Next Article in Journal
Sirt6 Activation Ameliorates Inflammatory Bone Loss in Ligature-Induced Periodontitis in Mice
Next Article in Special Issue
Current Therapeutical Approaches Targeting Lipid Metabolism in NAFLD
Previous Article in Journal
Kaempferol and Biomodified Kaempferol from Sophora japonica Extract as Potential Sources of Anti-Cancer Polyphenolics against High Grade Glioma Cell Lines
Previous Article in Special Issue
Non-Alcoholic Fatty Liver Disease: Translating Disease Mechanisms into Therapeutics Using Animal Models
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Innovative Therapeutic Approaches in Non-Alcoholic Fatty Liver Disease: When Knowing Your Patient Is Key

by
Marta Alonso-Peña
1,
Maria Del Barrio
1,
Ana Peleteiro-Vigil
1,
Carolina Jimenez-Gonzalez
1,
Alvaro Santos-Laso
1,
Maria Teresa Arias-Loste
1,
Paula Iruzubieta
1 and
Javier Crespo
1,2,*
1
Gastroenterology and Hepatology Department, Clinical and Translational Research in Digestive Diseases, Valdecilla Research Institute (IDIVAL), Marqués de Valdecilla University Hospital, 39011 Santander, Spain
2
Biomedical Research Networking Center in Hepatic and Digestive Diseases (CIBERehd), 28029 Madrid, Spain
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(13), 10718; https://doi.org/10.3390/ijms241310718
Submission received: 27 April 2023 / Revised: 21 June 2023 / Accepted: 24 June 2023 / Published: 27 June 2023

Abstract

:
Non-alcoholic fatty liver disease (NAFLD) encompasses a spectrum of disorders ranging from simple steatosis to non-alcoholic steatohepatitis (NASH). Hepatic steatosis may result from the dysfunction of multiple pathways and thus multiple molecular triggers involved in the disease have been described. The development of NASH entails the activation of inflammatory and fibrotic processes. Furthermore, NAFLD is also strongly associated with several extra-hepatic comorbidities, i.e., metabolic syndrome, type 2 diabetes mellitus, obesity, hypertension, cardiovascular disease and chronic kidney disease. Due to the heterogeneity of NAFLD presentations and the multifactorial etiology of the disease, clinical trials for NAFLD treatment are testing a wide range of interventions and drugs, with little success. Here, we propose a narrative review of the different phenotypic characteristics of NAFLD patients, whose disease may be triggered by different agents and driven along different pathophysiological pathways. Thus, correct phenotyping of NAFLD patients and personalized treatment is an innovative therapeutic approach that may lead to better therapeutic outcomes.

1. Introduction

Non-alcoholic fatty liver disease (NAFLD) includes a wide spectrum of liver injuries ranging from simple steatosis to non-alcoholic steatohepatitis (NASH), which is characterized by a variable grade of inflammation and hepatocellular damage [1,2] and may further progress to more severe hepatic disorders [3]. NAFLD is a growing contributor to end-stage liver disease and liver transplantation [4]. Additionally, NAFLD exhibits a robust correlation with numerous extra-hepatic metabolic conditions, including type 2 diabetes mellitus (T2DM), obesity, hypertension, cardiovascular disease and chronic kidney disease, among others. Consequently, this elevates the mortality rate associated with the condition [5,6]. Such attributes have led to suggestions for a nomenclature change to metabolic-associated fatty liver disease (MAFLD), which carries significant implications for patient management strategies [7]. However, besides metabolic dysfunction, other diseases result in hepatic steatosis, such as alcohol- and drug-induced liver injury, viral infections and chronic inflammatory diseases. Undoubtedly, this liver condition should no longer be considered a “histological disease” and moved away from the two-stage division into NAFLD and NASH as such categorization may not fully reflect the diverse range of disease progression in response to modifications in the underlying metabolic dysfunction or medical treatments [8]. Furthermore, consensus regarding the NAFLD or MAFLD name has not been achieved among experts. Here, we will use NAFLD nomenclature as a standard, except for research on MAFLD as specified by authors.
Two main events have been defined in the pathophysiology of NAFLD: lipid accumulation within the hepatocytes, especially free fatty acids (FFAs), and liver-related innate immune responses [9]. However, inflammation may precede steatosis as inflammatory events may lead to lipid accumulation [10]. Therefore, there are many factors influencing NAFLD initiation and progression: environmental exposure, lifestyle, genetic susceptibility, metabolic status and the microbiome [11]. All these factors could induce either steatosis or inflammation, which further triggers endoplasmic reticulum stress, expression of pro-inflammatory cytokines, oxidative stress, hepatic insulin resistance and apoptosis [9,12,13] (Figure 1). The complex interaction of all these mechanisms suggests the existence of different phenotypes within NAFLD that differ in the molecular pathways altered which could result in different natural history, disease course and clinical outcomes.
This review is aimed at discussing the potential stratification of NAFLD patients to provide personalized health care and treatment by summarizing current knowledge about the characteristics of NAFLD patients that could influence pathophysiology and both liver and extrahepatic manifestations of the disease. Understanding the different phenotypes encompassed under NAFLD diagnosis will improve treatment strategies and foster the identification of successful treatments for this pathology by improving clinical trials’ design and control for individual genetic predisposition, signal transduction, or metabolic profiles.

