Next Article in Journal
Small Molecule Compounds of Natural Origin Target Cellular Receptors to Inhibit Cancer Development and Progression
Previous Article in Journal
MAP3Kε1/2 Interact with MOB1A/1B and Play Important Roles in Control of Pollen Germination through Crosstalk with JA Signaling in Arabidopsis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Aldehyde Dehydrogenase 2 as a Therapeutic Target in Oxidative Stress-Related Diseases: Post-Translational Modifications Deserve More Attention

1
State Key Laboratory of Animal Nutrition, Institute of Animal Sciences, Chinese Academy of Agricultural Sciences, Beijing 100193, China
2
State Key Laboratory of Animal Nutrition, College of Animal Science and Technology, China Agricultural University, Beijing 100193, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(5), 2682; https://doi.org/10.3390/ijms23052682
Submission received: 21 January 2022 / Revised: 19 February 2022 / Accepted: 21 February 2022 / Published: 28 February 2022
(This article belongs to the Section Molecular Biology)

Abstract

:
Aldehyde dehydrogenase 2 (ALDH2) has both dehydrogenase and esterase activity; its dehydrogenase activity is closely related to the metabolism of aldehydes produced under oxidative stress (OS). In this review, we recapitulate the enzyme activity of ALDH2 in combination with its protein structure, summarize and show the main mechanisms of ALDH2 participating in metabolism of aldehydes in vivo as comprehensively as possible; we also integrate the key regulatory mechanisms of ALDH2 participating in a variety of physiological and pathological processes related to OS, including tissue and organ fibrosis, apoptosis, aging, and nerve injury-related diseases. On this basis, the regulatory effects and application prospects of activators, inhibitors, and protein post-translational modifications (PTMs, such as phosphorylation, acetylation, S-nitrosylation, nitration, ubiquitination, and glycosylation) on ALDH2 are discussed and prospected. Herein, we aimed to lay a foundation for further research into the mechanism of ALDH2 in oxidative stress-related disease and provide a basis for better use of the ALDH2 function in research and the clinic.

1. Introduction

Oxidative stress (OS) plays a key role in the pathogeneses of various etiologies. Reactive oxygen species (ROS) are mainly produced from the mitochondrial electron transport chain and enzymatic systems, such as xanthine oxidase and reduced nicotinamide adenine dinucleotide (NADH)/reduced nicotinamide adenine dinucleotide phosphate (NADPH) oxidase [1,2]. Under harmful stimulation, the production and accumulation of ROS exceed the ability of cells to remove them; this leads to oxidant–antioxidant imbalance and induces OS, resulting in oxidative attack on biological macromolecules that causes cell damage [3,4]. Membrane lipids are one of the most vulnerable targets of ROS damage [5], and increased ROS triggers lipid peroxidation. The main primary products of lipid peroxidation process are lipid hydroperoxides (LOOH), and the secondary products contain more than 200 types of highly active lipid-derived aldehydes (LDAs) [6], which can further aggravate cell damage [5]. Some aldehydes are strongly electrophilic, highly reactive, and toxic, including 4-hydroxy-2-nonenal (4-HNE), malondialdehyde (MDA), acrolein, 4-oxo-2-nonenal (4-ONE), and crotonaldehyde [5,7]. These LDAs further mediate the biological effects of free radicals, leading to the sustained generation of mitochondrial ROS and the consequent aggravation of tissue damage, thus amplifying the OS [8,9]. Reactive aldehydes easily form adducts with proteins, deoxyribonucleic acid (DNA), and phospholipids, inducing mutations in critical genes, the inactivation of key enzymes and consequent cell damage, and OS-related diseases such as cardiopathy, vasculopathy, cancer, diabetes, osteoporosis, and neurodegenerative diseases (NDDs) [10,11,12,13,14].
The aldehyde dehydrogenase (ALDH) family plays important roles in the metabolism of aldehydes in vivo. ALDH2 is an ALDH family member that catalyzes the oxidation of various aldehydes; it is widely distributed throughout the body as an oxidoreductase, but is mainly found in the liver, heart, lungs, kidneys, and other mitochondria-rich tissues. ALDH2 can remove aldehydes produced during cellular-substance metabolism and OS to reduce cytotoxicity and oxidative damage. ALDH2 can be activated by several pathways, its gene polymorphism and activity are related to a variety of pathophysiological processes, and it is widely involved in and affects the occurrence and development of various diseases and important metabolic processes. Consequently, it is attracting increasingly more attention as a key regulatory enzyme in OS. The structure and function of ALDH2, as well as the regulation of its dehydrogenase activity, are the subjects of this review.

2. ALDH2 in Ethanol and Aldehyde Metabolism

2.1. Human ALDH Protein Family

ALDH proteins, a class of nicotinamide adenine dinucleotide (phosphate) (NAD[P]+)-dependent enzymes [15,16], are mainly localized in subcellular compartments such as the cytoplasm, mitochondria, the endoplasmic reticulum (ER), and nuclei. They have common structural and functional characteristics (Table 1); they can catalyze the oxidation of various aliphatic and aromatic aldehydes, and can participate in the oxidation of many endogenous and exogenous aldehydes into corresponding carboxylic acids [15,16,17] to protect cells from oxidative damage caused by the accumulation of aldehydes during metabolism [18]. So far, the human genome has been found to contain 19 putatively functional ALDH genes (Table 1), which are located on different chromosomes and have different substrates and reactivities.

2.2. Effect of ALDH2 on Detoxification of Reactive Aldehyde

ALDH2 is the ALDH with the highest affinity (Km < 5 μM) for acetaldehyde [16,65], and has catalytic activity for various toxic aldehydes. The main toxic effect of aldehydes results from covalent modifications of several biological macromolecules, which produce adducts that interfere with the physiological functions of these biological macromolecules [66] and lead to protein inactivation and DNA damage. Reactive aldehydes can bind to nucleophilic amino acids (i.e., cysteine, histidine, and lysine), bases in nucleic acids, and lipids with amino groups through Michael addition or Schiff base formation [67,68,69], which in turn damages the macromolecules and impairs cellular functions. The C3 of C2=C3 double bond of 4-HNE can target thiol or amino groups of proteins by Michael additions, and the C1 carbonyl groups of 4-HNE react with primary amines of proteins to yield Schiff bases [5,67]; The Michael addition of a C=C double bond of 4-HNE to the NH2-group of deoxyguanosine also yields exocyclic adduct [5,67]. Similarly, MDA reacts with guanine, adenine, and cytosine to form adducts of deoxyguanosine and deoxyadenosine; then, exocyclic DNA adducts that are produced lead to DNA damage and mutation [68].
In addition, the mutual triggering of aldehydes and ROS (mainly refers to hydroxyl radical [HO] and hydroperoxyl [HO2]) aggravates damage to cellular components, viz., lipids, proteins, and nucleic acids [5]. On the one hand, ROS-triggered lipid oxidation is the most common source of toxic aldehydes, such as 4-HNE [67]; OS induced by ROS accumulation causes lipid peroxidation and the production of active aldehydes, which can interact with sulfhydryl and amino groups to eliminate endogenous antioxidants, such as glutathione [70,71]. On the other hand, the NADH/NADPH oxidase system is a major source of ROS [1], and the promoting effect of 4-HNE on ROS could be related to its regulation of the NADH/NADPH enzyme system. Under physiological conditions, low concentrations of aldehydes can induce antioxidant defense [2,72], but when the aldehydes accumulate past a certain threshold, they will continue to trigger the production of ROS. The oxidation process of ethanol and aldehydes increases the NADH/NAD+ ratio in both the cytosol and mitochondria, which improves xanthine oxidase activity and catalyzes the generation of superoxide free radicals [13,72]. 4-NHE was also reported to promote NADPH oxidase activity through increased membrane translocation of p47phox (subunit of NADPH oxidase) with subsequent enhancement of the production of ROS [73]. 4-HNE is capable of inducing mitochondrial ROS production through a mechanism that requires activities of NADH-ubiquinone oxidoreductase (Complex I) of the electron transport chain [71]. ALDH2 can decompose the highly toxic aldehydes produced during lipid peroxidation into non-toxic carboxylic acids [16], which is beneficial for reducing aldehyde-induced tissue damage [66,74,75] and protecting cells from OS [18]. Therefore, ALDH2 is a major protective enzyme.
ALDH2 plays an important role in the metabolism of toxic aldehydes directly and indirectly produced by ethanol metabolism (Figure 1). Ethanol is oxidized to acetaldehyde by three main pathways: (1) the vast majority of ethanol is oxidized by alcohol dehydrogenases (ADHs) in cytoplasm under the action of NAD+; (2) some of the ethanol is oxidized via the microsomal ethanol-oxidizing system (MEOS), which relies on cytochrome P4502E1 (CYP2E1); and (3) a small amount of ethanol is converted into acetaldehyde under the action of catalase (CAT) [76]. ALDH2 further metabolizes acetaldehyde into acetate, which can be transported out of the cell through a carrier or converted into acetyl-coenzyme A (acetyl-CoA) through the enzymatic action of cytosolic acetyl-CoA synthetase 2 (ACSS2) [66]. Then, acetyl-CoA enters different physiological pathways, including energy metabolism pathways such as the tricarboxylic acid (TCA) cycle or the oxidative phosphorylation (OXPHOS) pathway [77]. Eventually, it is metabolized into water and carbon dioxide. Acetyl-CoA also participates in the production of ketone bodies (KB), fatty acids (FA), and cholesterol (CHOL) [66,78]. In addition, the MDA produced in lipid peroxidation is converted into malonic semialdehyde under the action of ALDH2 and further generates malonic acid (MOA), or alternatively, a carboxyl group is removed and converted into acetaldehyde; ultimately, acetyl-CoA is generated and enters the TCA cycle [79]. Some of the 4-NHE produced during lipid peroxidation generates 4-hydroxy-2-nonenoic acid (NHA) under the action of ALDH2, while the remainder is catalyzed by other reductase enzymes such as glutathione-S-transferase (GST), aldo-keto reductases (AKRs)/ADHs into s glutathionyl-HNE (GS-HNE), and 1,4-dihydroxy-2-nonene (DHN) [80,81]. Loss or inhibition of ALDH2 biological activity interferes with its detoxification effect on aldehydes, which in turn triggers the production of ROS and amplifies oxidative damage to cells and leads to a variety of human diseases [75,82].

3. Structure and Gene Polymorphism of ALDH2

The ALDH2 gene was first identified in 1987 [83], and its X-ray crystal structure was completely analyzed in 1999 [84,85]. The gene encoding human ALDH2 is located on the long arm of chromosome 12 (12q24.2) and contains 13 exons. It encodes the polypeptide chain of the enzyme precursor protein, which consists of 517 amino acid residues; the first 17 amino acids are mitochondrial guide peptides, which are cut after they help the enzyme enter the mitochondrial matrix to obtain mature enzyme polypeptide [17,65,85,86]. The analysis of the protein structure of ALDH2 shows that the enzyme is a tetramer composed of four identical subunits, each of which consists of three main domains with different functions: catalytic, NAD+ coenzyme-binding, and oligomerization [12,87,88]. Glu487 forms hydrogen bonds with Arg264 residues of the same subunit in ALDH2 holoenzyme, and forms hydrogen bonds with Arg475 of the adjacent subunit through hydrogen bonds to form dimers, and the two dimers assemble into a tetramer [89,90]. The catalytic mechanism of ALDH2 is that NAD+ first binds to ALDH2 by multiple binding sites to activate the key active-site cysteine residue (Cys302). The sulfhydryl group of Cys302 has high nucleophilicity and conducts nucleophilic attacks on the carbonyl carbon of the substrate aldehyde to form a thiohemiacetal intermediate. Substrate hydride is transferred from the thiohemiacetal intermediate to the nicotinamide ring of NAD+ to form NADH, which is accompanied by the production of thioester intermediates. Then, the structure of the enzyme-substrate complex changes [89,91]; the adjacent glutamate residue Glu268 activates H2O to remove protons from Cys302, and the carbonyl carbon of the thioester intermediate undergoes nucleophilic attack, resulting in the destruction of the carbon–sulfur bond to regenerate the free enzyme. The final product, carboxylic acid, is obtained, followed by NADH release, completing the reaction [92,93,94]. Guanine at the 1510th nucleotide position (from the initiation codon) in exon 12 of the human ALDH2 gene is replaced by adenine (G → A), which changes codon 487 from GAA to AAA, leading to a glutamic to lysine amino acid change (Glu → Lys) [65,95]. The Glu487Lys SNP of ALDH2 (also known as Glu504Lys, rs671, c.1510G > A, p.E504K) [65,96] changes the wild-type (WT) allele ALDH2*1 into the mutant allele ALDH2*2, resulting in a decrease or loss of catalytic activity of the ALDH2 tetrameric enzyme (Figure 2B) [97]. As a result, three genotypes are produced: WT homozygous ALDH2*1/1, mutant heterozygous ALDH2*1/2, and mutant homozygous ALDH2*2/2 [98,99,100]. An ALDH2 rs671 mutant is found in 30–50% of East Asians, while the ALDH2*2 variant is essentially absent in people of European descent (<5%) [101,102].
Glu487 is located within the oligomerization domain at the dimer interface of the tetrameric enzyme; this location is critical for the formation of both dimer and tetramer. Glutamic acid is mainly negatively charged at physiological pH, while lysine is mostly positively charged. The structures of the mutant subunit and its dimer chaperone are altered once Glu487 is replaced by Lys487 [103], a mutation that causes disorder of the α-helices (Figure 2A). The normal formation of α-helices near the dimer interface of each subunit is inhibited, which affects the formation of local secondary ALDH2 tetramer structures [12,88,104]. These changes cause the catalytic and coenzyme-binding domains of a single subunit to rotate away from the interface to destroy the NAD+ binding sites, resulting in severe loss of ALDH2 enzyme activity [88].
Figure 2. Tetramer structure of ALDH2. (A) (a) Tetramer structure of wild-type (WT) ALDH2 (ALDH2*1, Protein Data Bank [PDB] Accession Code: 1O05). (A) (b) Tetramer structure of the inactive mutant of ALDH2 (ALDH2*2, PDB Accession Code: 1ZUM). Green circle: the α-helix structure missing from the mutant tetramer versus WT; modified from [105]. (B) (a) “E” = subunit encoded by the WT allele ALDH2*1; the first base of codon 487 is guanine (G), and the 487th amino acid residue is glutamic acid (Glu/E). “K” = subunit encoded by the mutant allele ALDH2*2; the first base of codon 487 is mutated to adenine (A), and the 487th amino acid residue is changed from Glu to lysine (Lys/K). (B) (b) E4 homotetramers reflect normal enzyme activity; heterologous tetramer E3K, E2K2, and EK3 enzyme activities are substantially reduced; and K4 homotetramer enzyme activities are almost lost.
Figure 2. Tetramer structure of ALDH2. (A) (a) Tetramer structure of wild-type (WT) ALDH2 (ALDH2*1, Protein Data Bank [PDB] Accession Code: 1O05). (A) (b) Tetramer structure of the inactive mutant of ALDH2 (ALDH2*2, PDB Accession Code: 1ZUM). Green circle: the α-helix structure missing from the mutant tetramer versus WT; modified from [105]. (B) (a) “E” = subunit encoded by the WT allele ALDH2*1; the first base of codon 487 is guanine (G), and the 487th amino acid residue is glutamic acid (Glu/E). “K” = subunit encoded by the mutant allele ALDH2*2; the first base of codon 487 is mutated to adenine (A), and the 487th amino acid residue is changed from Glu to lysine (Lys/K). (B) (b) E4 homotetramers reflect normal enzyme activity; heterologous tetramer E3K, E2K2, and EK3 enzyme activities are substantially reduced; and K4 homotetramer enzyme activities are almost lost.
Ijms 23 02682 g002

4. ALDH2 in Oxidative Stress-Related Physiological and Pathological Processes

The metabolism of various endobiotic and xenobiotic compounds produces a series of aldehydes. Endogenous aldehydes are formed when amino acids, carbohydrates, lipids, biogenic amines, vitamins, and steroids are metabolized [18,106,107]. In addition, the biotransformation of many drugs and environmental chemicals produces aldehydes [105]. Accumulation of these aldehydes in cells enhances the damage caused by OS and induces related diseases. As ALDH2 is an oxidoreductase, its catalytic deficiency is related to fibrosis, apoptosis, aging, and neurodegeneration (Figure 3) [108]. Epidemiological and functional studies have reported that ALDH2 Glu487Lys gene polymorphism and the alterations of its enzyme activity not only are related to adverse reactions to alcohol after drinking, such as severe facial flushing, nausea, headache, and tachycardia, but also play important roles in cardio- and cerebrovascular diseases (CCVDs), cancer, NDDs, osteoporosis and drug toxicity (Table 2, Figure 3) [109,110].

