Next Article in Journal
Boon and Bane of DNA Double-Strand Breaks
Next Article in Special Issue
Diagnosis and Therapeutic Management of Liver Fibrosis by MicroRNA
Previous Article in Journal
Molecular and Histologic Adaptation of Horned Gall Induced by the Aphid Schlechtendalia chinensis (Pemphigidae)
Previous Article in Special Issue
Expression of Interferons Lambda 3 and 4 Induces Identical Response in Human Liver Cell Lines Depending Exclusively on Canonical Signaling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Alcohol-Related Liver Disease: Basic Mechanisms and Clinical Perspectives

1
Department of Emergency Medicine, E-Da Hospital, I-Shou University, Kaohsiung 82445, Taiwan
2
School of Medicine for International Student, I-Shou University, Kaohsiung 82445, Taiwan
3
School of Chinese Medicine for Post Baccalaureate, I-Shou University, Kaohsiung 82445, Taiwan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(10), 5170; https://doi.org/10.3390/ijms22105170
Submission received: 17 April 2021 / Revised: 7 May 2021 / Accepted: 12 May 2021 / Published: 13 May 2021
(This article belongs to the Special Issue Pathophysiology of Chronic Liver Disease Development)

Abstract

:
Alcohol-related liver disease (ALD) refers to the liver damage occurring due to excessive alcohol consumption and involves a broad spectrum of diseases that includes liver steatosis, steatohepatitis, hepatitis, cirrhosis, and hepatocellular carcinoma (HCC). The progression of ALD is mainly associated with the amount and duration of alcohol usage; however, it is also influenced by genetic, epigenetic, and environmental factors. The definite diagnosis of ALD is based on a liver biopsy, although several non-invasive diagnostic tools and serum biomarkers have emerging roles in the early detection of ALD. While alcohol abstinence and nutritional support remain the cornerstone of ALD treatment, growing evidence has revealed that the therapeutic agents that target oxidative stress or gut-liver axis, inflammatory response inhibition, and liver regeneration enhancement also play a role in ALD management. Furthermore, microRNAs modulation and mesenchymal stem cell-based therapy have emerging potential as ALD therapeutic options. This review summarizes the updated understanding of the pathophysiology, diagnosis, and novel therapeutic approaches for ALD.

1. Introduction

Chronic alcohol consumption is one of the most common causes of morbidity and mortality worldwide and has an impact on more than 200 disease and injury outcomes [1,2]. According to the World Health Organization (WHO), 2.3 billion people are current drinkers, and approximately 1 billion people are classified as heavy episodic drinkers [3]. The alcohol-attributable fraction is the highest for liver cirrhosis, followed by road injury and other digestive diseases [3]. Alcohol is a well-recognized carcinogen in several types of cancers and has an established relationship with other liver-specific disease progression, such as chronic viral hepatitis and hepatocellular carcinoma (HCC) [4,5,6,7].
The liver is the primary organ responsible for ethanol metabolism, suffering from greater tissue injury through oxidative stress, acetaldehyde, and lipopolysaccharide (LPS) accumulation after excessive alcohol consumption [8,9]. Alcohol-related liver disease (ALD) refers to a broad spectrum of diseases, including asymptomatic early ALD (fatty liver or steatosis), steatohepatitis, advanced ALD (alcoholic hepatitis, cirrhosis), and HCC attributable to alcohol consumption [10,11].
Although there is a clear correlation between the amount and duration of alcohol intake and ALD progression, it is postulated that other co-factors (e.g., genetic, epigenetic, and environment factors) also play a role in ALD development, as only 10–20% of individuals with chronic alcohol use will progress to advanced ALD [12].
The prompt diagnosis of early ALD and complete alcohol abstinence is crucial in the ALD treatment strategy, as irreversible liver damage and hepatic decompensation have not occurred at this stage [12,13]. Nevertheless, no reliable symptoms, signs, or biochemical tests could aid in an early ALD diagnosis; additionally, the histological features of ALD may be similar to those of nonalcoholic fatty liver disease (NAFLD) [12,14]. Several non-invasive tests, including serum biomarkers and elastography, have been proposed for the detection and staging of ALD in a more accurate manner [2]. Moreover, a number of therapeutic agents that target oxidative stress [15], microRNA [16], gut microbiota modulation [17], and mesenchymal stem cell-based therapy [18] have been investigated in ALD prevention and treatment plans. In this review article, we summarize the basic mechanisms regarding the pathophysiology, diagnosis, and clinical management strategies, including molecule-based therapeutic agents in ALD.

2. Pathophysiology

2.1. Ethanol Metabolism

Alcohol is absorbed through the gastrointestinal tract into the blood circulation and is mainly metabolized by hepatocytes in the liver [19]. There are three main enzymatic metabolic pathways responsible for alcohol metabolism within the hepatocytes [20].
The first and the main pathway is hepatocyte cytoplasmic alcohol dehydrogenase (ADH), which uses nicotinamide adenine dinucleotide (NAD+) as a co-factor and oxidizes ethanol to acetaldehyde, which is highly toxic and causes DNA synthesis impairment [8]. The second pathway is the microsomal ethanol-oxidizing system (MEOS) in the smooth endoplasmic reticulum, which requires the cytochrome P450 2E1 (CYP2E1) enzyme to oxidize ethanol to acetaldehyde, generating reactive oxygen species (ROS), and triggering oxidative stress and inflammation [8,21]. It is noteworthy that CYP2E1 only catalyzes approximately 10% of ethanol into acetaldehyde under normal physiological conditions; however, it becomes more prominent in chronic alcohol consumption due to enhanced CYP2E1 expression [8,9]. The third and more minor pathway is via the heme-containing catalase in the peroxisomes, which can also oxidize ethanol to acetaldehyde [22]. The enzyme aldehyde dehydrogenase (ALDH) is located in the hepatocyte mitochondria and further oxidizes acetaldehyde to acetate, which is released into the circulation system and is further oxidized to carbon dioxide in various extrahepatic tissues [9].

2.2. Mechanisms in ALD Development

Although complex and not fully understood, the underlying molecular mechanisms responsible for ALD development include direct ethanol hepatotoxicity and lipid peroxidation, oxidative stress and ROS production, immune response activation and cytokine accumulation, and hepatic metabolism disorder [19,23,24].

2.2.1. Direct Ethanol Hepatotoxicity and Lipid Peroxidation

As mentioned earlier, chronic alcohol use upregulates CYP2E1 production, which leads to increased acetaldehyde concentration, diminished ALDH activity, and reduced acetaldehyde oxidation, resulting in acetaldehyde accumulation, which directly damages the mitochondria and microtubules of hepatocytes in the liver [19,20]. Furthermore, ethanol and acetaldehyde downregulate adiponectin, signal transducer and activator of transcription 3 (STAT3), and zinc levels, further inhibiting 5′-AMP-activated protein kinase (AMPK), peroxisome proliferator activated receptor α (PPARα), and its target gene activity, resulting in phospholipid peroxidation and lipid free radical production, and enhanced early growth response protein 1 (Egr-1) and adiponectin and acetyl-CoA carboxylase (ACC) expression, all of which cause fatty acid accumulation in the liver [25,26]. Recent studies also provided evidence that the lipogenesis process by lipolysis and free fatty acid flux to the liver from the small intestine further decreased the adipose tissue mass in an animal model with chronic alcohol consumption [27,28].

2.2.2. ROS Production and Oxidative Stress

CYP2E1-mediated or ethanol-induced inflammatory oxidative stress causes ROS generation (e.g., superoxide, hydroxyl radicals), which can bind to proteins and result in structural or functional alterations [15,26]. ROS can also bind directly to DNA and generate highly carcinogenic exocyclic ε-DNA adducts, which exhibit high mutagenic potential in numerous types of base pair substitutions and genetic damage in organisms, and has been analyzed as a representative lipid peroxidation-derived DNA damage marker in several studies [29,30]. In addition, acetaldehyde-mediated glutathione decreases and dysregulates the expression of antioxidant genes, including nuclear factor erythroid 2-related factor 2 (Nrf-2) and thioredoxin, leading to decreased antioxidant and detoxification enzyme production, and low activity of the antioxidant defense system [26,31].

2.2.3. Cytokines Activation and Advanced Fibrogenesis

Chronic alcohol consumption is a known factor of intestinal endotoxin accumulation and intestinal wall permeability increasing, facilitating the translocation of endotoxins from the intestines to the liver in the form of LPS, which are toxic to hepatocytes [9]. LPS can bind to different toll-like receptors (TLRs) and activate the synthesis and release of cytokines and inflammatory factors, such as tumor necrosis factor α (TNF-α), interlukin-1 (IL-1), interlukin-6 (IL-6), and platelet-derived growth factor (PDGF), further stimulating neutrophil and macrophage accumulation, and finally causing hepatic inflammation and systemic injury in hepatic Kupffer cells [32,33,34]. Additionally, liver injury activates hepatic stellate cell (HSCs) proliferation, which enhance transforming growth factor β (TGF-β) secretion and collagen synthesis, thus forming extracellular matrix deposition and advanced fibrogenesis [35,36].

2.2.4. Hepatic Metabolism Disorder

Numerous studies have revealed that alcohol consumption is correlated with iron overload and hepcidin synthesis downregulation in Kupffer cells and hepatocytes in the liver, while hepcidin was proposed as a key mediator in iron homeostasis [37,38]. Moreover, alcohol can abolish the protective effect of hepcidin in situations of iron overload by rendering hepatic hepcidin synthesis insensitive to total body iron levels [38]. Excessive iron can act synergistically with alcohol to induce the oxidative stress and lipid peroxidation, increasing transferrin receptor 1 (TfR1) expression to promote intestinal iron absorption; thus, the additive effect of iron absorption and deposition could potentiate progressive liver damage [26,39].

2.3. Genetic Factors

The genetic impact on alcohol use disorder (AUD) and ALD development was elucidated in previous studies, as individual variation exists after chronic alcohol consumption [40]. Several genome-wide association studies (GWAS) have identified several genetic risk loci for ALD development, including Patatin-like phospholipase domain-containing-3 (PNPLA3) gene, which is the major risk factor for the progression of ALD, and to a lesser extent, transmembrane 6 superfamily member 2 (TM6SF2) and membrane-bound O-acyltransferase domain-containing protein 7 (MBOAT7), are the key determinants of ALD progression [41,42,43]. Meanwhile, a splice variant in hydroxysteroid 17-β dehydrogenase 13 (HSD17B13) was identified as a candidate for protection against ALD in a recent study [44].

2.3.1. PNPLA3

PNPLA3 is predominantly expressed in adipose tissue and is closely related to lipid metabolism, regulation of energy homeostasis, and the maintenance of membrane integrity [45]. It is also highly synthesized in HSCs and is responsible for retinyl ester hydrolysis [46]. The rs738409 variant (C.444 C > G p.Ile148Met) in the PNPLA3 gene results in hydrolytic function reduction, fat accumulation, and further liver inflammation injury [47]. This variant is more frequent in the Hispanic population and particularly sensitive to fatty liver diseases, although it has also been investigated in NAFLD and HCC [47,48]. A previous meta-analysis provided evidence for a significant role for rs738409 in PNPLA3 in ALD progression [49]. In addition, the presence of rs738409 in PNPLA3 is associated with an increased risk of HCC development in patients with cirrhosis due to ALD, with an estimated risk of twofold after adjusting for other confounders (e.g., age, sex, body mass index) [50,51].

2.3.2. TM6SF2

TM6SF2 is mainly expressed in the liver and intestine and is presumed to be involved in lipid turnover and metabolism [52]. The rs58542926 (c.499C > T p.Glu167Lys) genetic variant is associated with reduced triglyceride-rich lipoprotein secretion and increased liver triglyceride concentration, which has been investigated as a determinant of NAFLD or advanced liver fibrosis development in previous studies [53,54,55]. Notably, TM6SF2 rs58542926 combined with PNPLA3 rs738409 may have additive value in predicting ALD cirrhosis progression [56]; additionally, both of them had a potential risk for HCC development in ALD patients [51,57]. It is postulated that lipid accumulation, followed by lipid peroxidation, inflammatory cell response, oxidative stress, and DNA damage are also important steps in HCC development [58].

2.3.3. MBOAT7

MBOAT7 is an enzyme that transfers fatty acids between phospholipids and lysophospholipids and is involved in the phospholipid acyl-chain remodeling pathway [59]. The rs641738 C > T variant in the MBOAT7-TMC4 locus is associated with phosphatidylinositol acetylation disturbance, further hepatic inflammation, and fibrosis progression, and has been identified as a risk factor for NAFLD or non-alcoholic steatohepatitis (NASH) formation [60,61,62]. Similarly, MBOAT7 combined with TM6SF2 and PNPLA3 gene variants were identified as risk factors in ALD cirrhosis development in a GWAS study [43], although inconsistent findings were revealed in a recent study, which may need more evidence for validation [63].

2.3.4. HSD17B13

HSD17B13 encodes hepatic lipid droplet protein and plays a role in lipid metabolism. A recent in vitro study indicated that HSD17B13 may regulate HSCs activity and participate in liver fibrosis development [64]. The rs72613567 T > A variant results in shortened protein production and reduced enzymatic activity, which is associated with both ALD and NAFLD progression to cirrhosis or HCC, and decreased liver enzyme levels [44]. These findings were validated in two recent large cohort studies in different ethnic groups with alcohol use, suggesting the therapeutic potential of HSD17B13 in the future [65,66].

2.4. Epigenetic Modifiers

Epigenetics refers to a process that changes the gene activity without DNA sequence alteration, such as DNA methylation, histone modification, or RNA silencing by microRNAs (miRNAs) [67,68]. miRNAs are single-stranded, 19–22 nucleotide non-coding sequences that bind to the complementary sequence of messenger RNA molecules, regulating their function by inhibiting or silencing translation [69]. As shown in Table 1, the dysregulation of miRNAs correlates with ALD severity and prognosis via regulation of multiple functions, including intestinal permeability change, liver steatosis and fibrosis, and oxidative stress [8,16,70]. Other epigenetic mechanisms include acetylation and methylation of DNA mediated by ethanol [71]. The ongoing metabolism of ethanol produces excessive ROS and depletes glutathione, diverting the reaction from the production of methionine and S-adenosylmethionine (SAM), which is the predominant methyl donor in DNA methylation [72,73]. The hypomethylation of DNA further facilitates hepatocyte proliferation and tumorigenesis [74]. In addition, alcohol-exposed hepatocytes show decreased NAD+ levels, which is also an important co-factor in histone acetylation [73]. Decreased histone acetylation further impairs the Sirtuin 1 (SIRT1)–AMPK pathway, a hepatic lipid metabolism pathway, and results in fatty liver and advanced fibrosis formation [75]. Collectively, epigenetic modifiers may influence the ALD therapy response and can help in treatment plan modification, which needs to be confirmed in future studies [69].

2.5. Environmental Factors

Numerous epidemiological factors affect ALD development and progression [12]. Women are more vulnerable to ethanol-related liver damage than men after the same amount of alcohol consumption, possibly due to their lower ADH activity and higher body fat composition; moreover, the estrogen-mediated inflammatory response increases the risk of ethanol-related liver damage [8,113]. Obesity is the most widely recognized environmental risk factor in ALD and has a close interaction and additive effect with alcohol [114,115]. Obesity can affect the ethanol lipid solubility and adipose tissue pro-inflammatory cytokine production, leading to alcoholic steatohepatitis, whereas alcoholic fatty liver induces insulin resistance and promotes obesity [8]. Meanwhile, multiple components of metabolic syndrome, including waist circumference, smoking, and alcohol use are the risk factors for severe liver disease in a recent population-based study [116]. Other known comorbidities, including viral hepatitis, hereditary hemochromatosis, and HIV coinfection in patients with concomitant alcohol use have a higher risk of accelerated liver fibrosis and increased mortality of liver-specific disease [10,117,118,119]. Notably, caffeine intake may protect against ALD cirrhosis in recent studies [120,121]; additionally, drinking two cups of coffee per day was estimated to decrease half the risk of ALD cirrhosis in a meta-analysis [122].

2.6. ALD Spectrum

2.6.1. Alcoholic Fatty Liver or Steatosis

The diagnosis of alcoholic fatty liver (AFL) disease is established in a patient with known AUD with hepatic steatosis seen on ultrasound combined with liver enzyme elevation and the absence of other causes of liver disease [12]. AFL development is regulated by several direct or indirect regulatory mechanisms, including PPARα and AMPK expression inhibition, and ACC activity enhancement, which results in increased fatty acid synthesis and deposition [22]. Fat vacuoles or macrovesicles can be observed in liver tissues under a microscope, which resolves rapidly after complete abstinence [123]. AFL are seldom diagnosed because of their asymptomatic or nonspecific symptoms [2].

2.6.2. Steatohepatitis due to ALD

Steatohepatitis due to ALD is presumed to be a progressive liver lesion, which has an increased risk of cirrhosis and HCC [2]. The common histological features of steatohepatitis due to ALD include steatosis, ballooned liver cells containing large Mallory-Denk bodies, sclerosing hyaline necrosis, and lobular inflammation predominated by neutrophils, which are rarely seen in NAFLD [19,115]. The inflammatory environment enables further leukocyte infiltration, ROS formation, and hepatocyte injury. As the injury continues, the release of damage-associated molecular patterns (DAMPs) activates multiple immune reactions and promotes liver fibrosis or malignancy [9,124]. Similar to AFL, mild steatohepatitis rarely presents with clinical symptoms and can only be diagnosed by liver biopsy; further, the development of novel non-invasive tests is urgent for this condition [12].