2. How to Phenotype NAFLD Patients

Despite more than a decade of extensive research focusing on NAFLD, there is currently no approved therapy for NASH. The complexity and heterogeneity of NAFLD represent important impediments to the discovery of highly effective drug treatments. In addition, clinical trials are not controlled for individual genetic predisposition or signal transduction or metabolic profiles. Trial recruitment is currently based on liver histologic involvement, but many pathological pathways can lead to the same histological phenotype. Therefore, clinical trial reporting for NAFLD is suboptimal, limiting our understanding [14]. The initial step is trying to change this simplistic view of NAFLD, both in clinical trials and in daily clinical practice. A multi-omics data integration approach for NAFLD patients could help us to properly subphenotype and stratify patients, paving the way for precision medicine in NAFLD.
The importance of the classification of NAFLD patients into different subtypes is reflected in several studies based on metabolomics [15,16]. These authors identified a unique serum metabolomic profile of Mat1a (methionine adenosyltransferase 1A) knockout (KO) mice and 0.1MCD (methionine and choline deficient) model and observed, using a large cohort of serum samples from biopsied NAFLD patients, that some of them showed this metabolic signature (M-subtype), identifying those patients that will likely benefit from therapy with S-adenosylmethionine (SAMe) or Aramchol. However, this approach also results in a certain number of unclassified patients (denominated indeterminate). Potential integration of other omics data as well as clinical parameters may improve this novel subtyping approach for NAFLD patients, allowing further interpretation of the complex and heterogeneous disease. Multi-omics observations could identify how genetic, epigenetic, or transcriptional changes lead to metabolic alterations in complex diseases such as NAFLD in a comprehensive manner [17]. Therefore, a comprehensive landscape of the main NAFLD drivers and patient outcome determinants may be obtained by integrating multi-omics and clinical data.
Following state-of-the-art research, in this section, we summarize the main factors to take into consideration when attending NAFLD patients, which could lead to better clinical management and treatment (Figure 2).

2.1. Histological Features

It is widely accepted that information about disease activity and, in particular, the extent of liver fibrosis is necessary to assess the severity of liver disease and provide prognosis in NAFLD [18]. Histological evaluation of NAFLD liver biopsies is the gold-standard technique for the assessment of disease phenotype and progression [4]. The histological scoring system for staging fibrosis ranges from stage 0 (no fibrosis) to stage 4 (cirrhosis) [18]. However, non-invasive, point-of-care techniques such as imaging modalities (i.e., vibration-controlled transient elastography [VCTE]) and index-based approaches (i.e., FIB-4, NAFLD fibrosis score [NFS]) have been proposed as better options for surveillance programs in at-risk or global populations [19].
It has been shown that the incidence of liver-related complications and all-cause mortality increased with the degree of fibrosis, with fibrosis grades F3 and F4 being associated with an increased risk of liver complications and all-cause mortality [4]. These findings provide support for the use of “progression to cirrhosis” as a generally accepted surrogate outcome for regulatory approval of therapeutic agents. Moreover, the higher rate of hepatic decompensation events (i.e., ascites, variceal bleeding, and encephalopathy) and hepatocellular carcinoma (HCC) among patients with F3 provides a rationale to test the hypothesis that a one-stage regression of fibrosis may translate to fewer hepatic decompensation events [4]. However, fibrosis worsened by one stage (from baseline stage 0 fibrosis) on average during 7.1 years for patients with NASH and by one stage over 14.3 years for patients with steatosis [18]. This long progression time precludes the usefulness of “progression to cirrhosis” as a surrogate outcome as clinical trials are not usually designed for the evaluation of impact over such long periods. In fact, one-stage regression of fibrosis has not been met in most clinical trials testing pharmacological treatments for NAFLD [20]. Furthermore, the incidence of non-hepatic cancers is similar across all fibrosis grades [4], although the leading cause of death in patients with NAFLD is cardiovascular disease, followed by extrahepatic malignancy [18].
On the other hand, the liver phenotype in NAFLD is defined by histological findings such as hepatocellular ballooning, steatosis grade and lobular inflammation [21]. Currently, histologically assessed hepatocyte ballooning is a key feature used in the discrimination of NASH from steatosis as it is considered a form of hepatocyte injury associated with fibrogenesis. This distinction is key to patient selection for trial enrolment, and it also serves as a surrogate endpoint for drug efficacy assessment [22]. Although it is a well-established way to categorize NAFLD patients, liver histology evaluation shows several limitations that might impact clinical trial efficacy and patient management. This includes great inter- and intra-observer variation in pathologists’ assessment of grade of activity in general and ballooning specifically [22]. In this sense, new techniques have been developed to better determine liver fat content, such as magnetic resonance-based methods [23], meanwhile others are still under development [24].
Vilar-Gómez et al. demonstrated the independent association between the presence of steatosis and all-cause mortality, observing that patients with steatosis had cardiovascular and malignant mortality rates comparable to those of patients with cirrhosis [25]. These results strongly suggest that patients with severe steatosis have a higher vascular risk than other patients. Thus, clinicians should pay attention not only to patients with advanced fibrosis but also to those patients with moderate or mild fibrosis who have a high degree of steatosis.
Therefore, although fibrosis assessment and liver phenotyping are essential for the prognosis of NAFLD, they are insufficient for the correct characterization and management of NAFLD patients, and new approaches should be developed to overcome the limitations of histology evaluation.