4.1. Fibrosis

The process of lipid peroxidation caused by OS produces aldehydes, which is related to tissue and organ fibrosis. Fibrosis is a common feature of most chronic diseases involving tissue damage [123], affecting multiple organs, such as the liver, kidneys, cardiovascular (CV) system, lungs, and skin. There are different harmful compounds and inducements that lead to the occurrence and development of fibrosis in different organs. The fibrotic response is characterized by an increase in extracellular matrix (ECM) components [124,125], such as fibronectin (FN), collagen, laminin, elastin, fibrillin, and proteoglycans, together with the proliferation, migration, and accumulation of mesenchymal cells [126]. Transforming growth factor-β (TGF-β) can regulate the migration, proliferation, and differentiation of endothelial cells, epithelial cells (ECs), and fibroblasts; promote the production of collagen and FN in ECs and mesenchymal cells; and lead to net accumulation of ECM during wound healing and fibrosis [127]. Increasing evidence shows that TGF-β acts as a key regulator of fibrosis.
ALDH2 plays a protective role in myocardial, liver, and other organ fibrosis caused by OS-related diseases and exposure to toxic chemicals [128]. Under the action of TGF-β, both Drosophila mothers against decapentaplegic protein 2 (Smad2) and Smad3 are activated by phosphorylation, and Smad2/3 is an upstream regulator of ALDH2. The activation of ALDH2 regulated by the TGF-β1/Smad signaling pathway inhibits the transformation of human cardiac fibroblasts (HCFs) into myofibroblasts, which is manifested by a decrease in α-smooth muscle actin (α-SMA) and periostin expression, as well as myofibroblastic proliferation, collagen formation, and contractility reduction [111]. Studies have shown that activation of ALDH2 can reduce expression of collagen types I and III (COL1, COL3) as well as α-SMA in rat hearts and reduce levels of β-catenin, phosphor–glycogen synthase kinase-3β (p-GSK-3β), and wingless-type mouse mammary tumor virus (MMTV) integration site family member 1 (Wnt-1), indicating that ALDH2 protects against myocardial infarction (MI)-related cardiac fibrosis by downregulating the Wnt/catenin signaling pathway [129]. In addition, ALDH2 is important in the pathogenesis of alcoholic cirrhosis. Feeding ethanol plus carbon tetrachloride (CCl4) to ALDH2 knockout (KO) mice leads to the accumulation of malondialdehyde–acetaldehyde adducts (MAAs) in the liver, stimulating cells to produce pro-inflammatory cytokines, such as interleukin-6 (IL-6). Hepatic expression levels of α-SMA, TGF-β, and collagen type I alpha 1 (COLIA1) messenger ribonucleic acid (mRNA) have been shown to be much higher in ALDH2−/− mice than in WT mice, which led to inflammation and accelerated liver fibrosis; this indicated that ALDH2 deficiency can aggravate liver inflammation and fibrosis in mice [114]. N-(1,3-Benzodioxol-5-ylmethyl)-2,6-dichlorobenzamide (Alda-1) can activate the nuclear factor erythroid 2-related factor 2/heme oxygenase-1 (Nrf2/HO-1) antioxidant pathway and mitosis regulated by Parkin protein to reduce CCl4-induced chronic liver fibrosis in ALDH2 KO mice [115]. Numerous studies have suggested that the inhibition of the TGF-β/Smad pathway and the production of ROS stimulated by TGF-β exerts an anti-fibrotic effect on pulmonary and kidney fibroses. Therefore, we speculated that inhibition of TGF-β/Smad signaling can reduce collagen and α-SMA expression by activating ALDH2 to delay OS-related fibrosis in the heart, liver, and other organs [130,131,132,133,134,135,136,137]. However, these hypotheses need to be confirmed by additional relevant studies.

4.2. Apoptosis

ROS produced during OS can induce and regulate apoptosis. Apoptosis (programmed cell death) is a programmed physiological process that removes superfluous, damaged, or altered cells to regulate homeostasis in the organism. Excessive apoptosis can lead to a decline in tissue and organ function and is implicated in a growing number of diseases, whereas insufficient apoptosis leads to carcinogenesis [138]. Apoptosis is a complex protease cascade reaction process mediated by members of the cysteinyl aspartate-specific proteinase (caspase) family [139]. The upstream caspases activate caspase-9 and further activate downstream caspases-3 and -7 under the action of apoptosis-inducing factor (AIF), hydrolyzing target substances in cells on a large scale and degrading intracellular proteins to induce cell death [140]. B-cell lymphoma-2 (Bcl-2) family members are important regulators of apoptosis; studies have demonstrated that balance of the expression levels of anti-apoptotic Bcl-2 and pro-apoptotic Bcl-2–associated X protein (Bax) play a major role in the regulation of apoptosis, and the Bcl-2/Bax ratio represents the degree of apoptosis [113]. Activation of mitogen-activated protein kinases (MAPKs), including extracellular signal-regulated kinase 1/2 (ERK1/2), p38 mitogen-activated protein kinase (p38 MAPK), and c-Jun NH2-terminal kinase (JNK), are closely related to apoptosis induced by OS. It has been found that the activation of p38 MAPK and JNK exert pro-apoptotic functions [141,142]. The expression of Bcl-2 is regulated by the upstream JNK and p38 MAPK under specific treatments. The activation of JNK can reduce Bcl-2 protein levels, and p38 MAPK can regulate the translocation of Bax to mitochondria, and promote the phosphorylation of Bcl-2 to accelerate its degradation, and subsequent activation of caspase-3 to lead to cell death [141,143]. Researchers have reported that upregulation of ALDH2 expression is accompanied by an increase in the myocardial Bcl-2/Bax ratio and a decrease in active caspase-3 levels, indicating that ALDH2 upregulation might help reduce the occurrence of myocardial-cell apoptosis [144].
Apoptosis of myocardial cells induced by OS is a main factor in the pathogenesis of heart failure. ALDH2 can protect myocardium by reducing myocardial-cell apoptosis caused by OS and oxidative damage to myocardium [145]. Studies suggest that ALDH2 can protect myocardial cells from OS and apoptosis induced by antimycin A by downregulating the MAPK signaling pathway [112]. In addition, ALDH2 can reduce renal cell apoptosis by inhibiting the MAPK pathway and increasing the Bcl-2/Bax ratio during ischemia/reperfusion (I/R) injury [146]. ALDH2 activated by Alda-1 has been shown to decrease the levels of MDA, 4-HNE, and cleaved caspase-3 expression, and increase the Bcl-2/Bax ratio in pancreatic acinar cells [147]. Overexpression of ALDH2 transgene prevents acetaldehyde-induced OS and inhibits the activation of key stress signaling molecules ERK1/2 and p38 MAPK as well as that of caspase-3, and also inhibits apoptosis of human umbilical-vein endothelial cells [70]. However, in the absence of ALDH2 protein from ALDH2 KO mice, expression of active caspases-9 and -3 in mouse myocardial cells is increased, which induces heart injury by activating ROS-dependent apoptosis as well as necroptosis medicated by receptor interacting protein kinases-1 and -3/mixed-lineage kinase domain-like protein (RIP1/RIP3/MLKL) [148]. It can be seen that the increase in ALDH2 expression is often accompanied by a reduction in MAPK signaling pathways or an increase in the Bcl-2/Bax ratio to reduce apoptosis [122,149,150]. Given the results of studies investigating the effect of MAPKs on the Bcl-2/Bax ratio, we speculate that ALDH2 inhibits apoptosis by increasing the Bcl-2/Bax ratio and reducing caspase-3 levels possibly via MAPK signaling pathway. However, the specific regulatory mechanism of each MAPK member on the Bcl-2/Bax ratio under the action of ALDH2 remains to be elucidated.

4.3. Aging

Evidence suggests that OS is a major underlying pathogenic factor in the aging process and related diseases. Cellular senescence is a stress response that links multiple pathologies, including atherosclerosis, stroke, and NDDs [151,152]. The general free-radical theory of aging attributes this process to the accumulation of damage from oxidative modifications of cellular components by ROS, which induces various diseases [153,154]. Some studies support a close relationship between ALDH2 deficiency and aging, especially cardiac and vascular diseases related to aging of ECs.
The activation of ALDH2 can increase SIRT1 levels to inhibit acetylation of the aging marker p53 and resist cell senescence. Studies have demonstrated that ALDH2 improves aging-related myocardial ischemic intolerance via a deacetylase sirtuin-dependent manner [155]. Sirtuin 1 (SIRT1) belongs to a class of NAD+-dependent histone deacetylases (sirtuins). It is widely involved in physiological activities such as fatty-acid oxidation and stress tolerance, and is closely related to apoptosis and aging [156,157]. Studies have found that activation or overexpression of ALDH2 can reduce levels of aging markers such as senescence-associated β-galactosidase (SA-β-gal) and p53/p21 in aging cells to reduce the risk of myocardial aging and atherosclerosis [158,159,160]. P53 is a downstream target of SIRT1 and plays a key role in cell aging [161,162]. ROS accumulated during OS can induce p53-dependent aging through the p38 MAPK and ataxia telangiectasia mutated protein (ATM) pathways. Meanwhile, 4-HNE produced by OS can combine with SIRT1 to form SIRT1–4-HNE adducts, which hinder the transfer of SIRT1 from cytoplasm to nucleus and reduce SIRT1 deacetylation activity, significantly increasing the level of acetylated p53 and promoting the occurrence of aging. ALDH2 affects the acetylation of p53 by regulating SIRT1 activity [158]. ALDH2 can also inactivate the EC p38 MAPK pathway and inhibit the ROS–MAPK–p53/p16 pathway to participate in the anti-aging process, as well as inhibit 4-HNE accumulation and protect endothelium by regulating SIRT1/p53-dependent EC senescence [120,161].

4.4. Nervous System Injury

Aldehydes produced by lipid peroxidation, especially 4-HNE, are related to nerve trauma, NDDs, and neuroinflammation [163]. The strong electrophilic properties of 4-HNE based on α,β-unsaturated structures make it highly reactive and liable to form stable protein adducts, causing cell damage [160,164]. Oxidative damage causes loss of nerve reserve and cognitive dysfunction. Alzheimer’s disease (AD) and Parkinson’s disease (PD) are both caused by progressive degeneration of the central nervous system (CNS).
OS is an important factor in the pathogenesis of AD. The main pathological characteristics of the AD brain are the presence of senile plaques (SPs) with abnormally aggregated β-amyloid (Aβ) as the main component, neurofibrillary tangles (NFTs) formed by hyperphosphorylated microtubule-associated protein tau, and neuronal loss. However, 4-HNE can significantly inhibit microtubule formation and neurite outgrowth and can react with phosphorylated tau to induce conformational changes in tau protein, thereby promoting NFT formation and causing neuronal death and synaptic dysfunction [12]. 4-HNE covalently modifies Aβ by 1,4 conjugates, altering Aβ’s function and impairing cell metabolism, signal transduction, and structural integrity [154]. Studies have demonstrated that the lack of ALDH2 activity caused by expression of the ALDH2*2 gene leads to accumulation of 4-HNE in the brains of transgenic mice and the death of neuronal cells, resulting in age-related neurodegeneration; this shows that ALDH2 can reduce 4-HNE–induced Aβ production and tau phosphorylation to prevent NFTs and SPs formation so as to inhibit neurotoxicity in AD [74].
The most prominent pathological changes in PD are the degeneration of dopaminergic neurons in the substantia nigra (SN) of the midbrain. The typical neuropathological feature is the aggregation of α-synuclein (α-Syn) in the remaining neurons into Lewy bodies (LBs) [165]. Studies have shown that 3,4-dihydroxyphenylacetaldehyde (DOPAL) is covalently modified with α-Syn by Lys residue, inducing the oligomerization of α-Syn and damaging synaptic vesicles of dopaminergic neurons [166]. ALDH2 either (1) catalyzes oxidation of the dopamine (DA) metabolite DOPAL to 3,4-dihydroxyphenylacetic acid (DOPAC), which can be further metabolized into homovanillic acid (HVA) or (2) catalyzes 3,4-dihydroxyphenylglycolaldehyde (DOPEGAL), the metabolite of norepinephrine (NE) and epinephrine (Epi), to 3,4-dihydroxymandelic acid (DOMA), which can be further metabolized into vanillylmandelic acid (VMA) so as to reduce ROS and caspase-3 protein expression as well as the aggregation of α-Syn to protect neurons from neurite damage [118,167,168]. That is, ALDH2 can inhibit α-Syn abnormal aggregation to form LBs in the SN of the brain by metabolizing the toxic DA metabolite DOPAL into DOPAC, thereby alleviating dopaminergic neuron injury in PD.

4.5. Other Oxidative Stress-Related Diseases

Studies have also demonstrated that ALDH2 deficiency is likely to cause the accumulation of aldehydes in the body under long-term adverse stimulation, which can induce gene mutations and tissue damage by reacting with biomolecules, such as proteins and nucleic acids, to generate various adducts. This increases the risk of hypertension, pulmonary hypertension and other vascular diseases, osteoporosis, and cancer [10,11,12,13,14,169] (Table 2). Among them, in addition to those on CCVDs, the mechanisms of ALDH2 involvement in cancer have been widely reported, including digestive system cancer, lung cancer, breast cancer, head and neck cancer [170,171,172]. ALDH2 is widely associated with cancer mediated by the metabolic disorder of alcohol and aldehydes, and it could be a possible prognostic biomarker for several types of cancer [173]. One of the most common studies is about the regulatory mechanism of ALDH2 gene polymorphism in digestive system cancer, such as esophageal cancer, colorectal cancer, gastric cancer, liver cancer, and pancreatic cancer [173,174,175,176,177]. ALDH2 was reported to protect against hepato-cellular carcinoma (HCC) metastasis by reducing acetaldehyde accumulation and activating the adenosine 5’-monophosphate-activated protein kinase (AMPK) pathway [169]. The expression of ALDH2 is decreased by silencing transcription factor FOXM1, which subsequently increases the level of Bax and cleaved-caspase-3 to promote the apoptosis of liver cancer stem cells (LCSCs) as well as inhibit tumorigenesis of LCSCs [178]. However, although there are many meta-analyses on the association between ALDH2 gene polymorphism and cancer risk, studies on the regulatory mechanism of ALDH2 involved in cancer pathogenesis and prognosis are still relatively scarce. ALDH2 plays an important role in the occurrence and development of OS-related diseases. Many diseases are accompanied by an increase in oxidation level or a decrease in antioxidant capacity to different degrees. The relationships between OS and the mechanisms of various diseases have received increasingly more attention [13], and research on the regulation of ALDH2 activity therefore deserves the same.

5. Regulation of ALDH2 Activity

5.1. Activators

Alda-1 is a specific agonist of ALDH2 [108,179,180,181,182]. It can activate WT ALDH2*1 and restore the activity of mutant ALDH2*2 to near-WT levels by acting as a structural chaperone. Alda-1 binds, not near the ALDH2*2 mutation site, but at the substrate entrance tunnel of the ALDH2 apo-enzyme, and extends to the active site. This mode of binding leaves the catalytic Cys302 and Glu268 residue unimpeded. After Alda-1 binds to ALDH2, dehydrogenase activity is activated by limiting substrate diffusion in the substrate-to-coenzyme tunnel, increasing the effective concentration of reactive groups within the active site. Alda-1 also reduces the accessibility of 4-HNE to Cys302 and its adjacent Cys301 and Cys303 residues and inhibits the formation of aldehyde adducts, reversing the inhibitory effect of 4-HNE on ALDH2 [176]. At the same time, the binding of Alda-1 restores the structures of the mutant ALDH2 (ALDH2*2) tetramer’s α-helix and active-site loop to near-similarity with those of their WT counterparts, improving structural stability and rescuing most of the functional defects [181,183].
Many studies have verified that Alda-1 is effective in activating ALDH2 activity. Some have found that Alda-1 improves brain injury and neurological function by increasing ALDH2 activity and reducing the accumulation of active aldehydes [184]. Alda-1, an activator of ALDH2, can significantly reverse the liver I/R injury of rats by inhibiting the production of ROS and reducing the accumulation of 4-HNE and MDA, all of which significantly decrease cell apoptosis and reduce liver pathological damage [185]. Another study found that levels of tumor necrosis factor-alpha (TNF-α) and IL-6 in the retinas of an Alda-1–treated group were lower; superoxide dismutase (SOD) activity was significantly higher; and expression of ALDH2, SIRT1, and Nrf2 was significantly increased. Meanwhile, expression of vascular endothelial growth factor-alpha (VEGF-α) was significantly decreased. These findings indicated that ALDH2 alleviates retinal damage [186]. In addition, Alda-1 can completely eliminate the increase in 4-HNE levels in neurons subjected to oxygen–glucose deprivation (OGD) and reduce neurotoxicity [37]. Other studies have found that Alda-1 treatment can reduce 4-HNE levels, reduce lung injury by activating ALDH2 activity, and improve the integrity of the lung epithelial barrier in rat models of severe hemorrhagic shock and resuscitation [187]. Alda-44 has also been shown to be a direct activator of ALDH2, which reduces I/R-induced cardiomyocytic apoptosis by decreasing the 4-HNE–mediated activation (phosphorylation) of stress-activated protein kinase (SAPK)/JNK [180]. Other studies have shown that activated protein kinase C epsilon (PKCε) can lead to activation of its downstream substrate ALDH2, which plays a protective role in heart and brain injury [37,85,88,188]. Moderate levels of ethanol [37,189], isoflurane [190], estrogen [140], heat shock factor 1 (HSF1) [85], and melatonin [188] have been reported to increase ALDH2 activity by enhancing the mitochondrial import of PKCε.
In addition to specific activators, antioxidants such as lipoic acid (LA) [191], resveratrol [192], and vitamins such as D and E indirectly activate ALDH2 activity via related signaling pathways, which can effectively reduce the risk of cell damage and disease. Previous studies showed that LA activated ALDH2 in animal models of heart failure, nitrate tolerance, and diabetes [193,194,195]. Consequently, these antioxidants and vitamins have been the subjects of further research and development [196,197].