2.6.3. Alcoholic Hepatitis

Alcoholic hepatitis (AH) is a clinical entity associated with severe steatohepatitis due to ALD and has a high short-term mortality risk [125]. In addition to steatohepatitis, the histological features of AH may also include megamitochondria, satellitosis, and cholestasis, which are related to the prognosis [126]. Both adaptive and innate immune dysfunctions are more prominent in patients with AH than in those with non-alcoholic liver disease, contributing to a higher risk of neutrophilia, liver dysfunction, and multi-organ failure [127]. The clinical symptoms of AH are characterized by the presence of jaundice with/without other hepatic decompensation events (e.g., ascites, hepatic encephalopathy) in patients with ongoing alcohol use [12]. It is noteworthy that despite complete abstinence, a significant proportion of patients had persistent AH and even progressed to cirrhosis [128,129].

2.6.4. Fibrosis/Cirrhosis due to Alcohol-Related Liver Disease

As the vicious cycle continues (i.e., liver injury and regeneration) in ALD patients with ongoing alcohol use, the acetaldehyde-protein adducts inactivate DNA repair, damage hepatocyte mitochondria, impair oxygen utilization, and further stimulate collagen band synthesis and deposition between central veins and portal areas, resulting in liver fibrosis [9,11]. Cirrhosis is further characterized by marked hepatic architectural distortion due to extensive fibrosis and regenerative nodule formation [2]. The clinical manifestations of the patients with fibrosis or cirrhosis due to ALD widely range from asymptomatic to various decompensation events (e.g., variceal bleeding, bacterial infection, and hepatorenal syndrome) [130]. Importantly, active excessive alcohol consumption was identified as the second most frequent triggering factor of acute-on-chronic liver failure (ACLF) in patients with chronic liver disease, including ALD [131,132].

3. Diagnosis

There is no unique presentation of ALD that can be completely distinguished from other etiologies of liver disease [133]. The European Association for the Study of the Liver (EASL) proposed that the diagnosis of ALD is suspected when the amount of regular alcohol consumption is >30 g/day in men or >20 g/day in women, combined with the presence of clinical and/or biological evidence suggestive of liver injury [2]. ALD should also be considered in patients with AUD presenting with extrahepatic manifestations, including symmetric peripheral neuropathy, pancreatitis, or cardiomyopathy [134]. For the purpose of early ALD detection because of their favorable prognosis, asymptomatic patients with heavy alcohol consumption are recommended for regular ALD screening transfer [135].
A liver biopsy is considered the gold standard to establish a definite ALD diagnosis, to assess the stage and prognosis of liver disease, and to exclude alternative causes of liver injury, as approximately 20% of the patients with chronic alcohol consumption and abnormal liver enzymes were proven to have other coexisting liver disease etiologies [125,136,137]. Nevertheless, the performance of liver biopsy may induce morbidities, including intrahepatic bleeding and pneumothorax in approximately 2% of patients and is generally not recommended in routine clinical practice in all patients with suspected ALD [136,138].

3.1. Non-Invasive Diagnostic Tools

Ultrasonography is a non-invasive, inexpensive, and widely available tool for early ALD screening, with a sensitivity of 60–94% and a specificity of 88–95% for detecting steatosis, although it varies significantly with the degree of fatty content [139,140]. Other limitations of ultrasound include its operator-dependent [140] and difficulty in differentiating fibrosis from steatosis [136]. Magnetic resonance imaging-based methods (e.g., magnetic resonance spectroscopy) are reliable methods for measuring hepatic fat with reproducibility; however, the cost and long examination time limit their use in routine examinations [2,141]. The controlled attenuation parameter (CAP) is a novel ultrasound-based elastography method used to measure hepatic steatosis [142]. CAP revealed modest discriminative ability in differentiating the steatosis severity in a recent ALD study, although further validation studies are warranted [143].
The measurement of liver stiffness by transient elastography (TE) has become a popular non-invasive method for screening liver fibrosis/cirrhosis in recent years [144]. The value of liver stiffness measured in ALD patients correlates well with the degree of fibrosis, portal pressure, and its complications [145,146]. Regardless of ethnicity, the implementation of TE was validated as a cost-effective screening tool and correlated with the prognosis in patients with different stages of liver disease [147]. Notably, several pathophysiological conditions may exist and interfere with the interpretation of TE values, such as hepatic neuroinflammation, congestion, and cholestasis [148,149,150]. Compared with the patients with other etiologies of cirrhosis, the patients with cirrhosis due to ALD had significantly higher liver stiffness values, presumably with a higher degree of fibrosis, although the serum aspartate aminotransferase (AST) and bilirubin levels may also be considered for value adjustment [151,152]. Recently, various elastographic methods, such as acoustic radiation force impulse (ARFI) and shear-wave elastography (SWE), have been developed to assess the degree of hepatic fibrosis [153]. Although limited ALD studies have compared these different elastographic methods, they may show a similar performance based on current evidence [136].

3.2. Blood Tests and Biomarkers

Various serum biomarkers have been proposed for the detection of chronic alcohol use in recent years, including indirect markers, such as alcohol-induced metabolic products (e.g., carbohydrate-deficient transferrin (CDT), 5-hydroxytryptophol (5-HTOL)), and direct markers, such as alcohol metabolites (e.g., ethyl glucuronide (EtG), ethyl sulfate (EtS), phosphatidylethanol (Peth), and fatty acid ethyl esters (FAEEs)) [154]. CDT has been the most commonly used biomarker to confirm chronic alcohol consumption for years, although the sensitivity and specificity vary considerably between different studies, and numerous confounders could influence the results, including age, sex, and the stage of liver disease [155,156]. EtG can be detected in the urine or hair, has a much longer detection window, and a higher specificity than CDT, although biological variability (e.g., medications and foods) still exists and may complicate the interpretation of the results [157,158]. Several studies have compared the detection ability of these biomarkers and revealed conflicting results [154,157,159,160]. It is proposed that the combination of these biomarkers, such as EtS in urine, FAEEs in hair, and PEth in serum, may improve the overall sensitivity and specificity and provide more reliable results [154]. Recently, the differential methylation of DNA in specific genes revealed its potential utility as a diagnostic marker of active alcohol consumption, and may provide another choice in molecule-based studies in ALD [161,162].
Conventional liver function tests (i.e., AST, alanine aminotransferase (ALT), γ-glutamyl transferase) are widely used for liver disease screening; additionally, an AST/ALT ratio greater than 1.5 is traditionally considered as a diagnostic biomarker in ALD [136,163]. However, this frequent lack of ethanol specificity or distinguishing of the disease pathology, while advanced fibrosis due to ALD may present with normal liver function tests [164]. In order to exclude alternative causes of liver injury, EASL recommended a series of blood examinations, including hepatitis B and C virus serology, autoimmune markers, transferrin, and α1-antitrypsin in the workup of ALD patients [2,125,155]. In patients with advanced fibrosis or cirrhosis due to ALD, serum albumin levels, coagulation function profiles, bilirubin levels, and white blood cell counts should also be collected to determine the severity of the liver injury [165].
As a distinct clinical syndrome widely ranges from a few signs or symptoms to liver failure, the need for novel diagnostic biomarkers in AH is urgent, including cytokines and microRNA [166]. Mallory-Denk bodies, the hallmark of alcoholic hepatitis and steatohepatitis, contain cytokeratin-18 (CK-18) and cytokeratin-19 (CK-19), while the serum CK-18 and CK-19 levels were increased in AH patients in previous studies [167]. Meanwhile, the caspase-cleaved CK18 fragment M30 and M65 had modest predictive ability in biopsy-proven AH patients [168]. Interleukin-22 (IL-22) is a member of the IL-10 cytokine family and plays a protective role against liver injury in AH by anti-apoptosis, anti-oxidation, anti-fibrosis, and liver regeneration promoting effects [169,170]. The potential role of IL-22 as a diagnostic biomarker of AH has been investigated in several current studies [171]. Circulating miRNAs are another emerging biomarker candidate, as they regulate the inflammatory response and Kupffer cell activity (e.g., miR-155), hepatocyte damage (e.g., miR-122), cell proliferation and apoptosis (e.g., miR-30a) [82,172,173]. The majority of these miRNAs are packaged into exosomes or extracellular vesicles in the circulation; therefore, the lack of standardized extracellular vesicle isolation and sample collection/handling methods is currently a major obstacle [163].
The development of biomarkers for excess connective tissue deposition activity and progression of fibrosis due to ALD is another important issue [155]. Type I and type III collagens are the principal collagen deposited in liver tissues in response to alcohol injury; additionally, their derivatives of procollagen (e.g., type III procollagen aminopropeptide (PIIINP), type I procollagen aminopropeptide (PINP), and type I procollagen carboxypropeptide (PICP)) have been investigated as the biomarkers of fibrosis progression in ALD patients [35]. Furthermore, hyaluronic acid (HA) synthesized by HSCs and matrix metalloproteinases (MMPs) together with tissue inhibitor of metalloproteinases (TIMPs) also play a key role in fibrogenesis and fibrolysis, extracellular matrix formation, and degradation [174]. Recently, the enhanced liver fibrosis (ELF) test combined with HA, PIIINP, and TIMP-1, which revealed a similar diagnostic accuracy to the histological examinations in fibrosis progression evaluation in ALD patients [175]. The ELF also demonstrated superior fibrosis discrimination ability compared with different biological tests, such as the AST to platelet ratio index (APRI) or fibrosis-4 (FIB-4) index [176,177]. In summary, a combination of non-invasive diagnostic tests with serum-based fibrosis biomarkers will likely emerge as the mainstay diagnostic tool in the future [136].

4. Management

4.1. General Consideration

Complete alcohol abstinence is the cornerstone and improves the clinical outcomes in the treatment of all ALD stages [2,12]. Multidisciplinary management with the use of pharmacological therapy and behavioral intervention, as well as lifestyle modification, is recommended for prolonged abstinence attainment [178,179]. Sudden discontinuation or reduction in alcohol consumption may lead to alcohol withdrawal syndrome (AWS) in alcohol-dependent patients within 24 h after the last drink [180]. AWS includes varying degrees of autonomic hyperactivity symptoms and may progress to life-threatening conditions such as seizures, coma, cardiac arrest, and even death [2,180]. Benzodiazepines are considered as the standard treatment for AWS, and other drugs, such as β-blocker or α2-agonists can be used as adjunctive therapy [180]. Regular physical exercise and nutritional assessments should be incorporated into the abstinence strategy [115]. Meanwhile, public health policies to reduce the general population exposed to alcohol, such as a price-based taxation, increasing the legal age for buying alcohol, alcohol availability limitations, and advertising restrictions, have shown their effectiveness [181].
Malnutrition is a common complication in ALD patients, defined as the loss of body weight, muscle or fat mass, muscle strength, visceral protein levels, and immune function [182]. Malnutrition was estimated to be present in 50% of outpatient ALD patients and nearly all hospitalized ALD patients, and is associated with poorer prognosis [183,184]. The reason for ALD patients with malnutrition is often multifactorial, including an altered olfactory and gustatory perception, appetite-related hormone alterations, and gut microbiota changes [184]. The nutritional support in hospitalized ALD patients mainly focuses on the increased protein/calorie intake via various routes (e.g., enteral or parenteral), replacement of amino acids, and micronutrients [182,183]. Patients with ALD should receive proper nutritional status assessment and nutritional support when malnutrition is recognized [12]. In addition, efforts should be made to ensure adequate outpatient follow-up [184].

4.2. Pharmacology Therapy in Relapse Prevention

The current medications for relapse prevention in ALD patients are highlighted in Table 2. Among these, only disulfiram, naltrexone, and acamprosate were both approved by the US and Europe [26,133]. Disulfiram inhibits the enzyme acetaldehyde dehydrogenase, producing high levels of acetaldehyde following alcohol consumption [185]. Naltrexone is an opioid receptor antagonist that can reduce dopamine release in the reward system [186]. Both disulfiram and naltrexone undergo hepatic metabolism; hence, they should be avoided in patients with advanced ALD because of the fear of hepatotoxicity [133]. Acamprosate is the calcium salt of N-acetyl homotaurine, which is mainly metabolized through kidney excretion [186]. Acamprosate was proposed as a treatment option for patients with AUD, although it failed to demonstrate the treatment efficacy in a recent network meta-analysis [186]. Nalmefene, another opioid receptor antagonist approved only in Europe, could be considered in patients with early stages of ALD, where abstinence is not feasible in these patients [187,188]. Baclofen, a gamma-aminobutyric acid-B (GABA-B) agonist, increases the abstinence rate and prevents relapse in patients with AUD [189]. Notably, baclofen is the only AUD pharmacotherapy tested in patients with established cirrhosis, although several studies have reported inconsistent results [190,191,192,193]. Nevertheless, a recent international consensus statement recommended the consideration use of baclofen to treat AUD in patients with advanced ALD [194].

4.3. Specific Treatment for Alcoholic Hepatitis

4.3.1. Corticosteroids

Corticosteroids are the most widely studied interventions in severe AH; they can change the cytokine balance, reduce pro-inflammatory cytokines, and increase anti-inflammatory cytokines [211]. The Maddrey discriminant function is the mostly widely used scoring system for AH, and a cut-off value of 32 identifies patients for initiating corticosteroid therapy [12,212]. Subsequently, the Lille score, a prognostic model calculated by baseline data and a change of serum bilirubin on day 7 of corticosteroid therapy, can be used to assess the response to corticosteroid [12,213]. A Lille score (ranges from 0 to 1) ≥ 0.45 indicates non-response to corticosteroid, and patients should discontinue therapy and consider other treatment strategies [12]. Numerous randomized trials regarding corticosteroid efficacy in AH have demonstrated conflicting results, likely due to the heterogeneity of the studied group and lack of power to detect survival differences [214]. A landmark multicenter randomized trial revealed that corticosteroids confer only modest 28-day survival benefits but not long-term outcomes, which was confirmed in a recent meta-analysis [215,216]. Infection is the most worrisome complication of corticosteroids, ranging approximately 20% in AH patients receiving corticosteroid treatment, and may offset its therapeutic benefit [217,218]. Therefore, the use of corticosteroids is restricted to the AH patients without infection, which eliminates a substantial proportion of patients [126].

4.3.2. Antioxidants

Since oxidative stress plays a central role in the hepatotoxicity and pathogenesis of ALD and AH, antioxidants are of theoretical interest in AH treatment [219]. N-acetylcysteine (NAC) restores the glutathione stores and limits oxidative stress, although no additional survival benefit was observed in AH patients treated with NAC alone [220,221]. Notably, the combination of NAC and prednisolone compared with prednisolone alone revealed an improvement in the one-month survival, as well as a decreased incidence of hepatorenal syndrome and infections [222]. Metadoxine is another promising antioxidant that can aid in glutathione metabolism and inhibit hepatic steatosis [223]. The combination of metadoxine with steroids or pentoxifylline showed a significant improvement in both short-and long-term survival, and prolonged abstinence maintenance was observed [224,225]. Future large studies are required to validate these findings and provide further suggestions [126].

4.3.3. Liver Regeneration

The counterbalance of cell death in AH is the liver regeneration capacity, which is supported by bone marrow-derived stem cells and hepatic progenitor cells [226,227]. Granulocyte-colony stimulating factor (G-CSF) can stimulate the bone marrow to produce and release stem cells into the circulation and facilitate liver progenitor cell proliferation [228]. Several small clinical studies have demonstrated that G-CSF administration is associated with increased survival rates and a decreased risk of infection in AH patients [229,230]. Furthermore, a recent Indian randomized trial showed the mortality benefit of G-CSF compared with a placebo in patients with steroid-nonresponsive AH [231]. Nevertheless, this favorable effect was not duplicated in the currently available US or European studies, which requires more evidence in future trials [232,233]. Obeticholic acid, the farnesoid X receptor (FXR) agonist, regulates bile acid homeostasis, decreases cholestasis, and further modulates liver regeneration, although the clinical trial was terminated due to its hepatotoxicity [228,234].

4.3.4. Anti-Inflammatory/Anti-Apoptosis

The highly inflammatory condition of AH involves crosstalk between various signaling pathways, including pro-inflammatory cytokine and chemokine production, such as TNF-α, IL-1, and IL-6 [126]. Pentoxifylline (PTX), a non-selective phosphodiesterase inhibitor, can decrease TNF-α and other pro-inflammatory cytokine production and was initially tested as a therapeutic agent in AH [235]. However, growing evidence has demonstrated that the use of PTX is associated with a reduction in the hepatorenal syndrome development, but not the mortality benefit [215,216]. Based on current evidence, the use of PTX in AH treatment was no longer recommended in the current guidelines [2,12,133]. Similarly, studies of TNF-α inhibitors, such as infliximab and etanercept, were terminated early due to infection-related mortality [236,237]. Other novel therapeutic agents, including IL-1β antagonists, IL-22 agonists, and C-C chemokine antagonists, have been investigated in various clinical trials [238,239].
One of the alcohol-induced hepatocyte injury pathways is the process of apoptosis and macrophage activation mediated by endoplasmic reticulum stress, the mitochondrial pathway, and the caspase-dependent pathway, which further triggers abnormal liver tissue repair and inflammation [24,240]. Emricasan is a pan-caspase inhibitor that has been studied in the animal models of liver injury [241]. A recent multicenter randomized trial showed an improved three-month survival after treatment with emricasan compared with a placebo in patients with cirrhosis, regardless of the etiologies [242]. Nevertheless, a dose-ranging study of emricasan may be needed in further AH trials, as patients with severe AH have altered hepatic metabolism and pharmacodynamics [243]. Selonsertib is an oral apoptosis signal regulating kinase-1 (ASK-1) enzyme inhibitor, which can theoretically attenuate apoptosis and cytokine signaling [238]. Selonsertib failed to show survival benefit or liver function improvement compared with steroids in a recent study of patients with severe AH [244].

