cholesterol in lipid droplets. The stored cholesterol is then subject to conversion into oxysterols, particularly 24(S)-hydroxycholesterol (24-OHC).

3. Conclusion and Perspective

CoA plays a crucial role in regulating lipid metabolism, influencing cellular energy status and signal transmission. Numerous reports have underscored the correlation between CoA-related lipid disorders and a spectrum of diseases, including cancer, neurodegenerative diseases, and CVDs [15, 174, 180]. The primary aspect of lipid metabolism involves the synthesis and degradation of FAs. During FA synthesis, cellular reactions involving acetyl-CoA and malonyl-CoA lead to the formation of long-chain FAs, such as palmitic acid [181]. This process primarily takes place in the liver and adipose tissue, contributing to the reduction of β-oxidation, as well as the utilization and synthesis of endogenous storage forms of neutral lipids, specifically TGs [182, 183]. Conversely, the degradation process mainly occurs in the mitochondria through the FAO pathway, a pivotal player in sustaining a balanced energy state within the body [184]. This pathway catalyzes the conversion of long-chain FAs into acetyl-CoA, ultimately generating energy through the TCA cycle [185].

AMPK plays a multifaceted role in cell metabolism, including lipid metabolism, which is implicated in cancer, CVD, and neurodegenerative diseases (Fig. 4). AMPK orchestrates the regulation of FFA concentration by activating FAO while concomitantly inhibiting lipolysis and fat synthesis [186, 187]. Within the FA metabolism, AMPK facilitates FA uptake and activates FAO by suppressing the de novo synthesis of FAS, cholesterol, and TGs [188]. The inhibition of FA synthesis by AMPK involves the modulation of the phosphorylation of ACC and SREBP1c [189]. Notably, ACC1 is highly expressed in adipose tissue, regulating the intracellular availability of acetyl-CoA involved in FA synthesis to meet the material and energy requirements necessary for disease development [190]. The synthesis of FAs is under the control of SREBPs, with SREBP1 primarily regulating FA synthesis and the expression of the low density lipoprotein receptor (LDLR), while SREBP2 primarily regulates the expression of cholesterol biosynthesis genes [191, 192]. Additionally, AMPK inhibits cholesterol synthesis by suppressing the phosphorylation of the key enzyme HMGCR [193]. The benefits of statins stem from their ability to inhibit HMGCR, subsequently upregulating the cholesterol uptake gene, increasing expression of the LDLR, and consequently reducing blood cholesterol levels [194]. Additionally, the cytoplasmic enzyme ACLY, highly expressed in adipose-synthesizing tissues, is positioned upstream of HMGCR [195]. ACLY is located upstream of ACC, and the acetyl-CoA produced by ACLY can be utilized for the synthesis of cholesterol and FAs [196, 197, 198]. Likewise, the utilization of the substrate N-acetyl-aspartate by ACSS2 is strongly involved in acetyl-CoA production in adipocytes. Importantly, this process is not limited to adipocytes and also occurs within the nervous system [199, 200, 201, 202, 203]. However, inhibition of ACLY can suppress cholesterol and FA synthesis while increasing the expression of LDLR in cells [204]. AMPK impacts TG synthesis by inhibiting glycerol-3-phosphate acyltransferase, the enzyme responsible for catalyzing the initial step in TG synthesis [205, 206]. Acetyl-CoA is transported intracellularly into mitochondria for β-oxidation by carnitine palmitoyl-transferase 1 (CPT1). AMPK enhances CPT1 activity by inhibiting the phosphorylation of ACC2, thereby elevating FAO levels. ACC2, an enzyme located in the outer membrane of mitochondria in close proximity to CPT1, effectively impedes malonyl-CoA production [207].

Fig. 4.

Lipid metabolism in cancer, CVDs, and neurodegenerative diseases (NDD). The activation of AMP-activated protein kinase (AMPK) inhibits the activity of sterol regulatory element-binding protein (SREBP), resulting in a subsequent decrease in the expression of two key enzymes involved in HMG-CoA biosynthesis, namely HMG-CoA reductase (HMGCR) and HMG-CoA synthase (HMGCS). This negative regulation of FA and cholesterol synthesis ultimately impact the overall process of lipid metabolism in various diseases.

In conclusion, this study underscores the significant role of CoA and its derivatives in the pathogenesis and progression of various ailments such as cancer, CVDs, and neurodegenerative diseases. Moreover, this study highlights the immense potential of targeting CoA-associated processes for disease prevention and treatment.

Abbreviations

ACAT1, Acetyl-CoA acetyltransferase 1; ACBP, Acyl-CoA binding protein; ACC, Acetyl-CoA carboxylase; ACLY, ATP-citrate lyase; ACOT, Acyl-CoA thioesterase; ACOT12, Acyl-CoA thioesterase 12; ACSS2, Acyl-CoA short chain synthetase family member 2; AKT, Serine/threonine-protein kinases; AML, Acute myeloid leukemia; AMPK, AMP-activated protein kinase; ASCVD, Atherosclerotic cardiovascular disease; BMI, Body mass index; BRD4, Bromodomain containing 4; CAF, Cancer-associated fibroblasts; CoA, Coenzyme A; COASY, CoA synthetase; CoPAN, CoA synthase protein associated neurodegeneration; CPT1, Carnitine palmitoyl transferase 1; CVD, Cardiovascular diseases; DGAT1, Diacylglycerol O-acyltransferase 1; DGAT2, Diacylglycerol acyltransferase 2; DNL, De novo lipogenesis; EMT, Epithelial-mesenchymal transition; FAO, Fatty acid oxidation; FAS, Fatty acid synthetase; GBM, Glioblastoma; GLUT1, Glucose transporter 1; GPAT, Glycerol-3-phosphate acyltransferase; HDL-C, High-density lipoprotein cholesterol; Hhat, Hedgehog acetyltransferase; HIF, Hypoxia inducible factor; HMGCL, 3-hydroxy-3-methylglutaryl CoA lyase; HMG-CoA, 3-hydroxy-3-methylglutaryl CoA; HMGCS1, 3-hydroxy-3-methylglutaryl CoA synthase 1; HSIL, High-grade squamous intraepithelial lesions; IRF4, Interferon regulatory factor 4; KAT2A, Lysine Acetyltransferase 2A; LCAD, long chain acyl CoA dehydrogenase; LDL-C, Low-density lipoprotein cholesterol; LDLR, Low-density lipoprotein receptor; MCAD, Medium chain acyl-CoA dehydrogenase; MePAN, Protein lipoylation; MTP, Mitochondrial trifunctional protein; MUFA, Monounsaturated fatty acids; MVA, Mevalonic acid; NBIA, Neurodegeneration with brain iron accumulation; PANK2, Pantothenate kinase 2; PDA, Pancreatic ductal adenocarcinoma; PDC, Pyruvate dehydrogenase complex; PDH-E2, Pyruvate dehydrogenase-E2; PDK, Pyruvate dehydrogenase kinase; PFK1, Phosphofructokinase 1; PFK2, Phosphofructokinase 2; PKAN, Pantothenate kinase-related neurodegeneration; SCD1, Stearoyl-CoA desaturase; SFA, Saturated fatty acids; Shh, Sonic hedgehog; SIRT3, Sirtuin3; SIRT5, Sirtuin 5; SREBP1c, Sterol regulatory element-binding protein 1c; SREBPs, Sterol regulatory element-binding proteins; SUCLA2, Succinate-CoA ligase ADP-forming subunit beta; TAG, Triglyceride; TCA, Tricarboxylic acid; TGIF, TG-interacting factor; TWIST2, Twist family BHLH transcription factor 2; α-KGDH, Alpha-ketoglutarate dehydrogenase; β-HAD, β-hydroxyl-CoA dehydrogenase; β-HB, β-hydroxybutyrate.