5.2. Inhibitors

Daidzin, a kind of isoflavone that predominantly exists in glucoside form [198,199], is a specific inhibitor of ALDH2 in vitro. As with Alda-1, daidzin binds to ALDH2 at the entrance of the substrate tunnel, occupying a site that partially overlaps with Alda-1 [181]. However, the additional phenol arm of daidzin reaches the catalytic site and contacts Cys302 and Glu268, thereby blocking the catalytic function of ALDH2. Kinetic studies have demonstrated that daidzin inhibition is competitive against aldehyde substrates but not against NAD+ due to the considerable conformational flexibility of the latter when binding to ALDH2 [199].
Studies have shown that daidzin increases ROS production in mice with angiotensin II (Ang II)–induced hypertension by inactivating ALDH2, which increases vasoconstriction and aggravates hypertensive organ damage [200]. Other studies have found that daidzin’s inhibition of ALDH2 activity can significantly enhance apoptosis and ROS production in rat myocardial cells by activating the MAPK signaling pathway [112]. In addition, daidzin can promote differentiation of HCFs induced by TGF-β1 and the fibrotic process in myocardial cells [111,128].
Cyanamide (CYA) [37,201], disulfiram (DSF) [199,201,202], and other non-specific inhibitors can also indirectly inhibit ALDH2 activity through related signaling pathways, giving them application value in both clinical treatment of diseases and scientific research. Studies have demonstrated that CYA can downregulate expression of ALDH2 protein in a rat acute kidney injury (AKI) model induced by cecal ligation and puncture (CLP). Researchers have found that levels of plasma creatinine (CRE), blood urea nitrogen (BUN), renal MDA, and NF-κB p65 protein expression increased while SOD activity decreased and glomerular atrophy was aggravated, indicating that CYA aggravates renal injury by inhibiting ALDH2 [121]. DSF, which was approved by the FDA to treat alcohol abuse and dependence in 1951 [203], is bioactivated to S-methyl N, N-diethylthiocarbamate sulfoxide (DETC-MeSO) to inhibit ALDH2. ALDH2 inhibition by DSF exacerbates the Ang II–induced decrease in coronary angiogenesis by lowering levels of VEGF receptors (VEGFR1 and VEGFR2) and 4-HNE adducts to ameliorate cardiometabolic diseases [204]. DSF can also be used in aversion therapy for alcohol abuse and dependence; it increases acetaldehyde in the blood and leads to disulfiram–alcohol reactions such as nausea, vomiting, headache, hypotension, and tachycardia, which deter patients from consuming alcohol.

5.3. Post-Translational Modifications (PTMs) of ALDH2

PTMs of some mitochondrial enzymes are involved in the regulation of ROS production and oxidative metabolism, thereby playing important roles in cellular functions. Different types of protein PTMs have different effects on ALDH2 activity [205].
PKCε phosphorylation is important in the regulation and protection of ALDH2 enzyme activity. PKCε can activate ALDH2 by inducing phosphorylation of serine/threonine in ALDH2, while PKCε can be activated by several pathways [88]. Studies have shown that ethanol can activate PKCε during I/R in vivo [180,206] and ALDH2 is further activated by phosphorylation of Thr185, Thr412, and possibly Ser279 to protect the heart [179]. Estrogen in females can also lead to ALDH2 phosphorylation by activating phosphatidylinositol 3-kinase (PI3K) and PKC so as to reduce production of ROS and toxic aldehydes; this might be why the degree of I/R injury is lower in female than in male rats [207]. Research has found that PKCε is activated after I/R and accelerates its transfer to mitochondria under the action of Hsp90, increasing the phosphorylation and activity of ALDH2 [208]. In addition, isoflurane, a volatile anesthetic, can mediate ALDH2 phosphorylation and protect the heart from I/R injury through PKCε translocation from cytoplasm to mitochondria [190]. Under different mechanisms, phosphorylation at different sites also inhibits ALDH2 enzyme activity. In contrast to PKCε-mediated phosphorylation and subsequent activation of ALDH2, one study indicated that CCl4 exposure activates JNK, which translocates to mitochondria and, there, phosphorylates and inhibits ALDH2 [209].
SIRT3, which is mainly located in the mitochondrial matrix, is a histone deacetylase that depends on NAD+ to regulate the acetylation level of lysine residues [88,205]. Lysines of ALDH2 are also targets of deacetylase SIRT3 [210]. SIRT3 can change the acetylation level of lysine sites in the domain that ALDH2 binds with its cofactor NAD+, thereby affecting the binding of these sites; acetylation changes the structure of homotetramer and affects its activity. However, reports differ on acetylation’s effect on ALDH2 activity. Most studies suggest that SIRT3-mediated deacetylation reduces the activity of ALDH2 dehydrogenase [16,88]. Some research has shown that under acute ethanol exposure, SIRT3 is inactivated due to the decrease in NAD+/NADH ratio induced by low-dose ethanol, leading to high acetylation and activation of ALDH2 [105,211]. Other studies have found that acetaminophen (N-acetyl-p-aminophenol [APAP]) promotes SIRT3-mediated ALDH2 deacetylation and deactivation [212]. Conversely, acetylation of ALDH2 in vitro leads to a significant decrease in enzymatic activity, while SIRT3 can partially restore ALDH2 activity [210]; this study was conducted completely in vitro using human recombinant ALDH2 protein, which cannot completely simulate the complex living environment of cells in vivo, but those findings also suggested that regulation of acetylation by ALDH2 activity involves a more complex mechanism, which is worthy of further exploration.
ALDH2 activity can be inhibited by such conditions or substances as chronic or excessive ethanol abuse [213,214], APAP [215], nitric oxide (NO) donor [215], and I/R injury [216] via cysteine S-nitrosylation and tyrosine nitration, because of the increase in the production of peroxynitrite (ONOO) [217]. Mitochondria-targeted lipophilic ubiquinone (MitoQ) can prevent alcohol-induced S-nitrosylation of cysteine residues and nitration of tyrosine residues from impairing ALDH2 activity, so as to restore ALDH activity, and reduce alcohol-induced liver and intestinal damage [214].
In the aggregate (Figure 4), phosphorylation is generally related to activation of ALDH2 enzyme activity, but phosphorylation through the JNK pathway leads to inactivation of ALDH2. Most studies have revealed the promotional effect of acetylation on ALDH2 enzyme activity, but this must be confirmed by additional studies. The more consistent conclusion is that S-nitrosylation and nitration generally cause a decrease in ALDH2 enzyme activity. In addition, ubiquitination and glycosylation are widely involved in the regulation of OS protein function in vivo, but few studies have been conducted on their relationship with the regulation of ALDH2 activity. There is a synergistic and competitive relationship between different modification types, such as the mutual stimulation and inhibition among phosphorylation, methylation, glycation, acetylation, and ubiquitination [218,219,220,221], which plays an important role in the function of the target protein. Moreover, different modifications play different roles such as activation, inhibition, and even degradation of the target protein (such as ubiquitination) [222,223,224,225,226]. The synergy or inhibition among different modifications, together with their impact on the activity of the target protein, is closely related to the types of target proteins and modification sites of the modification, which deserves more attention in follow-up research.

6. Conclusions

ALDH2 not only plays an important role in the prevention and treatment of alcohol abuse, but also widely participates in the key regulatory pathways of diseases related to apoptosis, aging, fibrosis, and neurodegeneration [105]. A certain degree of research on the activation of ALDH2 and improvement of its function, in particular the compensation for the functional defects of its mutant genotypes (ALDH2*1/2, ALDH2*2/2), has been carried out. Alda-1, an effective activator of ALDH2, has certain clinical value and application potentialities, but due to its poor solubility in most common solvents, its physical and chemical properties still need to be further explored to improve the drug-like properties. In addition to specific activators, other non-specific activators, such as antioxidant LA can indirectly activate ALDH2 through related signaling pathways and inhibit toxic aldehydes caused by OS and cytotoxicity. LA is an over-the-counter supplement that acts as a potent antioxidant and has been widely used as supplementary treatment of diabetic neuropathy [227,228]. However, further extensive and in-depth exploration of the regulation of non-specific activators on ALDH2 and related signaling pathway is still needed.
When exploring the regulation mechanism of ALDH2 activity, the role of PTMs cannot be ignored. ALDH isozymes can be modified by various forms of PTMs, including phosphorylation, acetylation, and S-nitrosylation, which affect its activity effectively [154]. Although most studies show that phosphorylation and acetylation are mainly involved in the activation mechanism of ALDH2 activity, nitrosylation and nitrification mainly damage its activity, it is not excluded that sometimes different upstream regulatory substances or different modification sites will also have the opposite effect. Regulation of ALDH2 by other PTMs and the synergy or inhibition between different types of PTMs are worthy of further research. Therefore, in addition to the exploration of safer and more efficient activators of ALDH2 and their regulatory pathways, the degree and regulatory mechanisms of its key regulatory PTMs and the interaction among different modification types are also worthy of attention. In the context of human diseases associated with ALDH2 activity, it will be worth exploring the roles played by PTMs of ALDH2 and their related regulatory pathway in the future.

Author Contributions

J.G. assisted in literature review, drafted the manuscript, composed the figures, and reviewed the final paper. X.G. contributed to the concept and design of this article, assessed the literature and revised the final manuscript. Y.H. and X.P. assessed the literature and revised the final manuscript. All authors read and approved the final manuscript.

Funding

This work was supported in part by the National Science Fund for Distinguished Young Scholars (32002208), the Central Public-interest Scientific Institution Basal Research Fund (2020-YWF-YTS-09), and the Agricultural Science and Technology Innovation Program (ASTIP-IAS07).