4.3.5. Gut-liver Axis Targeting

Alcohol-induced gut-liver axis dysfunction was initiated with an intestinal microbiome composition change and permeability increase, which stimulates bacterial translocation, LPS release, and endotoxemia [245]. An animal study demonstrated that the susceptibility to ALD can be manipulated by intestinal microbiome implantation in AH patients, further confirming the role of microbiota in AH pathogenesis [246]. The strategy to reverse these processes may be achieved by intestinal decontamination [238]. IMM-124E is a purified hyperimmune bovine colostrum enriched with IgG antibodies against LPS [247]. It can modulate immune cell function, thus alleviating liver injury in experimental models; additionally, the clinical trials of AH are ongoing [241]. The administration of probiotic or antibiotic therapy is another potential approach for AH management [241]. A pilot study in the patients with mild AH demonstrated that the administration of Bifidobacterium bifidum and Lactobacillus plantarum 8PA3 for five days significantly reduced liver biochemistry profiles and restored gut flora [248]. Other trials on probiotic or antibiotic, including amoxicillin clavulanate, Lactobacillus rhamnosus, and rifaximin, are under investigation [17]. Another promising therapeutic option is fecal microbiota transplantation (FMT). It revealed increased survival, reduced pathogen levels, and increased levels of beneficial bacterial strains in steroid-resistant patients with severe AH [249]. In summary, microbiome modulation has emerged as a novel and practical therapeutic approach for AH treatment [126,228,250].

4.4. Other Novel Therapies

4.4.1. MicroRNA Modulation

As mentioned earlier, miRNAs are packaged into exosomes or extracellular vesicles and are expressed as the regulators of target proteins involved in a variety of oxidative stress, inflammatory responses, and lipid metabolism during ALD development [8]. The most widely studied miRNAs in ALD are miR-122 and miR-155 [251]. In mature hepatocytes, miR-122 constitutes 70% of all miRNAs, and notably, it possibly has pleiotropic roles in ALD pathogenesis, as it could sensitize monocytes to LPS stimulation and increases the pro-inflammatory cytokine levels in ethanol-treated hepatocytes [88]. Moreover, it protects hepatocytes from ethanol-induced damage by reducing hypoxia inducible factor 1 α (HIF-1α) levels [90]. Miravirsen, an miR-122 inhibitor, was previously investigated in hepatitis C treatment and may also have therapeutic potential in ALD [252,253]. Another important miRNA is miR-155, a major regulator of increased Kupffer cell activation and TNF-α production, and is also involved in ethanol-induced liver fibrosis and steatohepatitis by mediating the peroxisome proliferator-activated receptor response element (PPRE) and PPARα pathway [97]. The inhibition of miR-155 can lead to decreased ethanol-induced sensitivity of Kupffer cells to LPS in vivo [98]. Currently, there are no clinical trials regarding miRNA targeting in ALD treatment. Also, their roles as a main or adjunct therapeutic regimen also need more evidence for validation in the future [251].

4.4.2. Mesenchymal Stem Cell

Stem cell transplantation therapy is another potential therapeutic option for ALD, especially in liver fibrosis [9]. Mesenchymal stem cells (MSCs) can provide support to hematopoietic stem cells and initiate the hematopoiesis process, and they also play an important role in organ homeostasis in past studies [254]. Regarding liver regeneration treatment, the benefits of MSCs intervention in ALD include parenchymal cell trans-differentiation and hepatocyte proliferation, promotion of regeneration ability, modulation of inflammatory responses, and inhibition of liver fibrosis [18]. Compared with miRNAs, a variety of clinical studies on MSCs therapy in different etiologies of liver disease have been conducted in recent years, including the ALD spectrum [255]. A phase 2 pilot study used bone marrow-derived MSCs for treating patients with cirrhosis due to ALD, which revealed a significant histological and quantitative improvement of hepatic fibrosis at 12 weeks after MSCs injection [256]. Another multicenter study used bone marrow-derived MSCs for the treatment of cirrhosis due to ALD; it showed significant improvement in histologic fibrosis and liver function after a longer follow-up period [257]. Taken together, the implementation of MSCs can be an attractive strategy in ALD treatment if their survival rate and activity could be further enhanced in the future field of regenerative medicine [258].

4.5. Liver Transplantation

Patients with end-stage ALD who respond poorly to medical therapies may be considered for liver transplantation (LT) [259]. A prior prospective multicenter study demonstrated that early LT improved the six-month survival probability in patients with severe alcoholic hepatitis, nonresponsive to standard corticosteroid therapy [260]. Notably, ALD is the leading indication of LT, accounting for 15–20% of all LTs in the US and Europe [261,262]. Although the survival rates of LT in ALD patients were poor in the 1980s, they have become comparable to those in patients transplanted for other indications [262]. It is noteworthy that long-term alcohol consumption often damages other organs and presents with extra-hepatic manifestations (e.g., cardiomyopathy, chronic kidney disease, pancreatitis, sarcopenia, and peripheral neuropathy), which should be evaluated before surgery as they may negatively impact post-transplantation outcomes [261]. In addition, complete abstinence is required before surgery, as it allows time for the liver to recover from alcohol-related toxic effects; also, the patient’s commitment to sobriety can be assessed [211]. Nevertheless, there is no consensus regarding the duration of abstinence in LT guidelines [2,12,133,261].
The major obstacles to LT in ALD patients include the scarcity of donors, immunologic rejection, complexity and costs of surgery, and most importantly, the ethical issue [263]. ALD is widely considered a self-inflicted disease in the general public, even practicing physicians, and the allocation of organs is more likely to prioritize patients with acquired diseases that are less directly related to behavior [2,263]. The concerns of relapse after LT is another consideration affecting their willingness to provide LT for ALD patients, as LT cures the liver disease but not the underlying AUD [264]. It was estimated that approximately 20–25% of ALD patients relapsed within five years of LT [265]. A multidisciplinary approach, including a relapse prevention program, may help reduce the risk of recidivism [266].

5. Conclusions

ALD is a major cause of liver disease worldwide, causing an extensive public health burden related to health, social, and economic harm. Growing evidence supports that ALD development is not solely explained by excessive alcohol consumption, but also comprises complex factor interactions, including genetic, epigenetic, and environmental modifier effects. Most early phase patients are asymptomatic, and the combination of non-invasive diagnostic tools and serum-based biomarkers has the potential to improve their prompt diagnosis. Complete abstinence and nutritional support remained the mainstay of ALD treatment, and a multidisciplinary implementation including psychosocial support and pharmaceutical intervention is necessary, which can help to prevent alcohol relapse. Moreover, the application of public health policies is a practical method for reducing the general population ALD risk. The miRNAs, mesenchymal stem cells, and other emerging novel therapies have potential roles in advanced-stage ALD treatment. Future translational science-based research and clinical trials are urgently required to identify the biomarkers for liver regeneration assessment, inflammation evaluation, and organ failure prediction in ALD patients [10].

Author Contributions

Conceptualization, Y.-C.H.; writing—original draft preparation, S.-Y.L., I.-T.T. and Y.-C.H.; writing—review and editing, S.-Y.L., I.-T.T. and Y.-C.H.; supervision, Y.-C.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

5-HTOL5-hydroxytryptophol
α-SMAα smooth muscle actin
ACCadiponectin and acetyl-CoA carboxylase
ACLFacute-on-chronic liver failure
ADHalcohol dehydrogenase
AFLalcoholic fatty liver
AFRIacoustic radiation force impulse
AHalcoholic hepatitis
ALDalcohol-related liver disease
ALDHaldehyde dehydrogenase
ALTalanine aminotransferase
AMPKAMP-activated protein kinase
APRIaspartate aminotransferase to platelet ratio index
ASK-1apoptosis signal regulating kinase-1
ASTaspartate aminotransferase
AUDalcohol use disorder
AWSalcohol withdrawal syndrome
Btg2b-cell translocation gene 2
CAPcontrolled attenuation parameter
CCl4carbon tetrachloride
CCL20C-C motif chemokine ligand 20
CDTcarbohydrate deficient transferrin
C/EBPβCCAAT/enhancer-binding protein β
CK-18cytokeratin-18
CK-19cytokeratin-19
CXCL1C-X-C motif chemokine ligand 1
CYP2E1cytochrome P450 2E1
DAMPsdamage-associated molecular patterns
DR5death receptor 5
DUSPdual specificity phosphatase
EASLEuropean association for the study of the liver
Egr-1early growth response protein 1
ELFenhanced liver fibrosis
ET-1endothelin-1
EtGethyl glucuronide
EtSethyl sulfate
FAEEsfatty acid ethyl esters
FASLGfas ligand
FIB-4fibrosis-4
FMTfecal microbiota transplantation
FXRfarnesoid X receptor
GABA-Bgamma-aminobutyric acid-B
G-CSFgranulocyte-colony stimulating factor
GSRglutathione reductase
GWASgenome-wide association studies
HAhyaluronic acid
HCChepatocellular carcinoma
HIF-1αhypoxia inducible factor 1α
HIVhuman immunodeficiency virus
HSCshepatic stellate cells
HSD17B13hydroxysteroid 17-β dehydrogenase 13
ILinterleukin
LPSlipopolysaccharide
LTliver transplantation
MBOAT7membrane-bound O-acyltransferase domain-containing protein 7
MeCP2methyl-CPG binding protein 2
MEOSmicrosomal ethanol-oxidizing system
miRNAmicroRNA
MMPmatrix metalloproteinases
MSCsmesenchymal stem cells
NACN-acetylcysteine
nAChRnicotinic acetylcholine receptors
NADnicotinamide adenine dinucleotide
NAFLDnonalcoholic fatty liver disease
NASHnonalcoholic steatohepatitis
NF-κBnuclear factor kappa-light-chain-enhancer of activated B cells
NMDAN-methyl-d-aspartate
Nrf-2nuclear factor erythroid 2-related factor 2
PICPtype I procollagen carboxypropeptide
PINPtype I procollagen aminopropeptide
PIIINPtype III procollagen aminopropeptide
PEthphosphatidylethanol
PDGFplatelet-derived growth factor
Phoxphagocytic oxidase
PI3Kphosphoinositide 3-kinase
PIGFplacenta growth factor
PNPLA3patatin-like phospholipase domain-containing-3
PORcytochrome P450 oxidoreductase
PPARαperoxisome proliferator activated receptor α
PTXpentoxifylline
RCTrandomized controlled trial
ROSreactive oxygen species
SAMs-adenosylmethionine
SIRT1sirtuin 1
STAT3signal transducer and activator of transcription 3
SSRIselective serotonin reuptake inhibitors
SWEshear-wave elastography
TEtransient elastography
TfR1transferrin receptor 1
TGF-βtransforming growth factor β
TGFβRIItype II transforming growth factor-β receptor
TIMPstissue inhibitor of metalloproteinases
TLRstoll-like receptors
TM6SF2transmembrane 6 superfamily member 2
TNF-αtumor necrosis factor α
VEGFvascular endothelial growth factor
WHOworld health organization
Yy1yin yang 1
ZEB2zinc finger E-box binding homeobox 2
ZO-1zonula occludens 1