Author Contributions

JX, XC, JZ, MZ, and HM conceived the review. JX and HM supervised the review. All authors wrote the draft. JX revised the manuscript. All authors read and agreed to publish the paper. All authors have participated sufficiently in the work and agreed to be accountable for all aspects of the work.

Ethics Approval and Consent to Participate

Not applicable.

Acknowledgment

Not applicable.

Funding

This study was supported by Key scientific research projects of Hubei polytechnic University (No. 23xjz08A).

Conflict of Interest

The authors declare no conflict of interest.

References
[1]
Dibble CC, Barritt SA, Perry GE, Lien EC, Geck RC, DuBois-Coyne SE, et al. PI3K drives the de novo synthesis of coenzyme A from vitamin B5. Nature. 2022; 608: 192–198.
[2]
Dobrzyn P. CoA in Health and Disease. International Journal of Molecular Sciences. 2022; 23: 4371.
[3]
Cai L, Sutter BM, Li B, Tu BP. Acetyl-CoA induces cell growth and proliferation by promoting the acetylation of histones at growth genes. Molecular Cell. 2011; 42: 426–437.
[4]
Sutendra G, Kinnaird A, Dromparis P, Paulin R, Stenson TH, Haromy A, et al. A nuclear pyruvate dehydrogenase complex is important for the generation of acetyl-CoA and histone acetylation. Cell. 2014; 158: 84–97.
[5]
Wellen KE, Hatzivassiliou G, Sachdeva UM, Bui TV, Cross JR, Thompson CB. ATP-citrate lyase links cellular metabolism to histone acetylation. Science (New York, N.Y.). 2009; 324: 1076–1080.
[6]
Sivanand S, Rhoades S, Jiang Q, Lee JV, Benci J, Zhang J, et al. Nuclear Acetyl-CoA Production by ACLY Promotes Homologous Recombination. Molecular Cell. 2017; 67: 252–265.e6.
[7]
Schwer B, Bunkenborg J, Verdin RO, Andersen JS, Verdin E. Reversible lysine acetylation controls the activity of the mitochondrial enzyme acetyl-CoA synthetase 2. Proceedings of the National Academy of Sciences of the United States of America. 2006; 103: 10224–10229.
[8]
Zhao S, Torres A, Henry RA, Trefely S, Wallace M, Lee JV, et al. ATP-Citrate Lyase Controls a Glucose-to-Acetate Metabolic Switch. Cell Reports. 2016; 17: 1037–1052.
[9]
Tong L. Acetyl-coenzyme A carboxylase: crucial metabolic enzyme and attractive target for drug discovery. Cellular and Molecular Life Sciences: CMLS. 2005; 62: 1784–1803.
[10]
Abu-Elheiga L, Matzuk MM, Abo-Hashema KA, Wakil SJ. Continuous fatty acid oxidation and reduced fat storage in mice lacking acetyl-CoA carboxylase 2. Science (New York, N.Y.). 2001; 291: 2613–2616.
[11]
Pietrocola F, Galluzzi L, Bravo-San Pedro JM, Madeo F, Kroemer G. Acetyl coenzyme A: a central metabolite and second messenger. Cell Metabolism. 2015; 21: 805–821.
[12]
Liu X, Si W, He L, Yang J, Peng Y, Ren J, et al. The existence of a nonclassical TCA cycle in the nucleus that wires the metabolic-epigenetic circuitry. Signal Transduction and Targeted Therapy. 2021; 6: 375.
[13]
DeBose-Boyd RA. Feedback regulation of cholesterol synthesis: sterol-accelerated ubiquitination and degradation of HMG CoA reductase. Cell Research. 2008; 18: 609–621.
[14]
Guertin DA, Wellen KE. Acetyl-CoA metabolism in cancer. Nature Reviews. Cancer. 2023; 23: 156–172.
[15]
He W, Li Q, Li X. Acetyl-CoA regulates lipid metabolism and histone acetylation modification in cancer. Biochimica et Biophysica Acta. Reviews on Cancer. 2023; 1878: 188837.
[16]
Carrer A, Trefely S, Zhao S, Campbell SL, Norgard RJ, Schultz KC, et al. Acetyl-CoA Metabolism Supports Multistep Pancreatic Tumorigenesis. Cancer Discovery. 2019; 9: 416–435.
[17]
Hwang CY, Choe W, Yoon KS, Ha J, Kim SS, Yeo EJ, et al. Molecular Mechanisms for Ketone Body Metabolism, Signaling Functions, and Therapeutic Potential in Cancer. Nutrients. 2022; 14: 4932.
[18]
Goudarzi A. The recent insights into the function of ACAT1: A possible anti-cancer therapeutic target. Life Sciences. 2019; 232: 116592.
[19]
An D, Zeng Q, Zhang P, Ma Z, Zhang H, Liu Z, et al. Alpha-ketoglutarate ameliorates pressure overload-induced chronic cardiac dysfunction in mice. Redox Biology. 2021; 46: 102088.
[20]
Feng X, Zhang L, Xu S, Shen AZ. ATP-citrate lyase (ACLY) in lipid metabolism and atherosclerosis: An updated review. Progress in Lipid Research. 2020; 77: 101006.
[21]
Cotter DG, Schugar RC, Crawford PA. Ketone body metabolism and cardiovascular disease. American Journal of Physiology. Heart and Circulatory Physiology. 2013; 304: H1060–H1076.
[22]
Yang SH, Li W, Sumien N, Forster M, Simpkins JW, Liu R. Alternative mitochondrial electron transfer for the treatment of neurodegenerative diseases and cancers: Methylene blue connects the dots. Progress in Neurobiology. 2017; 157: 273–291.
[23]
Mi Y, Qi G, Vitali F, Shang Y, Raikes AC, Wang T, et al. Loss of fatty acid degradation by astrocytic mitochondria triggers neuroinflammation and neurodegeneration. Nature Metabolism. 2023; 5: 445–465.
[24]
Casserly I, Topol E. Convergence of atherosclerosis and Alzheimer’s disease: inflammation, cholesterol, and misfolded proteins. Lancet (London, England). 2004; 363: 1139–1146.
[25]
Galper J, Dean NJ, Pickford R, Lewis SJG, Halliday GM, Kim WS, et al. Lipid pathway dysfunction is prevalent in patients with Parkinson’s disease. Brain: a Journal of Neurology. 2022; 145: 3472–3487.
[26]
Peng Y, Kim MJ, Hullinger R, O’Riordan KJ, Burger C, Pehar M, et al. Improved proteostasis in the secretory pathway rescues Alzheimer’s disease in the mouse. Brain: a Journal of Neurology. 2016; 139: 937–952.
[27]
Marcus R. What Is Huntington Disease? JAMA. 2023; 330: 1014.
[28]
Vaghari-Tabari M, Ferns GA, Qujeq D, Andevari AN, Sabahi Z, Moein S. Signaling, metabolism, and cancer: An important relationship for therapeutic intervention. Journal of Cellular Physiology. 2021; 236: 5512–5532.
[29]
Vander Heiden MG, Cantley LC, Thompson CB. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science (New York, N.Y.). 2009; 324: 1029–1033.