2.2. Metabolic Comorbidities

It has now been established that the primary causes of mortality in patients with NAFLD are cardiovascular disease (CVD) and malignancies, while liver-related mortality occupies the third position. This finding indicates that NAFLD functions as a systemic disorder, which is not unexpected considering its association with insulin resistance (IR) and metabolic syndrome (MetS) [26,27].
NAFLD is closely and bidirectionally associated with MetS and especially with T2DM, whose pathophysiological components are key modifiers of NAFLD development and progression [28]. A recent study published by Ajmera et al. [29] has shown that the prevalence of NAFLD rises to 65% in T2DM patients. Moreover, the global prevalence of NASH rises to 30–40% and significant fibrosis (F2–F4) to 12–20% [30]. Additionally, current clinical evidence highlights that NAFLD could be a precursor to the future development of MetS components and thus linked to an increased cardiovascular risk (CVR) independently of MetS risk factors [31]. Therefore, the association between NAFLD and T2DM brings an additional risk of both hepatic and cardiovascular adverse clinical outcomes; thus, hepatologists should routinely screen for T2DM and perform cardiovascular risk work-ups periodically [32].
We should note that the link between T2DM and NAFLD is complex [33]. These pathologies share many risk factors, such as impairment of glucose and lipid metabolism, and NAFLD is also predictive of T2DM (see review by Muzica et al. [34]). In contrast to other T2DM complications, screening and assessment of NAFLD are not usually routinely done [32]. However, as T2DM may promote progression to NASH and liver fibrosis in up to 20% of the patients [35,36,37], which can eventually turn into liver cirrhosis and/or HCC, screening for NAFLD should be done in these patients. The latest American Association of Clinical Endocrinology recommendation is that patients with T2DM or prediabetes and elevated liver enzymes or fatty liver disease on ultrasound should be evaluated for the presence of NAFLD [37]. These individuals should be screened for liver fibrosis using non-invasive methods, such as FIB-4 and NFS, followed by VCTE or analysis of patented serum biomarkers [34].

2.3. Weight

2.3.1. Obese Patients

NAFLD is the most frequent cause of chronic liver disease, with a global prevalence around 25% [38,39]. However, this percentage is increased in obese patients for whom the prevalence rises to 58% [39]. In obese children and adolescents, the prevalence is also high (44%), being greater in developed countries. Moreover, the prevalence increases as does the Body Mass Index (BMI): being 20.23% in overweight and 38.47% in obese patients [40].
The underlying pathophysiology is based on the presence of IR and adipocyte dysfunction, which lead to lipolysis, with an increase in circulating FFAs and leptin and a decrease in adiponectin, which favors intrahepatic fat accumulation. This fact is worsened with de novo lipogenesis because of fat and carbohydrates in the diet. Moreover, immune cells can infiltrate the liver, further producing a chronic low-grade intrahepatic inflammation. Lipotoxicity and glucotoxicity along with mitochondrial damage and oxidative stress lead to NASH progression and ultimately to the development of fibrosis [41].
Thus, obese patients have greater liver impairment with higher aspartate aminotransferase (AST) and alanine aminotransferase (ALT) levels and more liver fibrosis [42]. The presence of obesity in NAFLD is associated with the occurrence of hypertriglyceridemia (OR 1.51), MetS (OR 2.66), hypertension and T2DM (OR 1.35) [42]. That is, it is associated with a higher prevalence of other CVR factors [43].
As discussed later, diet and exercise are the main treatment of patients with NAFLD and obesity [44]. Weight loss of ≥10% induces high rates of improvement (>80%) not only of the comorbidities but also of all the histological lesions of NAFLD [25].