References

  1. Rehm, J.; Gmel, G.E.; Gmel, G.; Hasan, O.S.M.; Imtiaz, S.; Popova, S.; Probst, C.; Roerecke, M.; Room, R.; Samokhvalov, A.V.; et al. The relationship between different dimensions of alcohol use and the burden of disease-an update. Addiction 2017, 112, 968–1001. [Google Scholar] [CrossRef] [Green Version]
  2. Thursz, M.; Gual, A.; Lackner, C.; Mathurin, P.; Moreno, C.; Spahr, L.; Sterneck, M.; Cortez-Pinto, H. EASL Clinical Practice Guidelines: Management of alcohol-related liver disease. J. Hepatol. 2018, 69, 154–181. [Google Scholar] [CrossRef] [Green Version]
  3. Poznyak, V.; Rekve, D. Global Status Report on Alcohol and Health 2018; World Health Organization: Geneva, Switzerland, 2018; ISBN 978-92-4-156563-9. [Google Scholar]
  4. Dolganiuc, A. Alcohol and Viral Hepatitis: Role of Lipid Rafts. Alcohol Res. Curr. Rev. 2015, 37, 299–309. [Google Scholar]
  5. Sahlman, P.; Nissinen, M.; Pukkala, E.; Färkkilä, M. Cancer incidence among alcoholic liver disease patients in Finland: A retrospective registry study during years 1996–2013. Int. J. Cancer 2016, 138, 2616–2621. [Google Scholar] [CrossRef] [Green Version]
  6. LoConte, N.K.; Brewster, A.M.; Kaur, J.S.; Merrill, J.K.; Alberg, A.J. Alcohol and Cancer: A Statement of the American Society of Clinical Oncology. J. Clin. Oncol. 2018, 36, 83–93. [Google Scholar] [CrossRef] [Green Version]
  7. Ganesan, M.; Eikenberry, A.; Poluektova, L.Y.; Kharbanda, K.K.; Osna, N.A. Role of alcohol in pathogenesis of hepatitis B virus infection. World J. Gastroenterol. 2020, 26, 883–903. [Google Scholar] [CrossRef]
  8. Meroni, M.; Longo, M. Genetic and Epigenetic Modifiers of Alcoholic Liver Disease. Int. J. Mol. Sci. 2018, 19, 3857. [Google Scholar] [CrossRef] [Green Version]
  9. Kong, L.Z.; Chandimali, N.; Han, Y.H.; Lee, D.H.; Kim, J.S.; Kim, S.U.; Kim, T.D.; Jeong, D.K.; Sun, H.N.; Lee, D.; et al. Pathogenesis, Early Diagnosis, and Therapeutic Management of Alcoholic Liver Disease. Int. J. Mol. Sci. 2019, 20, 2712. [Google Scholar] [CrossRef] [Green Version]
  10. Thursz, M.; Kamath, P.S.; Mathurin, P.; Szabo, G.; Shah, V.H. Alcohol-related liver disease: Areas of consensus, unmet needs and opportunities for further study. J. Hepatol. 2019, 70, 521–530. [Google Scholar] [CrossRef] [Green Version]
  11. Chacko, K.R.; Reinus, J. Spectrum of Alcoholic Liver Disease. Clin. Liver Dis. 2016, 20, 419–427. [Google Scholar] [CrossRef]
  12. Singal, A.K.; Bataller, R.; Ahn, J.; Kamath, P.S.; Shah, V.H. ACG Clinical Guideline: Alcoholic Liver Disease. Am. J. Gastroenterol. 2018, 113, 175–194. [Google Scholar] [CrossRef] [PubMed]
  13. Addolorato, G.; Mirijello, A.; Barrio, P.; Gual, A. Treatment of alcohol use disorders in patients with alcoholic liver disease. J. Hepatol. 2016, 65, 618–630. [Google Scholar] [CrossRef] [Green Version]
  14. Toshikuni, N.; Tsutsumi, M.; Arisawa, T. Clinical differences between alcoholic liver disease and nonalcoholic fatty liver disease. World J. Gastroenterol. 2014, 20, 8393–8406. [Google Scholar] [CrossRef] [PubMed]
  15. Tan, H.K.; Yates, E.; Lilly, K.; Dhanda, A.D. Oxidative stress in alcohol-related liver disease. World J. Hepatol. 2020, 12, 332–349. [Google Scholar] [CrossRef] [PubMed]
  16. Xu, T.; Li, L.; Hu, H.Q.; Meng, X.M.; Huang, C.; Zhang, L.; Qin, J.; Li, J. MicroRNAs in alcoholic liver disease: Recent advances and future applications. J. Cell. Physiol. 2018, 234, 382–394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Meroni, M.; Longo, M. Alcohol or Gut Microbiota: Who Is the Guilty? Int. J. Mol. Sci. 2019, 20, 4568. [Google Scholar] [CrossRef] [Green Version]
  18. Ezquer, F.; Bruna, F.; Calligaris, S.; Conget, P.; Ezquer, M. Multipotent mesenchymal stromal cells: A promising strategy to manage alcoholic liver disease. World J. Gastroenterol. 2016, 22, 24–36. [Google Scholar] [CrossRef]
  19. Teschke, R. Alcoholic Liver Disease: Alcohol Metabolism, Cascade of Molecular Mechanisms, Cellular Targets, and Clinical Aspects. Biomedicines 2018, 6, 106. [Google Scholar] [CrossRef] [Green Version]
  20. Buchanan, R.; Sinclair, J.M.A. Alcohol use disorder and the liver. Addiction 2021, 116, 1270–1278. [Google Scholar] [CrossRef] [PubMed]
  21. Jiang, Y.; Zhang, T.; Kusumanchi, P.; Han, S.; Yang, Z.; Liangpunsakul, S. Alcohol Metabolizing Enzymes, Microsomal Ethanol Oxidizing System, Cytochrome P450 2E1, Catalase, and Aldehyde Dehydrogenase in Alcohol-Associated Liver Disease. Biomedicines 2020, 8, 50. [Google Scholar] [CrossRef] [Green Version]
  22. Ceni, E.; Mello, T.; Galli, A. Pathogenesis of alcoholic liver disease: Role of oxidative metabolism. World J. Gastroenterol. 2014, 20, 17756–17772. [Google Scholar] [CrossRef]
  23. Louvet, A.; Mathurin, P. Alcoholic liver disease: Mechanisms of injury and targeted treatment. Nat. Rev. Gastroenterol. Hepatol. 2015, 12, 231–242. [Google Scholar] [CrossRef] [PubMed]
  24. Nagy, L.E.; Ding, W.X.; Cresci, G.; Saikia, P.; Shah, V.H. Linking Pathogenic Mechanisms of Alcoholic Liver Disease With Clinical Phenotypes. Gastroenterology 2016, 150, 1756–1768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Liu, J. Ethanol and liver: Recent insights into the mechanisms of ethanol-induced fatty liver. World J. Gastroenterol. 2014, 20, 14672–14685. [Google Scholar] [CrossRef]
  26. Seitz, H.K.; Bataller, R.; Cortez-Pinto, H.; Gao, B.; Gual, A.; Lackner, C.; Mathurin, P.; Mueller, S.; Szabo, G.; Tsukamoto, H. Alcoholic liver disease. Nat. Rev. Dis. Primers 2018, 4, 16. [Google Scholar] [CrossRef]
  27. Wang, Z.G.; Dou, X.B.; Zhou, Z.X.; Song, Z.Y. Adipose tissue-liver axis in alcoholic liver disease. World J. Gastrointest. Pathophysiol. 2016, 7, 17–26. [Google Scholar] [CrossRef]
  28. Gao, B.; Xu, M.J.; Bertola, A.; Wang, H.; Zhou, Z.; Liangpunsakul, S. Animal Models of Alcoholic Liver Disease: Pathogenesis and Clinical Relevance. Gene Expr. 2017, 17, 173–186. [Google Scholar] [CrossRef]
  29. Linhart, K.; Bartsch, H.; Seitz, H.K. The role of reactive oxygen species (ROS) and cytochrome P-450 2E1 in the generation of carcinogenic etheno-DNA adducts. Redox Biol. 2014, 3, 56–62. [Google Scholar] [CrossRef] [Green Version]
  30. Mueller, S.; Peccerella, T.; Qin, H.; Glassen, K.; Waldherr, R.; Flechtenmacher, C.; Straub, B.K.; Millonig, G.; Stickel, F.; Bruckner, T.; et al. Carcinogenic Etheno DNA Adducts in Alcoholic Liver Disease: Correlation with Cytochrome P-4502E1 and Fibrosis. Alcohol. Clin. Exp. Res. 2018, 42, 252–259. [Google Scholar] [CrossRef]
  31. Seitz, H.K.; Stickel, F. Risk factors and mechanisms of hepatocarcinogenesis with special emphasis on alcohol and oxidative stress. Biol. Chem. 2006, 387, 349–360. [Google Scholar] [CrossRef]
  32. Niederreiter, L.; Tilg, H. Cytokines and fatty liver diseases. Liver Res. 2018, 2, 14–20. [Google Scholar] [CrossRef]
  33. Naseem, S.; Hussain, T.; Manzoor, S. Interleukin-6: A promising cytokine to support liver regeneration and adaptive immunity in liver pathologies. Cytokine Growth Factor Rev. 2018, 39, 36–45. [Google Scholar] [CrossRef] [PubMed]
  34. Barbier, L.; Ferhat, M.; Salamé, E.; Robin, A.; Herbelin, A.; Gombert, J.M.; Silvain, C.; Barbarin, A. Interleukin-1 Family Cytokines: Keystones in Liver Inflammatory Diseases. Front. Immunol. 2019, 10, 2014. [Google Scholar] [CrossRef] [PubMed]
  35. Lackner, C.; Tiniakos, D. Fibrosis and alcohol-related liver disease. J. Hepatol. 2019, 70, 294–304. [Google Scholar] [CrossRef]
  36. Parola, M.; Pinzani, M. Liver fibrosis: Pathophysiology, pathogenetic targets and clinical issues. Mol. Asp. Med. 2019, 65, 37–55. [Google Scholar] [CrossRef]
  37. Harrison-Findik, D.D.; Klein, E.; Crist, C.; Evans, J.; Timchenko, N.; Gollan, J. Iron-mediated regulation of liver hepcidin expression in rats and mice is abolished by alcohol. Hepatology 2007, 46, 1979–1985. [Google Scholar] [CrossRef]
  38. Harrison-Findik, D.D. Role of alcohol in the regulation of iron metabolism. World J. Gastroenterol. 2007, 13, 4925–4930. [Google Scholar] [CrossRef]
  39. Silva, I.; Rausch, V.; Seitz, H.K.; Mueller, S. Does Hypoxia Cause Carcinogenic Iron Accumulation in Alcoholic Liver Disease (ALD)? Cancers 2017, 9, 145. [Google Scholar] [CrossRef] [Green Version]
  40. Stickel, F.; Moreno, C.; Hampe, J.; Morgan, M.Y. The genetics of alcohol dependence and alcohol-related liver disease. J. Hepatol. 2017, 66, 195–211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Stickel, F.; Buch, S.; Lau, K.; Meyer zu Schwabedissen, H.; Berg, T.; Ridinger, M.; Rietschel, M.; Schafmayer, C.; Braun, F.; Hinrichsen, H.; et al. Genetic variation in the PNPLA3 gene is associated with alcoholic liver injury in Caucasians. Hepatology 2011, 53, 86–95. [Google Scholar] [CrossRef]
  42. Salameh, H.; Raff, E.; Erwin, A.; Seth, D.; Nischalke, H.D.; Falleti, E.; Burza, M.A.; Leathert, J.; Romeo, S.; Molinaro, A.; et al. PNPLA3 Gene Polymorphism Is Associated With Predisposition to and Severity of Alcoholic Liver Disease. Am. J. Gastroenterol. 2015, 110, 846–856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Buch, S.; Stickel, F.; Trépo, E. A genome-wide association study confirms PNPLA3 and identifies TM6SF2 and MBOAT7 as risk loci for alcohol-related cirrhosis. Nat. Genet. 2015, 47, 1443–1448. [Google Scholar] [CrossRef] [PubMed]
  44. Abul-Husn, N.S.; Cheng, X.; Li, A.H.; Xin, Y.; Schurmann, C.; Stevis, P.; Liu, Y.; Kozlitina, J.; Stender, S.; Wood, G.C.; et al. A Protein-Truncating HSD17B13 Variant and Protection from Chronic Liver Disease. N. Engl. J. Med. 2018, 378, 1096–1106. [Google Scholar] [CrossRef] [PubMed]
  45. BasuRay, S.; Wang, Y.; Smagris, E.; Cohen, J.C.; Hobbs, H.H. Accumulation of PNPLA3 on lipid droplets is the basis of associated hepatic steatosis. Proc. Natl. Acad. Sci. USA 2019, 116, 9521–9526. [Google Scholar] [CrossRef] [Green Version]
  46. Pirazzi, C.; Valenti, L.; Motta, B.M.; Pingitore, P.; Hedfalk, K.; Mancina, R.M.; Burza, M.A.; Indiveri, C.; Ferro, Y.; Montalcini, T.; et al. PNPLA3 has retinyl-palmitate lipase activity in human hepatic stellate cells. Hum. Mol. Genet. 2014, 23, 4077–4085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Grimaudo, S.; Pipitone, R.M.; Pennisi, G.; Celsa, C.; Cammà, C.; Di Marco, V.; Barcellona, M.R.; Boemi, R.; Enea, M.; Giannetti, A.; et al. Association Between PNPLA3 rs738409 C>G Variant and Liver-Related Outcomes in Patients With Nonalcoholic Fatty Liver Disease. Clin. Gastroenterol. Hepatol. 2020, 18, 935–944. [Google Scholar] [CrossRef]
  48. Ali, M.; Yopp, A.; Gopal, P.; Beg, M.S.; Zhu, H.; Lee, W.; Singal, A.G. A Variant in PNPLA3 Associated With Fibrosis Progression but not Hepatocellular Carcinoma in Patients With Hepatitis C Virus Infection. Clin. Gastroenterol. Hepatol. 2016, 14, 295–300. [Google Scholar] [CrossRef] [Green Version]
  49. Chamorro, A.J.; Torres, J.L.; Mirón-Canelo, J.A.; González-Sarmiento, R.; Laso, F.J.; Marcos, M. Systematic review with meta-analysis: The I148M variant of patatin-like phospholipase domain-containing 3 gene (PNPLA3) is significantly associated with alcoholic liver cirrhosis. Aliment. Pharmacol. Ther. 2014, 40, 571–581. [Google Scholar] [CrossRef]
  50. Trépo, E.; Nahon, P.; Bontempi, G.; Valenti, L.; Falleti, E.; Nischalke, H.D.; Hamza, S.; Corradini, S.G.; Burza, M.A.; Guyot, E.; et al. Association between the PNPLA3 (rs738409 C>G) variant and hepatocellular carcinoma: Evidence from a meta-analysis of individual participant data. Hepatology 2014, 59, 2170–2177. [Google Scholar] [CrossRef]
  51. Falleti, E.; Cussigh, A.; Cmet, S.; Fabris, C.; Toniutto, P. PNPLA3 rs738409 and TM6SF2 rs58542926 variants increase the risk of hepatocellular carcinoma in alcoholic cirrhosis. Dig. Liver Dis. 2016, 48, 69–75. [Google Scholar] [CrossRef]
  52. O’Hare, E.A.; Yang, R.; Yerges-Armstrong, L.M.; Sreenivasan, U.; McFarland, R.; Leitch, C.C.; Wilson, M.H.; Narina, S.; Gorden, A.; Ryan, K.A.; et al. TM6SF2 rs58542926 impacts lipid processing in liver and small intestine. Hepatology 2017, 65, 1526–1542. [Google Scholar] [CrossRef] [Green Version]
  53. Liu, Y.L.; Reeves, H.L.; Burt, A.D.; Tiniakos, D.; McPherson, S.; Leathart, J.B.; Allison, M.E.; Alexander, G.J.; Piguet, A.C.; Anty, R.; et al. TM6SF2 rs58542926 influences hepatic fibrosis progression in patients with non-alcoholic fatty liver disease. Nat. Commun. 2014, 5, 4309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Goffredo, M.; Caprio, S.; Feldstein, A.E.; D’Adamo, E.; Shaw, M.M.; Pierpont, B.; Savoye, M.; Zhao, H.; Bale, A.E.; Santoro, N. Role of TM6SF2 rs58542926 in the pathogenesis of nonalcoholic pediatric fatty liver disease: A multiethnic study. Hepatology 2016, 63, 117–125. [Google Scholar] [CrossRef] [Green Version]
  55. Musso, G.; Cipolla, U.; Cassader, M.; Pinach, S.; Saba, F.; De Michieli, F.; Paschetta, E.; Bongiovanni, D.; Framarin, L.; Leone, N.; et al. TM6SF2 rs58542926 variant affects postprandial lipoprotein metabolism and glucose homeostasis in NAFLD. J. Lipid Res. 2017, 58, 1221–1229. [Google Scholar] [CrossRef] [Green Version]
  56. Mancina, R.M.; Ferri, F.; Farcomeni, A.; Molinaro, A.; Maffongelli, A.; Mischitelli, M.; Poli, E.; Parlati, L.; Burza, M.A.; De Santis, A.; et al. A two gene-based risk score predicts alcoholic cirrhosis development in males with at-risk alcohol consumption. Appl. Clin. Genet. 2019, 12, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Yang, J.; Trépo, E.; Nahon, P.; Cao, Q.; Moreno, C.; Letouzé, E.; Imbeaud, S.; Gustot, T.; Deviere, J.; Debette, S.; et al. PNPLA3 and TM6SF2 variants as risk factors of hepatocellular carcinoma across various etiologies and severity of underlying liver diseases. Int. J. Cancer 2019, 144, 533–544. [Google Scholar] [CrossRef]
  58. Hill-Baskin, A.E.; Markiewski, M.M.; Buchner, D.A.; Shao, H.; DeSantis, D.; Hsiao, G.; Subramaniam, S.; Berger, N.A.; Croniger, C.; Lambris, J.D.; et al. Diet-induced hepatocellular carcinoma in genetically predisposed mice. Hum. Mol. Genet. 2009, 18, 2975–2988. [Google Scholar] [CrossRef] [Green Version]
  59. Scott, E.; Anstee, Q.M. Genetics of alcoholic liver disease and non-alcoholic steatohepatitis. Clin. Med. 2018, 18, S54–S59. [Google Scholar] [CrossRef]
  60. Thabet, K.; Asimakopoulos, A.; Shojaei, M.; Romero-Gomez, M.; Mangia, A.; Irving, W.L.; Berg, T.; Dore, G.J.; Grønbæk, H.; Sheridan, D.; et al. MBOAT7 rs641738 increases risk of liver inflammation and transition to fibrosis in chronic hepatitis C. Nat. Commun. 2016, 7, 12757. [Google Scholar] [CrossRef] [Green Version]
  61. Mancina, R.M.; Dongiovanni, P.; Petta, S.; Pingitore, P.; Meroni, M.; Rametta, R.; Borén, J.; Montalcini, T.; Pujia, A.; Wiklund, O.; et al. The MBOAT7-TMC4 Variant rs641738 Increases Risk of Nonalcoholic Fatty Liver Disease in Individuals of European Descent. Gastroenterology 2016, 150, 1219–1230. [Google Scholar] [CrossRef] [Green Version]