[30]
Vaupel P, Schmidberger H, Mayer A. The Warburg effect: essential part of metabolic reprogramming and central contributor to cancer progression. International Journal of Radiation Biology. 2019; 95: 912–919.
[31]
Huang Z, Zhang M, Plec AA, Estill SJ, Cai L, Repa JJ, et al. ACSS2 promotes systemic fat storage and utilization through selective regulation of genes involved in lipid metabolism. Proceedings of the National Academy of Sciences of the United States of America. 2018; 115: E9499–E9506.
[32]
Currie E, Schulze A, Zechner R, Walther TC, Farese RV, Jr. Cellular fatty acid metabolism and cancer. Cell Metabolism. 2013; 18: 153–161.
[33]
Yoshii Y, Waki A, Furukawa T, Kiyono Y, Mori T, Yoshii H, et al. Tumor uptake of radiolabeled acetate reflects the expression of cytosolic acetyl-CoA synthetase: implications for the mechanism of acetate PET. Nuclear Medicine and Biology. 2009; 36: 771–777.
[34]
Li X, Yu W, Qian X, Xia Y, Zheng Y, Lee JH, et al. Nucleus-Translocated ACSS2 Promotes Gene Transcription for Lysosomal Biogenesis and Autophagy. Molecular Cell. 2017; 66: 684–697.e9.
[35]
Liu M, Liu N, Wang J, Fu S, Wang X, Chen D. Acetyl-CoA Synthetase 2 as a Therapeutic Target in Tumor Metabolism. Cancers. 2022; 14: 2896.
[36]
Schug ZT, Peck B, Jones DT, Zhang Q, Grosskurth S, Alam IS, et al. Acetyl-CoA synthetase 2 promotes acetate utilization and maintains cancer cell growth under metabolic stress. Cancer Cell. 2015; 27: 57–71.
[37]
Comerford SA, Huang Z, Du X, Wang Y, Cai L, Witkiewicz AK, et al. Acetate dependence of tumors. Cell. 2014; 159: 1591–1602.
[38]
Mashimo T, Pichumani K, Vemireddy V, Hatanpaa KJ, Singh DK, Sirasanagandla S, et al. Acetate is a bioenergetic substrate for human glioblastoma and brain metastases. Cell. 2014; 159: 1603–1614.
[39]
Liang Y, Pi H, Liao L, Tan M, Deng P, Yue Y, et al. Cadmium promotes breast cancer cell proliferation, migration and invasion by inhibiting ACSS2/ATG5-mediated autophagy. Environmental Pollution (Barking, Essex: 1987). 2021; 273: 116504.
[40]
Shah A, Melhuish TA, Fox TE, Frierson HF, Jr, Wotton D. TGIF transcription factors repress acetyl CoA metabolic gene expression and promote intestinal tumor growth. Genes & Development. 2019; 33: 388–402.
[41]
Bae JM, Kim JH, Oh HJ, Park HE, Lee TH, Cho NY, et al. Downregulation of acetyl-CoA synthetase 2 is a metabolic hallmark of tumor progression and aggressiveness in colorectal carcinoma. Modern Pathology: an Official Journal of the United States and Canadian Academy of Pathology, Inc. 2017; 30: 267–277.
[42]
Li Z, Liu H, He J, Wang Z, Yin Z, You G, et al. Acetyl-CoA Synthetase 2: A Critical Linkage in Obesity-Induced Tumorigenesis in Myeloma. Cell Metabolism. 2021; 33: 78–93.e7.
[43]
Covarrubias AJ, Aksoylar HI, Yu J, Snyder NW, Worth AJ, Iyer SS, et al. Akt-mTORC1 signaling regulates Acly to integrate metabolic input to control of macrophage activation. eLife. 2016; 5: e11612.
[44]
Verschueren KHG, Blanchet C, Felix J, Dansercoer A, De Vos D, Bloch Y, et al. Structure of ATP citrate lyase and the origin of citrate synthase in the Krebs cycle. Nature. 2019; 568: 571–575
[45]
Bose S, Ramesh V, Locasale JW. Acetate Metabolism in Physiology, Cancer, and Beyond. Trends in Cell Biology. 2019; 29: 695–703.
[46]
Icard P, Wu Z, Fournel L, Coquerel A, Lincet H, Alifano M. ATP citrate lyase: A central metabolic enzyme in cancer. Cancer Letters. 2020; 471: 125–134.
[47]
Lee JV, Carrer A, Shah S, Snyder NW, Wei S, Venneti S, et al. Akt-dependent metabolic reprogramming regulates tumor cell histone acetylation. Cell Metabolism. 2014; 20: 306–319.
[48]
Zheng W, Tasselli L, Li TM, Chua KF. Mammalian SIRT6 Represses Invasive Cancer Cell Phenotypes through ATP Citrate Lyase (ACLY)-Dependent Histone Acetylation. Genes. 2021; 12: 1460.
[49]
Kumari R, Deshmukh RS, Das S. Caspase-10 inhibits ATP-citrate lyase-mediated metabolic and epigenetic reprogramming to suppress tumorigenesis. Nature Communications. 2019; 10: 4255.
[50]
Khwairakpam AD, Shyamananda MS, Sailo BL, Rathnakaram SR, Padmavathi G, Kotoky J, et al. ATP citrate lyase (ACLY): a promising target for cancer prevention and treatment. Current Drug Targets. 2015; 16: 156–163.
[51]
Zhang C, Liu J, Huang G, Zhao Y, Yue X, Wu H, et al. Cullin3-KLHL25 ubiquitin ligase targets ACLY for degradation to inhibit lipid synthesis and tumor progression. Genes & Development. 2016; 30: 1956–1970.
[52]
Jin X, Zhang KJ, Guo X, Myers R, Ye Z, Zhang ZP, et al. Fatty acid synthesis pathway genetic variants and clinical outcome of non-small cell lung cancer patients after surgery. Asian Pacific Journal of Cancer Prevention: APJCP. 2014; 15: 7097–7103.
[53]
Zhang C, Wang XY, Zhang P, He TC, Han JH, Zhang R, et al. Cancer-derived exosomal HSPC111 promotes colorectal cancer liver metastasis by reprogramming lipid metabolism in cancer-associated fibroblasts. Cell Death & Disease. 2022; 13: 57.
[54]
Li C, Zhang L, Qiu Z, Deng W, Wang W. Key Molecules of Fatty Acid Metabolism in Gastric Cancer. Biomolecules. 2022; 12: 706.
[55]
Hanai JI, Doro N, Seth P, Sukhatme VP. ATP citrate lyase knockdown impacts cancer stem cells in vitro. Cell Death & Disease. 2013; 4: e696.
[56]
Veprik A, Denwood G, Liu D, Bany Bakar R, Morfin V, McHugh K, et al. Acetyl-CoA-carboxylase 1 (ACC1) plays a critical role in glucagon secretion. Communications Biology. 2022; 5: 238.
[57]
Yeudall S, Upchurch CM, Seegren PV, Pavelec CM, Greulich J, Lemke MC, et al. Macrophage acetyl-CoA carboxylase regulates acute inflammation through control of glucose and lipid metabolism. Science Advances. 2022; 8: eabq1984.
[58]
Yu Y, Nie Q, Wang Z, Di Y, Chen X, Ren K. Targeting acetyl-CoA carboxylase 1 for cancer therapy. Frontiers in Pharmacology. 2023; 14: 1129010.
[59]