Acknowledgments

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Seno, T.; Inoue, N.; Gao, D.; Okuda, M.; Sumi, Y.; Matsui, K.; Yamada, S.; Hirata, K.-I.; Kawashima, S.; Tawa, R.; et al. Involvement of NADH/NADPH oxidase in human platelet ROS production. Thromb. Res. 2001, 103, 399–409. [Google Scholar] [CrossRef]
  2. Breitzig, M.; Bhimineni, C.; Lockey, R.; Kolliputi, N. 4-Hydroxy-2-nonenal: A critical target in oxidative stress? Am. J. Physiol. Cell Physiol. 2016, 311, C537–C543. [Google Scholar] [CrossRef]
  3. Klaunig, J.E.; Xu, Y.; Isenberg, J.S.; Bachowski, S.; Kolaja, K.L.; Jiang, J.; Stevenson, D.E.; Walborg, E.F., Jr. The role of oxidative stress in chemical carcinogenesis. Environ. Health Perspect. 1998, 106, 289–295. [Google Scholar] [CrossRef] [PubMed]
  4. Seddon, M.; Looi, Y.H.; Shah, A.M. Oxidative stress and redox signalling in cardiac hypertrophy and heart failure. Heart Off. J. Br. Card. Soc. 2007, 93, 903–907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Ayala, A.; Muñoz, M.F.; Argüelles, S. Lipid peroxidation: Production, metabolism, and signaling mechanisms of malondialdehyde and 4-Hydroxy-2-Nonenal. Oxid. Med. Cell. Longev. 2014, 2014, 360438. [Google Scholar] [CrossRef]
  6. Marchitti, S.A.; Brocker, C.; Stagos, D.; Vasiliou, V. Non-P450 aldehyde oxidizing enzymes: The aldehyde dehydrogenase superfamily. Expert Opin. Drug Metab. Toxicol. 2008, 4, 697–720. [Google Scholar] [CrossRef] [Green Version]
  7. Singh, S.; Brocker, C.; Koppaka, V.; Ying, C.; Jackson, B.; Matsumoto, A.; Thompson, D.C.; Vasiliou, V. Aldehyde dehydrogenases in cellular responses to oxidative/electrophilicstress. Free Radic. Biol. Med. 2013, 56, 89–101. [Google Scholar] [CrossRef] [Green Version]
  8. Moazamian, R.; Polhemus, A.; Connaughton, H.; Fraser, B.; Whiting, S.; Gharagozloo, P.; Aitken, R.J. Oxidative stress and human spermatozoa: Diagnostic and functional significance of aldehydes generated as a result of lipid peroxidation. Mol. Hum. Reprod. 2015, 21, 502–515. [Google Scholar] [CrossRef] [Green Version]
  9. Zhang, R.; Liu, B.; Fan, X.; Wang, W.; Xu, T.; Wei, S.; Zheng, W.; Yuan, Q.; Gao, L.; Yin, X.; et al. Aldehyde Dehydrogenase 2 Protects Against Post-Cardiac Arrest Myocardial Dysfunction Through a Novel Mechanism of Suppressing Mitochondrial Reactive Oxygen Species Production. Front. Pharmacol. 2020, 11, 373. [Google Scholar] [CrossRef]
  10. Hoshi, H.; Hao, W.; Fujita, Y.; Funayama, A.; Miyauchi, Y.; Hashimoto, K.; Miyamoto, K.; Iwasaki, R.; Sato, Y.; Kobayashi, T.; et al. Aldehyde-stress resulting from Aldh2 mutation promotes osteoporosis due to impaired osteoblastogenesis. J. Bone Miner. Res. 2012, 27, 2015–2023. [Google Scholar] [CrossRef]
  11. Mittal, M.; Khan, K.; Pal, S.; Porwal, K.; China, S.P.; Barbhuyan, T.K.; Baghe, K.S.; Rawat, T.; Sanyal, S.; Bhadauria, S.; et al. The thiocarbamate disulphide drug, disulfiram induces osteopenia in rats by inhibition of osteoblast function due to suppression of acetaldehyde dehydrogenase activity. Toxicol. Sci. 2014, 139, 257–270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Zhao, Y.; Wang, C.C. Glu504Lys single nucleotide polymorphism of aldehyde dehydrogenase 2 gene and the risk of human diseases. Biomed Res. Int. 2015, 2015, 174050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Orywal, K.; Szmitkowski, M. Alcohol dehydrogenase and aldehyde dehydrogenase in malignant neoplasms. Clin. Exp. Med. 2017, 17, 131–139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Zhao, Y.; Wang, B.L.; Zhang, J.; He, D.Y.; Zhang, Q.; Pan, C.; Yuan, Q.H.; Shi, Y.N.; Tang, H.Y.; Xu, F.; et al. ALDH2 (aldehyde dehydrogenase 2) protects against hypoxia-induced pulmonary hypertension. Arterioscler. Thromb. Vasc. Biol. 2019, 39, 2303–2319. [Google Scholar] [CrossRef]
  15. Vasiliou, V.; Pappa, A. Polymorphisms of human aldehyde dehydrogenases. Pharmacology 2000, 61, 192–198. [Google Scholar] [CrossRef]
  16. Wei, S.J.; Xing, J.H.; Wang, B.L.; Xue, L.; Wang, J.L.; Li, R.; Qin, W.D.; Wang, J.; Wang, X.P.; Zhang, M.X.; et al. Poly (ADP-ribose) polymerase inhibition prevents reactive oxygen species induced inhibition of aldehyde dehydrogenase2 activity. Biochim. Biophys. Acta (BBA) Mol. Cell Res. 2013, 1833, 479–486. [Google Scholar] [CrossRef] [Green Version]
  17. Song, W.; Zou, Z.Y.; Xu, F.; Gu, X.X.; Xu, X.F.; Zhao, Q.S. Molecular cloning and expression of a second zebrafish aldehyde dehydrogenase 2 gene (aldh2b). DNA Seq. 2006, 17, 262–269. [Google Scholar] [CrossRef]
  18. Lee, D.J.; Lee, H.M.; Kim, J.H.; Park, I.S.; Rho, Y.S. Heavy alcohol drinking downregulates ALDH2 gene expression but heavy smoking up-regulates SOD2 gene expression in head and neck squamous cell carcinoma. World J. Surg. Oncol. 2017, 15, 163. [Google Scholar] [CrossRef] [Green Version]
  19. Molotkov, A.; Duester, G. Genetic evidence that retinaldehyde dehydrogenase Raldh1 (Aldh1a1) functions downstream of alcohol dehydrogenase Adh1 in metabolism of retinol to retinoic acid. J. Biol. Chem. 2003, 278, 36085–36090. [Google Scholar] [CrossRef] [Green Version]
  20. Vasiliou, V.; Nebert, D.W. Analysis and update of the human aldehyde dehydrogenase (ALDH) gene family. Hum. Genom. 2005, 2, 138–143. [Google Scholar] [CrossRef] [Green Version]
  21. Paterson, E.K.; Ho, H.; Kapadia, R.; Ganesan, A.K. 9-cis retinoic acid is the ALDH1A1 product that stimulates melanogenesis. Exp. Dermatol. 2013, 22, 202–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Hwang, P.H.; Lian, L.; Athanasios, I.; Zavras, A.I. Alcohol intake and folate antagonism via CYP2E1 and ALDH1: Effects on oral carcinogenesis. Med. Hypotheses 2012, 78, 197–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Fan, X.; Molotkov, A.; Manabe, S.-I.; Donmoyer, C.M.; Deltour, L.; Foglio, M.H.; Cuenca, A.E.; Blaner, W.S.; Lipton, S.A.; Duester, G. Targeted disruption of Aldh1a1 (Aaldh1) provides evidence for a complex mechanism of retinoic acid synthesis in the developing retina. Mol. Cell. Biol. 2003, 23, 4637–4648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Bchini, R.; Vasiliou, V.; Branlant, G.; Talfournier, F.; Rahuel-Clermont, S. Retinoic acid biosynthesis catalyzed by retinal dehydrogenases relies on a rate-limiting conformational transition associated with substrate recognition. Chem.-Biol. Interact. 2013, 202, 78–84. [Google Scholar] [CrossRef] [Green Version]
  25. Pai, J.; Haenisch, M.; Muller, C.H.; Goldstein, A.S.; Arnold, S.; Isoherranen, N.; Brabb, T.; Treuting, P.M.; Amory, J.K. Inhibition of retinoic acid biosynthesis by the bisdichloroacetyldiamine WIN 18,446 markedly suppresses spermatogenesis and alters retinoid metabolism in mice. J. Biol. Chem. 2014, 289, 15104–15117. [Google Scholar] [CrossRef] [Green Version]
  26. Trasion, S.E.; Harrison, E.H.; Wang, T.T.Y. Androgen regulation of aldehyde dehydrogenase 1A3 (ALDH1A3) in the androgen-responsive human prostate cancer cell line LNCaP. Exp. Biol. Med. 2007, 232, 762–771. [Google Scholar] [CrossRef]
  27. Sullivan, K.E.; Rojas, K.; Cerione, R.A.; Nakano, I.; Wilson, K.F. The stem cell/cancer stem cell marker ALDH1A3 regulates the expression of the survival factor tissue transglutaminase, in mesenchymal glioma stem cells. Oncotarget 2017, 8, 22325–22343. [Google Scholar] [CrossRef] [Green Version]
  28. Jackson, B.C.; Holmes, R.S.; Backos, D.S.; Reigan, P.; Thompson, D.C.; Vasiliou, V. Comparative genomics, molecular evolution and computational modeling of ALDH1B1 and ALDH2. Chem.-Biol. Interact. 2013, 202, 11–21. [Google Scholar] [CrossRef] [Green Version]
  29. Jackson, B.C.; Reigan, P.; Miller, B.; Thompson, D.C.; Vasiliou, V. Human ALDH1B1 polymorphisms may affect the metabolism of acetaldehyde and all-trans retinaldehyde—in vitro studies and computational modeling. Pharm. Res. 2015, 32, 1648–1662. [Google Scholar] [CrossRef] [Green Version]
  30. Hoeferlin, L.A.; Oleinik, N.V.; Krupenko, N.I.; Krupenko, S.A. Activation of p21-dependent G1/G2 arrest in the absence of DNA damage as an antiapoptotic response to metabolic stress. Genes Cancer 2011, 2, 889–899. [Google Scholar] [CrossRef] [Green Version]
  31. Khan, Q.A.; Pediaditakis, P.; Malakhau, Y.; Esmaeilniakooshkghazi, A.; Ashkavand, Z.; Sereda, V.; Krupenko, N.I.; Krupenko, S.A. CHIP E3 ligase mediates proteasomal degradation of the proliferation regulatory protein ALDH1L1 during the transition of NIH3T3 fibroblasts from G0/G1 to S-phase. PLoS ONE 2018, 13, e0199699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Krupenko, S.A.; Krupenko, N.I. Loss of ALDH1L1 folate enzyme confers a selective metabolic advantage for tumor progression. Chem.-Biol. Interact. 2019, 302, 149–155. [Google Scholar] [CrossRef] [PubMed]
  33. Krupenko, N.I.; Dubard, M.E.; Strickland, K.C.; Moxley, K.M.; Oleinik, N.V.; Krupenko, S.A. ALDH1L2 is the mitochondrial homolog of 10-formyltetrahydrofolate dehydrogenase. J. Biol. Chem. 2010, 285, 23056–23063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Strickland, K.C.; Krupenko, N.I.; Dubard, M.E.; Hu, C.J.; Tsybovsky, Y.; Krupenko, S.A. Enzymatic properties of ALDH1L2, a mitochondrial 10-formyltetrahydrofolate dehydrogenase. Chem.-Biol. Interact. 2011, 191, 129–136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Peng, G.S.; Wang, M.F.; Chen, C.Y.; Luu, S.U.; Chou, H.C.; Li, T.K.; Yin, S.J. Involvement of acetaldehyde for full protection against alcoholism by homozygosity of the variant allele of mitochondrial aldehyde dehydrogenase gene in Asians. Pharmacogenetics 1999, 9, 463–476. [Google Scholar] [CrossRef] [PubMed]
  36. Dollé, L.; Boulter, L.; Leclercq, I.A.; Grunsven, L.A.v. Next generation of ALDH substrates and their potential to study maturational lineage biology in stem and progenitor cells. Am. J. Physiol. Gastrointest. Liver Physiol. 2015, 308, G573–G578. [Google Scholar] [CrossRef] [Green Version]
  37. Guo, J.M.; Liu, A.J.; Zang, P.; Dong, W.Z.; Ying, L.; Wang, W.; Xu, P.; Song, X.R.; Cai, J.; Zhang, S.Q.; et al. ALDH2 protects against stroke by clearing 4-HNE. Cell Res. 2013, 23, 915–930. [Google Scholar] [CrossRef]
  38. Roy, B.; Sundar, K.; Palaniyandi, S.S. 4-hydroxy-2-nonenal decreases coronary endothelial cell migration: Potentiation by aldehyde dehydrogenase 2 inhibition. Vasc. Pharmacol. 2020, 131, 106762. [Google Scholar] [CrossRef]
  39. Choi, J.-W.; Kim, J.-H.; Cho, S.-C.; Ha, M.-K.; Song, K.-Y.; Youn, H.-D.; Park, S.C. Malondialdehyde inhibits an AMPK-mediated nuclear translocation and repression activity of ALDH2 in transcription. Biochem. Biophys. Res. Commun. 2011, 404, 400–406. [Google Scholar] [CrossRef]
  40. Townsend, A.J.; Leone-Kabler, S.; Haynes, R.L.; Wu, Y.; Szweda, L.; Bunting, K.D. Selective protection by stably transfected human ALDH3A1 (but not human ALDH1A1) against toxicity of aliphatic aldehydes in V79 cells. Chem.-Biol. Interact. 2001, 130–132, 261–273. [Google Scholar] [CrossRef]
  41. Wymore, T.; Hempel, J.; Cho, S.S.; Alexander, D.; MacKerell, A.D.; Hugh, B.; Nicholas, H.B., Jr.; Deerfield, W.D. Molecular recognition of aldehydes by aldehyde dehydrogenase and mechanism of nucleophile activation. Proteins Struct. Funct. Bioinform. 2010, 57, 758–771. [Google Scholar] [CrossRef] [PubMed]
  42. Nakahara, K.; Ohkuni, A.; Kitamura, T.; Abe, K.; Naganuma, T.; Ohno, Y.u.; Zoeller, R.A.; Kihara, A. The Sjögren-Larsson syndrome gene encodes a hexadecenal dehydrogenase of the sphingosine 1-phosphate degradation pathway. Mol. Cell 2012, 46, 461–471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Akio, K. Sphingosine 1-phosphate is a key metabolite linking sphingolipids to glycerophospholipids. Biochim. Biophys. Acta 2014, 1841, 766–772. [Google Scholar] [CrossRef]
  44. Neuber, C.; Schumacher, F.; Gulbins, E.; Kleuser, B. Method to simultaneously determine the sphingosine 1-phosphate breakdown product (2E)-hexadecenal and its fatty acid derivatives using isotope-dilution HPLC–electrospray ionization–quadrupole/time-of-flight mass spectrometry. Anal. Chem. 2014, 86, 9065–9073. [Google Scholar] [CrossRef] [PubMed]
  45. Marchitti, S.A.; Brocker, C.; Orlicky, D.J.; Vasiliou, V. Molecular characterization, expression analysis, and role of ALDH3B1 in the cellular protection against oxidative stress. Free Radic. Biol. Med. 2010, 49, 1432–1443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Marchitti, S.A.; Orlicky, D.J.; Vasiliou, V. Expression and initial characterization of human ALDH3B1. Biochem. Biophys. Res. Commun. 2007, 356, 792–798. [Google Scholar] [CrossRef] [Green Version]
  47. Kitamura, T.; Takagi, S.; Naganuma, T.; Kihara, A. Mouse aldehyde dehydrogenase ALDH3B2 is localized to lipid droplets via two C-terminal tryptophan residues and lipid modification. Biochem. J. 2015, 465, 79–87. [Google Scholar] [CrossRef] [Green Version]
  48. Pemberton, T.A.; Srivastava, D.; Sanyal, N.; Henzl, M.T.; Becker, D.F.; Tanner, J.J. Structural studies of yeast Δ1-pyrroline-5-carboxylate dehydrogenase (ALDH4A1): Active site flexibility and oligomeric state. Biochemistry 2014, 53, 1350–1359. [Google Scholar] [CrossRef]
  49. Tanner, J.J. Structural biology of proline catabolic enzymes. Antioxid. Redox Signal. 2017, 30, 650–673. [Google Scholar] [CrossRef]
  50. Bogner, A.N.; Stiers, K.M.; McKay, C.M.; Becker, D.F.; Tanner, J.J. Structural basis for the stereospecific inhibition of the dual proline/hydroxyproline catabolic enzyme ALDH4A1 by trans-4-hydroxy-L-proline. Protein Sci. 2021, 30, 1714–1722. [Google Scholar] [CrossRef]
  51. Campbell, A.C.; Bogner, A.N.; Mao, Y.; Becker, D.F.; Tanner, J.J. Structural analysis of prolines and hydroxyprolines binding to the L-glutamate-γ-semialdehyde dehydrogenase active site of bifunctional proline utilization A. Arch. Biochem. Biophys. 2021, 698, 108727. [Google Scholar] [CrossRef] [PubMed]
  52. Leo, S.; Capo, C.; Ciminelli, B.M.; Iacovelli, F.; Menduti, G.; Funghini, S.; Donati, M.A.; Falconi, M.; Rossi, L.; Malaspina, P. SSADH deficiency in an Italian family: A novel ALDH5A1 gene mutation affecting the succinic semialdehyde substrate binding site. Metab. Brain Dis. 2017, 32, 1383–1388. [Google Scholar] [CrossRef] [PubMed]
  53. Vogel, K.R.; Ainslie, G.R.; Jansen, E.E.W.; Salomons, G.S.; Gibson, K.M. Therapeutic relevance of mTOR inhibition in murine succinate semialdehyde dehydrogenase deficiency (SSADHD), a disorder of GABA metabolism. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 2017, 1863, 33–42. [Google Scholar] [CrossRef] [PubMed]
  54. Marcadier, J.L.; Smith, A.M.; Pohl, D.; Schwartzentruber, J.; Al-Dirbashi, O.Y.; Consortium, F.C.; Majewski, J.; Ferdinandusse, S.; Wanders, R.J.; Bulman, D.E.; et al. Mutations in ALDH6A1 encoding methylmalonate semialdehyde dehydrogenase are associated with dysmyelination and transient methylmalonic aciduria. Orphanet J. Rare Dis. 2013, 8, 98. [Google Scholar] [CrossRef] [Green Version]
  55. Brocker, C.; Cantore, M.; Failli, P.; Vasiliou, V. Aldehyde dehydrogenase 7A1 (ALDH7A1) attenuates reactive aldehyde and oxidative stress induced cytotoxicity. Chem.-Biol. Interact. 2011, 191, 269–277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Coulter-Mackie, M.B.; Tiebout, S.; Karnebeek, C.v.; Stockler, S. Overexpression of recombinant human antiquitin in E. coli: Partial enzyme activity in selected ALDH7A1 missense mutations associated with pyridoxine-dependent epilepsy. Mol. Genet. Metab. 2014, 111, 462–466. [Google Scholar] [CrossRef]