2.3.2. Lean Patients

Although NAFLD has a strong association with obesity, there is a proportion of cases with a normal BMI, which is usually called “non-obese NAFLD” or “lean NAFLD”. These two terms are not exactly synonyms and vary between studies: the first includes both overweight and normal weight patients and the second, normally, only those with normal BMI [45].
In a recent meta-analysis, the global prevalence of non-obese NAFLD was 40.8% within the NAFLD population and 12.1% within the general population. However, the prevalence varied by BMI, being more frequent in overweight than in normal weight people. Thus, the prevalence of lean NAFLD turned out to be 19.2% in the NAFLD population and 5.1% within the general population [46].
It is important to mention that a normal BMI is not a synonym for a metabolically healthy condition. People with lean MAFLD have lower waist circumference, diastolic blood pressure and serum triglycerides (TG) compared to overweight/obese MAFLD patients. Moreover, they have a different body composition, with significantly lower fatty tissue index, lean tissue index and total body water. However, compared to lean healthy controls, lean MAFLD had a worse metabolic profile, characterized by a higher percentage of hypertension, BMI, TG, low-density lipoprotein (LDL), glucose, HbA1C and lower HDL [47].
It has been stated that lean NAFLD patients have a less severe disease than obese NAFLD ones. In a prospective study, non-obese NAFLD patients seem to have lower NAFLD activity scores (NAS) than obese NAFLD ones, mainly attributable to lesser steatosis and a smaller proportion of ballooning. Moreover, non-obese NAFLD patients had less fibrosis. When analyzing the different parameters of MetS, it was observed that only TG levels were an independent predictor of disease severity [48]. In a retrospective Italian study, similar results were found, with significantly lower proportions of NASH (17% vs. 40%) patients and significant fibrosis (i.e., >F2) (17 vs. 42%) among lean NAFLD patients in comparison to overweight or obese NAFLD patients [49]. However, the presence of MetS seems to be associated with the progression to NASH and significant fibrosis in patients with NAFLD regardless of BMI [50]. In a large retrospective cross-sectional study in Asia, it was also found that MetS in non-obese NAFLD was associated with NASH (OR 1.59) and advanced fibrosis (OR 1.88) [51].
On the other hand, it was classically considered that lean NAFLD patients had a more benign course of illness, with a lower incidence of new onset CVD. However, it has been observed that all-cause cardiovascular and hepatic mortality are not negligible in these patients (incidence per 1000 person–years of 12.1%, 4% and 4.1%, respectively). Although it should be noted that non-obese NAFLD (i.e, patients with normal weight but also overweight) were included in this meta-analysis [46]. In a population-based cross-sectional study carried out in Korea, lean NAFLD patients seemed to have a significantly higher atherosclerotic cardiovascular disease (ASCVD) score and prevalence of a high ASCVD risk compared to obese NAFLD patients. However, this study had some limitations: first, being a cross-sectional study, longitudinal follow-up is not possible; second, the severity of the disease was measured by indirect methods, without biopsy; and third, cardiovascular events were not evaluated [52]. In a retrospective study, one-third of the lean NAFLD patients had carotid atherosclerosis [49].
The pathophysiology of lean NAFLD is not fully understood yet, with multiple factors that can have influence having been described [53]. It may be set by the genetic background and early alterations in bile acid (BA) and gut microbiota profile. Thus, lean NAFLD had a higher chance of carrying at least one PNPLA3 risk allele compared to lean healthy controls [42,54]. Lean NAFLD patients had higher total, primary and secondary BA levels than overweight–obese NAFLD ones, although it was only significant for secondary BAs. Moreover, the composition was different as lean NAFLD patients had lower deoxycholic acid, glycochenodeoxycholic acid and chenodeoxycholic acid but more glycocholic acid [55]. On the other hand, a lower level of various lysophosphatidylcholines, which is linked to obesity and hypertriglyceridemia, has been described in lean NAFLD patients [54].
All in all, the latest guidelines recommend that even with a normal weight, lean NAFLD patients should undergo lifestyle intervention, including exercise, diet modification, and avoidance of fructose- and sugar-sweetened drinks, to target a modest weight loss of 3–5% [56].

2.4. Cardiovascular Diseases

NAFLD patients are at risk of cardiovascular and cardiac diseases. They have more subclinical atherosclerosis, arrhythmias, cardiovascular events, conduction defects, aortic-valve sclerosis and heart failure, which increase disease-related morbimortality [57].
It seems that the metabolic dysfunction that defines MAFLD is associated with higher risks of all-cause mortality and cardiovascular mortality in MAFLD patients compared to patients with NAFLD [58].
In a recent study, three subtypes of NAFLD patients were identified. Subtype A, which phenocopied the metabolome of mice with impaired VLDL-TG secretion, had the lowest CVR measured by Framingham risk score. Moreover, it had lower serum TG, cholesterol, VLDL, small dense LDL and remnant lipoprotein cholesterol compared to type B (intermediate) and C (normal VLDL-TG) [59]. Thus, CVR assessment is a priority. ASCVD can be an easy tool as an ASCVD score ≥ 7.5% was associated with a higher risk of overall and cardiac-specific mortality [60].
In a large cohort study with more than 10,000 NAFLD patients, it was described that NAFLD subjects tended to meet a lower number of ideal health metrics (BMI, smoking, physical activity, diet, blood pressure, cholesterol and glycemia). If these modifiable risk factors were addressed, 66% of all-cause deaths and 83% of cardiovascular deaths were preventable [61].

2.5. Inflammatory Dysfunction

Growing evidence has pointed towards a disproportionately high prevalence of NAFLD and advanced fibrosis in patients with immune-mediated inflammatory diseases (IMID) such as psoriasis [62], inflammatory bowel disease (IBD) [63] and hidradenitis suppurativa [64]. Although this may be explained by an interplay between the distinctive chronic inflammation of IMIDs and the co-existence of metabolic risk factors, the IMID diagnosis acts as an independent risk factor for NAFLD. For instance, in IBD patients, advanced fibrosis was particularly prevalent, regardless of the influence of metabolic risk factors [63]. Therefore, while advanced fibrosis was found to be three times more common in NAFLD–IBD individuals than in the general population with NAFLD, the disparities were even greater when obesity was not present, with a four-fold higher prevalence. Furthermore, in the absence of T2DM, the prevalence of advanced fibrosis was nearly five times higher in the IBD population, and in individuals without both obesity and T2DM, the difference was almost seven times as great [63].
Thus, NAFLD has a disproportionately high tendency to develop in IMID populations, which may be explained by the distinctive chronic inflammatory burden of these conditions [63]. Interestingly, NASH and most IMIDs share some molecular characteristics such the activation of the tumor pathways depending on tumor necrosis factor (TNF)-α or the imbalance in T-cell subtypes such as Th17/Treg. This common pathogenesis may explain, at least in a subset of patients, the development of NASH in the absence of classic metabolic risk factors [64]. Spanish guidelines are moving towards the inclusion of these patients in NAFLD screening strategies [65]. However, international guidelines have not yet recognized the need for NAFLD evaluation in IMID patients [66] and studies regarding the link between NAFLD and IMID pathogenesis are still pending.