Rios Garcia M, Steinbauer B, Srivastava K, Singhal M, Mattijssen F, Maida A, et al. Acetyl-CoA Carboxylase 1-Dependent Protein Acetylation Controls Breast Cancer Metastasis and Recurrence. Cell Metabolism. 2017; 26: 842–855.e5.
[60]
Makarević J, Rutz J, Juengel E, Maxeiner S, Mani J, Vallo S, et al. HDAC Inhibition Counteracts Metastatic Re-Activation of Prostate Cancer Cells Induced by Chronic mTOR Suppression. Cells. 2018; 7: 129.
[61]
Pang B, Zhang J, Zhang X, Yuan J, Shi Y, Qiao L. Inhibition of lipogenesis and induction of apoptosis by valproic acid in prostate cancer cells via the C/EBPα/SREBP-1 pathway. Acta Biochimica et Biophysica Sinica. 2021; 53: 354–364.
[62]
Lally JSV, Ghoshal S, DePeralta DK, Moaven O, Wei L, Masia R, et al. Inhibition of Acetyl-CoA Carboxylase by Phosphorylation or the Inhibitor ND-654 Suppresses Lipogenesis and Hepatocellular Carcinoma. Cell Metabolism. 2019; 29: 174–182.e5.
[63]
Chiu CF, Hsu MI, Yeh HY, Park JM, Shen YS, Tung TH, et al. Eicosapentaenoic Acid Inhibits KRAS Mutant Pancreatic Cancer Cell Growth by Suppressing Hepassocin Expression and STAT3 Phosphorylation. Biomolecules. 2021; 11: 370.
[64]
Berger SL. The complex language of chromatin regulation during transcription. Nature. 2007; 447: 407–412.
[65]
Koprinarova M, Schnekenburger M, Diederich M. Role of Histone Acetylation in Cell Cycle Regulation. Current Topics in Medicinal Chemistry. 2016; 16: 732–744.
[66]
Sivanand S, Viney I, Wellen KE. Spatiotemporal Control of Acetyl-CoA Metabolism in Chromatin Regulation. Trends in Biochemical Sciences. 2018; 43: 61–74.
[67]
Carrer A, Wellen KE. Metabolism and epigenetics: a link cancer cells exploit. Current Opinion in Biotechnology. 2015; 34: 23–29.
[68]
Xu D, Shao F, Bian X, Meng Y, Liang T, Lu Z. The Evolving Landscape of Noncanonical Functions of Metabolic Enzymes in Cancer and Other Pathologies. Cell Metabolism. 2021; 33: 33–50.
[69]
Chen J, Guccini I, Di Mitri D, Brina D, Revandkar A, Sarti M, et al. Compartmentalized activities of the pyruvate dehydrogenase complex sustain lipogenesis in prostate cancer. Nature Genetics. 2018; 50: 219–228.
[70]
Kopp B, Khoury L, Audebert M. Validation of the γH2AX biomarker for genotoxicity assessment: a review. Archives of Toxicology. 2019; 93: 2103–2114.
[71]
Assante G, Chandrasekaran S, Ng S, Tourna A, Chung CH, Isse KA, et al. Acetyl-CoA metabolism drives epigenome change and contributes to carcinogenesis risk in fatty liver disease. Genome Medicine. 2022; 14: 67.
[72]
Chung S, Hwang JT, Park JH, Choi HK. Free fatty acid-induced histone acetyltransferase activity accelerates lipid accumulation in HepG2 cells. Nutrition Research and Practice. 2019; 13: 196–204.
[73]
Lu M, Zhu WW, Wang X, Tang JJ, Zhang KL, Yu GY, et al. ACOT12-Dependent Alteration of Acetyl-CoA Drives Hepatocellular Carcinoma Metastasis by Epigenetic Induction of Epithelial-Mesenchymal Transition. Cell Metabolism. 2019; 29: 886–900.e5.
[74]
Halbrook CJ, Nelson BS, Lysstiotis CA. Metabolism Drives Carcinogenesis and Maintenance of Pancreatic Tumors. Cancer Discovery. 2019; 9: 326–328.
[75]
Jiang Y, Hu T, Wang T, Shi X, Kitano A, Eagle K, et al. AMP-activated protein kinase links acetyl-CoA homeostasis to BRD4 recruitment in acute myeloid leukemia. Blood. 2019; 134: 2183–2194.
[76]
Triff K, McLean MW, Callaway E, Goldsby J, Ivanov I, Chapkin RS. Dietary fat and fiber interact to uniquely modify global histone post-translational epigenetic programming in a rat colon cancer progression model. International Journal of Cancer. 2018; 143: 1402–1415.
[77]
Koundouros N, Poulogiannis G. Reprogramming of fatty acid metabolism in cancer. British Journal of Cancer. 2020; 122: 4–22.
[78]
Beloribi-Djefaflia S, Vasseur S, Guillaumond F. Lipid metabolic reprogramming in cancer cells. Oncogenesis. 2016; 5: e189.
[79]
Chen L, Chen XW, Huang X, Song BL, Wang Y, Wang Y. Regulation of glucose and lipid metabolism in health and disease. Science China-Life Sciences. 2019; 62: 1420–1458
[80]
Steensels S, Ersoy BA. Fatty acid activation in thermogenic adipose tissue. Biochimica et Biophysica Acta. Molecular and Cell Biology of Lipids. 2019; 1864: 79–90.
[81]
Edmunds LR, Sharma L, Kang A, Lu J, Vockley J, Basu S, et al. c-Myc programs fatty acid metabolism and dictates acetyl-CoA abundance and fate. The Journal of Biological Chemistry. 2014; 289: 25382–25392.
[82]
Loo SY, Toh LP, Xie WH, Pathak E, Tan W, Ma S, et al. Fatty acid oxidation is a druggable gateway regulating cellular plasticity for driving metastasis in breast cancer. Science Advances. 2021; 7: eabh2443.
[83]
Cheng X, Geng F, Pan M, Wu X, Zhong Y, Wang C, et al. Targeting DGAT1 Ameliorates Glioblastoma by Increasing Fat Catabolism and Oxidative Stress. Cell Metabolism. 2020; 32: 229–242.e8.
[84]
Mallet RT, Sun J. Mitochondrial metabolism of pyruvate is required for its enhancement of cardiac function and energetics. Cardiovascular Research. 1999; 42: 149–161.
[85]
Patel MS, Roche TE. Molecular biology and biochemistry of pyruvate dehydrogenase complexes. FASEB Journal: Official Publication of the Federation of American Societies for Experimental Biology. 1990; 4: 3224–3233.
[86]
Pan C, Li B, Simon MC. Moonlighting functions of metabolic enzymes and metabolites in cancer. Molecular Cell. 2021; 81: 3760–3774.
[87]
Stacpoole PW. Therapeutic Targeting of the Pyruvate Dehydrogenase Complex/Pyruvate Dehydrogenase Kinase (PDC/PDK) Axis in Cancer. Journal of the National Cancer Institute. 2017; 109: djx071.
[88]
Pavlova NN, Zhu J, Thompson CB. The hallmarks of cancer metabolism: Still emerging. Cell Metabolism. 2022; 34: 355–377.
[89]
Jeoung NH. Pyruvate Dehydrogenase Kinases: Therapeutic Targets for Diabetes and Cancers. Diabetes & Metabolism Journal. 2015; 39: 188–197.
[90]
Fan J, Shan C, Kang HB, Elf S, Xie J, Tucker M, et al. Tyr phosphorylation of PDP1 toggles recruitment between ACAT1 and SIRT3 to regulate the pyruvate dehydrogenase complex. Molecular Cell. 2014; 53: 534–548.