  62. Krawczyk, M.; Rau, M.; Schattenberg, J.M.; Bantel, H.; Pathil, A.; Demir, M.; Kluwe, J.; Boettler, T.; Lammert, F.; Geier, A. Combined effects of the PNPLA3 rs738409, TM6SF2 rs58542926, and MBOAT7 rs641738 variants on NAFLD severity: A multicenter biopsy-based study. J. Lipid Res. 2017, 58, 247–255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Basyte-Bacevice, V.; Skieceviciene, J.; Valantiene, I.; Sumskiene, J.; Petrenkiene, V.; Kondrackiene, J.; Petrauskas, D.; Lammert, F.; Kupcinskas, J. TM6SF2 and MBOAT7 Gene Variants in Liver Fibrosis and Cirrhosis. Int. J. Mol. Sci. 2019, 20, 1277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Ma, Y.; Belyaeva, O.V.; Brown, P.M.; Fujita, K.; Valles, K.; Karki, S.; de Boer, Y.S.; Koh, C.; Chen, Y.; Du, X.; et al. 17-Beta Hydroxysteroid Dehydrogenase 13 Is a Hepatic Retinol Dehydrogenase Associated With Histological Features of Nonalcoholic Fatty Liver Disease. Hepatology 2019, 69, 1504–1519. [Google Scholar] [CrossRef] [PubMed]
  65. Stickel, F.; Lutz, P.; Buch, S.; Nischalke, H.D.; Silva, I.; Rausch, V.; Fischer, J.; Weiss, K.H.; Gotthardt, D.; Rosendahl, J.; et al. Genetic Variation in HSD17B13 Reduces the Risk of Developing Cirrhosis and Hepatocellular Carcinoma in Alcohol Misusers. Hepatology 2020, 72, 88–102. [Google Scholar] [CrossRef]
  66. Chen, H.; Zhang, Y.; Guo, T.; Yang, F.; Mao, Y.; Li, L.; Liu, C.; Gao, H.; Jin, Y.; Che, Y.; et al. Genetic variant rs72613567 of HSD17B13 gene reduces alcohol-related liver disease risk in Chinese Han population. Liver Int. 2020, 40, 2194–2202. [Google Scholar] [CrossRef] [PubMed]
  67. Berkel, T.D.; Pandey, S.C. Emerging Role of Epigenetic Mechanisms in Alcohol Addiction. Alcohol. Clin. Exp. Res. 2017, 41, 666–680. [Google Scholar] [CrossRef] [Green Version]
  68. Hardy, T.; Mann, D.A. Epigenetics in liver disease: From biology to therapeutics. Gut 2016, 65, 1895–1905. [Google Scholar] [CrossRef] [PubMed]
  69. Choudhary, N.S.; Duseja, A. Genetic and epigenetic disease modifiers: Non-alcoholic fatty liver disease (NAFLD) and alcoholic liver disease (ALD). Transl. Gastroenterol. Hepatol. 2021, 6, 2. [Google Scholar] [CrossRef]
  70. Li, H.D.; Du, X.S.; Huang, H.M.; Chen, X.; Yang, Y.; Huang, C.; Meng, X.M.; Li, J. Noncoding RNAs in alcoholic liver disease. J. Cell. Physiol. 2019, 234, 14709–14720. [Google Scholar] [CrossRef]
  71. Shukla, S.D.; Velazquez, J.; French, S.W.; Lu, S.C.; Ticku, M.K.; Zakhari, S. Emerging role of epigenetics in the actions of alcohol. Alcohol. Clin. Exp. Res. 2008, 32, 1525–1534. [Google Scholar] [CrossRef]
  72. Lu, S.C.; Huang, Z.Z.; Yang, H.; Mato, J.M.; Avila, M.A.; Tsukamoto, H. Changes in methionine adenosyltransferase and S-adenosylmethionine homeostasis in alcoholic rat liver. Am. J. Physiol. Gastrointest. Liver Physiol. 2000, 279, G178–G185. [Google Scholar] [CrossRef] [Green Version]
  73. Zakhari, S. Alcohol metabolism and epigenetics changes. Alcohol Res. Curr. Rev. 2013, 35, 6–16. [Google Scholar]
  74. Sidharthan, S.; Kottilil, S. Mechanisms of alcohol-induced hepatocellular carcinoma. Hepatol. Int. 2014, 8, 452–457. [Google Scholar] [CrossRef] [Green Version]
  75. Mandrekar, P. Epigenetic regulation in alcoholic liver disease. World J. Gastroenterol. 2011, 17, 2456–2464. [Google Scholar] [CrossRef]
  76. Brandon-Warner, E.; Feilen, N.A.; Culberson, C.R.; Field, C.O.; de Lemos, A.S.; Russo, M.W.; Schrum, L.W. Processing of miR17–92 Cluster in Hepatic Stellate Cells Promotes Hepatic Fibrogenesis During Alcohol-Induced Injury. Alcohol. Clin. Exp. Res. 2016, 40, 1430–1442. [Google Scholar] [CrossRef] [Green Version]
  77. Dippold, R.P.; Vadigepalli, R.; Gonye, G.E.; Hoek, J.B. Chronic ethanol feeding enhances miR-21 induction during liver regeneration while inhibiting proliferation in rats. Am. J. Physiol. Gastrointest. Liver Physiol. 2012, 303, G733–G743. [Google Scholar] [CrossRef] [Green Version]
  78. Francis, H.; McDaniel, K.; Han, Y.; Liu, X.; Kennedy, L.; Yang, F.; McCarra, J.; Zhou, T.; Glaser, S.; Venter, J.; et al. Regulation of the extrinsic apoptotic pathway by microRNA-21 in alcoholic liver injury. J. Biol. Chem. 2014, 289, 27526–27539. [Google Scholar] [CrossRef] [Green Version]
  79. Juskeviciute, E.; Dippold, R.P.; Antony, A.N.; Swarup, A.; Vadigepalli, R. Inhibition of miR-21 rescues liver regeneration after partial hepatectomy in ethanol-fed rats. Am. J. Physiol. Gastrointest. Liver Physiol. 2016, 311, G794–G806. [Google Scholar] [CrossRef] [Green Version]
  80. Han, W.; Fu, X.; Xie, J.; Meng, Z.; Gu, Y.; Wang, X.; Li, L.; Pan, H.; Huang, W. MiR-26a enhances autophagy to protect against ethanol-induced acute liver injury. J. Mol. Med. 2015, 93, 1045–1055. [Google Scholar] [CrossRef] [Green Version]
  81. Saha, B.; Bruneau, J.C.; Kodys, K.; Szabo, G. Alcohol-induced miR-27a regulates differentiation and M2 macrophage polarization of normal human monocytes. J. Immunol. 2015, 194, 3079–3087. [Google Scholar] [CrossRef] [Green Version]
  82. Saha, B.; Momen-Heravi, F.; Kodys, K.; Szabo, G. MicroRNA Cargo of Extracellular Vesicles from Alcohol-exposed Monocytes Signals Naive Monocytes to Differentiate into M2 Macrophages. J. Biol. Chem. 2016, 291, 149–159. [Google Scholar] [CrossRef] [Green Version]
  83. Roderburg, C.; Urban, G.W.; Bettermann, K.; Vucur, M.; Zimmermann, H.; Schmidt, S.; Janssen, J.; Koppe, C.; Knolle, P.; Castoldi, M.; et al. Micro-RNA profiling reveals a role for miR-29 in human and murine liver fibrosis. Hepatology 2011, 53, 209–218. [Google Scholar] [CrossRef]
  84. Meng, F.; Glaser, S.S.; Francis, H.; Yang, F.; Han, Y.; Stokes, A.; Staloch, D.; McCarra, J.; Liu, J.; Venter, J.; et al. Epigenetic regulation of miR-34a expression in alcoholic liver injury. Am. J. Pathol. 2012, 181, 804–817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Dippold, R.P.; Vadigepalli, R.; Gonye, G.E.; Patra, B.; Hoek, J.B. Chronic ethanol feeding alters miRNA expression dynamics during liver regeneration. Alcohol. Clin. Exp. Res. 2013, 37, E59–E69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Yin, H.; Hu, M.; Liang, X.; Ajmo, J.M.; Li, X.; Bataller, R.; Odena, G.; Stevens, S.M.J.; You, M. Deletion of SIRT1 from hepatocytes in mice disrupts lipin-1 signaling and aggravates alcoholic fatty liver. Gastroenterology 2014, 146, 801–811. [Google Scholar] [CrossRef] [Green Version]
  87. Wan, Y.; McDaniel, K.; Wu, N.; Ramos-Lorenzo, S.; Glaser, T.; Venter, J.; Francis, H.; Kennedy, L.; Sato, K.; Zhou, T.; et al. Regulation of Cellular Senescence by miR-34a in Alcoholic Liver Injury. Am. J. Pathol. 2017, 187, 2788–2798. [Google Scholar] [CrossRef] [Green Version]
  88. Momen-Heravi, F.; Bala, S.; Kodys, K.; Szabo, G. Exosomes derived from alcohol-treated hepatocytes horizontally transfer liver specific miRNA-122 and sensitize monocytes to LPS. Sci. Rep. 2015, 5, 9991. [Google Scholar] [CrossRef] [Green Version]
  89. Ambade, A.; Satishchandran, A.; Szabo, G. Alcoholic hepatitis accelerates early hepatobiliary cancer by increasing stemness and miR-122-mediated HIF-1α activation. Sci. Rep. 2016, 6, 21340. [Google Scholar] [CrossRef] [Green Version]
  90. Satishchandran, A.; Ambade, A.; Rao, S.; Hsueh, Y.C.; Iracheta-Vellve, A.; Tornai, D.; Lowe, P.; Gyongyosi, B.; Li, J.; Catalano, D.; et al. MicroRNA 122, Regulated by GRLH2, Protects Livers of Mice and Patients From Ethanol-Induced Liver Disease. Gastroenterology 2018, 154, 238–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Alpini, G.; Glaser, S.S.; Zhang, J.P.; Francis, H.; Han, Y.; Gong, J.; Stokes, A.; Francis, T.; Hughart, N.; Hubble, L.; et al. Regulation of placenta growth factor by microRNA-125b in hepatocellular cancer. J. Hepatol. 2011, 55, 1339–1345. [Google Scholar] [CrossRef] [Green Version]
  92. Ladeiro, Y.; Couchy, G.; Balabaud, C.; Bioulac-Sage, P.; Pelletier, L.; Rebouissou, S.; Zucman-Rossi, J. MicroRNA profiling in hepatocellular tumors is associated with clinical features and oncogene/tumor suppressor gene mutations. Hepatology 2008, 47, 1955–1963. [Google Scholar] [CrossRef]
  93. Guo, C.J.; Pan, Q.; Xiong, H.; Qiao, Y.Q.; Bian, Z.L.; Zhong, W.; Sheng, L.; Li, H.; Shen, L.; Hua, J.; et al. Dynamic expression of miR-126* and its effects on proliferation and contraction of hepatic stellate cells. FEBS Lett. 2013, 587, 3792–3801. [Google Scholar] [CrossRef] [Green Version]
  94. Bala, S.; Marcos, M.; Kodys, K.; Csak, T.; Catalano, D.; Mandrekar, P.; Szabo, G. Up-regulation of MicroRNA-155 in Macrophages Contributes to Increased Tumor Necrosis Factor α (TNFα) Production via Increased mRNA Half-life in Alcoholic Liver Disease. J. Biol. Chem. 2011, 286, 1436–1444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Lippai, D.; Bala, S.; Catalano, D.; Kodys, K.; Szabo, G. Micro-RNA-155 deficiency prevents alcohol-induced serum endotoxin increase and small bowel inflammation in mice. Alcohol. Clin. Exp. Res. 2014, 38, 2217–2224. [Google Scholar] [CrossRef] [Green Version]
  96. Csak, T.; Bala, S.; Lippai, D.; Kodys, K.; Catalano, D.; Iracheta-Vellve, A.; Szabo, G. MicroRNA-155 Deficiency Attenuates Liver Steatosis and Fibrosis without Reducing Inflammation in a Mouse Model of Steatohepatitis. PLoS ONE 2015, 10, e0129251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Bala, S.; Csak, T.; Saha, B.; Zatsiorsky, J.; Kodys, K.; Catalano, D.; Satishchandran, A.; Szabo, G. The pro-inflammatory effects of miR-155 promote liver fibrosis and alcohol-induced steatohepatitis. J. Hepatol. 2016, 64, 1378–1387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Bala, S.; Csak, T.; Kodys, K.; Catalano, D.; Ambade, A.; Furi, I.; Lowe, P.; Cho, Y.; Iracheta-Vellve, A.; Szabo, G. Alcohol-induced miR-155 and HDAC11 inhibit negative regulators of the TLR4 pathway and lead to increased LPS responsiveness of Kupffer cells in alcoholic liver disease. J. Leukoc. Biol. 2017, 102, 487–498. [Google Scholar] [CrossRef] [Green Version]
  99. Saikia, P.; Bellos, D.; McMullen, M.R.; Pollard, K.A.; de la Motte, C.; Nagy, L.E. MicroRNA 181b-3p and its target importin α5 regulate toll-like receptor 4 signaling in Kupffer cells and liver injury in mice in response to ethanol. Hepatology 2017, 66, 602–615. [Google Scholar] [CrossRef]
  100. Dolganiuc, A.; Petrasek, J.; Kodys, K.; Catalano, D.; Mandrekar, P.; Velayudham, A.; Szabo, G. MicroRNA expression profile in Lieber-DeCarli diet-induced alcoholic and methionine choline deficient diet-induced nonalcoholic steatohepatitis models in mice. Alcohol. Clin. Exp. Res. 2009, 33, 1704–1710. [Google Scholar] [CrossRef] [Green Version]
  101. Blaya, D.; Coll, M.; Rodrigo-Torres, D.; Vila-Casadesús, M.; Altamirano, J.; Llopis, M.; Graupera, I.; Perea, L.; Aguilar-Bravo, B.; Díaz, A.; et al. Integrative microRNA profiling in alcoholic hepatitis reveals a role for microRNA-182 in liver injury and inflammation. Gut 2016, 65, 1535–1545. [Google Scholar] [CrossRef]
  102. Yeligar, S.; Tsukamoto, H.; Kalra, V.K. Ethanol-induced expression of ET-1 and ET-BR in liver sinusoidal endothelial cells and human endothelial cells involves hypoxia-inducible factor-1alpha and microrNA-199. J. Immunol. 2009, 183, 5232–5243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Huang, G.-H.; Shan, H.; Li, D.; Zhou, B.; Pang, P.-F. MiR-199a-5p suppresses tumorigenesis by targeting clathrin heavy chain in hepatocellular carcinoma. Cell Biochem. Funct. 2017, 35, 98–104. [Google Scholar] [CrossRef] [PubMed]
  104. Tang, Y.; Banan, A.; Forsyth, C.B.; Fields, J.Z.; Lau, C.K.; Zhang, L.J.; Keshavarzian, A. Effect of alcohol on miR-212 expression in intestinal epithelial cells and its potential role in alcoholic liver disease. Alcohol. Clin. Exp. Res. 2008, 32, 355–364. [Google Scholar] [CrossRef]
  105. Tang, Y.; Zhang, L.; Forsyth, C.B.; Shaikh, M.; Song, S.; Keshavarzian, A. The Role of miR-212 and iNOS in Alcohol-Induced Intestinal Barrier Dysfunction and Steatohepatitis. Alcohol. Clin. Exp. Res. 2015, 39, 1632–1641. [Google Scholar] [CrossRef]
  106. Dong, X.; Liu, H.; Chen, F.; Li, D.; Zhao, Y. MiR-214 promotes the alcohol-induced oxidative stress via down-regulation of glutathione reductase and cytochrome P450 oxidoreductase in liver cells. Alcohol. Clin. Exp. Res. 2014, 38, 68–77. [Google Scholar] [CrossRef]
  107. Yin, H.; Hu, M.; Zhang, R.; Shen, Z.; Flatow, L.; You, M. MicroRNA-217 promotes ethanol-induced fat accumulation in hepatocytes by down-regulating SIRT1. J. Biol. Chem. 2012, 287, 9817–9826. [Google Scholar] [CrossRef] [Green Version]
  108. Yin, H.; Liang, X.; Jogasuria, A.; Davidson, N.O.; You, M. miR-217 regulates ethanol-induced hepatic inflammation by disrupting sirtuin 1-lipin-1 signaling. Am. J. Pathol. 2015, 185, 1286–1296. [Google Scholar] [CrossRef] [Green Version]
  109. Li, M.; He, Y.; Zhou, Z.; Ramirez, T.; Gao, Y.; Gao, Y.; Ross, R.A.; Cao, H.; Cai, Y.; Xu, M.; et al. MicroRNA-223 ameliorates alcoholic liver injury by inhibiting the IL-6-p47(phox)-oxidative stress pathway in neutrophils. Gut 2017, 66, 705–715. [Google Scholar] [CrossRef] [Green Version]
  110. Saikia, P.; Roychowdhury, S.; Bellos, D.; Pollard, K.A.; McMullen, M.R.; McCullough, R.L.; McCullough, A.J.; Gholam, P.; de la Motte, C.; Nagy, L.E. Hyaluronic acid 35 normalizes TLR4 signaling in Kupffer cells from ethanol-fed rats via regulation of microRNA291b and its target Tollip. Sci. Rep. 2017, 7, 15671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Hyun, J.; Wang, S.; Kim, J.; Rao, K.M.; Park, S.Y.; Chung, I.; Ha, C.S.; Kim, S.W.; Yun, Y.H.; Jung, Y. MicroRNA-378 limits activation of hepatic stellate cells and liver fibrosis by suppressing Gli3 expression. Nat. Commun. 2016, 7, 10993. [Google Scholar] [CrossRef] [Green Version]
  112. Kim, Y.D.; Hwang, S.L.; Lee, E.J.; Kim, H.M.; Chung, M.J.; Elfadl, A.K.; Lee, S.E.; Nedumaran, B.; Harris, R.A.; Jeong, K.S. Melatonin ameliorates alcohol-induced bile acid synthesis by enhancing miR-497 expression. J. Pineal Res. 2017, 62. [Google Scholar] [CrossRef]
  113. Wilsnack, R.W.; Wilsnack, S.C.; Gmel, G.; Kantor, L.W. Gender Differences in Binge Drinking. Alcohol Res. Curr. Rev. 2018, 39, 57–76. [Google Scholar]
  114. Harman, D.J.; Ryder, S.D.; James, M.W.; Wilkes, E.A.; Card, T.R.; Aithal, G.P.; Guha, I.N. Obesity and type 2 diabetes are important risk factors underlying previously undiagnosed cirrhosis in general practice: A cross-sectional study using transient elastography. Aliment. Pharmacol. Ther. 2018, 47, 504–515. [Google Scholar] [CrossRef] [Green Version]
  115. Ntandja Wandji, L.C.; Gnemmi, V.; Mathurin, P.; Louvet, A. Combined alcoholic and non-alcoholic steatohepatitis. JHEP Rep. Innov. Hepatol. 2020, 2, 100101. [Google Scholar] [CrossRef] [PubMed]
  116. Åberg, F.; Helenius-Hietala, J.; Puukka, P.; Färkkilä, M.; Jula, A. Interaction between alcohol consumption and metabolic syndrome in predicting severe liver disease in the general population. Hepatology 2018, 67, 2141–2149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Hagström, H. Alcohol Consumption in Concomitant Liver Disease: How Much is Too Much? Curr. Hepatol. Rep. 2017, 16, 152–157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Marshall, B.D.L.; Tate, J.P.; McGinnis, K.A.; Bryant, K.J.; Cook, R.L.; Edelman, E.J.; Gaither, J.R.; Kahler, C.W.; Operario, D.; Fiellin, D.A.; et al. Long-term alcohol use patterns and HIV disease severity. AIDS 2017, 31, 1313–1321. [Google Scholar] [CrossRef]
  119. Duko, B.; Ayalew, M.; Ayano, G. The prevalence of alcohol use disorders among people living with HIV/AIDS: A systematic review and meta-analysis. Subst. Abus. Treat. Prev. Policy 2019, 14, 52. [Google Scholar] [CrossRef]