2.6. Genetic Factors

Genetic and epidemiological studies indicate strong heritability of hepatic fat content and the key role of genes in the development and progression of liver diseases [67]. In particular, genome-wide association studies (GWAS) have shown a significant relation between several single nucleotide polymorphisms (SNPs) and an increased risk of chronic liver disease [68]. In fact, there are several polymorphisms in different genes associated with NAFLD development and progression. The five more strongly related with disease severity are: patatin-like phospholipase domain (PNPLA3), transmembrane 6 superfamily member 2 (TM6SF2), glucokinase regulator (GCKR), membrane bound O-acyltransferase domain-containing 7 (MBOAT7) and hydroxysteroid 17β-dehydrogenase (HSD17B13) [69] (Table 1).
One of the most described and robustly validated associations is a missense variant in PNPLA3. The substitution of cytosine by guanine in codon 148 results in an amino acid change from isoleucine to methionine in PNPLA3 (rs738409; p.Ile148Met), which is strongly associated with hepatic fat content and inflammation as described by Romeo, S. et al. [70]. PNPLA3 protein is implicated in lipid regulation in hepatocytes and stellate cells. In hepatocytes, PNPLA3 acts as a triacylglycerol lipase and acylglycerol O-acyltransferase which involves catalyzing the transfer of polyunsaturated fatty acids (PUFA) from di- and tri-acylglycerols to phosphocholines [88].
PNPLA3 is degraded through ubiquitination of lysine and subsequent proteosome degradation. The lack of function in PNPLA3 rs738409 and its loss of accessibility to be ubiquitinated leads to a retention of TG and PUFA in the liver [89]. In NAFLD patients, the phenotypic manifestations of this polymorphism are higher TG levels, elevated ALT and AST ratio, severity of steatohepatitis and increased fibrosis [68].
A missense variant in TM6SF2, encoding transmembrane 6 superfamily member 2, is associated with NAFLD. TM6SF2 is mainly expressed in the liver and small intestine, although its exact function is not well known; TM6SF2 regulates fat metabolism, specifically cholesterol synthesis and lipoprotein secretion [90]. The rs58542926 polymorphism encodes a substitution of glutamic acid to lysine at position 167 (p.Glu167Lys). This amino acid change results in a loss-of-function, inducing higher liver TG levels and lower circulating lipoproteins [69].
In vitro studies revealed that TM6SF2 siRNA inhibition was associated with the reduced secretion of very-low density lipoproteins (VLDLs) and increased cellular TG concentrations and lipid droplet levels, whereas TM6SF2 overexpression reduced liver cell steatosis [76]. Apparently, TM6SF2 rs58542926 polymorphism elevates the risk of liver disease but reduces cardiovascular event risk [91].
GCKR controls de novo lipogenesis by regulating the influx of glucose into hepatocytes [78]. Loss-of-function in GCKR (rs1260326; p.Pro446Leu) regulates glucokinase in response to fructose-6-phosphate, activating hepatic glucose uptake. This leads to decreased circulating fasting glucose and insulin levels but increases the production of malonyl-CoA. In fact, a higher concentration of malonyl-CoA favors hepatic fat accumulation by serving as a substrate for lipogenesis and by blocking fatty-acid β-oxidation in the mitochondria [68,92]. Thus, rs1260326 has been related to NAFLD [78].
Under the hypothesis that alcoholic liver disease (ALD) and NAFLD share common genetic determinants, Mancina et al. [93] identified a significant locus for both pathologies using GWAS. MBOAT7 is highly expressed in inflammatory and immune cells. It encodes lysophosphatidylinositol acyltransferase 1 (LPIAT1), which is involved in remodeling arachidonic acid to phosphatidynositol in the Lands cycle [72]. A study in a European descent cohort demonstrated the association between MBOAT7 rs641738 and the development and severity of NAFLD [93].
The latest addition to the list of genes that are involved in NAFLD, based on GWAS studies, was HSD17B13, a liver-specific enzyme that regulates lipid homeostasis. Its aberrant expression and high enzyme activity have been confirmed to promote the development of NAFLD [73]. In contrast, the polymorphism HSD17B13 rs72613567 results in a loss-of-function truncated protein, thus attenuating the progression of NAFLD. Furthermore, HSD17B13 rs72613567 has been associated with reduced serum AST and ALT levels, lower inflammation and NAS including ballooning and fibrosis [94]. These results allow researchers to conclude that the truncated protein has a protective role against liver diseases [95].