[91]
Prochownik EV, Wang H. The Metabolic Fates of Pyruvate in Normal and Neoplastic Cells. Cells. 2021; 10: 762.
[92]
Mennerich D, Kubaichuk K, Kietzmann T. DUBs, Hypoxia, and Cancer. Trends in Cancer. 2019; 5: 632–653.
[93]
Ratcliffe PJ. Oxygen sensing and hypoxia signalling pathways in animals: the implications of physiology for cancer. The Journal of Physiology. 2013; 591: 2027–2042.
[94]
Semenza GL. HIF-1: upstream and downstream of cancer metabolism. Current Opinion in Genetics & Development. 2010; 20: 51–56.
[95]
Kim JW, Tchernyshyov I, Semenza GL, Dang CV. HIF-1-mediated expression of pyruvate dehydrogenase kinase: a metabolic switch required for cellular adaptation to hypoxia. Cell Metabolism. 2006; 3: 177–185.
[96]
Moffett JR, Puthillathu N, Vengilote R, Jaworski DM, Namboodiri AM. Acetate Revisited: A Key Biomolecule at the Nexus of Metabolism, Epigenetics, and Oncogenesis - Part 2: Acetate and ACSS2 in Health and Disease. Frontiers in Physiology. 2020; 11: 580171.
[97]
Chen R, Xu M, Nagati J, Garcia JA. Coordinate regulation of stress signaling and epigenetic events by Acss2 and HIF-2 in cancer cells. PLoS ONE. 2017; 12: e0190241.
[98]
Chen R, Xu M, Nagati JS, Hogg RT, Das A, Gerard RD, et al. The acetate/ACSS2 switch regulates HIF-2 stress signaling in the tumor cell microenvironment. PLoS ONE. 2015; 10: e0116515.
[99]
Gao X, Lin SH, Ren F, Li JT, Chen JJ, Yao CB, et al. Acetate functions as an epigenetic metabolite to promote lipid synthesis under hypoxia. Nature Communications. 2016; 7: 11960.
[100]
Longo R, Peri C, Cricrì D, Coppi L, Caruso D, Mitro N, et al. Ketogenic Diet: A New Light Shining on Old but Gold Biochemistry. Nutrients. 2019; 11: 2497.
[101]
Grabacka M, Pierzchalska M, Dean M, Reiss K. Regulation of Ketone Body Metabolism and the Role of PPARα. International Journal of Molecular Sciences. 2016; 17: 2093.
[102]
De Oliveira MP, Liesa M. The Role of Mitochondrial Fat Oxidation in Cancer Cell Proliferation and Survival. Cells. 2020; 9: 2600.
[103]
Cowman SJ, Koh MY. Revisiting the HIF switch in the tumor and its immune microenvironment. Trends in Cancer. 2022; 8: 28–42.
[104]
Trefely S, Lovell CD, Snyder NW, Wellen KE. Compartmentalised acyl-CoA metabolism and roles in chromatin regulation. Molecular Metabolism. 2020; 38: 100941.
[105]
Wu F, Zhang Y, Sun B, McMahon AP, Wang Y. Hedgehog Signaling: From Basic Biology to Cancer Therapy. Cell Chemical Biology. 2017; 24: 252–280.
[106]
Pak E, Segal RA. Hedgehog Signal Transduction: Key Players, Oncogenic Drivers, and Cancer Therapy. Developmental Cell. 2016; 38: 333–344.
[107]
Pepinsky RB, Zeng C, Wen D, Rayhorn P, Baker DP, Williams KP, et al. Identification of a palmitic acid-modified form of human Sonic hedgehog. The Journal of Biological Chemistry. 1998; 273: 14037–14045.
[108]
Long J, Tokhunts R, Old WM, Houel S, Rodgriguez-Blanco J, Singh S, et al. Identification of a family of fatty-acid-speciated sonic hedgehog proteins, whose members display differential biological properties. Cell Reports. 2015; 10: 1280–1287.
[109]
Sun L, Zhang H, Gao P. Metabolic reprogramming and epigenetic modifications on the path to cancer. Protein & Cell. 2022; 13: 877–919.
[110]
Wang Y, Guo YR, Liu K, Yin Z, Liu R, Xia Y, et al. KAT2A coupled with the α-KGDH complex acts as a histone H3 succinyltransferase. Nature. 2017; 552: 273–277.
[111]
Serrano-Villar S, Tincati C, Raju SC, Sáenz JS, Moreno E, Bargiela R, et al. Microbiome-derived cobalamin and succinyl-CoA as biomarkers for improved screening of anal cancer. Nature Medicine. 2023; 29: 1738–1749.
[112]
Boese AC, Kang J, Hwang JS, Kim J, Eun K, Malin CM, et al. Succinyl-CoA ligase ADP-forming subunit beta promotes stress granule assembly to regulate redox and drive cancer metastasis. Proceedings of the National Academy of Sciences of the United States of America. 2023; 120: e2217332120.
[113]
Zhou C, Wang Z, Cao Y, Zhao L. Pan-cancer analysis reveals the oncogenic role of 3-hydroxy-3-methylglutaryl-CoA synthase 1. Cancer Reports (Hoboken, N.J.). 2022; 5: e1562.
[114]
Bathaie SZ, Ashrafi M, Azizian M, Tamanoi F. Mevalonate Pathway and Human Cancers. Current Molecular Pharmacology. 2017; 10: 77–85.
[115]
van Besien H, Sassano A, Altman JK, Platanias LC. Antileukemic properties of 3-hydroxy-3-methylglutaryl-coenzyme A reductase inhibitors. Leukemia & Lymphoma. 2013; 54: 2601–2605.
[116]
Gouirand V, Gicquel T, Lien EC, Jaune-Pons E, Da Costa Q, Finetti P, et al. Ketogenic HMG-CoA lyase and its product β-hydroxybutyrate promote pancreatic cancer progression. The EMBO Journal. 2022; 41: e110466.
[117]
Czumaj A, Szrok-Jurga S, Hebanowska A, Turyn J, Swierczynski J, Sledzinski T, et al. The Pathophysiological Role of CoA. International Journal of Molecular Sciences. 2020; 21: 9057.
[118]
Gaidai O, Cao Y, Loginov S. Global Cardiovascular Diseases Death Rate Prediction. Current Problems in Cardiology. 2023; 48: 101622.
[119]
Sirtori CR. The pharmacology of statins. Pharmacological Research. 2014; 88: 3–11.
[120]
Ussher JR, Elmariah S, Gerszten RE, Dyck JRB. The Emerging Role of Metabolomics in the Diagnosis and Prognosis of Cardiovascular Disease. Journal of the American College of Cardiology. 2016; 68: 2850–2870.
[121]
Maas SL, Donners MMPC, van der Vorst EPC. ADAM10 and ADAM17, Major Regulators of Chronic Kidney Disease Induced Atherosclerosis? International Journal of Molecular Sciences. 2023; 24: 7309.
[122]
Emdin CA, Anderson SG, Callender T, Conrad N, Salimi-Khorshidi G, Mohseni H, et al. Usual blood pressure, peripheral arterial disease, and vascular risk: cohort study of 4.2 million adults. BMJ (Clinical Research Ed.). 2015; 351: h4865.
[123]
Mohammadifard N, Haghighatdoost F, Rahimlou M, Rodrigues APS, Gaskarei MK, Okhovat P, et al. The Effect of Ketogenic Diet on Shared Risk Factors of Cardiovascular Disease and Cancer. Nutrients. 2022; 14: 3499.