  57. Davis, I.; Yang, Y.; Wherritt, D.; Liu, A. Reassignment of the human aldehyde dehydrogenase ALDH8A1 (ALDH12) to the kynurenine pathway in tryptophan catabolism. J. Biol. Chem. 2018, 293, 9594–9603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Končitíková, R.; Vigouroux, A.; Kopečná, M.; Šebela, M.; Moréra, S.; Kopečný, D. Kinetic and structural analysis of human ALDH9A1. Biosci. Rep. 2019, 39, BSR20190558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Wyatt, J.W.; Korasick, D.A.; Qureshi, I.A.; Campbell, A.C.; Gates, K.S.; Tanner, J.J. Inhibition, crystal structures, and in-solution oligomeric structure of aldehyde dehydrogenase 9A1. Arch. Biochem. Biophys. 2020, 691, 108477. [Google Scholar] [CrossRef]
  60. Vasiliou, V.; Sandoval, M.; Backos, D.S.; Jackson, B.C.; Chen, Y.; Reigan, P.; Lanaspa, M.A.; Johnson, R.J.; Koppaka, V.; Thompson, D.C. ALDH16A1 is a novel non-catalytic enzyme that may be involved in the etiology of gout via protein–protein interactions with HPRT1. Chem.-Biol. Interact. 2013, 202, 22–31. [Google Scholar] [CrossRef] [Green Version]
  61. Bicknell, L.S.; Pitt, J.; Aftimos, S.; Ramadas, R.; Maw, M.A.; Robertson, S.P. A missense mutation in ALDH18A1, encoding Δ1-pyrroline-5-carboxylate synthase (P5CS), causes an autosomal recessive neurocutaneous syndrome. Eur. J. Hum. Genet. 2008, 16, 1176–1186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Wolthuis, D.F.G.J.; Asbeck, E.V.; Mohamed, M.; Gardeitchik, T.; Lim-Melia, E.R.; Wevers, R.A.; Morava, E. Cutis laxa, fat pads and retinopathy due to ALDH18A1 mutation and review of the literature. Eur. J. Paediatr. Neurol. 2014, 18, 511–515. [Google Scholar] [CrossRef] [PubMed]
  63. Holmes, R.S. Comparative and evolutionary studies of ALDH18A1 genes and proteins. Chem.-Biol. Interact. 2016, 276, 2–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Steenhof, M.; Kibæk, M.; Larsen, M.J.; Christensen, M.; Lund, A.M.; Brusgaard, K.; Hertz, J.M. Compound heterozygous mutations in two different domains of ALDH18A1 do not affect the amino acid levels in a patient with hereditary spastic paraplegia. Neurogenetics 2018, 19, 145–149. [Google Scholar] [CrossRef] [PubMed]
  65. Oota, H.; Pakstis, A.J.; Bonne-Tamir, B.; Goldman, D.; Grigorenko, E.; Kajuna, S.L.B.; Karoma, N.J.; Kungulilo, S.; Lu, R.B.; Odunsi, K.; et al. The evolution and population genetics of the ALDH2 locus: Random genetic drift, selection, and low levels of recombination. Ann. Hum. Genet. 2004, 68, 93–109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Serio, R.N.; Lu, C.; Gross, S.S.; Gudas, L.J. Different effects of knockouts in ALDH2 and ACSS2 on embryonic stem cell differentiation. Alcohol. Clin. Exp. Res. 2019, 43, 1859–1871. [Google Scholar] [CrossRef] [PubMed]
  67. Dalleau, S.; Baradat, M.; Guéraud, F.; Huc, L. Cell death and diseases related to oxidative stress: 4-hydroxynonenal (HNE) in the balance. Cell Death Differ. 2013, 20, 1615–1630. [Google Scholar] [CrossRef] [Green Version]
  68. Tudek, B.; Zdżalik-Bielecka, D.; Tudek, A.; Kosicki, K.; Fabisiewicz, A.; Speina, E. Lipid peroxidation in face of DNA damage, DNA repair and other cellular processes. Free Radic. Biol. Med. 2016, 107, 77–89. [Google Scholar] [CrossRef]
  69. Adav, S.S.; Sze, S.K. Hypoxia-induced degenerative protein modifications associated with aging and age-associated disorders. Aging Dis. 2020, 11, 341–364. [Google Scholar] [CrossRef] [Green Version]
  70. Li, S.Y.; Gomelsky, M.; Duan, J.; Zhang, Z.; Gomelsky, L.; Zhang, X.; Epstein, P.N.; Ren, J. Overexpression of aldehyde dehydrogenase-2 (ALDH2) transgene prevents acetaldehyde-induced cell injury in human umbilical vein endothelial cells: Role of ERK and p38 mitogen-activated protein kinase. J. Mol. Cell. Cardiol. 2004, 40, 283–294. [Google Scholar] [CrossRef]
  71. Lee, J.Y.; Jung, G.Y.; Heo, H.J.; Yun, M.R.; Park, J.Y.; Bae, S.S.; Hong, K.W.; Lee, W.S.; Kim, C.D. 4-Hydroxynonenal induces vascular smooth muscle cell apoptosis through mitochondrial generation of reactive oxygen species. Toxicol. Lett. 2006, 166, 212–221. [Google Scholar] [CrossRef] [PubMed]
  72. Yalcinkaya, T.; Uzilday, B.; Ozgur, R.; Turkan, I. The roles of reactive carbonyl species in induction of antioxidant defence and ROS signalling in extreme halophytic model Eutrema parvulum and glycophytic model Arabidopsis thaliana. Environ. Exp. Bot. 2019, 160, 81–91. [Google Scholar] [CrossRef]
  73. Yun, M.R.; Park, H.M.; Seo, K.W.; Lee, S.J.; Im, D.S.; Kim, C.D. 5-Lipoxygenase plays an essential role in 4-HNE-enhanced ROS production in murine macrophages via activation of NADPH oxidase. Free Radic. Res. 2010, 44, 742–750. [Google Scholar] [CrossRef]
  74. Endo, J.; Sano, M.; Katayama, T.; Hishiki, T.; Shinmur, K.; Morizane, S.; Matsuhash, T.; Katsumata, Y.; Zhang, Y.; Ito, H.; et al. Metabolic remodeling induced by mitochondrial aldehyde stress stimulates tolerance to oxidative stress in the heart. Circ. Res. 2009, 105, 1118–1127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Zhai, X.X.; Zhang, Z.X.; Liu, W.W.; Liu, B.S.; Zhang, R.; Wang, W.J.; Zheng, W.; Xu, F.; Wang, J.L.; Chen, Y.G. Protective effect of ALDH2 against cyclophosphamide-induced acute hepatotoxicity via attenuating oxidative stress and reactive aldehydes. Biochem. Biophys. Res. Commun. 2018, 499, 93–98. [Google Scholar] [CrossRef]
  76. Manzo-Avalos, S.; Saavedra-Molina, A. Cellular and mitochondrial effects of alcohol consumption. Int. J. Environ. Res. Public Health 2010, 7, 4281–4304. [Google Scholar] [CrossRef] [Green Version]
  77. Shi, L.; Tu, B.P. Acetyl-CoA and the regulation of metabolism: Mechanisms and consequences. Curr. Opin. Cell Biol. 2015, 33, 125–131. [Google Scholar] [CrossRef] [Green Version]
  78. Zhang, Y.M.; Ren, J. ALDH2 in alcoholic heart diseases: Molecular mechanism and clinical implications. Pharmacol. Ther. 2011, 132, 86–95. [Google Scholar] [CrossRef] [Green Version]
  79. Siu, G.M.; Draper, H.H. Metabolism of malonaldehyde in vivo and in vitro. Lipids 1982, 17, 349–355. [Google Scholar] [CrossRef]
  80. Grune, T.; Siems, W.; Kowalewski, J.; Zollner, H.; Esterbauer, H. Identification of metabolic pathways of the lipid peroxidation product 4-hydroxynonenal by enterocytes of rat small intestine. Biochem. Int. 1991, 25, 963–971. [Google Scholar] [CrossRef]
  81. Zhong, H.Q.; Yin, H.Y. Role of lipid peroxidation derived 4-hydroxynonenal (4-HNE) in cancer: Focusing on mitochondria. Redox Biol. 2015, 4, 193–199. [Google Scholar] [CrossRef] [Green Version]
  82. Lv, L.Z.; Ye, W.J.; Song, P.Y.; Chen, Y.B.; Yang, J.; Zhang, C.M.; Chen, X.P.; Luo, F.Y. Relationship between ALDH2 genotype and in-stent restenosis in Chinese Han patients after percutaneous coronary intervention. BMC Cardiovasc. Disord. 2019, 19, 176. [Google Scholar] [CrossRef] [PubMed]
  83. Braun, T.; Bober, E.; Singh, S.; Agarwal, D.P.; Goedde, H.W. Evidence for a signal peptide at the amino-terminal end of human mitochondrial aldehyde dehydrogenase. FEBS Lett. 1987, 215, 233–236. [Google Scholar] [CrossRef] [Green Version]
  84. Ni, L.; Zhou, J.Z.; Hurley, T.D.; Weiner, H. Human liver mitochondrial aldehyde dehydrogenase: Three-dimensional structure and the restoration of solubility and activity of chimeric forms. Protein Sci. 2010, 8, 2784–2790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Ji, E.D.; Jiao, T.T.; Shen, Y.L.; Xu, Y.J.; Sun, Y.Q.; Cai, Z.C.; Zhang, Q.; Li, J.M. Molecular mechanism of HSF1-upregulated ALDH2 by PKC in ameliorating pressure overload-induced heart failure in mice. BioMed Res. Int. 2020, 2020, 3481623. [Google Scholar] [CrossRef]
  86. Hsu, L.C.; Bendel, R.E.; Yoshida, A. Genomic structure of the human mitochondrial aldehyde dehydrogenase gene. Genomics 1988, 2, 57–65. [Google Scholar] [CrossRef]
  87. Weiner, H.; Wei, B.X.; Zhou, J.Z. Subunit communication in tetrameric class 2 human liver aldehyde dehydrogenase as the basis for half-of-the-site reactivity and the dominance of the oriental subunit in a heterotetramer. Chem. Biol. Interact. 2001, 130–132, 47–56. [Google Scholar] [CrossRef]
  88. Gong, D.X.; Zhang, H.; Hu, S.S. Mitochondrial aldehyde dehydrogenase 2 activation and cardioprotection. J. Mol. Cell. Cardiol. 2013, 55, 58–63. [Google Scholar] [CrossRef]
  89. Steinmetz, C.G.; Xie, P.G.; Weiner, H.; Hurley, T.D. Structure of mitochondrial aldehyde dehydrogenase: The genetic component of ethanol aversion. Structure 1997, 5, 701–711. [Google Scholar] [CrossRef]
  90. Larson, H.N.; Zhou, J.Z.; Chen, Z.Q.; Stamler, J.S.; Weiner, H.; Hurley, T.D. Structural and functional consequences of coenzyme binding to the inactive Asian variant of mitochondrial aldehyde dehydrogenase. J. Biol. Chem. 2007, 282, 12940–12950. [Google Scholar] [CrossRef] [Green Version]
  91. González-Segura, L.; Ho, K.K.; Perez-Miller, S.; Weine, H.; Hurley, T.D. Catalytic contribution of threonine 244 in human ALDH2. Chem. Biol. Interact. 2013, 202, 32–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Moore, S.A.; Baker, H.M.; Blythe, T.J.; Kitson, K.E.; Kitson, T.M.; Baker, E.N. Sheep liver cytosolic aldehyde dehydrogenase: The structure reveals the basis for the retinal specificity of class 1 aldehyde dehydrogenases. Structure 1998, 6, 1541–1551. [Google Scholar] [CrossRef] [Green Version]
  93. Perez-Miller, S.J.; Hurley, T.D. Coenzyme isomerization is integral to catalysis in aldehyde dehydrogenase. Biochemistry 2003, 42, 7100–7109. [Google Scholar] [CrossRef]
  94. Wenzl, M.V.; Beretta, M.; Gorren, A.C.F.; Zeller, A.; Baral, P.K.; Gruber, K.; Russwurm, M.; Koesling, D.; Schmidt, K.; Mayer, B. Role of the general base Glu-268 in nitroglycerin bioactivation and superoxide formation by aldehyde dehydrogenase-2. J. Biol. Chem. 2009, 284, 19878–19886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Li, R.; Zhao, Z.H.; Sun, M.Y.; Luo, J.C.; Xiao, Y.C. ALDH2 gene polymorphism in different types of cancers and its clinical significance. Life Sci. 2016, 147, 59–66. [Google Scholar] [CrossRef]
  96. Yang, L.Y.; Li, L.L.; Tang, H.Y.; Ma, T.B.; Li, Y.L.; Zhang, X.W.; Shi, X.L.; Liu, J.Y. Alcohol-aggravated episodic pain in humans with SCN11A mutation and ALDH2 polymorphism. Pain 2020, 161, 1470–1482. [Google Scholar] [CrossRef] [PubMed]
  97. Sasaki, T.; Nishimoto, Y.; Hirata, T.; Abe, Y.; Takebayashi, T.; Arai, Y. ALDH2 p.E504K variation and sex are major factors associated with current and quitting alcohol drinking in Japanese Oldest Old. Genes 2021, 12, 799. [Google Scholar] [CrossRef]
  98. Wang, X.P.; Sheikh, S.; Saigal, D.; Robinson, L.; Weiner, H. Heterotetramers of human liver mitochondrial (class 2) aldehyde dehydrogenase expressed in Escherichia coli: A model to study the heterotetramers expected to be found in Oriental people. J. Biol. Chem. 1996, 271, 31172–31178. [Google Scholar] [CrossRef] [Green Version]
  99. Lai, C.L.; Yao, C.T.; Chau, G.Y.; Yang, L.F.; Kuo, T.Y.; Chiang, C.P.; Yin, S.J. Dominance of the inactive Asian variant over activity and protein contents of mitochondrial aldehyde dehydrogenase 2 in human liver. Alcohol. Clin. Exp. Res. 2014, 38, 44–50. [Google Scholar] [CrossRef]
  100. Zuo, W.; Zhan, Z.Y.; Ma, L.; Bai, W.; Zeng, S.G. Effect of ALDH2 polymorphism on cancer risk in Asians. Medicine 2019, 98, e14855. [Google Scholar] [CrossRef]
  101. Idewaki, Y.; Iwase, M.; Fujii, H.; Ohkuma, T.; Ide, H.; Kaizu, S.; Jodai, T.; Kikuchi, Y.; Hirano, A.; Nakamura, U.; et al. Association of genetically determined aldehyde dehydrogenase 2 activity with diabetic complications in relation to alcohol consumption in japanese patients with type 2 diabetes mellitus: The fukuoka diabetes registry. PLoS ONE 2015, 10, e0143288. [Google Scholar] [CrossRef]
  102. Zhong, S.S.; Li, L.X.; Zhang, Y.L.; Zhang, L.L.; Lu, J.H.; Guo, S.Y.; Liang, N.N.; Ge, J.; Zhu, M.J.; Tao, Y.Z.; et al. Acetaldehyde dehydrogenase 2 interactions with LDLR and AMPK regulate foam cell formation. J. Clin. Investig. 2019, 129, 252–267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Ohta, S.; Ohsawa, I.; Kamino, K.; Ando, F.; Shimokata, H. Mitochondrial ALDH2 deficiency as an oxidative stress. Ann. New York Acad. Sci. 2004, 1011, 36–44. [Google Scholar] [CrossRef] [PubMed]
  104. Gross, E.R.; Zambelli, V.O.; Small, B.A.; Ferreira, J.C.B.; Chen, C.-H.; Mochly-Rosen, D. A personalized medicine approach for Asian Americans with the aldehyde dehydrogenase 2*2 variant. Annu. Rev. Pharmacol. Toxicol. 2015, 55, 107–127. [Google Scholar] [CrossRef] [Green Version]
  105. Chen, C.H.; Ferreira, J.C.B.; Gross, E.R.; Mochly-Rosen, D. Targeting aldehyde dehydrogenase 2: New therapeutic opportunities. Physiol. Rev. 2014, 94, 1–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Vasiliou, V.; Pappa, A.; Petersen, D.R. Role of aldehyde dehydrogenases in endogenous and xenobiotic metabolism. Chem.-Biol. Interact. 2000, 129, 1–19. [Google Scholar] [CrossRef]
  107. Wen, Y.Q.; Wang, X.P.; Xiao, S.Y.; Wang, Y.J. Ectopic expression of VpALDH2B4, a novel aldehyde dehydrogenase gene from Chinese wild grapevine (Vitis pseudoreticulata), enhances resistance to mildew pathogens and salt stress in Arabidopsis. Planta 2012, 236, 525–539. [Google Scholar] [CrossRef]
  108. Khan, M.; Qiao, F.; Kumar, P.; Islam, S.M.T.; Singh, A.K.; Won, J.; Sing, I. Neuroprotective effects of Alda-1 mitigate spinal cord injury in mice: Involvement of Alda-1-induced ALDH2 activation-mediated suppression of reactive aldehyde mechanisms. Neural Regen. Res. 2021, 17, 185–193. [Google Scholar] [CrossRef]
  109. Kim, J.; Chen, C.H.; Yang, J.Y.; Mochly-Rosen, D. Aldehyde dehydrogenase 2*2 knock-in mice show increased reactive oxygen species production in response to cisplatin treatment. J. Biomed. Sci. 2017, 24, 33. [Google Scholar] [CrossRef] [Green Version]
  110. Yuan Hou, J.; Xiong Zhong, Z.; Ting Deng, Q.; Dong Liu, S.; Fang Lin, L. Association between the polymorphism of aldehyde dehydrogenase 2 gene and cerebral infarction in a Hakka population in southern China. Biochem. Genet. 2020, 58, 322–334. [Google Scholar] [CrossRef]
  111. Yuan, Q.H.; Cao, S.C.; Dong, Q.Q.; Wang, Z.; Xu, Y.S.; Han, Q.; Ma, J.J.; Wei, S.J.; Pang, J.J.; Yang, F.H.; et al. ALDH2 activation inhibited cardiac fibroblast-to-myofibroblast transformation via the TGF-β1/Smad signaling pathway. J. Cardiovasc. Pharmacol. 2019, 73, 9. [Google Scholar] [CrossRef] [PubMed]
  112. Zhang, P.; Xu, D.L.; Wang, S.J.; Fu, H.; Wang, K.Q.; Zou, Y.Z.; Sun, A.J.; Ge, J.B. Inhibition of aldehyde dehydrogenase 2 activity enhances antimycin-induced rat cardiomyocytes apoptosis through activation of MAPK signaling pathway. Biomed. Pharmacother. 2011, 65, 590–593. [Google Scholar] [CrossRef] [PubMed]