2.7. Microbiome

The research carried out in recent years points to the fundamental role of the gut microbiome (GM) in the development of NAFLD as well as in multiple physiological processes such as energy metabolism and immune functions [96].
The human GM is dominated by four bacterial phyla: Bacteroidetes, Firmicutes, Proteobacteria and Actinobacteria [97]. Recent studies show that lean and obese individuals differ in gut bacterial composition [98]. In fact, obesity has been associated with phylum-level changes in the GM, reduced bacterial diversity and altered representation of bacterial genes and metabolic pathways [99]. In NAFLD, it has been demonstrated that alterations in the GM go through an increase in Gram negatives and a decrease in Gram positives, which translates to an increase in Proteobacteria and a decrease in Bacteroidetes-Firmicutes ratio at the phylum level [100].
Dysbiosis has been described as an imbalance in the microbiota composition and function, resulting in a negative effect on the physiology of the host. It could be caused by several environmental factors, such as diet, physical activity, medication and geographical localization [101,102].
The GM interaction with the liver via the “gut–liver axis”, established by the portal vein, enables the transport of GM-derived products directly to the liver and bile and antibody secretion in the opposite way [103]. Alterations in the gut–liver axis caused by GM imbalance and mucosa permeability changes may allow metabolic bacterial products and components to cross the intestinal barrier and reach the liver, causing inflammatory and oxidative responses and exacerbating NAFLD pathology [104]. Some bacterial metabolites could interfere with glucose and lipid metabolism, triggering liver disease; whereas microbial components, pathogen-associated molecular patterns such as lipopolysaccharide and peptidoglycan, can activate pattern recognition receptors (PRRs) in Kupffer cells and hepatic stellate cells, inducing inflammatory responses and contributing to liver injury and fibrosis [103,105,106]. All these conditions are involved in NAFLD progression [102]. Murine models have given further support to the role of microbiota in liver fibrosis as high-fat diet microbiota transplantation to control mice resulted in an increase in liver injury [107].
Focusing on the role of metabolites in alcohol fermentative pathways, acetaldehyde and acetate have been involved in the degradation of intestinal tight junctions [108,109]. In fact, the genes that encode for these enzymes in the gut microbiome are overexpressed in NAFLD, which suggests that alcohol metabolism could be a trigger in this pathology [110].
The GM is also involved in the fermentation of other compounds such as complex carbohydrates to produce short-chain fatty acids including acetate, propionate and butyrate [111]. It has been described that butyrate contributes to the maintenance of the intestinal barrier and its reduction is related to the weakening of tight junctions and an increase in permeability [112,113].
Intestinal metabolic dysregulation has been associated with the metabolism of aromatic amino acids (AAA). In healthy conditions, the AAA tryptophan can be metabolized by the GM, producing indole. Further studies have demonstrated the beneficial effect of indole and its derivatives in upregulating endothelial tight junctions, downregulating pro-inflammatory cytokines production and modulating the secretion of glucagon-like peptide-1 (GLP-1) [114,115].
Thus, accumulated evidence supports the significant role of microbiome dysbiosis in NAFLD onset and progression and it constitutes an important factor to consider for patient management. Furthermore, therapies targeting dysbiosis are under investigation and NAFLD patients showing this condition are expected to be most benefited by its treatment [116].

2.8. Toxics Consumption

2.8.1. Alcohol

NAFLD diagnosis is based on the exclusion of harmful alcohol intake, which has been set as daily ingestion below 20 g (women) and 30 g (men) of pure ethanol by European and American guidelines [117,118]. However, the World Health Organization reported that the average pure ethanol use exceeds those limits (it rises to 32.8 g ethanol/day among women, and more than 40 g ethanol/day among men) [119] and is associated with significant health risks. In fact, moderate (20–40 g ethanol/ day) or heavy (>40 g ethanol/day) alcohol use causes additional liver damage and hepatic steatosis in more than 25% of patients with presumed NAFLD [120,121]. In contrast, there is conflicting evidence of a slightly protective effect of low and moderate alcohol consumption in NAFLD (see review by Petroni et al., 2019 [122]).
The main challenge limiting an accurate diagnosis relies on the fact that NAFLD and ALD have not been reliably distinguished by well-established diagnostic means. Using non-invasive biomarkers, such as ethyl glucuronide (EtG, a metabolite of alcohol), in hair and urine can accurately detect potentially harmful alcohol consumption in patients with NAFLD [120,123]. Hence, according to hair EtG levels, presumed NAFLD patients can be reclassified with regard to their risk of alcohol-related liver damage due to repeated moderate–excessive alcohol consumption [120]. Lifestyle intervention in NAFLD patients with low alcohol consumption should include the recommendation of total abstinence.

2.8.2. Drugs

Drug-induced steatosis (DIS) is usually associated with the prolonged intake of a medication at a specific dose, and it is relatively rare as just 2% of NAFLD cases are estimated to be drug-induced [124]. This phenomenon involves an acute energy crisis through inhibited fatty acids β-oxidation and other impaired mitochondrial and peroxisomal functions, initially resulting in microvesicular steatosis that can usually be reversed (reviewed by Dash et al. [125]). Table 2 summarizes the most commonly used drugs known to cause steatosis.
Imaging methods can estimate hepatic fat content, but these techniques are unreliable when distinguishing steatosis from steatohepatitis [151,152]. Furthermore, serum transaminases’ usefulness as noninvasive indicators of DIS is limited since these proteins are within the reference range in most individuals with hepatic steatosis [153,154,155]. Hence, finding sensitive and specific serum biomarkers is key to assessing the contribution of drugs to NAFLD. Cytokeratin 18 (CK18), fibroblast growth factor 21 (FGF21), insulin-like growth factor binding protein 1 (IGFBP1), several microRNAs and forkhead box protein A1 (FOXA1) have been identified as potential biomarkers for DIS detection and prognosis, as reviewed by Pavlik et al. [156].