[124]
Shukla SK, Liu W, Sikder K, Addya S, Sarkar A, Wei Y, et al. HMGCS2 is a key ketogenic enzyme potentially involved in type 1 diabetes with high cardiovascular risk. Scientific Reports. 2017; 7: 4590.
[125]
Li J, Viswanadha S, Loor JJ. Hepatic Metabolic, Inflammatory, and Stress-Related Gene Expression in Growing Mice Consuming a Low Dose of Trans-10, cis-12-Conjugated Linoleic Acid. Journal of Lipids. 2012; 2012: 571281.
[126]
Vilà-Brau A, De Sousa-Coelho AL, Mayordomo C, Haro D, Marrero PF. Human HMGCS2 regulates mitochondrial fatty acid oxidation and FGF21 expression in HepG2 cell line. The Journal of Biological Chemistry. 2011; 286: 20423–20430.
[127]
Takahashi Y, Terada S, Banjo M, Seike K, Nakano S, Hatta H. Effects of β-hydroxybutyrate treatment on glycogen repletion and its related signaling cascades in epitrochlearis muscle during 120 min of postexercise recovery. Applied Physiology, Nutrition, and Metabolism = Physiologie Appliquee, Nutrition et Metabolisme. 2019; 44: 1311–1319.
[128]
Wei S, Binbin L, Yuan W, Zhong Z, Donghai L, Caihua H. β-Hydroxybutyrate in Cardiovascular Diseases: A Minor Metabolite of Great Expectations. Frontiers in Molecular Biosciences. 2022; 9: 823602.
[129]
Song JP, Chen L, Chen X, Ren J, Zhang NN, Tirasawasdichai T, et al. Elevated plasma β-hydroxybutyrate predicts adverse outcomes and disease progression in patients with arrhythmogenic cardiomyopathy. Science Translational Medicine. 2020; 12: eaay8329.
[130]
Al-Zaid NS, Dashti HM, Mathew TC, Juggi JS. Low carbohydrate ketogenic diet enhances cardiac tolerance to global ischaemia. Acta Cardiologica. 2007; 62: 381–389.
[131]
Matsushita K, Ballew SH, Wang AYM, Kalyesubula R, Schaeffner E, Agarwal R. Epidemiology and risk of cardiovascular disease in populations with chronic kidney disease. Nature Reviews. Nephrology. 2022; 18: 696–707.
[132]
Wu X, Huang T. Recent development in acetyl-CoA carboxylase inhibitors and their potential as novel drugs. Future Medicinal Chemistry. 2020; 12: 533–561.
[133]
ALJohani AM, Syed DN, Ntambi JM. Insights into Stearoyl-CoA Desaturase-1 Regulation of Systemic Metabolism. Trends in Endocrinology and Metabolism: TEM. 2017; 28: 831–842.
[134]
Goedeke L, Bates J, Vatner DF, Perry RJ, Wang T, Ramirez R, et al. Acetyl-CoA Carboxylase Inhibition Reverses NAFLD and Hepatic Insulin Resistance but Promotes Hypertriglyceridemia in Rodents. Hepatology (Baltimore, Md.). 2018; 68: 2197–2211.
[135]
Mika A, Sikorska-Wiśniewska M, Małgorzewicz S, Stepnowski P, Dębska-Ślizień A, Śledziński T, et al. Potential contribution of monounsaturated fatty acids to cardiovascular risk in chronic kidney disease. Polish Archives of Internal Medicine. 2018; 128: 755–763.
[136]
Chmielewski M, Carrero JJ, Nordfors L, Lindholm B, Stenvinkel P. Lipid disorders in chronic kidney disease: reverse epidemiology and therapeutic approach. Journal of Nephrology. 2008; 21: 635–644.
[137]
Zhang Y, Gu Z, Wan J, Lou X, Liu S, Wang Y, et al. Stearoyl-CoA Desaturase-1 dependent lipid droplets accumulation in cancer-associated fibroblasts facilitates the progression of lung cancer. International Journal of Biological Sciences. 2022; 18: 6114–6128.
[138]
Kamal S, Saleem A, Rehman S, Bibi I, Iqbal HMN. Protein engineering: Regulatory perspectives of stearoyl CoA desaturase. International Journal of Biological Macromolecules. 2018; 114: 692–699.
[139]
Neubauer S. The failing heart–an engine out of fuel. The New England Journal of Medicine. 2007; 356: 1140–1151.
[140]
Lopaschuk GD, Ussher JR, Folmes CDL, Jaswal JS, Stanley WC. Myocardial fatty acid metabolism in health and disease. Physiological Reviews. 2010; 90: 207–258.
[141]
Chen MS, Lee RT, Garbern JC. Senescence mechanisms and targets in the heart. Cardiovascular Research. 2022; 118: 1173–1187.
[142]
Bonnet D, Martin D, Pascale De Lonlay, Villain E, Jouvet P, Rabier D, et al. Arrhythmias and conduction defects as presenting symptoms of fatty acid oxidation disorders in children. Circulation. 1999; 100: 2248–2253.
[143]
Miklas JW, Clark E, Levy S, Detraux D, Leonard A, Beussman K, et al. TFPa/HADHA is required for fatty acid beta-oxidation and cardiolipin re-modeling in human cardiomyocytes. Nature Communications. 2019; 10: 4671.
[144]
Bernardo BC, Weeks KL, Pongsukwechkul T, Gao X, Kiriazis H, Cemerlang N, et al. Gene delivery of medium chain acyl-coenzyme A dehydrogenase induces physiological cardiac hypertrophy and protects against pathological remodelling. Clinical Science (London, England: 1979). 2018; 132: 381–397.
[145]
Fukushima A, Lopaschuk GD. Acetylation control of cardiac fatty acid β-oxidation and energy metabolism in obesity, diabetes, and heart failure. Biochimica et Biophysica Acta. 2016; 1862: 2211–2220.
[146]
Rasmussen BB, Wolfe RR. Regulation of fatty acid oxidation in skeletal muscle. Annual Review of Nutrition. 1999; 19: 463–484.
[147]
Abo Alrob O, Lopaschuk GD. Role of CoA and acetyl-CoA in regulating cardiac fatty acid and glucose oxidation. Biochemical Society Transactions. 2014; 42: 1043–1051.
[148]
Maher AC, McFarlan J, Lally J, Snook LA, Bonen A. TBC1D1 reduces palmitate oxidation by inhibiting β-HAD activity in skeletal muscle. American Journal of Physiology. Regulatory, Integrative and Comparative Physiology. 2014; 307: R1115–R1123.
[149]
Kirkby B, Roman N, Kobe B, Kellie S, Forwood JK. Functional and structural properties of mammalian acyl-coenzyme A thioesterases. Progress in Lipid Research. 2010; 49: 366–377.
[150]
He W, Berthiaume JM, Previs S, Kasumov T, Zhang GF. Ischemia promotes acyl-CoAs dephosphorylation and propionyl-CoA accumulation. Metabolomics: Official Journal of the Metabolomic Society. 2023; 19: 12.
[151]