  120. Alferink, L.J.M.; Fittipaldi, J.; Kiefte-de Jong, J.C.; Taimr, P.; Hansen, B.E.; Metselaar, H.J.; Schoufour, J.D.; Ikram, M.A.; Janssen, H.L.A.; Franco, O.H.; et al. Coffee and herbal tea consumption is associated with lower liver stiffness in the general population: The Rotterdam study. J. Hepatol. 2017, 67, 339–348. [Google Scholar] [CrossRef]
  121. Whitfield, J.B.; Masson, S.; Liangpunsakul, S.; Mueller, S.; Aithal, G.P.; Eyer, F.; Gleeson, D.; Thompson, A.; Stickel, F.; Soyka, M.; et al. Obesity, Diabetes, Coffee, Tea, and Cannabis Use Alter Risk for Alcohol-Related Cirrhosis in 2 Large Cohorts of High-Risk Drinkers. Am. J. Gastroenterol. 2021, 116, 106–115. [Google Scholar] [CrossRef]
  122. Kennedy, O.J.; Roderick, P.; Buchanan, R.; Fallowfield, J.A.; Hayes, P.C.; Parkes, J. Systematic review with meta-analysis: Coffee consumption and the risk of cirrhosis. Aliment. Pharmacol. Ther. 2016, 43, 562–574. [Google Scholar] [CrossRef] [Green Version]
  123. Gao, B.; Bataller, R. Alcoholic liver disease: Pathogenesis and new therapeutic targets. Gastroenterology 2011, 141, 1572–1585. [Google Scholar] [CrossRef] [Green Version]
  124. Gao, B.; Ahmad, M.F.; Nagy, L.E.; Tsukamoto, H. Inflammatory pathways in alcoholic steatohepatitis. J. Hepatol. 2019, 70, 249–259. [Google Scholar] [CrossRef] [Green Version]
  125. Crabb, D.W.; Bataller, R.; Chalasani, N.P.; Kamath, P.S.; Lucey, M.; Mathurin, P.; McClain, C.; McCullough, A.; Mitchell, M.C.; Morgan, T.R.; et al. Standard Definitions and Common Data Elements for Clinical Trials in Patients With Alcoholic Hepatitis: Recommendation From the NIAAA Alcoholic Hepatitis Consortia. Gastroenterology 2016, 150, 785–790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Sehrawat, T.S.; Liu, M.; Shah, V.H. The knowns and unknowns of treatment for alcoholic hepatitis. Lancet. Gastroenterol. Hepatol. 2020, 5, 494–506. [Google Scholar] [CrossRef]
  127. Fung, P.; Pyrsopoulos, N. Emerging concepts in alcoholic hepatitis. World J. Hepatol. 2017, 9, 567–585. [Google Scholar] [CrossRef] [PubMed]
  128. Lourens, S.; Sunjaya, D.B.; Singal, A.; Liangpunsakul, S.; Puri, P.; Sanyal, A.; Ren, X.; Gores, G.J.; Radaeva, S.; Chalasani, N.; et al. Acute Alcoholic Hepatitis: Natural History and Predictors of Mortality Using a Multicenter Prospective Study. Mayo Clinic Proc. Innov. Qual. Outcomes 2017, 1, 37–48. [Google Scholar] [CrossRef] [Green Version]
  129. Hosseini, N.; Shor, J.; Szabo, G. Alcoholic Hepatitis: A Review. Alcohol Alcohol. 2019, 54, 408–416. [Google Scholar] [CrossRef]
  130. Mellinger, J.L.; Shedden, K.; Winder, G.S.; Tapper, E.; Adams, M.; Fontana, R.J.; Volk, M.L.; Blow, F.C.; Lok, A.S.F. The high burden of alcoholic cirrhosis in privately insured persons in the United States. Hepatology 2018, 68, 872–882. [Google Scholar] [CrossRef] [Green Version]
  131. Moreau, R.; Jalan, R.; Gines, P.; Pavesi, M.; Angeli, P.; Cordoba, J.; Durand, F.; Gustot, T.; Saliba, F.; Domenicali, M.; et al. Acute-on-chronic liver failure is a distinct syndrome that develops in patients with acute decompensation of cirrhosis. Gastroenterology 2013, 144, 1426–1437. [Google Scholar] [CrossRef]
  132. Gustot, T.; Jalan, R. Acute-on-chronic liver failure in patients with alcohol-related liver disease. J. Hepatol. 2019, 70, 319–327. [Google Scholar] [CrossRef] [Green Version]
  133. Crabb, D.W.; Im, G.Y.; Szabo, G.; Mellinger, J.L.; Lucey, M.R. Diagnosis and Treatment of Alcohol-Associated Liver Diseases: 2019 Practice Guidance From the American Association for the Study of Liver Diseases. Hepatology 2020, 71, 306–333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Torruellas, C.; French, S.W.; Medici, V. Diagnosis of alcoholic liver disease. World J. Gastroenterol. 2014, 20, 11684–11699. [Google Scholar] [CrossRef]
  135. Askgaard, G.; Leon, D.A.; Kjaer, M.S.; Deleuran, T.; Gerds, T.A.; Tolstrup, J.S. Risk for alcoholic liver cirrhosis after an initial hospital contact with alcohol problems: A nationwide prospective cohort study. Hepatology 2017, 65, 929–937. [Google Scholar] [CrossRef] [Green Version]
  136. Moreno, C.; Mueller, S.; Szabo, G. Non-invasive diagnosis and biomarkers in alcohol-related liver disease. J. Hepatol. 2019, 70, 273–283. [Google Scholar] [CrossRef] [Green Version]
  137. Louvet, A.; Labreuche, J.; Artru, F.; Boursier, J.; Kim, D.J.; O’Grady, J.; Trépo, E.; Nahon, P.; Ganne-Carrié, N.; Naveau, S.; et al. Combining Data From Liver Disease Scoring Systems Better Predicts Outcomes of Patients With Alcoholic Hepatitis. Gastroenterology 2015, 149, 398–406. [Google Scholar] [CrossRef] [Green Version]
  138. Filingeri, V.; Francioso, S.; Sforza, D.; Santopaolo, F.; Oddi, F.M.; Tisone, G. A retrospective analysis of 1.011 percutaneous liver biopsies performed in patients with liver transplantation or liver disease: Ultrasonography can reduce complications? Eur. Rev. Med. Pharmacol. Sci. 2016, 20, 3609–3617. [Google Scholar]
  139. Lupşor-Platon, M.; Stefănescu, H.; Mureșan, D.; Florea, M.; Szász, M.E.; Maniu, A.; Badea, R. Noninvasive assessment of liver steatosis using ultrasound methods. Med. Ultrason. 2014, 16, 236–245. [Google Scholar]
  140. Dietrich, C.F.; Bamber, J.; Berzigotti, A.; Bota, S.; Cantisani, V.; Castera, L.; Cosgrove, D.; Ferraioli, G.; Friedrich-Rust, M.; Gilja, O.H.; et al. EFSUMB Guidelines and Recommendations on the Clinical Use of Liver Ultrasound Elastography, Update 2017 (Long Version). Ultraschall Med. 2017, 38, e16–e47. [Google Scholar]
  141. Noureddin, M.; Lam, J.; Peterson, M.R.; Middleton, M.; Hamilton, G.; Le, T.A.; Bettencourt, R.; Changchien, C.; Brenner, D.A.; Sirlin, C.; et al. Utility of magnetic resonance imaging versus histology for quantifying changes in liver fat in nonalcoholic fatty liver disease trials. Hepatology 2013, 58, 1930–1940. [Google Scholar] [CrossRef]
  142. Karlas, T.; Petroff, D.; Sasso, M.; Fan, J.G.; Mi, Y.Q.; de Lédinghen, V.; Kumar, M.; Lupsor-Platon, M.; Han, K.H.; Cardoso, A.C.; et al. Individual patient data meta-analysis of controlled attenuation parameter (CAP) technology for assessing steatosis. J. Hepatol. 2017, 66, 1022–1030. [Google Scholar] [CrossRef] [PubMed]
  143. Thiele, M.; Rausch, V.; Fluhr, G.; Kjærgaard, M.; Piecha, F.; Mueller, J.; Straub, B.K.; Lupșor-Platon, M.; De-Ledinghen, V.; Seitz, H.K.; et al. Controlled attenuation parameter and alcoholic hepatic steatosis: Diagnostic accuracy and role of alcohol detoxification. J. Hepatol. 2018, 68, 1025–1032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Pavlov, C.S.; Casazza, G.; Nikolova, D.; Tsochatzis, E.; Burroughs, A.K.; Ivashkin, V.T.; Gluud, C. Transient elastography for diagnosis of stages of hepatic fibrosis and cirrhosis in people with alcoholic liver disease. Cochrane Database Syst. Rev. 2015, 1, Cd010542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Janssens, F.; de Suray, N.; Piessevaux, H.; Horsmans, Y.; de Timary, P.; Stärkel, P. Can transient elastography replace liver histology for determination of advanced fibrosis in alcoholic patients: A real-life study. J. Clin. Gastroenterol. 2010, 44, 575–582. [Google Scholar] [CrossRef]
  146. Castera, L.; Pinzani, M. Biopsy and non-invasive methods for the diagnosis of liver fibrosis: Does it take two to tango? Gut 2010, 59, 861–866. [Google Scholar] [CrossRef]
  147. Serra-Burriel, M.; Graupera, I.; Torán, P.; Thiele, M.; Roulot, D.; Wai-Sun Wong, V.; Neil Guha, I.; Fabrellas, N.; Arslanow, A.; Expósito, C.; et al. Transient elastography for screening of liver fibrosis: Cost-effectiveness analysis from six prospective cohorts in Europe and Asia. J. Hepatol. 2019, 71, 1141–1151. [Google Scholar] [CrossRef]
  148. Dechêne, A.; Sowa, J.P.; Gieseler, R.K.; Jochum, C.; Bechmann, L.P.; El Fouly, A.; Schlattjan, M.; Saner, F.; Baba, H.A.; Paul, A.; et al. Acute liver failure is associated with elevated liver stiffness and hepatic stellate cell activation. Hepatology 2010, 52, 1008–1016. [Google Scholar] [CrossRef]
  149. Millonig, G.; Friedrich, S.; Adolf, S.; Fonouni, H.; Golriz, M.; Mehrabi, A.; Stiefel, P.; Pöschl, G.; Büchler, M.W.; Seitz, H.K.; et al. Liver stiffness is directly influenced by central venous pressure. J. Hepatol. 2010, 52, 206–210. [Google Scholar] [CrossRef] [PubMed]
  150. Zayed, N.; Darweesh, S.K.; Mousa, S.; Atef, M.; Ramzy, E.; Yosry, A. Liver stiffness measurement by acoustic radiation forced impulse and transient elastography in patients with intrahepatic cholestasis. Eur. J. Gastroenterol. Hepatol. 2019, 31, 520–527. [Google Scholar] [CrossRef]
  151. Michalak, S.; Rousselet, M.C.; Bedossa, P.; Pilette, C.; Chappard, D.; Oberti, F.; Gallois, Y.; Calès, P. Respective roles of porto-septal fibrosis and centrilobular fibrosis in alcoholic liver disease. J. Pathol. 2003, 201, 55–62. [Google Scholar] [CrossRef]
  152. Nguyen-Khac, E.; Thiele, M.; Voican, C.; Nahon, P.; Moreno, C.; Boursier, J.; Mueller, S.; de Ledinghen, V.; Stärkel, P.; Gyune Kim, S.; et al. Non-invasive diagnosis of liver fibrosis in patients with alcohol-related liver disease by transient elastography: An individual patient data meta-analysis. Lancet Gastroenterol. Hepatol. 2018, 3, 614–625. [Google Scholar] [CrossRef]
  153. Dhyani, M.; Xiang, F.; Li, Q.; Chen, L.; Li, C.; Bhan, A.K.; Anthony, B.; Grajo, J.R.; Samir, A.E. Ultrasound Shear Wave Elastography: Variations of Liver Fibrosis Assessment as a Function of Depth, Force and Distance from Central Axis of the Transducer with a Comparison of Different Systems. Ultrasound Med. Biol. 2018, 44, 2209–2222. [Google Scholar] [CrossRef] [PubMed]
  154. Bortolotti, F.; Sorio, D.; Bertaso, A.; Tagliaro, F. Analytical and diagnostic aspects of carbohydrate deficient transferrin (CDT): A critical review over years 2007–2017. J. Pharm. Biomed. Anal. 2018, 147, 2–12. [Google Scholar] [CrossRef] [PubMed]
  155. Niemelä, O. Biomarker-Based Approaches for Assessing Alcohol Use Disorders. Int. J. Environ. Res. Public Health 2016, 13, 166. [Google Scholar] [CrossRef] [PubMed]
  156. Helander, A.; Wielders, J.; Anton, R.; Arndt, T.; Bianchi, V.; Deenmamode, J.; Jeppsson, J.O.; Whitfield, J.B.; Weykamp, C.; Schellenberg, F. Reprint of Standardisation and use of the alcohol biomarker carbohydrate-deficient transferrin (CDT). Clin. Chim. Acta Int. J. Clin. Chem. 2017, 467, 15–20. [Google Scholar] [CrossRef]
  157. Bianchi, V.; Premaschi, S.; Raspagni, A.; Secco, S.; Vidali, M. A comparison between serum carbohydrate-deficient transferrin and hair ethyl glucuronide in detecting chronic alcohol consumption in routine. Alcohol Alcohol. 2015, 50, 266–270. [Google Scholar] [CrossRef] [Green Version]
  158. Van de Luitgaarden, I.A.T.; Beulens, J.W.J.; Schrieks, I.C.; Kieneker, L.M.; Touw, D.J.; van Ballegooijen, A.J.; van Oort, S.; Grobbee, D.E.; Bakker, S.J.L. Urinary Ethyl Glucuronide Can Be Used as a Biomarker of Habitual Alcohol Consumption in the General Population. J. Nutr. 2019, 149, 2199–2205. [Google Scholar] [CrossRef] [PubMed]
  159. Nguyen, V.L.; Paull, P.; Haber, P.S.; Chitty, K.; Seth, D. Evaluation of a novel method for the analysis of alcohol biomarkers: Ethyl glucuronide, ethyl sulfate and phosphatidylethanol. Alcohol 2018, 67, 7–13. [Google Scholar] [CrossRef]
  160. Reisfield, G.M.; Teitelbaum, S.A.; Opie, S.O.; Jones, J.; Morrison, D.G.; Lewis, B. The roles of phosphatidylethanol, ethyl glucuronide, and ethyl sulfate in identifying alcohol consumption among participants in professionals health programs. Drug Test. Anal. 2020, 12, 1102–1108. [Google Scholar] [CrossRef]
  161. Liu, C.; Marioni, R.E.; Hedman, Å.K.; Pfeiffer, L.; Tsai, P.C.; Reynolds, L.M.; Just, A.C.; Duan, Q.; Boer, C.G.; Tanaka, T.; et al. A DNA methylation biomarker of alcohol consumption. Mol. Psychiatry 2018, 23, 422–433. [Google Scholar] [CrossRef] [PubMed]
  162. Mahna, D.; Puri, S.; Sharma, S. DNA methylation signatures: Biomarkers of drug and alcohol abuse. Mutat. Res. 2018, 777, 19–28. [Google Scholar] [CrossRef] [PubMed]
  163. Szabo, G.; Momen-Heravi, F. Extracellular vesicles in liver disease and potential as biomarkers and therapeutic targets. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 455–466. [Google Scholar] [CrossRef] [PubMed]
  164. Niemelä, O.; Alatalo, P. Biomarkers of alcohol consumption and related liver disease. Scand. J. Clin. Lab. Investig. 2010, 70, 305–312. [Google Scholar] [CrossRef] [PubMed]
  165. Osna, N.A.; Donohue, T.M.J.; Kharbanda, K.K. Alcoholic Liver Disease: Pathogenesis and Current Management. Alcohol Res. Curr. Rev. 2017, 38, 147–161. [Google Scholar]
  166. Sanyal, A.J.; Gao, B.; Szabo, G. Gaps in Knowledge and Research Priorities for Alcoholic Hepatitis. Gastroenterology 2015, 149, 4–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Gonzalez-Quintela, A.; García, J.; Campos, J.; Perez, L.F.; Alende, M.R.; Otero, E.; Abdulkader, I.; Tomé, S. Serum cytokeratins in alcoholic liver disease: Contrasting levels of cytokeratin-18 and cytokeratin-19. Alcohol 2006, 38, 45–49. [Google Scholar] [CrossRef]
  168. Bissonnette, J.; Altamirano, J.; Devue, C.; Roux, O.; Payancé, A.; Lebrec, D.; Bedossa, P.; Valla, D.; Durand, F.; Ait-Oufella, H.; et al. A prospective study of the utility of plasma biomarkers to diagnose alcoholic hepatitis. Hepatology 2017, 66, 555–563. [Google Scholar] [CrossRef]
  169. Wu, Y.; Min, J.; Ge, C.; Shu, J.; Tian, D.; Yuan, Y.; Zhou, D. Interleukin 22 in Liver Injury, Inflammation and Cancer. Int. J. Biol. Sci. 2020, 16, 2405–2413. [Google Scholar] [CrossRef]
  170. Gao, B.; Xiang, X. Interleukin-22 from bench to bedside: A promising drug for epithelial repair. Cell. Mol. Immunol. 2019, 16, 666–667. [Google Scholar] [CrossRef] [Green Version]
  171. Xiang, X.; Hwang, S.; Feng, D.; Shah, V.H.; Gao, B. Interleukin-22 in alcoholic hepatitis and beyond. Hepatol. Int. 2020, 14, 667–676. [Google Scholar] [CrossRef]
  172. Momen-Heravi, F.; Saha, B.; Kodys, K.; Catalano, D.; Satishchandran, A.; Szabo, G. Increased number of circulating exosomes and their microRNA cargos are potential novel biomarkers in alcoholic hepatitis. J. Transl. Med. 2015, 13, 261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Jiang, L.H.; Zhang, H.D.; Tang, J.H. MiR-30a: A Novel Biomarker and Potential Therapeutic Target for Cancer. J. Oncol. 2018, 2018, 5167829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Roeb, E. Matrix metalloproteinases and liver fibrosis (translational aspects). Matrix Biol. 2018, 68–69, 463–473. [Google Scholar] [CrossRef]
  175. Thiele, M.; Madsen, B.S.; Hansen, J.F.; Detlefsen, S.; Antonsen, S.; Krag, A. Accuracy of the Enhanced Liver Fibrosis Test vs FibroTest, Elastography, and Indirect Markers in Detection of Advanced Fibrosis in Patients With Alcoholic Liver Disease. Gastroenterology 2018, 154, 1369–1379. [Google Scholar] [CrossRef] [Green Version]
  176. Xie, Q.; Zhou, X.; Huang, P.; Wei, J.; Wang, W.; Zheng, S. The performance of enhanced liver fibrosis (ELF) test for the staging of liver fibrosis: A meta-analysis. PLoS ONE 2014, 9, e92772. [Google Scholar] [CrossRef]