2.9. Concomitant Infections

Since the improvement of antiretroviral regimens, NAFLD has emerged as a growing concern in the long-term management of patients with HIV mono-infection. HIV infection itself induces various metabolic alterations that can lead to steatosis by disrupting fatty acid β-oxidation in the liver and adipose tissue [157,158]. Nevertheless, and as mentioned above, it should also be noted that antiretroviral therapy may be partially responsible for this steatogenic effect. Moreover, HIV invasion of hepatic stellate cells triggers fibrosis [159,160], favoring liver damage progression to NASH.
Several studies have attempted to determine the prevalence of NAFLD or NASH in HIV-infected patients, and those have been reviewed in detail by authors such as Verna [161], Squillace et al. [162], and Morrison et al. [163]. However, the definition of NAFLD, study populations, group matching criteria, and methods for fatty liver assessment are heterogeneous, making the outcomes difficult to interpret or even contradictory among different studies. Previous results suggest that NAFLD prevalence in HIV-infected individuals is higher (30–50%) and progresses at an increased rate compared to the general population, with antiretroviral exposure being an additional risk factor [164,165]. Lipodystrophy syndrome, which includes abnormal fat distribution and increased visceral adiposity, is usually present in HIV-positive patients [166,167] and it directly contributes to NAFLD development [164,165].
HIV-induced steatosis and/or fibrosis can be detected through imaging techniques [168,169] and ultrasound/transient elastography [170,171,172]. Regarding blood tests, HIV-positive patients at risk of NASH may show increased levels of serum transaminases [173,174,175] and, more precisely, higher scores for non-invasive markers of fibrosis (FIB-4, APRI) [169,176,177,178,179,180] compared to HIV-negative patients. However, these tests might not be accurate enough, so there is an urgent need to develop and validate new non-invasive biomarkers and imaging assessments for liver disease in HIV-positive patients. Recently, several proteins have aroused interest as biomarkers for detecting steatosis (FGF21 [181], IL-18 [182]) and fibrosis (CK18 [183]), as well as other proteins involved in tissue repair and immune response pathways [184] in these individuals. In addition, some polymorphisms may predict NAFLD development in HIV-infected patients [185], and there is an increasing focus on circulating miRNAs as a non-invasive reflection of liver disease progression in people living with HIV (thoroughly reviewed by Martinez et al. [186]).

3. NAFLD Treatment

3.1. Treatments to Rule Them All

Lifestyle intervention, with modifications in diet and physical activity, has become the first line treatment for patients with NAFLD. A greater extent of weight loss, induced by lifestyle changes, is associated with the level of improvement in histologic features of NASH. The highest rates of NAS reduction, NASH resolution and fibrosis regression occurred in patients with weight losses of 10% or more [25]. Furthermore, age, sex, T2DM and genes impact on the effect of diet in weight loss and NASH resolution [187]. These factors are integrated in the so called ‘nutritional geometry’, which considers the relevance of nutrition, science and the environment to understand how food components interact to regulate the properties of diets [188]. Stratifying patients according to the geometry of nutrition could improve the rate of response [187].
Regarding weight loss, therapies such as bariatric surgery and metabolic endoscopic techniques can be useful alternatives as only 10% of participants achieved enough body weight reduction through lifestyle interventions [189]. Bariatric surgery has also been shown to improve obesity, its metabolic consequences and NASH [190,191]. However, given the surgical risk, it cannot be considered a first-line therapy, especially in those patients with decompensated cirrhosis or portal hypertension. In this context, endoscopic bariatric techniques have emerged as a potential treatment option as they can reproduce those benefits in a minimally invasive manner [192]. However, this kind of intervention might be eligible only in obese patients.
On the other hand, a recent expert meeting has gathered strong evidence that regular physical activity plays an important role in preventing NAFLD and improving intermediate clinical outcomes [193]. Various studies have demonstrated an improvement in NAFLD with personalized physical exercise programs [194], even in the absence of significant weight loss [195,196], and a reduction of the hepatic venous pressure gradient in cirrhotic patients [197].
The prescription of an appropriate diet and the indication of physical exercise in proportion to the disease and the characteristics of the patients is the main curative option for all NAFLD patients, including lean NAFLD [55,198]. A randomized controlled trial from Asia demonstrates using MR spectroscopy that almost half of non-obese individuals achieved NAFLD remission with 3–5% weight loss [198].
Thus, an innovative therapeutic strategy in this setting would be the constitution of multidisciplinary units integrating clinicians (hepatologists, endocrinologists, cardiologists, internists), physiotherapists, nutritionists, nurses, social educators and, of course, patients, where a health-promoting diet, avoidance of tobacco, alcohol and other toxins, and sustainable, inclusive and adapted physical activity is prescribed considering the needs of each patient. Furthermore, such multidisciplinary units allow the integration of adequate assessments for the risk of both significant liver and vascular disease, macro and/or microvascular complications of T2DM, the risk of hepatic and extrahepatic neoplasms and other potential comorbidities.

3.2. Targeted Therapy

Thanks to the continuous research on NAFLD pathogenesis, several druggable targets have been identified and thus targeted therapies for NAFLD treatment have entered clinical trials, which have been recently and extensively reviewed by Santos-Laso et al. [199]. However, limited impact has been achieved for now, probably due to the extensive placebo effect in NAFLD [200], the complexity of the pathological pathways involved [189], the short follow-up time for the expected outcomes to be evaluated and the limited characterization of NAFLD patients before inclusion in clinical trials [201].
Table 3 summarizes the main targeted therapies for NAFLD which are currently under evaluation in Phase III clinical trials.