  113. Yu, Y.; Jia, X.J.; Zong, Q.F.; Zhang, G.J.; Ye, H.W.; Hu, J.; Gao, Q.; Guan, S.D. Remote ischemic postconditioning protects the heart by upregulating ALDH2 expression levels through the PI3K/Akt signaling pathway. Mol. Med. Rep. 2014, 10, 536–542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Kwon, H.J.; Won, Y.S.; Park, O.; Chang, B.; Duryee, M.J.; Thiele, G.E.; Matsumoto, A.; Singh, S.; Abdelmegeed, M.A.; Song, B.J.; et al. Aldehyde dehydrogenase 2 deficiency ameliorates alcoholic fatty liver but worsens liver inflammation and fibrosis in mice. Hepatology 2014, 60, 146–157. [Google Scholar] [CrossRef] [Green Version]
  115. Ma, X.; Luo, Q.; Zhu, H.; Liu, X.; Dong, Z.; Zhang, K.; Zou, Y.; Wu, J.; Ge, J.; Sun, A. Aldehyde dehydrogenase 2 activation ameliorates CCl4-induced chronic liver fibrosis in mice by up-regulating Nrf2/HO-1 antioxidant pathway. J. Cell. Mol. Med. 2017, 22, 3965–3978. [Google Scholar] [CrossRef]
  116. D’Souza, Y.; Elharram, A.; Soon-Shiong, R.; Andrew, R.D.; Bennett, B.M. Characterization of Aldh2−/− mice as an age-related model of cognitive impairment and Alzheimer’s disease. Mol. Brain 2015, 8, 27. [Google Scholar] [CrossRef]
  117. Ghoweria, A.O.; Gagolewiczb, P.; Fraziera, H.N.; Ganta, J.C.; Andrewb, R.D.; Bennettb, B.M.; Thibault, O. Neuronal calcium imaging, excitability, and plasticity changes in the aldh2−/− mouse model of sporadic Alzheimer’s disease. J. Alzheimer’s Dis. 2020, 77, 1623–1637. [Google Scholar] [CrossRef]
  118. Grünblatt, E.; Riederer, P. Aldehyde dehydrogenase (ALDH) in Alzheimer’s and Parkinson’s disease. J. Neural Transm. 2016, 123, 83–90. [Google Scholar] [CrossRef]
  119. Yu, R.L.; Tu, S.C.; Wu, R.M.; Lu, P.A.; Tan, C.H. Interactions of COMT and ALDH2 genetic polymorphisms on symptoms of parkinson’s disease. Brain Sci. 2021, 11, 361. [Google Scholar] [CrossRef]
  120. Xue, L.; Xue, L.; Yang, F.H.; Han, Z.Q.; Cui, S.M.; Dai, S.; Xu, F.; Zhang, C.X.; Wang, X.P.; Pang, J.J.; et al. ALDH2 mediates the dose-response protection of chronic ethanol against endothelial senescence through SIRT1/p53 pathway. Biochem. Biophys. Res. Commun. 2018, 504, 777–783. [Google Scholar] [CrossRef]
  121. Hu, J.F.; Wang, H.X.; Li, H.H.; Hu, J.; Yu, Y.; Gao, Q. Inhibition of ALDH2 expression aggravates renal injury in a rat sepsis syndrome model. Exp. Ther. Med. 2017, 14, 2249–2254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Zhu, Q.K.; He, G.Z.; Wang, J.; Wang, Y.K.; Chen, W. Pretreatment with the ALDH2 agonist Alda-1 reduces intestinal injury induced by ischaemia and reperfusion in mice. Clin. Sci. 2017, 131, 1123–1136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Henderson, N.C.; Rieder, F.; Wynn, T.A. Fibrosis: From mechanisms to medicines. Nature 2020, 587, 555–566. [Google Scholar] [CrossRef] [PubMed]
  124. Santibañez, J.F.; Quintanilla, M.; Bernabeu, C. TGF-β/TGF-β receptor system and its role in physiological and pathological conditions. Clin. Sci. 2011, 121, 233–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Rauchman, M.; Griggs, D. Emerging strategies to disrupt the central TGF-β axis in kidney fibrosis. Transl. Res. 2019, 209, 90–104. [Google Scholar] [CrossRef]
  126. Akhurst, R.J. Targeting TGF-β signaling for therapeutic gain. Cold Spring Harb. Perspect. Biol. 2017, 9, a022301. [Google Scholar] [CrossRef] [Green Version]
  127. Fernandez, I.E.; Eickelberg, O. The Impact of TGF-β on lung fibrosis. Proc. Am. Thorac. Soc. 2012, 9, 111–116. [Google Scholar] [CrossRef]
  128. Gu, X.Y.; Fang, T.T.; Kang, P.F.; Hu, J.F.; Yu, Y.; Li, Z.H.; Cheng, X.Y.; Gao, Q. Effect of ALDH2 on high glucose-induced cardiac fibroblast oxidative stress, apoptosis, and fibrosis. Oxid. Med. Cell. Longev. 2017, 2017, 9257967. [Google Scholar] [CrossRef] [Green Version]
  129. Zhao, X.J.; Hua, Y.; Chen, H.M.; Yang, H.Y.; Zhang, T.; Huang, G.Q.; Fan, H.J.; Tan, Z.B.; Huang, X.F.; Liu, B.; et al. Aldehyde dehydrogenase-2 protects against myocardial infarction-related cardiac fibrosis through modulation of the Wnt/β-catenin signaling pathway. Ther. Clin. Risk Manag. 2015, 11, 1371–1381. [Google Scholar] [CrossRef] [Green Version]
  130. Gu, L.; Zhu, Y.J.; Yang, X.; Guo, Z.J.; Xu, W.B.; Tian, X.L. Effect of TGF-β/Smad signaling pathway on lung myofibroblast differentiation. Acta Pharmacol. Sin. 2007, 28, 382–391. [Google Scholar] [CrossRef] [Green Version]
  131. Yu, H.Y.; Königshoff, M.; Jayachandran, A.; Handley, D.; Seeger, W.; Kaminski, N.; Eickelberg, O. Transgelin is a direct target of TGF-β/Smad3-dependent epithelial cell migration in lung fibrosis. FASEB J. 2008, 22, 1778–1789. [Google Scholar] [CrossRef] [PubMed]
  132. Cheng, X.Y.; Gao, W.K.; Dang, Y.Y.; Liu, X.; Li, Y.J.; Peng, X.; Ye, X.Y. Both ERK/MAPK and TGF-Beta/Smad signaling pathways play a role in the kidney fbrosis of diabetic mice accelerated by blood glucose fluctuation. J. Diabetes Res. 2013, 2013, 463740. [Google Scholar] [CrossRef] [Green Version]
  133. Meng, X.M.; Huang, X.R.; Xiao, J.; Chung, A.C.K.; Qin, W.; Chen, H.Y.; Lan, H.Y. Disruption of Smad4 impairs TGF-β/Smad3 and Smad7 transcriptional regulation during renal inflammation and fibrosis in vivo and in vitro. Kidney Int. 2012, 81, 266–279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Harris, W.T.; Kelly, D.R.; Zhou, Y.; Wang, D.; Macewen, M.; Hagood, J.S.; Clancy, J.P.; Ambalavanan, N.; Sorscher, E.J. Myofibroblast differentiation and enhanced Tgf-B signaling in cystic fibrosis lung disease. PLoS ONE 2013, 8, e70196. [Google Scholar] [CrossRef]
  135. Xiao, Z.C.; Zhang, J.; Peng, X.G.; Dong, Y.J.; Jia, L.X.; Li, H.H.; Du, J. The Notch γ-secretase inhibitor ameliorates kidney fibrosis via inhibition of TGF-β/Smad2/3 signaling pathway activation. Int. J. Biochem. Cell Biol. 2014, 55, 65–71. [Google Scholar] [CrossRef]
  136. Chu, H.Y.; Shi, Y.; Jiang, S.; Zhong, Q.C.; Zhao, Y.Q.; Liu, Q.M.; Ma, Y.Y.; Shi, X.G.; Ding, W.F.; Zhou, X.D.; et al. Treatment effects of the traditional Chinese medicine Shenks in bleomycin-induced lung fibrosis through regulation of TGF-beta/Smad3 signaling and oxidative stress. Sci. Rep. 2017, 7, 2252. [Google Scholar] [CrossRef] [PubMed]
  137. Louzada, R.A.; Corre, R.; Hassani, R.A.E.; Meziani, L.; Jaillet, M.; Cazes, A.; Crestani, B.; Deutsch, E.; Dupuy, C. NADPH oxidase DUOX1 sustains TGF-β1 signalling and promotes lung fibrosis. Eur. Respir. J. 2020, 57, 1901949. [Google Scholar] [CrossRef] [PubMed]
  138. Donepudi, M.; Grutter, M.G. Structure and zymogen activation of caspases. Biophys. Chem. 2002, 101–102, 145–153. [Google Scholar] [CrossRef]
  139. Chakrabarty, S.; Verhelst, S.H.L. Controlled inhibition of apoptosis by photoactivatable caspase inhibitors. Cell Chem. Biol. 2020, 27, 1434–1440.e10. [Google Scholar] [CrossRef]
  140. Li, D.R.; Chen, J.; Ai, Y.W.; Gu, X.Q.; Li, L.; Che, D.; Jiang, Z.D.; Li, L.; Chen, S.; Huang, H.W.; et al. Estrogen-related hormones induce apoptosis by stabilizing schlafen-12 protein turnover. Mol. Cell 2019, 75, 1103–1116. [Google Scholar] [CrossRef]
  141. Torcia, M.; Chiara, G.D.; Nencioni, L.; Ammendola, S.; Labardi, D.; Lucibello, M.; Rosini, P.; Marlier, L.N.J.L.; Bonini, P.; Sbarba, P.D.; et al. Nerve growth factor inhibits apoptosis in memory B lymphocytes via inactivation of p38 MAPK, prevention of Bcl-2 phosphorylation, and cytochrome c release. J. Biol. Chem. 2001, 276, 39027–39036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Wang, D.D.; Weng, Q.J.; Zhang, L.; He, Q.J.; Yang, B. VEGF and Bcl-2 interact via MAPKs signaling pathway in the response to hypoxia in neuroblastoma. Cell. Mol. Neurobiol. 2009, 29, 391–401. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, Y.Y.; Wang, R.; Wang, Y.J.; Peng, R.Q.; Wu, Y.; Yuan, Y.F. Ginkgo biloba extract mitigates liver fibrosis and apoptosis by regulating p38 MAPK, NF-κB/IκBα, and Bcl-2/Bax signaling. Drug Des. Dev. Ther. 2015, 9, 6303–6317. [Google Scholar] [CrossRef] [Green Version]
  144. Andreka, G.; Vertesaljai, M.; Szantho, G.; Font, G.; Piroth, Z.; Fontos, G.; Juhasz, E.D.; Szekely, L.; Szelid, Z.; Turner, M.S.; et al. Remote ischaemic postconditioning protects the heart during acute myocardial infarction in pigs. Heart 2007, 93, 749–752. [Google Scholar] [CrossRef] [PubMed]
  145. Li, S.Y.; Gilbert, S.A.B.; Li, Q.; Ren, J. Aldehyde dehydrogenase-2 (ALDH2) ameliorates chronic alcohol ingestion-induced myocardial insulin resistance and endoplasmic reticulum stress. J. Mol. Cell. Cardiol. 2009, 47, 247–255. [Google Scholar] [CrossRef] [Green Version]
  146. Zhong, Z.B.; Hu, Q.C.; Fu, Z.; Wang, R.; Xiong, Y.; Zhang, Y.; Liu, Z.Z.; Wang, Y.F.; Ye, Q.F. Increased expression of aldehyde dehydrogenase 2 reduces renal cell apoptosis during ischemia/reperfusion injury after hypothermic machine perfusion. Artif. Organs 2016, 40, 596–603. [Google Scholar] [CrossRef]
  147. Cao, S.C.; Bian, Y.; Zhou, X.; Yuan, Q.H.; Wei, S.J.; Xue, L.; Yang, F.H.; Dong, Q.Q.; Wang, W.J.; Zheng, B.Y.; et al. A small-molecule activator of mitochondrial aldehyde dehydrogenase 2 reduces the severity of cerulein-induced acute pancreatitis. Biochem. Biophys. Res. Commun. 2019, 522, 518–524. [Google Scholar] [CrossRef]
  148. Shen, C.; Wang, C.; Han, S.S.; Wang, Z.J.; Dong, Z.; Zhao, X.N.; Wang, P.; Zhu, H.; Sun, X.L.; Ma, X.; et al. Aldehyde dehydrogenase 2 deficiency negates chronic low-to-moderate alcohol consumption-induced cardioprotecion possibly via ROS-dependent apoptosis and RIP1/RIP3/MLKL-mediated necroptosis. Biochim. Biophys. Acta 2017, 1863, 1912–1918. [Google Scholar] [CrossRef]
  149. Xu, D.; Guthrie, J.R.; Mabry, S.; Sack, T.M.; Truog, W.E. Mitochondrial aldehyde dehydrogenase attenuates hyperoxia-induced cell death through activation of ERK/MAPK and PI3K-Akt pathways in lung epithelial cells. Am. J. Physiol. Lung Cell. Mol. Physiol. 2016, 291, L966–L975. [Google Scholar] [CrossRef]
  150. Kim, G.-J.; Song, D.H.; Yoo, H.S.; Chung, K.-H.; Lee, K.J.; An, J.H. Hederagenin supplementation alleviates the pro-inflammatory and apoptotic response to alcohol in rats. Nutrients 2017, 9, 41. [Google Scholar] [CrossRef] [Green Version]
  151. Tian, X.L.; Li, Y. Endothelial cell senescence and age-related vascular diseases. J. Genet. Genom. 2014, 41, 485–495. [Google Scholar] [CrossRef] [PubMed]
  152. Yan, J.H.; Wang, J.L.; Huang, H.J.; Huang, Y.; Mi, T.; Zhang, C.T.; Zhang, L. Fibroblast growth factor 21 delayed endothelial replicative senescence and protected cells from H2O2-induced premature senescence through SIRT1. Am. J. Transl. Res. 2017, 9, 4492–4501. [Google Scholar] [PubMed]
  153. Koc, A.; Gasch, A.P.; Rutherford, J.C.; Kim, H.Y.; Gladyshev, V.N. Methionine sulfoxide reductase regulation of yeast lifespan reveals reactive oxygen species-dependent and -independent components of aging. Proc. Natl. Acad. Sci. USA 2004, 101, 7999–8004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Nannelli, G.; Ziche, M.; Donnini, S.; Morbidelli, L. Endothelial aldehyde dehydrogenase 2 as a target to maintain vascular wellness and function in ageing. Biomedicines 2020, 8, 4. [Google Scholar] [CrossRef] [Green Version]
  155. Gu, C.H.; Xing, Y.; Jiang, L.; Chen, M.; Xu, M.; Yin, Y.; Li, C.; Yang, Z.; Yu, L.; Ma, H. Impaired cardiac SIRT1 activity by carbonyl stress contributes to aging-related ischemic intolerance. PLoS ONE 2013, 8, e74050. [Google Scholar] [CrossRef] [Green Version]
  156. Haigis, M.C.; Sinclair, D.A. Mammalian sirtuins: Biological insights and disease relevance. Annu. Rev. Pathol. Mech. Dis. 2010, 5, 253–295. [Google Scholar] [CrossRef] [Green Version]
  157. Prola, A.; Silva, J.P.D.; Guilbert, A.; Lecru, L.; Piquereau, J.; Ribeiro, M.; Mateo, P.; Gressette, M.; Fortin, D.; Boursier, C.; et al. SIRT1 protects the heart from ER stress-induced cell death through eIF2α deacetylation. Cell Death Differ. 2016, 24, 343–356. [Google Scholar] [CrossRef] [Green Version]
  158. Xue, L.; Cui, S.M.; Cui, Z.Q.; Yang, F.H.; Pang, J.J.; Xu, F.; Chen, Y.G. ALDH2: A new protector against age-independent myocardial senescence. Int. J. Cardiol. 2016, 210, 38–40. [Google Scholar] [CrossRef]
  159. Nannelli, G.; Terzuoli, E.; Giorgio, V.; Donnini, S.; Lupetti, P.; Giachetti, A.; Bernardi, P.; Ziche, M. ALDH2 activity reduces mitochondrial oxygen reserve capacity in endothelial cells and induces senescence properties. Oxid. Med. Cell. Longev. 2018, 2018, 9765027. [Google Scholar] [CrossRef] [Green Version]
  160. Zhu, H.; Wang, Z.; Dong, Z.; Wang, C.; Cao, Q.; Fan, F.; Zhao, J.J.; Liu, X.W.; Yuan, M.; Sun, X.L.; et al. Aldehyde dehydrogenase 2 deficiency promotes atherosclerotic plaque instability through accelerating mitochondrial ROS-mediated vascular smooth muscle cell senescence. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 2018, 1865, 1782–1792. [Google Scholar] [CrossRef]
  161. Xue, L.; Zhu, W.Y.; Yang, F.H.; Dai, S.; Han, Z.Q.; Xu, F.; Guo, P.; Chen, Y.G. Appropriate dose of ethanol exerts anti-senescence and anti-atherosclerosis protective effects by activating ALDH2. Biochem. Biophys. Res. Commun. 2019, 512, 319–325. [Google Scholar] [CrossRef] [PubMed]
  162. Bartoli-Leonard, F.; Saddic, L.; Aikawa, E. Double-edged sword of ALDH2 mutations: One polymorphism can both benefit and harm the cardiovascular system. Eur. Heart J. 2020, 41, 2453–2455. [Google Scholar] [CrossRef] [PubMed]
  163. Chen, C.H.; Joshi, A.U.; Mochly-Rosen, D. The role of mitochondrial aldehyde dehydrogenase 2 (ALDH2) in neuropathology and neurodegeneration. Acta Neurol. Taiwanica 2016, 25, 111–123. [Google Scholar]
  164. Gęgotek, A.; Skrzydlewska, E. Biological effect of protein modifications by lipid peroxidation products. Chem. Phys. Lipids 2019, 221, 46–52. [Google Scholar] [CrossRef]
  165. Burke, W.J.; Kumar, V.B.; Pandey, N.; Panneton, W.M.; Gan, Q.; Franko, M.W.; O’Dell, M.; Li, S.W.; Pan, Y.; Chung, H.D.; et al. Aggregation of α-synuclein by DOPAL, the monoamine oxidase metabolite of dopamine. Acta Neuropathol. 2008, 115, 193–203. [Google Scholar] [CrossRef]
  166. Plotegher, N.; Berti, G.; Ferrari, E.; Tessari, I.; Zanetti, M.; Lunelli, L.; Greggio, E.; Bisaglia, M.; Veronesi, M.; Girotto, S. DOPAL derived alpha-synuclein oligomers impair synaptic vesicles physiological function. Sci. Rep. 2017, 7, 40699. [Google Scholar] [CrossRef]
  167. Doorn, J.A.; Florang, V.R.; Schamp, J.H.; Vanle, B.C. Aldehyde dehydrogenase inhibition generates a reactive dopamine metabolite autotoxic to dopamine neurons. Parkinsonism Relat. Disord. 2014, 20, S73–S75. [Google Scholar] [CrossRef] [Green Version]
  168. Goldsteina, D.S.; Kopin, I.J.; Sharabi, Y. Catecholamine autotoxicity. Implications for pharmacology and therapeutics of Parkinson disease and related disorders. Pharmacol. Ther. 2014, 144, 268–282. [Google Scholar] [CrossRef] [Green Version]
  169. Hou, G.J.; Chen, L.; Liu, G.; Li, L.; Yang, Y.; Yan, H.X.; Zhang, H.L.; Tang, J.; Yang, Y.C.; Lin, X.M.; et al. Aldehyde dehydrogenase-2 (ALDH2) opposes hepatocellular carcinoma progression by regulating AMP-activated protein kinase signaling in mice. Hepatology 2017, 65, 1628–1644. [Google Scholar] [CrossRef] [Green Version]
  170. Choi, J.Y.; Abel, J.; Neuhaus, T.; Ko, Y.; Harth, V.; Hamajima, N.; Tajima, K.; Yoo, K.Y.; Park, S.K.; Noh, D.Y.; et al. Role of alcohol and genetic polymorphisms of CYP2E1 and ALDH2 in breast cancer development. Pharmacogenetics 2003, 13, 67–72. [Google Scholar] [CrossRef]