  177. Naveau, S.; Essoh, B.M.; Ghinoiu, M.; Marthey, L.; Njiké-Nakseu, M.; Balian, A.; Lachgar, M.; Prévot, S.; Perlemuter, G. Comparison of Fibrotest and PGAA for the diagnosis of fibrosis stage in patients with alcoholic liver disease. Eur. J. Gastroenterol. Hepatol. 2014, 26, 404–411. [Google Scholar] [CrossRef]
  178. Lackner, C.; Spindelboeck, W.; Haybaeck, J.; Douschan, P.; Rainer, F.; Terracciano, L.; Haas, J.; Berghold, A.; Bataller, R.; Stauber, R.E. Histological parameters and alcohol abstinence determine long-term prognosis in patients with alcoholic liver disease. J. Hepatol. 2017, 66, 610–618. [Google Scholar] [CrossRef] [PubMed]
  179. Louvet, A.; Labreuche, J.; Artru, F.; Bouthors, A.; Rolland, B.; Saffers, P.; Lollivier, J.; Lemaître, E.; Dharancy, S.; Lassailly, G.; et al. Main drivers of outcome differ between short term and long term in severe alcoholic hepatitis: A prospective study. Hepatology 2017, 66, 1464–1473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Mirijello, A.; D’Angelo, C.; Ferrulli, A.; Vassallo, G.; Antonelli, M.; Caputo, F.; Leggio, L.; Gasbarrini, A.; Addolorato, G. Identification and management of alcohol withdrawal syndrome. Drugs 2015, 75, 353–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Sheron, N. Alcohol and liver disease in Europe—Simple measures have the potential to prevent tens of thousands of premature deaths. J. Hepatol. 2016, 64, 957–967. [Google Scholar] [CrossRef]
  182. Dasarathy, S. Nutrition and Alcoholic Liver Disease: Effects of Alcoholism on Nutrition, Effects of Nutrition on Alcoholic Liver Disease, and Nutritional Therapies for Alcoholic Liver Disease. Clin. Liver Dis. 2016, 20, 535–550. [Google Scholar] [CrossRef] [Green Version]
  183. McClain, C.; Vatsalya, V.; Cave, M. Role of Zinc in the Development/Progression of Alcoholic Liver Disease. Curr. Treat. Options Gastroenterol. 2017, 15, 285–295. [Google Scholar] [CrossRef]
  184. Styskel, B.; Natarajan, Y.; Kanwal, F. Nutrition in Alcoholic Liver Disease: An Update. Clin. Liver Dis. 2019, 23, 99–114. [Google Scholar] [CrossRef] [PubMed]
  185. Krampe, H.; Ehrenreich, H. Supervised disulfiram as adjunct to psychotherapy in alcoholism treatment. Curr. Pharm. Des. 2010, 16, 2076–2090. [Google Scholar] [CrossRef] [PubMed]
  186. Palpacuer, C.; Duprez, R.; Huneau, A.; Locher, C.; Boussageon, R.; Laviolle, B.; Naudet, F. Pharmacologically controlled drinking in the treatment of alcohol dependence or alcohol use disorders: A systematic review with direct and network meta-analyses on nalmefene, naltrexone, acamprosate, baclofen and topiramate. Addiction 2018, 113, 220–237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Castera, P.; Stewart, E.; Großkopf, J.; Brotons, C.; Brix Schou, M.; Zhang, D.; Steiniger Brach, B.; Meulien, D. Nalmefene, Given as Needed, in the Routine Treatment of Patients with Alcohol Dependence: An Interventional, Open-Label Study in Primary Care. Eur. Addict. Res. 2018, 24, 293–303. [Google Scholar] [CrossRef]
  188. Miyata, H.; Takahashi, M.; Murai, Y.; Tsuneyoshi, K.; Hayashi, T.; Meulien, D.S.P.; Higuchi, S. Nalmefene in alcohol-dependent patients with a high drinking risk: Randomized controlled trial. Psychiatry Clin. Neurosci. 2019, 73, 697–706. [Google Scholar] [CrossRef]
  189. Addolorato, G.; Leggio, L. Safety and efficacy of baclofen in the treatment of alcohol-dependent patients. Curr. Pharm. Des. 2010, 16, 2113–2117. [Google Scholar] [CrossRef] [PubMed]
  190. Addolorato, G.; Leggio, L.; Ferrulli, A.; Cardone, S.; Vonghia, L.; Mirijello, A.; Abenavoli, L.; D’Angelo, C.; Caputo, F.; Zambon, A.; et al. Effectiveness and safety of baclofen for maintenance of alcohol abstinence in alcohol-dependent patients with liver cirrhosis: Randomised, double-blind controlled study. Lancet 2007, 370, 1915–1922. [Google Scholar] [CrossRef]
  191. Bschor, T.; Henssler, J.; Müller, M.; Baethge, C. Baclofen for alcohol use disorder-a systematic meta-analysis. Acta Psychiatr. Scand. 2018, 138, 232–242. [Google Scholar] [CrossRef] [PubMed]
  192. Pierce, M.; Sutterland, A.; Beraha, E.M.; Morley, K.; van den Brink, W. Efficacy, tolerability, and safety of low-dose and high-dose baclofen in the treatment of alcohol dependence: A systematic review and meta-analysis. Eur. Neuropsychopharmacol. 2018, 28, 795–806. [Google Scholar] [CrossRef]
  193. Rose, A.K.; Jones, A. Baclofen: Its effectiveness in reducing harmful drinking, craving, and negative mood. A meta-analysis. Addiction 2018, 113, 1396–1406. [Google Scholar] [CrossRef]
  194. Agabio, R.; Sinclair, J.M.; Addolorato, G.; Aubin, H.J.; Beraha, E.M.; Caputo, F.; Chick, J.D.; de La Selle, P.; Franchitto, N.; Garbutt, J.C.; et al. Baclofen for the treatment of alcohol use disorder: The Cagliari Statement. Lancet. Psychiatry 2018, 5, 957–960. [Google Scholar] [CrossRef] [Green Version]
  195. Skinner, M.D.; Lahmek, P.; Pham, H.; Aubin, H.J. Disulfiram efficacy in the treatment of alcohol dependence: A meta-analysis. PLoS ONE 2014, 9, e87366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Ray, L.A.; Green, R.; Roche, D.J.O.; Magill, M.; Bujarski, S. Naltrexone effects on subjective responses to alcohol in the human laboratory: A systematic review and meta-analysis. Addict. Biol. 2019, 24, 1138–1152. [Google Scholar] [CrossRef]
  197. Witkiewitz, K.; Saville, K.; Hamreus, K. Acamprosate for treatment of alcohol dependence: Mechanisms, efficacy, and clinical utility. Ther. Clin. Risk Manag. 2012, 8, 45–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Paille, F.; Martini, H. Nalmefene: A new approach to the treatment of alcohol dependence. Subst. Abus. Rehabil. 2014, 5, 87–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Minozzi, S.; Saulle, R.; Rösner, S. Baclofen for alcohol use disorder. Cochrane Database Syst. Rev. 2018, 11, Cd012557. [Google Scholar] [CrossRef]
  200. De Beaurepaire, R.; Sinclair, J.M.A.; Heydtmann, M.; Addolorato, G.; Aubin, H.J.; Beraha, E.M.; Caputo, F.; Chick, J.D.; de La Selle, P.; Franchitto, N.; et al. The Use of Baclofen as a Treatment for Alcohol Use Disorder: A Clinical Practice Perspective. Front. Psychiatry 2018, 9, 708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Manhapra, A.; Chakraborty, A.; Arias, A.J. Topiramate Pharmacotherapy for Alcohol Use Disorder and Other Addictions: A Narrative Review. J. Addict. Med. 2019, 13, 7–22. [Google Scholar] [CrossRef] [PubMed]
  202. Wetherill, R.R.; Spilka, N.; Jagannathan, K.; Morris, P.; Romer, D.; Pond, T.; Lynch, K.G.; Franklin, T.R.; Kranzler, H.R. Effects of topiramate on neural responses to alcohol cues in treatment-seeking individuals with alcohol use disorder: Preliminary findings from a randomized, placebo-controlled trial. Neuropsychopharmacology 2021. [Google Scholar] [CrossRef] [PubMed]
  203. Kranzler, H.R.; Morris, P.E.; Pond, T.; Crist, R.C.; Kampman, K.M.; Hartwell, E.E.; Lynch, K.G. Prospective randomized pharmacogenetic study of topiramate for treating alcohol use disorder. Neuropsychopharmacology 2021. [Google Scholar] [CrossRef]
  204. Mason, B.J.; Quello, S.; Shadan, F. Gabapentin for the treatment of alcohol use disorder. Expert Opin. Investig. Drugs 2018, 27, 113–124. [Google Scholar] [CrossRef]
  205. Kranzler, H.R.; Feinn, R.; Morris, P.; Hartwell, E.E. A meta-analysis of the efficacy of gabapentin for treating alcohol use disorder. Addiction 2019, 114, 1547–1555. [Google Scholar] [CrossRef]
  206. Anton, R.F.; Latham, P.; Voronin, K.; Book, S.; Hoffman, M.; Prisciandaro, J.; Bristol, E. Efficacy of Gabapentin for the Treatment of Alcohol Use Disorder in Patients With Alcohol Withdrawal Symptoms: A Randomized Clinical Trial. JAMA 2020, 180, 728–736. [Google Scholar] [CrossRef]
  207. Sherwood Brown, E.; McArdle, M.; Palka, J.; Bice, C.; Ivleva, E.; Nakamura, A.; McNutt, M.; Patel, Z.; Holmes, T.; Tipton, S. A randomized, double-blind, placebo-controlled proof-of-concept study of ondansetron for bipolar and related disorders and alcohol use disorder. Eur. Neuropsychopharmacol. 2021, 43, 92–101. [Google Scholar] [CrossRef]
  208. Gasparyan, A.; Navarrete, F.; Manzanares, J. The administration of sertraline plus naltrexone reduces ethanol consumption and motivation in a long-lasting animal model of post-traumatic stress disorder. Neuropharmacology 2021, 189, 108552. [Google Scholar] [CrossRef] [PubMed]
  209. Caputo, F.; Vignoli, T.; Tarli, C.; Domenicali, M.; Zoli, G.; Bernardi, M.; Addolorato, G. A Brief Up-Date of the Use of Sodium Oxybate for the Treatment of Alcohol Use Disorder. Int. J. Environ. Res. Public Health 2016, 13, 290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Hurt, R.T.; Ebbert, J.O.; Croghan, I.T.; Schroeder, D.R.; Hurt, R.D.; Hays, J.T. Varenicline for tobacco-dependence treatment in alcohol-dependent smokers: A randomized controlled trial. Drug Alcohol Depend. 2018, 184, 12–17. [Google Scholar] [CrossRef]
  211. Saberi, B.; Dadabhai, A.S.; Jang, Y.Y.; Gurakar, A.; Mezey, E. Current Management of Alcoholic Hepatitis and Future Therapies. J. Clin. Transl. Hepatol. 2016, 4, 113–122. [Google Scholar]
  212. Maddrey, W.C.; Boitnott, J.K.; Bedine, M.S.; Weber, F.L.J.; Mezey, E.; White, R.I.J. Corticosteroid therapy of alcoholic hepatitis. Gastroenterology 1978, 75, 193–199. [Google Scholar] [CrossRef]
  213. Louvet, A.; Naveau, S.; Abdelnour, M.; Ramond, M.J.; Diaz, E.; Fartoux, L.; Dharancy, S.; Texier, F.; Hollebecque, A.; Serfaty, L.; et al. The Lille model: A new tool for therapeutic strategy in patients with severe alcoholic hepatitis treated with steroids. Hepatology 2007, 45, 1348–1354. [Google Scholar] [CrossRef]
  214. Lieber, S.R.; Rice, J.P.; Lucey, M.R.; Bataller, R. Controversies in clinical trials for alcoholic hepatitis. J. Hepatol. 2018, 68, 586–592. [Google Scholar] [CrossRef]
  215. Thursz, M.R.; Richardson, P.; Allison, M.; Austin, A.; Bowers, M.; Day, C.P.; Downs, N.; Gleeson, D.; MacGilchrist, A.; Grant, A.; et al. Prednisolone or pentoxifylline for alcoholic hepatitis. N. Engl. J. Med. 2015, 372, 1619–1628. [Google Scholar] [CrossRef] [Green Version]
  216. Louvet, A.; Thursz, M.R.; Kim, D.J.; Labreuche, J.; Atkinson, S.R.; Sidhu, S.S.; O’Grady, J.G.; Akriviadis, E.; Sinakos, E.; Carithers, R.L.; et al. Corticosteroids Reduce Risk of Death Within 28 Days for Patients With Severe Alcoholic Hepatitis, Compared With Pentoxifylline or Placebo-a Meta-analysis of Individual Data From Controlled Trials. Gastroenterology 2018, 155, 458–468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Vergis, N.; Atkinson, S.R.; Knapp, S.; Maurice, J.; Allison, M.; Austin, A.; Forrest, E.H.; Masson, S.; McCune, A.; Patch, D.; et al. In Patients With Severe Alcoholic Hepatitis, Prednisolone Increases Susceptibility to Infection and Infection-Related Mortality, and Is Associated With High Circulating Levels of Bacterial DNA. Gastroenterology 2017, 152, 1068–1077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Dao, A.; Rangnekar, A.S. Steroids for Severe Alcoholic Hepatitis: More Risk Than Reward? Clin. Liver Dis. 2018, 12, 151–153. [Google Scholar] [CrossRef]
  219. Li, S.; Tan, H.Y.; Wang, N.; Zhang, Z.J.; Lao, L.; Wong, C.W.; Feng, Y. The Role of Oxidative Stress and Antioxidants in Liver Diseases. Int. J. Mol. Sci. 2015, 16, 26087–26124. [Google Scholar] [CrossRef] [Green Version]
  220. Moreno, C.; Langlet, P.; Hittelet, A.; Lasser, L.; Degré, D.; Evrard, S.; Colle, I.; Lemmers, A.; Devière, J.; Le Moine, O. Enteral nutrition with or without N-acetylcysteine in the treatment of severe acute alcoholic hepatitis: A randomized multicenter controlled trial. J. Hepatol. 2010, 53, 1117–1122. [Google Scholar] [CrossRef]
  221. Thursz, M.; Morgan, T.R. Treatment of Severe Alcoholic Hepatitis. Gastroenterology 2016, 150, 1823–1834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Nguyen-Khac, E.; Thevenot, T.; Piquet, M.A.; Benferhat, S.; Goria, O.; Chatelain, D.; Tramier, B.; Dewaele, F.; Ghrib, S.; Rudler, M.; et al. Glucocorticoids plus N-acetylcysteine in severe alcoholic hepatitis. N. Engl. J. Med. 2011, 365, 1781–1789. [Google Scholar] [CrossRef] [Green Version]
  223. Shipley, L.C.; Kodali, S.; Singal, A.K. Recent updates on alcoholic hepatitis. Dig. Liver Dis. 2019, 51, 761–768. [Google Scholar] [CrossRef] [PubMed]
  224. Higuera-de la Tijera, F.; Servín-Caamaño, A.I.; Serralde-Zúñiga, A.E.; Cruz-Herrera, J.; Pérez-Torres, E.; Abdo-Francis, J.M.; Salas-Gordillo, F.; Pérez-Hernández, J.L. Metadoxine improves the three- and six-month survival rates in patients with severe alcoholic hepatitis. World J. Gastroenterol. 2015, 21, 4975–4985. [Google Scholar] [CrossRef] [PubMed]
  225. Higuera-de la Tijera, F.; Servín-Caamaño, A.I.; Cruz-Herrera, J.; Serralde-Zúñiga, A.E.; Abdo-Francis, J.M.; Gutiérrez-Reyes, G.; Pérez-Hernández, J.L. Treatment with metadoxine and its impact on early mortality in patients with severe alcoholic hepatitis. Ann. Hepatol. 2014, 13, 343–352. [Google Scholar] [CrossRef]
  226. Van Haele, M.; Snoeck, J.; Roskams, T. Human Liver Regeneration: An Etiology Dependent Process. Int. J. Mol. Sci. 2019, 20, 2332. [Google Scholar] [CrossRef] [Green Version]
  227. Kim, A.; Wu, X.; Allende, D.S.; Nagy, L.E. Gene deconvolution reveals aberrant liver regeneration and immune cell infiltration in alcohol-associated hepatitis. Hepatology 2021. [Google Scholar] [CrossRef]
  228. Tornai, D.; Szabo, G. Emerging medical therapies for severe alcoholic hepatitis. Clin. Mol. Hepatol. 2020, 26, 686–696. [Google Scholar] [CrossRef] [PubMed]
  229. Singh, V.; Sharma, A.K.; Narasimhan, R.L.; Bhalla, A.; Sharma, N.; Sharma, R. Granulocyte colony-stimulating factor in severe alcoholic hepatitis: A randomized pilot study. Am. J. Gastroenterol. 2014, 109, 1417–1423. [Google Scholar] [CrossRef]
  230. Singh, V.; Keisham, A.; Bhalla, A.; Sharma, N.; Agarwal, R.; Sharma, R.; Singh, A. Efficacy of Granulocyte Colony-Stimulating Factor and N-Acetylcysteine Therapies in Patients With Severe Alcoholic Hepatitis. Clin. Gastroenterol. Hepatol. 2018, 16, 1650–1656. [Google Scholar] [CrossRef] [Green Version]
  231. Shasthry, S.M.; Sharma, M.K.; Shasthry, V.P.A. Sarin SK. Efficacy of Granulocyte Colony-stimulating Factor in the Management of Steroid-Nonresponsive Severe Alcoholic Hepatitis: A Double-Blind Randomized Controlled Trial. Hepatology 2019, 70, 802–811. [Google Scholar] [CrossRef]
  232. Marot, A.; Singal, A.K.; Moreno, C.; Deltenre, P. Granulocyte colony-stimulating factor for alcoholic hepatitis: A systematic review and meta-analysis of randomised controlled trials. JHEP Rep. 2020, 2, 100139. [Google Scholar] [CrossRef]
  233. Kedarisetty, C.K.; Kumar, A.; Sarin, S.K. Insights into the Role of Granulocyte Colony-Stimulating Factor in Severe Alcoholic Hepatitis. Semin. Liver Dis. 2021, 41, 67–78. [Google Scholar] [PubMed]
  234. Brandl, K.; Hartmann, P.; Jih, L.J.; Pizzo, D.P.; Argemi, J.; Ventura-Cots, M.; Coulter, S.; Liddle, C.; Ling, L.; Rossi, S.J.; et al. Dysregulation of serum bile acids and FGF19 in alcoholic hepatitis. J. Hepatol. 2018, 69, 396–405. [Google Scholar] [CrossRef]
  235. Lee, Y.-S.; Kim, H.J.; Kim, J.H.; Yoo, Y.J.; Kim, T.S.; Kang, S.H.; Suh, S.J.; Joo, M.K.; Jung, Y.K.; Lee, B.J.; et al. Treatment of Severe Alcoholic Hepatitis With Corticosteroid, Pentoxifylline, or Dual Therapy: A Systematic Review and Meta-Analysis. J. Clin. Gastroenterol. 2017, 51, 364–377. [Google Scholar] [CrossRef] [PubMed]