Tillander V, Alexson SEH, Cohen DE. Deactivating Fatty Acids: Acyl-CoA Thioesterase-Mediated Control of Lipid Metabolism. Trends in Endocrinology and Metabolism: TEM. 2017; 28: 473–484.
[152]
Zhang Y, Wen P, Luo J, Ding H, Cao H, He W, et al. Sirtuin 3 regulates mitochondrial protein acetylation and metabolism in tubular epithelial cells during renal fibrosis. Cell Death & Disease. 2021; 12: 847.
[153]
Jin L, Geng L, Ying L, Shu L, Ye K, Yang R, et al. FGF21-Sirtuin 3 Axis Confers the Protective Effects of Exercise Against Diabetic Cardiomyopathy by Governing Mitochondrial Integrity. Circulation. 2022; 146: 1537–1557.
[154]
Pinterić M, Podgorski II, Popović Hadžija M, Tartaro Bujak I, Tadijan A, Balog T, et al. Chronic High Fat Diet Intake Impairs Hepatic Metabolic Parameters in Ovariectomized Sirt3 KO Mice. International Journal of Molecular Sciences. 2021; 22: 4277.
[155]
Pownall HJ, Rosales C, Gillard BK, Gotto AM, Jr. High-density lipoproteins, reverse cholesterol transport and atherogenesis. Nature Reviews. Cardiology. 2021; 18: 712–723.
[156]
Sandesara PB, Virani SS, Fazio S, Shapiro MD. The Forgotten Lipids: Triglycerides, Remnant Cholesterol, and Atherosclerotic Cardiovascular Disease Risk. Endocrine Reviews. 2019; 40: 537–557.
[157]
Montégut L, Joseph A, Chen H, Abdellatif M, Ruckenstuhl C, Motiño O, et al. High plasma concentrations of acyl-coenzyme A binding protein (ACBP) predispose to cardiovascular disease: Evidence for a phylogenetically conserved proaging function of ACBP. Aging Cell. 2023; 22: e13751.
[158]
Casula M, Colpani O, Xie S, Catapano AL, Baragetti A. HDL in Atherosclerotic Cardiovascular Disease: In Search of a Role. Cells. 2021; 10: 1869.
[159]
Libby P, Buring JE, Badimon L, Hansson GK, Deanfield J, Bittencourt MS, et al. Atherosclerosis. Nature Reviews. Disease Primers. 2019; 5: 56.
[160]
Virani SS, Alonso A, Benjamin EJ, Bittencourt MS, Callaway CW, Carson AP, et al. Heart Disease and Stroke Statistics-2020 Update: A Report From the American Heart Association. Circulation. 2020; 141: e139–e596.
[161]
Zahedipour F, Butler AE, Eid AH, Sahebkar A. Pleiotropic properties of statins via angiogenesis modulation in cardiovascular disease. Drug Discovery Today. 2022; 27: 103325.
[162]
Nagakubo T, Kumano T, Hashimoto Y, Kobayashi M. Hemoglobin catalyzes CoA degradation and thiol addition to flavonoids. Scientific Reports. 2018; 8: 1282.
[163]
Aoi W, Iwasa M, Marunaka Y. Metabolic functions of flavonoids: From human epidemiology to molecular mechanism. Neuropeptides. 2021; 88: 102163.
[164]
Xu Y, Jiang H, Li L, Chen F, Liu Y, Zhou M, et al. Branched-Chain Amino Acid Catabolism Promotes Thrombosis Risk by Enhancing Tropomodulin-3 Propionylation in Platelets. Circulation. 2020; 142: 49–64.
[165]
Gan L, Cookson MR, Petrucelli L, La Spada AR. Converging pathways in neurodegeneration, from genetics to mechanisms. Nature Neuroscience. 2018; 21: 1300–1309.
[166]
Lambrechts RA, Schepers H, Yu Y, van der Zwaag M, Autio KJ, Vieira-Lara MA, et al. CoA-dependent activation of mitochondrial acyl carrier protein links four neurodegenerative diseases. EMBO Molecular Medicine. 2019; 11: e10488.
[167]
Di Meo I, Carecchio M, Tiranti V. Inborn errors of coenzyme A metabolism and neurodegeneration. Journal of Inherited Metabolic Disease. 2019; 42: 49–56.
[168]
Thakur N, Klopstock T, Jackowski S, Kuscer E, Tricta F, Videnovic A, et al. Rational Design of Novel Therapies for Pantothenate Kinase-Associated Neurodegeneration. Movement Disorders: Official Journal of the Movement Disorder Society. 2021; 36: 2005–2016.
[169]
Meyer E, Kurian MA, Hayflick SJ. Neurodegeneration with Brain Iron Accumulation: Genetic Diversity and Pathophysiological Mechanisms. Annual Review of Genomics and Human Genetics. 2015; 16: 257–279.
[170]
Luo J, Yang H, Song BL. Mechanisms and regulation of cholesterol homeostasis. Nature Reviews. Molecular Cell Biology. 2020; 21: 225–245.
[171]
Lu XY, Shi XJ, Hu A, Wang JQ, Ding Y, Jiang W, et al. Feeding induces cholesterol biosynthesis via the mTORC1-USP20-HMGCR axis. Nature. 2020; 588: 479–484
[172]
Li D, Zhang J, Liu Q. Brain cell type-specific cholesterol metabolism and implications for learning and memory. Trends in Neurosciences. 2022; 45: 401–414.
[173]
Zhang J, Liu Q. Cholesterol metabolism and homeostasis in the brain. Protein & Cell. 2015; 6: 254–264.
[174]
Dai L, Zou L, Meng L, Qiang G, Yan M, Zhang Z. Cholesterol Metabolism in Neurodegenerative Diseases: Molecular Mechanisms and Therapeutic Targets. Molecular Neurobiology. 2021; 58: 2183–2201.
[175]
Pingale TD, Gupta GL. Novel therapeutic approaches for Parkinson’s disease by targeting brain cholesterol homeostasis. The Journal of Pharmacy and Pharmacology. 2021; 73: 862–873.
[176]
Nieweg K, Schaller H, Pfrieger FW. Marked differences in cholesterol synthesis between neurons and glial cells from postnatal rats. Journal of Neurochemistry. 2009; 109: 125–134.
[177]
Birolini G, Verlengia G, Talpo F, Maniezzi C, Zentilin L, Giacca M, et al. SREBP2 gene therapy targeting striatal astrocytes ameliorates Huntington’s disease phenotypes. Brain: a Journal of Neurology. 2021; 144: 3175–3190.
[178]
Kao YC, Ho PC, Tu YK, Jou IM, Tsai KJ. Lipids and Alzheimer’s Disease. International Journal of Molecular Sciences. 2020; 21: 1505.
[179]
Huang X, Sterling NW, Du G, Sun D, Stetter C, Kong L, et al. Brain cholesterol metabolism and Parkinson’s disease. Movement Disorders: Official Journal of the Movement Disorder Society. 2019; 34: 386–395.
[180]
Bedi KC, Jr, Snyder NW, Brandimarto J, Aziz M, Mesaros C, Worth AJ, et al. Evidence for Intramyocardial Disruption of Lipid Metabolism and Increased Myocardial Ketone Utilization in Advanced Human Heart Failure. Circulation. 2016; 133: 706–716.
[181]
He Q, Chen Y, Wang Z, He H, Yu P. Cellular Uptake, Metabolism and Sensing of Long-Chain Fatty Acids. Frontiers in Bioscience (Landmark Edition). 2023; 28: 10.