  171. Eom, S.Y.; Zhang, Y.W.; Kim, S.H.; Choe, K.H.; Lee, K.Y.; Park, J.D.; Hong, Y.C.; Kim, Y.D.; Kang, J.W.; Kim, H. Influence of NQO1, ALDH2, and CYP2E1 genetic polymorphisms, smoking, and alcohol drinking on the risk of lung cancer in Koreans. Cancer Causes Control. 2009, 20, 137–145. [Google Scholar] [CrossRef] [PubMed]
  172. Hsiao, J.R.; Lee, W.T.; Ou, C.Y.; Huang, C.C.; Chang, C.C.; Tsai, S.T.; Chen, K.C.; Huang, J.S.; Wong, T.Y.; Lai, Y.H.; et al. Validation of alcohol flushing questionnaire to identify ALDH 2 status in a case–control study of head and neck cancer. Alcohol. Clin. Exp. Res. 2019, 43, 1225–1233. [Google Scholar] [CrossRef] [PubMed]
  173. Chen, X.; Legrand, A.J.; Cunniffe, S.; Hume, S.; Poletto, M.; Vaz, B.; Ramadan, K.; Yao, D.; Dianov, G.L. Interplay between base excision repair protein XRCC1 and ALDH2 predicts overall survival in lung and liver cancer patients. Cell. Oncol. 2018, 41, 527–539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Cui, R.; Kamatani, Y.; Takahashi, A.; Usami, M.; Hosono, N.; Kawaguchi, T.; Tsunoda, T.; Kamatani, N.; Kubo, M.; Nakamura, Y.; et al. Functional variants in ADH1B and ALDH2 coupled with alcohol and smoking synergistically enhance esophageal cancer risk. Gastroenterology 2009, 137, 1768–1775. [Google Scholar] [CrossRef] [PubMed]
  175. Miyasaka, K.; Hosoya, H.; Tanaka, Y.; Uegaki, S.; Kino, K.; Shimokata, H.; Kawanami, T.; Funakoshi, A. Association of aldehyde dehydrogenase 2 gene polymorphism with pancreatic cancer but not colon cancer. Geriatr. Gerontol. Int. 2010, 10, S120–S126. [Google Scholar] [CrossRef]
  176. Shin, C.M.; Kim, N.; Cho, S.-I.; Kim, J.S.; Jung, H.C.; Song, I.S. Association between alcohol intake and risk for gastric cancer with regard to ALDH2 genotype in the Korean population. Int. J. Epidemiol. 2011, 40, 1047–1055. [Google Scholar] [CrossRef] [Green Version]
  177. Yang, Y.; Huycke, M.; Herman, T.; Wang, X. Glutathione S-transferase alpha 4 induction by activator protein 1 in colorectal cancer. Oncogene 2016, 35, 5795–5806. [Google Scholar] [CrossRef]
  178. Chen, L.; Wu, M.; Ji, C.; Yuan, M.; Liu, C.; Yin, Q. Silencing transcription factor FOXM1 represses proliferation, migration, and invasion while inducing apoptosis of liver cancer stem cells by regulating the expression of ALDH2. IUBMB Life 2020, 72, 285–295. [Google Scholar] [CrossRef]
  179. Chen, C.H.; Budas, G.R.; Churchill, E.N.; Disatnik, M.H.; Hurley, T.D.; Mochly-Rosen, D. Activation of aldehyde dehydrogenase-2 reduces ischemic damage to the heart. Science 2008, 321, 1493–1495. [Google Scholar] [CrossRef] [Green Version]
  180. Budas, G.R.; Disatnik, M.H.; Chen, C.H.; Mochly-Rosen, D. Activation of aldehyde dehydrogenase 2 (ALDH2) confers cardioprotection in protein kinase C epsilon (PKCε) knockout mice. J. Mol. Cell. Cardiol. 2010, 48, 757–764. [Google Scholar] [CrossRef] [Green Version]
  181. Perez-Miller, S.; Younus, H.; Vanam, R.; Chen, C.H.; Mochly-Rosen, D.; Hurley, T.D. Alda-1 is an agonist and chemical chaperone for the common human aldehyde dehydrogenase 2 variant. Nat. Struct. Mol. Biol. 2010, 17, 159–165. [Google Scholar] [CrossRef] [Green Version]
  182. Yan, W.; Long, P.; Wei, D.; Yan, W.; Zheng, X.; Chen, G.; Wang, J.; Zhang, Z.; Chen, T.; Chen, M. Protection of retinal function and morphology in MNU-induced retinitis pigmentosa rats by ALDH2: An in-vivo study. BMC Ophthalmol. 2020, 20, 55. [Google Scholar] [CrossRef] [Green Version]
  183. Chen, C.H.; Sun, L.; Mochly-Rosen, D. Mitochondrial aldehyde dehydrogenase and cardiac diseases. Cardiovasc. Res. 2010, 88, 51–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Fu, S.H.; Zhang, H.F.; Yang, Z.B.; Li, T.B.; Liu, B.; Lou, Z.; Ma, Q.L.; Luo, X.J.; Peng, J. Alda-1 reduces cerebral ischemia/reperfusion injury in rat through clearance of reactive aldehydes. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2014, 387, 87–94. [Google Scholar] [CrossRef] [PubMed]
  185. Zhang, T.; Zhao, Q.; Ye, F.; Huang, C.Y.; Chen, W.M.; Huang, W.Q. Alda-1, an ALDH2 activator, protects against hepatic ischemia/reperfusion injury in rats via inhibition of oxidative stress. Free Radic Res. 2018, 52, 629–638. [Google Scholar] [CrossRef] [PubMed]
  186. He, M.S.; Long, P.; Yan, W.M.; Chen, T.; Guo, L.F.; Zhang, Z.M.; Wang, S.W. ALDH2 attenuates early-stage STZ-induced aged diabetic rats retinas damage via Sirt1/Nrf2 pathway. Life Sci. 2018, 215, 227–235. [Google Scholar] [CrossRef]
  187. Hua, T.F.; Yang, M.; Zhou, Y.Y.; Chen, L.M.; Wu, H.M.; Liu, R.Y. Alda-1 prevents pulmonary epithelial barrier dysfunction following severe hemorrhagic shock through clearance of reactive aldehydes. BioMed Res. Int. 2019, 2019, 2476252. [Google Scholar] [CrossRef] [Green Version]
  188. Wang, S.Y.; Wang, L.; Qin, X.; Turdi, S.; Sun, D.D.; Culver, B.; Reiter, R.J.; Wang, X.M.; Zhou, H.; Ren, J. ALDH2 contributes to melatonin-induced protection against APP/PS1 mutation-prompted cardiac anomalies through cGAS-STING-TBK1-mediated regulation of mitophagy. Signal Transduct. Target. Ther. 2020, 5, 1335–1347. [Google Scholar] [CrossRef]
  189. Chen, C.H.; Gray, M.O.; Mochly-Rosen, D. Cardioprotection from ischemia by a brief exposure to physiological levels of ethanol: Role of epsilon protein kinase C. Proc. Natl. Acad. Sci. USA 1999, 96, 12784–12789. [Google Scholar] [CrossRef] [Green Version]
  190. Lang, X.E.; Wang, X.; Zhang, K.R.; Lv, J.Y.; Jin, J.H.; Li, Q.S. Isoflurane preconditioning confers cardioprotection by cctivation of ALDH2. PLoS ONE 2013, 8, e52469. [Google Scholar] [CrossRef] [Green Version]
  191. Li, W.J.; Yin, L.; Sun, X.L.; Wu, J.; Dong, Z.; Hu, K.; Sun, A.J.; Ge, J.B. Alpha-lipoic acid protects against pressure overload-induced heart failure via ALDH2-dependent Nrf1-FUNDC1 signaling. Cell Death Dis. 2020, 11, 599. [Google Scholar] [CrossRef] [PubMed]
  192. Yan, Y.; Yang, J.Y.; Mou, Y.H.; Wang, L.H.; Zhou, Y.N.; Wu, C.F. Differences in the activities of resveratrol and ascorbic acid in protection of ethanol-induced oxidative DNA damage in human peripheral lymphocytes. Food Chem. Toxicol. 2012, 50, 168–174. [Google Scholar] [CrossRef] [PubMed]
  193. Wenzel, P.; Hink, U.; Oelze, M.; Schuppan, S.; Schaeuble, K.; Schildknecht, S.; Ho, K.K.; Weiner, H.; Bachschmid, M.; Münzel, T.; et al. Role of reduced lipoic acid in the redox regulation of mitochondrial aldehyde dehydrogenase (ALDH-2) activity. Implications for mitochondrial oxidative stress and nitrate tolerance. J. Biol. Chem. 2007, 282, 792–799. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Wang, J.; Wang, H.; Hao, P.; Xue, L.; Chen, Y. Inhibition of aldehyde dehydrogenase 2 by oxidative stress is associated with cardiac dysfunction in diabetic rats. Mol. Med. 2011, 17, 172–179. [Google Scholar] [CrossRef] [PubMed]
  195. He, L.; Liu, B.; Dai, Z.; Zhang, H.F.; Zhang, Y.S.; Luo, X.J.; Ma, Q.L.; Peng, J. Alpha lipoic acid protects heart against myocardial ischemia-reperfusion injury through a mechanism involving aldehyde dehydrogenase 2 activation. Eur. J. Pharmacol. 2012, 678, 32–38. [Google Scholar] [CrossRef]
  196. Zhang, H.; Xue, L.; Li, B.Y.; Zhang, Z.L.; Tao, S.S. Vitamin D protects against alcohol-induced liver cell injury within an NRF2-ALDH2 feedback loop. Mol. Nutr. Food Res. 2019, 63, 1801014. [Google Scholar] [CrossRef]
  197. Kobayashi, H.; Nakamura, S.; Sato, Y.; Kobayashi, T.; Miyamoto, K.; Oya, A.; Matsumoto, M.; Nakamura, M.; Kanaji, A.; Miyamoto, T. ALDH2 mutation promotes skeletal muscle atrophy in mice via accumulation of oxidative stress. Bone 2020, 142, 115739. [Google Scholar] [CrossRef]
  198. Gao, G.Y.; Li, D.J.; Keung, W.M. Synthesis of potential antidipsotropic isoflavones: Inhibitors of the mitochondrial monoamine oxidase-aldehyde dehydrogenase pathway. J. Med. Chem. 2001, 44, 3320–3328. [Google Scholar] [CrossRef]
  199. Lowe, E.D.; Gao, G.Y.; Johnson, L.N.; Keung, W.M. Structure of daidzin, a naturally occurring anti-alcohol-addiction agent, in complex with human mitochondrial aldehyde dehydrogenase. J. Med. Chem. 2008, 51, 4482–4487. [Google Scholar] [CrossRef]
  200. Choi, H.; Tostes, R.C.; Webb, R.C. Mitochondrial aldehyde dehydrogenase prevents ROS-induced vascular contraction in angiotensin II hypertensive mice. J. Am. Soc. Hypertens. 2011, 5, 154–160. [Google Scholar] [CrossRef] [Green Version]
  201. Rivera-Meza, M.; Vásquez, D.; Quintanilla, M.E.; Lagos, D.; Rojas, B.; Herrera-Marschitz, M.; Israel, Y. Activation of mitochondrial aldehyde dehydrogenase (ALDH2) by ALDA-1 reduces both the acquisition and maintenance of ethanol intake in rats: A dual mechanism? Neuropharmacology 2019, 146, 175–183. [Google Scholar] [CrossRef] [PubMed]
  202. Kitagawa, K.; Kawamoto, T.; Kunugita, N.; Tsukiyama, T.; Okamoto, K.; Yoshida, A.; Nakayama, K.; Nakayama, K.-i. Aldehyde dehydrogenase (ALDH) 2 associates with oxidation of methoxyacetaldehyde; in vitro analysis with liver subcellular fraction derived from human and Aldh2 gene targeting mouse. FEBS Lett. 2000, 476, 306–311. [Google Scholar] [CrossRef] [Green Version]
  203. Omran, Z.; Sheikh, R.; Baothman, O.A.; Zamzami, M.A.; Alarjah, M. Repurposing disulfiram as an anti-obesity drug: Treating and preventing obesity in high-fat-fed rats. Diabetes Metab. Syndr. Obes. Targets Ther. 2020, 13, 1473–1480. [Google Scholar] [CrossRef] [PubMed]
  204. Roy, B.; Palaniyandi, S.S. A role for aldehyde dehydrogenase (ALDH) 2 in angiotensin II-mediated decrease in angiogenesis of coronary endothelial cells. Microvasc. Res. 2021, 135, 104133. [Google Scholar] [CrossRef]
  205. Song, B.J.; Abdelmegeed, M.A.; Yoo, S.H.; Kim, B.J.; Jo, S.A.; Jo, I.; Moon, K.H. Post-translational modifications of mitochondrial aldehyde dehydrogenase and biomedical implications. J. Proteom. 2011, 74, 2691–2702. [Google Scholar] [CrossRef] [Green Version]
  206. Churchill, E.N.; Disatnik, M.H.; Mochly-Rosen, D. Time-dependent and ethanol-induced cardiac protection from ischemia mediated by mitochondrial translocation of εPKC and activation of aldehyde dehydrogenase 2. J. Mol. Cell. Cardiol. 2009, 46, 278–284. [Google Scholar] [CrossRef] [Green Version]
  207. Lagranha, C.J.; Deschamps, A.; Aponte, A.; Steenbergen, C.; Murphy, E. Sex differences in the phosphorylation of mitochondrial proteins result in reduced production of reactive oxygen species and cardioprotection in females. Circ. Res. 2010, 106, 1681–1691. [Google Scholar] [CrossRef] [Green Version]
  208. Budas, G.R.; Churchill, E.N.; Disatnik, M.H.; Sun, L.; Mochly-Rosen, D. Mitochondrial import of PKCε is mediated by HSP90: A role in cardioprotection from ischaemia and reperfusion injury. Cardiovasc. Res. 2010, 88, 83–92. [Google Scholar] [CrossRef] [Green Version]
  209. Moon, K.H.; Lee, Y.M.; Song, B.J. Inhibition of hepatic mitochondrial aldehyde dehydrogenase by carbon tetrachloride through JNK-mediated phosphorylation. Free Radic. Biol. Med. 2010, 48, 391–398. [Google Scholar] [CrossRef] [Green Version]
  210. Harris, P.S.; Gomez, J.D.; Backos, D.S.; Fritz, K.S. Characterizing sirtuin 3 deacetylase affinity for aldehyde dehydrogenase 2. Chem. Res. Toxicol. 2017, 30, 785–793. [Google Scholar] [CrossRef] [Green Version]
  211. Xue, L.; Xu, F.; Meng, L.J.; Wei, S.J.; Wang, J.L.; Hao, P.P.; Bian, Y.; Zhang, Y.; Chen, Y.G. Acetylation-dependent regulation of mitochondrial ALDH2 activation by SIRT3 mediates acute ethanol-induced eNOS activation. FEBS Lett. 2011, 586, 137–142. [Google Scholar] [CrossRef] [PubMed]
  212. Lu, Z.P.; Bourdi, M.; Li, J.H.; Aponte, A.M.; Chen, Y.; Lombard, D.B.; Gucek, M.; Pohl, L.R.; Sack, M.N. SIRT3-dependent deacetylation exacerbates acetaminophen hepatotoxicity. EMBO Rep. 2011, 12, 840–846. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Moon, K.H.; Hood, B.L.; Kim, B.J.; Hardwick, J.P.; Conrads, T.P.; Veenstra, T.D.; Song, B.J. Inactivation of oxidized and S-nitrosylated mitochondrial proteins in alcoholic fatty liver of rats. Hepatology 2006, 44, 1218–1230. [Google Scholar] [CrossRef] [PubMed]
  214. Hao, L.Y.; Sun, Q.; Zhong, W.; Zhang, W.L.; Sun, X.G.; Zhou, Z.X. Mitochondria-targeted ubiquinone (MitoQ) enhances acetaldehyde clearance by reversing alcohol-induced posttranslational modification of aldehyde dehydrogenase 2: A molecular mechanism of protection against alcoholic liver disease. Redox Biol. 2018, 14, 626–636. [Google Scholar] [CrossRef]
  215. Moon, K.H.; Kim, B.J.; Song, B.J. Inhibition of mitochondrial aldehyde dehydrogenase by nitric oxide-mediated S-nitrosylation. FEBS Lett. 2005, 579, 6115–6120. [Google Scholar] [CrossRef] [Green Version]
  216. Moon, K.H.; Hood, B.L.; Mukhopadhyay, P.; Mohanraj, R.; Abdelmegeed, M.A.; Kwon, Y.I.; Conrads, T.P.; Veenstra, T.D.; Song, B.J.; Pacher, P. Oxidative inactivation of key mitochondrial proteins leads to dysfunction and injury in hepatic ischemia reperfusion. Gastroenterology 2008, 135, 1344–1357. [Google Scholar] [CrossRef] [Green Version]
  217. Oelz, M.; Knor, M.; Schell, R.; Kamuf, J.; Pautz, A.; Art, J.; Wenzel, P.; Münzel, T.; Kleinert, H.; Daiber, A. Regulation of human mitochondrial aldehyde dehydrogenase (ALDH-2) activity by electrophiles in vitro. J. Biol. Chem. 2011, 286, 8893–8900. [Google Scholar] [CrossRef] [Green Version]
  218. Latorre-Muro, P.; Baeza, J.; Armstrong, E.A.; Hurtado-Guerrero, R.; Corzana, F.; Wu, L.E.; Sinclair, D.A.; López-Buesa, P.; Carrodeguas, J.A.; Denu, J.M. Dynamic acetylation of phosphoenolpyruvate carboxykinase toggles enzyme activity between gluconeogenic and anaplerotic reactions. Mol. Cell 2018, 71, 718–732. [Google Scholar] [CrossRef] [Green Version]
  219. Wang, C.M.; Yang, W.H.; Liu, R.; Wang, L.; Yang, W.H. FOXP3 activates SUMO-conjugating UBC9 gene in MCF7 breast cancer cells. Int. J. Mol. Sci. 2018, 19, 2036. [Google Scholar] [CrossRef] [Green Version]
  220. Sun, F.X.; Suttapitugsakul, S.; Xiao, H.P.; Wu, R.H. Comprehensive analysis of protein glycation reveals its potential impacts on protein degradation and gene expression in human cells. J. Am. Soc. Mass Spectrom. 2019, 30, 2480–2490. [Google Scholar] [CrossRef]
  221. Valencia-Sánchez, M.I.; Ioannes, P.D.; Wang, M.; Truong, D.M.; Lee, R.; Armache, J.-P.; Boeke, J.D.; Armache, K.-J. Regulation of the Dot1 histone H3K79 methyltransferase by histone H4K16 acetylation. Science 2021, 371, eabc6663. [Google Scholar] [CrossRef]
  222. Chi, Z.X.; Byeon, H.E.; Seo, E.; Nguyen, Q.A.T.; Lee, W.; Jeong, Y.; Choi, J.; Pandey, D.; Berkowitz, D.E.; Kim, J.H.; et al. Histone deacetylase 6 inhibitor tubastatin A attenuates angiotensin II-induced hypertension by preventing cystathionine γ-lyase protein degradation. Pharmacol. Res. 2019, 146, 104281. [Google Scholar] [CrossRef]
  223. Patel, N.; Wang, J.H.; Shiozawa, K.; Jones, K.B.; Zhang, Y.F.; Prokop, J.W.; Davenport, G.G.; Nihira, N.T.; Hao, Z.Y.; Wong, D.; et al. HDAC2 regulates site-specific acetylation of MDM2 and its ubiquitination signaling in tumor suppression. Iscience 2019, 13, 43–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Zhang, L.; Li, D.Q. MORC2 regulates DNA damage response through a PARP1-dependent pathway. Nucleic Acids Res. 2019, 47, 8502–8520. [Google Scholar] [CrossRef] [PubMed]