  236. Naveau, S.; Chollet-Martin, S.; Dharancy, S.; Mathurin, P.; Jouet, P.; Piquet, M.A.; Davion, T.; Oberti, F.; Broët, P.; Emilie, D. A double-blind randomized controlled trial of infliximab associated with prednisolone in acute alcoholic hepatitis. Hepatology 2004, 39, 1390–1397. [Google Scholar] [CrossRef]
  237. Boetticher, N.C.; Peine, C.J.; Kwo, P.; Abrams, G.A.; Patel, T.; Aqel, B.; Boardman, L.; Gores, G.J.; Harmsen, W.S.; McClain, C.J.; et al. A randomized, double-blinded, placebo-controlled multicenter trial of etanercept in the treatment of alcoholic hepatitis. Gastroenterology 2008, 135, 1953–1960. [Google Scholar] [CrossRef] [Green Version]
  238. Wong, V.W.; Singal, A.K. Emerging medical therapies for non-alcoholic fatty liver disease and for alcoholic hepatitis. Transl. Gastroenterol. Hepatol. 2019, 4, 53. [Google Scholar] [CrossRef] [PubMed]
  239. Arab, J.P.; Sehrawat, T.S.; Simonetto, D.A.; Verma, V.K.; Feng, D.; Tang, T.; Dreyer, K.; Yan, X.; Daley, W.L.; Sanyal, A.; et al. An Open-Label, Dose-Escalation Study to Assess the Safety and Efficacy of IL-22 Agonist F-652 in Patients With Alcohol-associated Hepatitis. Hepatology 2020, 72, 441–453. [Google Scholar] [CrossRef]
  240. Verma, V.K.; Li, H.; Wang, R.; Hirsova, P.; Mushref, M.; Liu, Y.; Cao, S.; Contreras, P.C.; Malhi, H.; Kamath, P.S.; et al. Alcohol stimulates macrophage activation through caspase-dependent hepatocyte derived release of CD40L containing extracellular vesicles. J. Hepatol. 2016, 64, 651–660. [Google Scholar] [CrossRef] [Green Version]
  241. Singal, A.K.; Shah, V.H. Current trials and novel therapeutic targets for alcoholic hepatitis. J. Hepatol. 2019, 70, 305–313. [Google Scholar] [CrossRef] [Green Version]
  242. Frenette, C.T.; Morelli, G.; Shiffman, M.L.; Frederick, R.T.; Rubin, R.A.; Fallon, M.B.; Cheng, J.T.; Cave, M.; Khaderi, S.A.; Massoud, O.; et al. Emricasan Improves Liver Function in Patients With Cirrhosis and High Model for End-Stage Liver Disease Scores Compared With Placebo. Clin. Gastroenterol. Hepatol. 2019, 17, 774–783. [Google Scholar] [CrossRef]
  243. Szabo, G. Clinical Trial Design for Alcoholic Hepatitis. Semin. Liver Dis. 2017, 37, 332–342. [Google Scholar] [CrossRef] [PubMed]
  244. Mathurin, P.; Dufour, J.-F.; Bzowej, N.H.; Shiffman, M.L.; Arterburn, S.; Nguyen, T.; Billin, A.; Chung, C.; Subramanian, M.; Myers, R.P. A Phase 2 Randomized Controlled Trial. In Selonsertib in Combination with Prednisolone for the Treatment of Severe Alcoholic Hepatitis; Wiley: Hoboken, NJ, USA, 2018; pp. 8A–9A. [Google Scholar]
  245. Szabo, G. Gut-liver axis in alcoholic liver disease. Gastroenterology 2015, 148, 30–36. [Google Scholar] [CrossRef] [Green Version]
  246. Llopis, M.; Cassard, A.M.; Wrzosek, L.; Boschat, L.; Bruneau, A.; Ferrere, G.; Puchois, V.; Martin, J.C.; Lepage, P.; Le Roy, T.; et al. Intestinal microbiota contributes to individual susceptibility to alcoholic liver disease. Gut 2016, 65, 830–839. [Google Scholar] [CrossRef]
  247. Spalinger, M.R.; Atrott, K.; Baebler, K.; Schwarzfischer, M.; Melhem, H.; Peres, D.R.; Lalazar, G.; Rogler, G.; Scharl, M.; Frey-Wagner, I. Administration of the Hyper-immune Bovine Colostrum Extract IMM-124E Ameliorates Experimental Murine Colitis. J. Crohn’s Colitis 2019, 13, 785–797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Kirpich, I.A.; Solovieva, N.V.; Leikhter, S.N.; Shidakova, N.A.; Lebedeva, O.V.; Sidorov, P.I.; Bazhukova, T.A.; Soloviev, A.G.; Barve, S.S.; McClain, C.J.; et al. Probiotics restore bowel flora and improve liver enzymes in human alcohol-induced liver injury: A pilot study. Alcohol 2008, 42, 675–682. [Google Scholar] [CrossRef] [Green Version]
  249. Philips, C.A.; Pande, A.; Shasthry, S.M.; Jamwal, K.D.; Khillan, V.; Chandel, S.S.; Kumar, G.; Sharma, M.K.; Maiwall, R.; Jindal, A.; et al. Healthy Donor Fecal Microbiota Transplantation in Steroid-Ineligible Severe Alcoholic Hepatitis: A Pilot Study. Clin. Gastroenterol. Hepatol. 2017, 15, 600–602. [Google Scholar] [CrossRef] [PubMed]
  250. Bajaj, J.S. Alcohol, liver disease and the gut microbiota. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 235–246. [Google Scholar] [CrossRef] [PubMed]
  251. Torres, J.L.; Novo-Veleiro, I.; Manzanedo, L.; Alvela-Suárez, L.; Macías, R.; Laso, F.J.; Marcos, M. Role of microRNAs in alcohol-induced liver disorders and non-alcoholic fatty liver disease. World J. Gastroenterol. 2018, 24, 4104–4118. [Google Scholar] [CrossRef] [PubMed]
  252. Janssen, H.L.; Reesink, H.W.; Lawitz, E.J.; Zeuzem, S.; Rodriguez-Torres, M.; Patel, K.; van der Meer, A.J.; Patick, A.K.; Chen, A.; Zhou, Y.; et al. Treatment of HCV infection by targeting microRNA. N. Engl. J. Med. 2013, 368, 1685–1694. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Bandiera, S.; Pfeffer, S.; Baumert, T.F.; Zeisel, M.B. miR-122—A key factor and therapeutic target in liver disease. J. Hepatol. 2015, 62, 448–457. [Google Scholar] [CrossRef] [Green Version]
  254. Pittenger, M.F.; Discher, D.E.; Péault, B.M.; Phinney, D.G.; Hare, J.M.; Caplan, A.I. Mesenchymal stem cell perspective: Cell biology to clinical progress. NPJ Regen. Med. 2019, 4, 22. [Google Scholar] [CrossRef] [Green Version]
  255. Hu, C.; Zhao, L.; Duan, J.; Li, L. Strategies to improve the efficiency of mesenchymal stem cell transplantation for reversal of liver fibrosis. J. Cell. Mol. Med. 2019, 23, 1657–1670. [Google Scholar] [CrossRef]
  256. Jang, Y.O.; Kim, Y.J.; Baik, S.K.; Kim, M.Y.; Eom, Y.W.; Cho, M.Y.; Park, H.J.; Park, S.Y.; Kim, B.R.; Kim, J.W.; et al. Histological improvement following administration of autologous bone marrow-derived mesenchymal stem cells for alcoholic cirrhosis: A pilot study. Liver Int. 2014, 34, 33–41. [Google Scholar] [CrossRef] [PubMed]
  257. Suk, K.T.; Yoon, J.H.; Kim, M.Y.; Kim, C.W.; Kim, J.K.; Park, H.; Hwang, S.G.; Kim, D.J.; Lee, B.S.; Lee, S.H.; et al. Transplantation with autologous bone marrow-derived mesenchymal stem cells for alcoholic cirrhosis: Phase 2 trial. Hepatology 2016, 64, 2185–2197. [Google Scholar] [CrossRef] [PubMed]
  258. Tsuchiya, A.; Takeuchi, S.; Watanabe, T.; Yoshida, T.; Nojiri, S.; Ogawa, M.; Terai, S. Mesenchymal stem cell therapies for liver cirrhosis: MSCs as “conducting cells” for improvement of liver fibrosis and regeneration. Inflamm. Regen. 2019, 39, 18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Lee, B.P.; Mehta, N.; Platt, L.; Gurakar, A.; Rice, J.P.; Lucey, M.R.; Im, G.Y.; Therapondos, G.; Han, H.; Victor, D.W.; et al. Outcomes of Early Liver Transplantation for Patients With Severe Alcoholic Hepatitis. Gastroenterology 2018, 155, 422–430. [Google Scholar] [CrossRef] [Green Version]
  260. Mathurin, P.; Moreno, C.; Samuel, D.; Dumortier, J.; Salleron, J.; Durand, F.; Castel, H.; Duhamel, A.; Pageaux, G.P.; Leroy, V.; et al. Early liver transplantation for severe alcoholic hepatitis. N. Engl. J. Med. 2011, 365, 1790–1800. [Google Scholar] [CrossRef] [Green Version]
  261. Patrizia, B.; Andrew, B.; Ivo, G.; Jacques, P.; García-Valdecasas, J.C.; Paolo, M.; Didier, S.; Xavier, F.; Andrew, B. EASL Clinical Practice Guidelines: Liver transplantation. J. Hepatol. 2016, 64, 433–485. [Google Scholar]
  262. Lee, B.P.; Vittinghoff, E.; Dodge, J.L.; Cullaro, G.; Terrault, N.A. National Trends and Long-term Outcomes of Liver Transplant for Alcohol-Associated Liver Disease in the United States. JAMA Intern. Med. 2019, 179, 340–348. [Google Scholar] [CrossRef] [PubMed]
  263. Marroni, C.A.; Fleck, A.M.J.; Fernandes, S.A.; Galant, L.H.; Mucenic, M.; de Mattos Meine, M.H.; Mariante-Neto, G.; Brandão, A.B.M. Liver transplantation and alcoholic liver disease: History, controversies, and considerations. World J. Gastroenterol. 2018, 24, 2785–2805. [Google Scholar] [CrossRef] [PubMed]
  264. Addolorato, G.; Mirijello, A.; Leggio, L.; Ferrulli, A.; D’Angelo, C.; Vassallo, G.; Cossari, A.; Gasbarrini, G.; Landolfi, R.; Agnes, S.; et al. Liver transplantation in alcoholic patients: Impact of an alcohol addiction unit within a liver transplant center. Alcohol. Clin. Exp. Res. 2013, 37, 1601–1608. [Google Scholar] [CrossRef] [PubMed]
  265. Fleming, M.F.; Smith, M.J.; Oslakovic, E.; Lucey, M.R.; Vue, J.X.; Al-Saden, P.; Levitsky, J. Phosphatidylethanol Detects Moderate-to-Heavy Alcohol Use in Liver Transplant Recipients. Alcohol. Clin. Exp. Res. 2017, 41, 857–862. [Google Scholar] [CrossRef]
  266. Brière, M.; Tocanier, L.; Allain, P.; Le Gal, D.; Allet, G.; Gorwood, P.; Gohier, B. Decision-Making Measured by the Iowa Gambling Task in Patients with Alcohol Use Disorders Choosing Harm Reduction versus Relapse Prevention Program. Eur. Addict. Res. 2019, 25, 182–190. [Google Scholar] [CrossRef] [PubMed]
Table 1. Current experimental models of miRNAs involved in ALD pathogenesis.
Table 1. Current experimental models of miRNAs involved in ALD pathogenesis.
miRNASample SourceALD SeverityDysregulationFunction
miR17-92 cluster [76]Human HSC cell line Rat modelCirrhosis and fibrosis due to ALDVarious (e.g., miR-17a decreased, miR-92 increased)HSC activation. Inhibited MeCP2/TGFβRII expression
miR-21 [77,78,79]Mice and rat model Human HSC and hepatocyte cell line Human HCC cell lineAlcoholic liver injury HCCIncreasedHSC activation, hepatocyte survival, transformation, and remodeling Increased α-SMA, FASLG, DR5 expression
miR-26a [80]Mice model, Human HCC cell lineAlcoholic liver injury HCCDecreasedPromote cytoprotective Autophagy, Target Beclin-1, DUSP4, DUSP5
miR-27a [81,82]Human blood, Human monocyteAlcoholic liver injury Alcoholic hepatitisIncreasedM2 Monocyte polarization, Downregulate sprouty2, Increase CD206
miR-29 [83]Human blood, Murine modelCirrhosis due to ALD CCl4-induced hepatic fibrogenesisDecreasedDownregulate HSC and collagen expression Modulate intestinal permeability
miR-34a [84,85,86,87]Human hepatocyte and cholangiocyte, Mice modelFibrosis due to ALD HCCIncreasedHepatocyte steatosis, inflammation and fibrosis Decrease caspase 2 and SIRT1
miR-122 [88,89,90]Human blood, Mice modelSteatohepatitis, fibrosis, or cirrhosis due to ALD HCCDecreasedDownregulate HIF-1α, cyclinG1, Bcl-w, Reprogram monocyte to LPS stimulation, Regulate lipid metabolism
miR-125b [85,91]Human HCC cell line and tissues, Rat modelHCC, Alcoholic liver injuryDecreasedDecrease PIGF expression, Distort MMP-2, MMP-9 expression
miR-126 [92,93]Human HCC tissues Rat modelHCC, CCl4-induced hepatic fibrogenesisDecreasedDownregulate HSC activity, Regulate VEGF-A, PI3K, p-AKT, cyclin D1 activity
miR-155 [94,95,96,97,98]Mice model, Murine hepatocytes and Kupffer cellsSteatohepatitis, fibrosis, or cirrhosis due to ALDIncreasedKupffer cells activation Induce TNF-α and NF κB activity, Mediate PPAR-α pathway, Induce C/EBPβ activity
miR-181b-3p [99]Mice model, Rat model Murine Kupffer cellsAlcoholic liver injuryDecreasedSensitize Kupffer cells to TLR4-mediated cytokine production, Modulate importin α5 expression
miR-182 [100,101]Human liver tissues Mice modelAlcoholic hepatitis, Alcoholic liver injuryIncreasedPromote hepatocyte inflammation, Upregulate CCL20, CXCL1, IL-8, Cyclin D1
miR-199 [102,103]Human HCC tissues Rat liver sinusoidal endothelial cellsHCC, Alcoholic liver injuryDecreasedRegulate hepatocyte inflammation and immune cells infiltration, Attenuate HIF-1α and ET-1 expression
miR-200a [103]Mice hepatocyte cell line, Mice modelAlcoholic liver injuryIncreasedModulate hepatocyte apoptosis, Decrease ZEB2 expression
miR-212 [104,105]Human gut epithelial cells, Mice modelAdvanced ALD Alcoholic liver injuryIncreasedDisrupt tight junctions integrity, gut leakage Downregulate ZO-1 expression
miR-214 [106]Human HCC cell line Rat hepatocyte, Rat modelHCC Alcoholic liver injuryIncreasedInduce hepatocyte oxidative stress, Repress GSR and POR activity
miR-217 [107,108]Murine macrophage Mice modelAlcoholic liver injuryIncreasedHepatocyte steatosis and inflammation, Downregulate SIRT1
miR-223 [109]Human blood, Mice modelChronic alcohol use, Alcoholic liver injuryDecreasedLimit neutrophil infiltration and ROS production Inhibits IL-6–p47phox–ROS pathway
miR-291b [110]Rat model, Human monocyteAlcoholic liver injury Alcoholic hepatitisIncreasedSensitize monocyte to TLR4 signaling, Downregulate Tollip expression
miR-378 [111]Mice model, Human HCC tissuesCCl4-induced hepatic Fibrogenesis, HCCDecreasedSuppress HSC activation Decrease Gli3 expression
miR-497 [112]Mice model, Mice hepatocyteAlcoholic liver injuryDecreasedAlleviate bile acid synthesis, Reduce Btg2, Yy1 levels
Table 2. Relapse prevention medications in patients with ALD.
Table 2. Relapse prevention medications in patients with ALD.
MedicationDoseMechanismsAdverse EffectsComment
Disulfiram [186,195]250–500 mg/dayInhibit acetaldehyde dehydrogenaseHepatotoxicity, metallic taste, polyneuritis, skin allergyEffective treatment, No studies in advanced ALD
Naltrexone [187,196]50 mg/dayOpioid receptor antagonistHepatotoxicity, headache, nervousness, abdominal cramps, myalgiaEffective treatment, No studies in advanced ALD
Acamprosate [187,197]1998 mg/dayNMDA receptor antagonist Glutamatergic receptor modulatorDiarrhea, insomnia, anorexia, astheniaAvoid in severe renal impairment, No studies in advanced ALD
Nalmefene [189,198]10–20 mg/dayOpioid receptor antagonistNausea, vomiting, dizzinessReduction of heavy drinking
Baclofen [191,192,193,194,199,200]15–60 mg/dayGABA-B receptor agonistDrowsiness, fatigue, headache, dry mouthOff-label use, Consider in advanced ALD, Low-dosage preferred
Topiramate [201,202,203]75–200 mg/dayGABA receptor agonist, glutamate receptor antagonistDrowsiness, dizziness, loss of coordination, anorexiaReduction of heavy drinking, No studies in ALD
Gabapentin [204,205,206]600–1800 mg/dayInhibit presynaptic calcium channel Influence GABA and glutamate activityDizziness, fatigue, ataxia, drowsiness, diplopiaA recent RCT showed good efficacy, Consider as second-line medication
Ondansetron [207]4–8 µg/kg/day5-HT3 receptor antagonistConstipation, headache, drowsinessNo recommendations in guidelines
Sertraline [208]50–200 mg/daySSRIAnorexia, dry mouth, dyspepsia, insomniaMay be helpful in selective patient group
Sodium oxybate [209]50–100 mg/kg/dayGABA receptor agonistDizziness, sedation, astheniaApproved in Italy and Austria, Risk of abuse
Varenicline [210]0.5–2 mg/dayPartial nAChR agonistNausea, vomiting, insomnia, headacheMay be effective in smokers with AUD
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, S.-Y.; Tsai, I.-T.; Hsu, Y.-C. Alcohol-Related Liver Disease: Basic Mechanisms and Clinical Perspectives. Int. J. Mol. Sci. 2021, 22, 5170. https://doi.org/10.3390/ijms22105170

AMA Style

Liu S-Y, Tsai I-T, Hsu Y-C. Alcohol-Related Liver Disease: Basic Mechanisms and Clinical Perspectives. International Journal of Molecular Sciences. 2021; 22(10):5170. https://doi.org/10.3390/ijms22105170

Chicago/Turabian Style

Liu, Szu-Yi, I-Ting Tsai, and Yin-Chou Hsu. 2021. "Alcohol-Related Liver Disease: Basic Mechanisms and Clinical Perspectives" International Journal of Molecular Sciences 22, no. 10: 5170. https://doi.org/10.3390/ijms22105170

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