[182]
O’Donnell JM, Fields AD, Sorokina N, Lewandowski ED. The absence of endogenous lipid oxidation in early stage heart failure exposes limits in lipid storage and turnover. Journal of Molecular and Cellular Cardiology. 2008; 44: 315–322.
[183]
Lahey R, Wang X, Carley AN, Lewandowski ED. Dietary fat supply to failing hearts determines dynamic lipid signaling for nuclear receptor activation and oxidation of stored triglyceride. Circulation. 2014; 130: 1790–1799.
[184]
Nakamura MT, Yudell BE, Loor JJ. Regulation of energy metabolism by long-chain fatty acids. Progress in Lipid Research. 2014; 53: 124–144.
[185]
de Carvalho CCCR, Caramujo MJ. The Various Roles of Fatty Acids. Molecules (Basel, Switzerland). 2018; 23: 2583.
[186]
Kampe K, Sieber J, Orellana JM, Mundel P, Jehle AW. Susceptibility of podocytes to palmitic acid is regulated by fatty acid oxidation and inversely depends on acetyl-CoA carboxylases 1 and 2. American Journal of Physiology. Renal Physiology. 2014; 306: F401–F409.
[187]
Frier BC, Jacobs RL, Wright DC. Interactions between the consumption of a high-fat diet and fasting in the regulation of fatty acid oxidation enzyme gene expression: an evaluation of potential mechanisms. American Journal of Physiology. Regulatory, Integrative and Comparative Physiology. 2011; 300: R212–R221.
[188]
Jeon SM. Regulation and function of AMPK in physiology and diseases. Experimental & Molecular Medicine. 2016; 48: e245.
[189]
Li Y, Xu S, Mihaylova MM, Zheng B, Hou X, Jiang B, et al. AMPK phosphorylates and inhibits SREBP activity to attenuate hepatic steatosis and atherosclerosis in diet-induced insulin-resistant mice. Cell Metabolism. 2011; 13: 376–388.
[190]
Chen L, Duan Y, Wei H, Ning H, Bi C, Zhao Y, et al. Acetyl-CoA carboxylase (ACC) as a therapeutic target for metabolic syndrome and recent developments in ACC1/2 inhibitors. Expert Opinion on Investigational Drugs. 2019; 28: 917–930.
[191]
Horton JD, Shah NA, Warrington JA, Anderson NN, Park SW, Brown MS, et al. Combined analysis of oligonucleotide microarray data from transgenic and knockout mice identifies direct SREBP target genes. Proceedings of the National Academy of Sciences of the United States of America. 2003; 100: 12027–12032.
[192]
Im SS, Yousef L, Blaschitz C, Liu JZ, Edwards RA, Young SG, et al. Linking lipid metabolism to the innate immune response in macrophages through sterol regulatory element binding protein-1a. Cell Metabolism. 2011; 13: 540–549.
[193]
Lee MKS, Cooney OD, Lin X, Nadarajah S, Dragoljevic D, Huynh K, et al. Defective AMPK regulation of cholesterol metabolism accelerates atherosclerosis by promoting HSPC mobilization and myelopoiesis. Molecular Metabolism. 2022; 61: 101514.
[194]
Goldstein JL, Brown MS. A century of cholesterol and coronaries: from plaques to genes to statins. Cell. 2015; 161: 161–172.
[195]
Bilen O, Ballantyne CM. Bempedoic Acid (ETC-1002): an Investigational Inhibitor of ATP Citrate Lyase. Current Atherosclerosis Reports. 2016; 18: 61.
[196]
Pinkosky SL, Groot PHE, Lalwani ND, Steinberg GR. Targeting ATP-Citrate Lyase in Hyperlipidemia and Metabolic Disorders. Trends in Molecular Medicine. 2017; 23: 1047–1063.
[197]
Batchuluun B, Pinkosky SL, Steinberg GR. Lipogenesis inhibitors: therapeutic opportunities and challenges. Nature Reviews. Drug Discovery. 2022; 21: 283–305.
[198]
Ameer F, Scandiuzzi L, Hasnain S, Kalbacher H, Zaidi N. De novo lipogenesis in health and disease. Metabolism: Clinical and Experimental. 2014; 63: 895–902.
[199]
Huber K, Hofer DC, Trefely S, Pelzmann HJ, Madreiter-Sokolowski C, Duta-Mare M, et al. N-acetylaspartate pathway is nutrient responsive and coordinates lipid and energy metabolism in brown adipocytes. Biochimica et Biophysica Acta. Molecular Cell Research. 2019; 1866: 337–348.
[200]
Bogner-Strauss JG. N-Acetylaspartate Metabolism Outside the Brain: Lipogenesis, Histone Acetylation, and Cancer. Frontiers in Endocrinology. 2017; 8: 240.
[201]
Hofer DC, Zirkovits G, Pelzmann HJ, Huber K, Pessentheiner AR, Xia W, et al. N-acetylaspartate availability is essential for juvenile survival on fat-free diet and determines metabolic health. FASEB Journal: Official Publication of the Federation of American Societies for Experimental Biology. 2019; 33: 13808–13824.
[202]
Pessentheiner AR, Pelzmann HJ, Walenta E, Schweiger M, Groschner LN, Graier WF, et al. NAT8L (N-acetyltransferase 8-like) accelerates lipid turnover and increases energy expenditure in brown adipocytes. The Journal of Biological Chemistry. 2013; 288: 36040–36051.
[203]
Prokesch A, Pelzmann HJ, Pessentheiner AR, Huber K, Madreiter-Sokolowski CT, Drougard A, et al. N-acetylaspartate catabolism determines cytosolic acetyl-CoA levels and histone acetylation in brown adipocytes. Scientific Reports. 2016; 6: 23723.
[204]
Duarte Lau F, Giugliano RP. Adenosine Triphosphate Citrate Lyase and Fatty Acid Synthesis Inhibition: A Narrative Review. JAMA Cardiology. 2023; 8: 879–887.
[205]
Zhang W, Li JY, Wei XC, Wang Q, Yang JY, Hou H, et al. Effects of dibutyl phthalate on lipid metabolism in liver and hepatocytes based on PPARα/SREBP-1c/FAS/GPAT/AMPK signal pathway. Food and Chemical Toxicology: an International Journal Published for the British Industrial Biological Research Association. 2021; 149: 112029.
[206]
Huang YQ, Wang Y, Hu K, Lin S, Lin XH. Hippocampal Glycerol-3-Phosphate Acyltransferases 4 and BDNF in the Progress of Obesity-Induced Depression. Frontiers in Endocrinology. 2021; 12: 667773.
[207]
Raud B, Roy DG, Divakaruni AS, Tarasenko TN, Franke R, Ma EH, et al. Etomoxir Actions on Regulatory and Memory T Cells Are Independent of Cpt1a-Mediated Fatty Acid Oxidation. Cell Metabolism. 2018; 28: 504–515.e7.

Publisher’s Note: IMR Press stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share
Back to top