  225. Chi, Z.X.; Byeon, H.E.; Seo, E.; Nguyen, Q.A.T.; Lee, W.; Jeong, Y.Y.; Choi, J.Y.; Pandey, D.; Berkowitz, D.E.; Kim, J.H.; et al. Post-translational modification analysis of Saccharomyces cerevisiae histone methylation enzymes reveals phosphorylation sites of regulatory potential. J. Biol. Chem. 2020, 296, 100192. [Google Scholar] [CrossRef]
  226. Zhou, F.; Liu, Q.Q.; Zhang, L.L.; Zhu, Q.; Wang, S.S.; Zhu, K.C.; Deng, R.Y.; Liu, Y.; Yuan, G.Y.; Wang, X.; et al. Selective inhibition of CBP/p300 HAT by A-485 results in suppression of lipogenesis and hepatic gluconeogenesis. Cell Death Dis. 2020, 11, 745. [Google Scholar] [CrossRef] [PubMed]
  227. Karaarslan, U.; İşgüder, R.; Bağ, Ö.; Kışla, M.; Ağın, H.; Ünal, N. Alpha lipoic acid intoxication, treatment and outcome. Clin. Toxicol. 2013, 51, 522. [Google Scholar] [CrossRef]
  228. Sarezky, D.; Raquib, A.R.; Dunaief, J.L.; Kim, B.J. Tolerability in the elderly population of high-dose alpha lipoic acid: A potential antioxidant therapy for the eye. Clin. Ophthalmol. 2016, 10, 1899–1903. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Pathway by which aldehyde dehydrogenase 2 (ALDH2) catalyzes aldehyde metabolism in ethanol metabolism. ALDH2 metabolizes acetaldehyde into acetate, metabolizes the malondialdehyde (MDA) into malonic acid (MOA) or acetaldehyde, and metabolizes the 4-hydroxy-2-nonenal (4-NHE) to generate 4-hydroxy-2-nonenoic acid (NHA) to reduce the accumulation of toxic aldehydes produced during metabolism. CHOL = cholesterol; FA = fatty acids; KB = ketone bodies.
Figure 1. Pathway by which aldehyde dehydrogenase 2 (ALDH2) catalyzes aldehyde metabolism in ethanol metabolism. ALDH2 metabolizes acetaldehyde into acetate, metabolizes the malondialdehyde (MDA) into malonic acid (MOA) or acetaldehyde, and metabolizes the 4-hydroxy-2-nonenal (4-NHE) to generate 4-hydroxy-2-nonenoic acid (NHA) to reduce the accumulation of toxic aldehydes produced during metabolism. CHOL = cholesterol; FA = fatty acids; KB = ketone bodies.
Ijms 23 02682 g001
Figure 3. The main regulatory mechanism of aldehyde dehydrogenase 2 (ALDH2) in certain OS-related physiological and pathological processes. ALDH2 activation inhibits the transformation of human cardiac fibroblasts (HCFs) into myofibroblasts via the transforming growth factor-β1 (TGF-β)/Drosophila mothers against decapentaplegic protein (Smad) signaling pathway to inhibit fibrosis; ALDH2 inhibits mitogen-activated protein kinase (MAPK) signaling pathways and upregulates the B-cell lymphoma-2 (Bcl-2)/Bcl-2–associated X protein (Bax) ratio to reduce the cysteinyl aspartate–specific proteinase 3 (caspase-3) level and inhibit apoptosis; ALDH2 can also inactivate inhibit 4-hydroxy-2-nonenal (4-HNE) accumulation and regulate sirtuin 1 (SIRT1)/p53-dependent aging; ALDH2 can eliminate the toxic metabolites of neurotransmitters (dopamine [DA], epinephrine [EPI], and norepinephrine [NE]) and can inhibit α-Synuclein (α-Syn) abnormal aggregation as well as Lewy bodies (LBs) formation, and reduce neurofibrillary tangles (NFTs) formed by hyperphosphorylated microtubule-associated protein tau, so as to protect neurons from the damage caused by 4-HNE and reactive oxygen species (ROS). (+) and (−) mainly reflect upregulation/activation or downregulation/inhibition, respectively, related to ALDH2 function. Red = plays a key role in the realization of corresponding physiological/pathological processes.
Figure 3. The main regulatory mechanism of aldehyde dehydrogenase 2 (ALDH2) in certain OS-related physiological and pathological processes. ALDH2 activation inhibits the transformation of human cardiac fibroblasts (HCFs) into myofibroblasts via the transforming growth factor-β1 (TGF-β)/Drosophila mothers against decapentaplegic protein (Smad) signaling pathway to inhibit fibrosis; ALDH2 inhibits mitogen-activated protein kinase (MAPK) signaling pathways and upregulates the B-cell lymphoma-2 (Bcl-2)/Bcl-2–associated X protein (Bax) ratio to reduce the cysteinyl aspartate–specific proteinase 3 (caspase-3) level and inhibit apoptosis; ALDH2 can also inactivate inhibit 4-hydroxy-2-nonenal (4-HNE) accumulation and regulate sirtuin 1 (SIRT1)/p53-dependent aging; ALDH2 can eliminate the toxic metabolites of neurotransmitters (dopamine [DA], epinephrine [EPI], and norepinephrine [NE]) and can inhibit α-Synuclein (α-Syn) abnormal aggregation as well as Lewy bodies (LBs) formation, and reduce neurofibrillary tangles (NFTs) formed by hyperphosphorylated microtubule-associated protein tau, so as to protect neurons from the damage caused by 4-HNE and reactive oxygen species (ROS). (+) and (−) mainly reflect upregulation/activation or downregulation/inhibition, respectively, related to ALDH2 function. Red = plays a key role in the realization of corresponding physiological/pathological processes.
Ijms 23 02682 g003
Figure 4. Effect of protein post-translational modifications (PTMs) on ALDH activity. Ethanol (acute/low-dose), isoflurane, heat shock factor 1 (HSF1), and melatonin can lead to phosphorylation and activation of aldehyde dehydrogenase 2 (ALDH2) by activating phosphatidylinositol protein kinase C epsilon (PKCε). ALDH2 can also be activated by its acetylation increased by the inactivation of sirtuin 3 (SIRT3) stimulated by ethanol (acute/low-dose). The phosphorylation of ALDH2 promoted by c-Jun NH2-terminal kinase (JNK) under carbon tetrachloride (CCl4) exposure as well as the S-nitrosylation and tyrosine nitration of ALDH2 stimulated by ethanol (chronic or excessive abuse), acetaminophen (N-acetyl-p-aminophenol [APAP]), nitric oxide (NO) donor, and ischemia/reperfusion (I/R) injury can inhibit ALDH2 activity. P = phosphorylation; Ac = acetylation; SNO = S-nitrosylation; NO2 = nitration.
Figure 4. Effect of protein post-translational modifications (PTMs) on ALDH activity. Ethanol (acute/low-dose), isoflurane, heat shock factor 1 (HSF1), and melatonin can lead to phosphorylation and activation of aldehyde dehydrogenase 2 (ALDH2) by activating phosphatidylinositol protein kinase C epsilon (PKCε). ALDH2 can also be activated by its acetylation increased by the inactivation of sirtuin 3 (SIRT3) stimulated by ethanol (acute/low-dose). The phosphorylation of ALDH2 promoted by c-Jun NH2-terminal kinase (JNK) under carbon tetrachloride (CCl4) exposure as well as the S-nitrosylation and tyrosine nitration of ALDH2 stimulated by ethanol (chronic or excessive abuse), acetaminophen (N-acetyl-p-aminophenol [APAP]), nitric oxide (NO) donor, and ischemia/reperfusion (I/R) injury can inhibit ALDH2 activity. P = phosphorylation; Ac = acetylation; SNO = S-nitrosylation; NO2 = nitration.
Ijms 23 02682 g004
Table 1. Human aldehyde dehydrogenase (ALDH) gene superfamily.
Table 1. Human aldehyde dehydrogenase (ALDH) gene superfamily.
Gene NameChromosomal LocationProtein Subcellular LocationExpressionSubstrate/Ref.
ALDH1A19q21.13Cytoplasm, cytosolLiver, duodenum, 18 other tissuesRetinaldehyde [19,20,21]; Acetaldehyde [22]
ALDH1A215q21.3Cytoplasm, cytosolEndometrium, testes, 8 other tissuesRetinaldehyde [23,24,25]
ALDH1A315q26.3Cytoplasm, cytosolProstate, bladder, 14 other tissuesRetinaldehyde [26,27]
ALDH1B19p13.1MitochondrionLiver, kidneys, 19 other tissuesAcetaldehyde [28,29]
ALDH1L13q21.3Cytoplasm, cytosolLiver, kidneys, 8 other tissues10-Formyltetrahydrofolate [30,31,32]
ALDH1L212q23.3MitochondrionPancreas, salivary gland, 22 other tissues10-Formyltetrahydrofolate [33,34]
ALDH212q24.12MitochondrionFat, liver, 19 other tissuesAcetaldehyde [35,36]; 4-Hydroxy-2-nonenal (4-HNE) [37,38]; Malondialdehyde (MDA) [39]
ALDH3A117p11.2Cytoplasm, cytosolEsophagus, stomach, skinBenzaldehyde [36,40,41]
ALDH3A217p11.2Endoplasmic reticulumSkin, adrenal glands, 22 other tissuesHexadecenal [42,43,44]
ALDH3B111q13.2Plasma membraneLungs, bone marrow, 17 other tissuesOctaldehyde [36]; Benzaldehyde [45]; 4-HNE [45,46]; Hexanal [45,46]; Nonanal [46]
ALDH3B211q13.2Lipid dropletSkin, esophagus, 3 other tissuesMedium-chain to long-chain aldehydes [47]
ALDH4A11p36.13MitochondrionKidneys, liver, 8 other tissuesGlutamate γ-semialdehyde [48,49,50,51]
ALDH5A16p22.3MitochondrionLiver, brain, 23 other tissuesSuccinic semialdehyde [52,53]; 4-HNE [53]
ALDH6A114q24.3MitochondrionKidneys, liver, 11 other tissuesMethylmalonate-semialdehyde [54]
ALDH7A15q23.2Cytoplasm, cytosol, mitochondrion, nucleusKidneys, liver, 24 other tissuesα-Aminoadipic-semialdehyde [55,56]
ALDH8A16q23.3Cytoplasm, cytosolLiver, kidneysRetinaldehyde [36,57]
ALDH9A11q24.1Cytoplasm, cytosolFat, thyroid gland, 25 other tissuesγ-Trimethylaminobutyraldehyde [58,59]
ALDH16A119q13.33MembraneSpleen, duodenum, 25 other tissuesUnknown, lacks measurable catalytic activity [60]
ALDH18A110q24.1MitochondrionDuodenum, small intestine, 25 other tissuesGlutamate; γ-Glutamyl phosphate [61,62,63,64]
Basic gene information was acquired from the NCBI (National Center for Biotechnology Information) and Uniprot (Universal Protein) databases.
Table 2. Function and regulatory pathways of ALDH2 in oxidative stress-related diseases and injury.
Table 2. Function and regulatory pathways of ALDH2 in oxidative stress-related diseases and injury.
OrganTreatmentTissue/Cell, SpeciesDisease/
Injury
Molecular MechanismALDH2 FunctionNotesRef.
HeartALDH2 (indirectly activated)Human cardiac fibroblasts (HCFs), humanMyocardial fibrosis(−) TGF-β1, Smad → (+) ALDH2 → (−) α-SMA, myofibroblastic proliferation, collagenInhibit myocardial fibrosis induced by TGF-β1 TGF-β1: transforming growth factor-β1
ALDH: aldehyde dehydrogenase
α-SMA: α-smooth muscle actin
Smad: Drosophila mothers against decapentaplegic protein
[111]
ALDH2 (inhibited by daidzin)Heart, ratCardiomyocytic apoptosisDaidzin → (−) ALDH2 → (+) ERK1/2, JNK, p38 MAPK → (+) ROS, 4-HNE → (+) cardiomyocytic apoptosisInhibit cardiomyocytic apoptosis by downregulating the MAPK pathwayERK1/2: extracellular signal–regulated kinase 1/2
JNK: c-Jun NH2-terminal kinase
p38 MAPK: p38 mitogen-activated protein kinase
4-HNE: 4-hydroxy-2-nonenal
ROS: reactive oxygen species
[112]
ALDH2 (activated by RIPostC)Heart, ratMyocardial ischemia/reperfusion (I/R) injuryRIPostC → (+) ALDH2, Bcl-2/Bax → p-Akt/Akt → (−) cleaved caspase-3 → (−) cardiomyocytic apoptosisHelp mediate the cardio-protection of RIPostC via the PI3K—Akt pathwayRIPostC: remote ischemic postconditioning
Bcl-2: B-cell lymphoma-2
Bax: Bcl-2–associated X protein
Akt: protein kinase B
PI3K: phosphatidylinositol 3-kinase
caspase-3: cysteinyl aspartate–specific proteinase 3
[113]
LiverALDH2/ (KO)Liver, mouseAlcoholic liver injuryAlcohol, CCl4, ALDH2/ → (+) MAAs, acetaldehyde → (+) IL-6 → (+) ERK1/2, p38 MAPK, STAT3, TGF-β, α-SMA, COLIA1, TIMP-1 → liver inflammation, fibrosisALDH2 deficiency accelerates alcohol-induced liver inflammatory response and fibrosisCCl4: carbon tetrachloride
MAAs: malondialdehyde–acetaldehyde adducts
IL-6: interleukin-6
STAT3: signal transducer and activator of transcription 3
COLIA1: collagen type I alpha 1
TIMP-1: tissue inhibitor of metalloproteinase-1
[114]
ALDH2/ (KO)Liver, mouseChronic liver fibrosisCCl4, ALDH2/ → (+) TGF-β1, ROS → Bcl-2/Bax ↓, Nrf2/HO-1 ↓; (+) TIMP-1, p62, α-SMA, COLIA1; (−) Parkin → (+) fibrosisAlleviate CCl4-induced hepatic fibrosis via Nrf2/HO-1 pathwayNrf2: nuclear factor erythroid 2-related factor 2
HO-1: heme oxygenase-1
[115]
NerveALDH2−/− (KO)Brain, mouseAlzheimer disease (AD)ALDH2−/− → (+) 4-HNE → (+) p-tau, Aβ, activated caspases → (+) NFTs → synaptic and mitochondrial dysfunction → neurodegenerationDefects in ALDH2 activity kill neurons by stimulating the accumulation of 4-HNE due to OStau: microtubule-associated protein tau
NFTs: neurofibrillary tangles
Aβ: β-amyloid
[116,117]
ALDH2Pheochromocytoma (PC12), ratParkinson disease (PD)DA, NE, EPI → DOPAL, DOPEGAL + ALDH2 → DOPAC, DOMA → HVA, VMA, (−) α-Syn, (−) LBs → (−) neurotoxicityPromote removal of neurotoxic metabolites produced in the metabolism of monoamine neurotransmitters and reduce neurotoxicityDA: dopamine
NE: norepinephrine
EPI: epinephrine
DOPAL: 3,4-dihydroxyphenylacetaldehyde
DOPEGAL: 3,4-dihydroxyphenylglycolaldehyde
DOPAC: 3,4-dihydroxyphenylacetic acid
DOMA: 3,4-dihydroxymandelic acid
HVA: homovanillic acid
VMA: vanillylmandelic acid
α-Syn: α-synuclein
LBs: Lewy bodies
[118,119]
Blood vesselALDH2 (transgenic over-expression)Human umbilical-vein endothelial cells (HUVECs), humanApoptosis in HUVECsAlcohol, (+) ALDH2 → (−) ROS, ERK1/2, p38 MAPK, caspase-3 → (−) HUVEC apoptosisAlleviate OS and apoptosis of HUVECs under acetaldehyde exposure via MAPK pathway[70]
ALDH2/ (KO)Human aorta epithelial cells (HAECs), humanEndothelial senescenceALDH2/ → (+) 4-HNE → (−) SIRT1 → (−) p53 deacetylation → (+) endothelial senescencePromote SIRT1 nuclear translocation by inhibiting 4-HNE accumulation to alleviate senescenceSIRT1: sirtuin 1[120]
OthersALDH2 (indirectly activated)Kidney, ratAcute kidney injury (AKI)CYA → (−) ALDH2, SOD; (+) plasma CRE, BUN, MDA, p65, NF-κB → aggravate glomerular atrophyInhibition of ALDH2 aggravated the renal injuryCYA: cyanamide
CRE: creatinine
BUN: blood urea nitrogen
MDA: malondialdehyde
SOD: superoxide dismutase
[121]
ALDH2 (activated by Alda-1)Intestine, mouseIntestinal I/R injuryAlda-1 → (+) ALDH2 → (−) 4-HNE, MDA, MPO, NO, iNOS, H2O2, caspase-3, NF-κBα, TNF-α, IL-6, IL-1β; (+) IκBα, Bcl-2/Bax → alleviate inflammatory response, OSALDH2 activation can alleviate intestinal I/R injury by relieving inflammatory response and OS Alda-1: N-(1,3-benzodioxol-5-ylmethyl)-2,6-dichlorobenzamide
MPO: myeloperoxidase
iNOS: inducible NO synthase
H2O2: hydrogen peroxide
NF-κB: nuclear factor κ-light-chain-enhancer of activated B cells
TNF-α: tumor necrosis factor-alpha
IκBα: inhibitor of NF-κB
IL-1β: interleukin-1β
[122]
ALDH2*2 (transfected with the ALDH2*2 gene)Osteoblasts, mouseOsteoporosisALDH2*2 → (+) 4-HNE, acetaldehyde → (+) PPARγ → (−) osteoblastic differentiation → (+) osteoblastic apoptosis → osteoporosisALDH2*2 as a dominant-negative form of ALDH2 promotes osteoporosis by impairing osteoblastogenesisPPARγ: peroxisome proliferator–activated receptor gamma[10]
(+) = activation/upregulation/ALDH2 overexpression; (−) = inhibition/downregulation; −/− = knockout/deletion; KO = knockout; ↑ = increase of ratio; ↓ = decrease of ratio; p- = phosphorylation; A + B = substance A interacts with substance B.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gao, J.; Hao, Y.; Piao, X.; Gu, X. Aldehyde Dehydrogenase 2 as a Therapeutic Target in Oxidative Stress-Related Diseases: Post-Translational Modifications Deserve More Attention. Int. J. Mol. Sci. 2022, 23, 2682. https://doi.org/10.3390/ijms23052682

AMA Style

Gao J, Hao Y, Piao X, Gu X. Aldehyde Dehydrogenase 2 as a Therapeutic Target in Oxidative Stress-Related Diseases: Post-Translational Modifications Deserve More Attention. International Journal of Molecular Sciences. 2022; 23(5):2682. https://doi.org/10.3390/ijms23052682

Chicago/Turabian Style

Gao, Jie, Yue Hao, Xiangshu Piao, and Xianhong Gu. 2022. "Aldehyde Dehydrogenase 2 as a Therapeutic Target in Oxidative Stress-Related Diseases: Post-Translational Modifications Deserve More Attention" International Journal of Molecular Sciences 23, no. 5: 2682. https://doi.org/10.3390/ijms23052682

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