3.2.1. Targeting Lipid Metabolism

Peroxisome proliferator-activated receptors (PPARs) are ligand-activated transcription factors which include PPARα, PPARγ and PPARβ/δ. The pan-PPAR agonist lanifibranor acts through the activation of all PPAR isoforms, reducing TG levels and increasing insulin sensitization, glucose metabolism and fatty acid metabolism. Lanifibranor has successfully completed a 24-week phase IIb trial, meeting its primary endpoint of a reduction of two points or more in the SAF (steatosis, activity and fibrosis) score, with no increase in fibrosis, and the secondary endpoint of reducing fibrosis by at least one stage without worsening NASH. Lanifibranor is well tolerated, although side effects include mild weight gain. Thus, a phase III study is underway [20].
Statins have been demonstrated to reduce steatosis in those with NASH and it prevents liver events in patients with metabolic syndrome with advanced NASH. Furthermore, a possible protective role of statin treatment against NAFLD progression to HCC was also demonstrated in observational studies. Interestingly, it has been shown that statins reduce CVR more in NAFLD vs. non-NAFLD high-risk individuals [32]. Thus, evaluation in a phase III clinical trial of rosuvastatin treatment for NAFLD is about to start recruitment (Table 3).
Oltipraz is a synthetic dithiolethione that functions as an anti-steatogenic agent against NAFLD by inhibiting LXR-α activity, which decreases the expression of SREBP-1c within the liver, reducing the synthesis of fatty acids but enhancing lipid oxidation [202]. Although two phase III clinical trials have been completed, no data are available yet (Table 3). However, the results from the phase II trial showed that oltipraz decreased liver fat content and BMI while absolute changes in insulin resistance, liver enzymes, lipids and cytokines were not significant [202].
Resmetirom is a liver-directed, orally active, selective thyroid hormone receptor-β agonist designed to improve NASH by increasing hepatic fat metabolism and reducing lipotoxicity. In phase II clinical trials, Resmetirom treatment resulted in significant reductions in hepatic fat after 12 weeks and 36 weeks of treatment in patients with NASH, while the adverse events were transient mild diarrhea and nausea [203]. Two phase III trials are currently recruiting participants (Table 3).
It could be hypothesized that obese patients, those with genetic predisposition to lipid accumulation or increased circulating levels of TG due to metabolic syndrome, might benefit from therapies targeting lipid metabolism.

3.2.2. Targeting Glucose Metabolism

Pioglitazone is a PPARγ agonist used in the treatment of T2DM due to its properties as an insulin sensitizer [204]. Moreover, results have been published regarding pioglitazone evaluation in non-diabetic NAFLD patients in comparison to vitamin E supplementation [205]. There was no benefit of pioglitazone compared to placebo for the primary outcome; however, significant benefits of pioglitazone were observed for some of the secondary outcomes such as steatohepatitis resolution, decrease in mean AST and ALT levels and improvement in insulin resistance [205].
GLP-1 receptor agonists (GLP-1RA) are widely used in the treatment of T2DM. Studies have found that GLP-1R has multiple biological effects, such as neuroinflammation reduction, nerve growth promotion, heart function improvement, appetite suppression, gastric emptying delay, blood lipid metabolism regulation and fat deposition reduction. Thus, effects of GLP-1RA include neuroprotection, cardiovascular protection and metabolic regulation [206]. Semaglutide has completed a phase II trial showing resolution of NASH with no worsening of fibrosis. Although it was unable to achieve its secondary outcome of improvement of fibrosis with no worsening of NASH, the drug induces a significant weight loss and most common adverse events were gastrointestinal [20]. Encouraging pilot results of the evaluation of exenatide, another GLP-1RA, for the treatment of NAFLD have been released [207]. Moreover, cotadutide, a dual GLP-1RA and glucagon receptor agonist, has been evaluated in a phase IIb clinical trial showing improved glycemic control and weight loss, along with improvements in hepatic parameters such as reduction in ALT, AST and gamma-glutamyltransferase (GGT) levels, as well as improvements in NFS and FIB-4 index [208].
Dipeptidyl peptidase 4 (DPP-4) inhibitors work by blocking the enzymatic inactivation of endogenous incretin hormones, resulting in glucose-dependent insulin release and a decrease in glucagon secretion [32]. Early results evaluating DPP-4 inhibitor vildagliptin in NAFLD patients with T2DM have shown significant improvement in blood sugar regulation, BMI, ALT, liver fibrosis and steatosis indices [209].
SGLT-2 inhibitors promote urinary excretion of glucose by inhibiting its renal proximal tubular reabsorption [32]. Dapagliflozin and empagliflozin are SGLT-2 inhibitors undergoing phase III clinical trials for the treatment of NASH. To date, it has been demonstrated that dapagliflozin can markedly reduce hepatic enzymes and metabolic indicators and improve body composition [210].
These drugs are prescribed in the treatment of T2DM. Thus, clinical guidelines have already included the preferential use of drugs with effects on the liver in the management of T2DM patients with NAFLD [211].
Table 3. Targeted drugs for the treatment of adult NAFLD being evaluated in phase III clinical trials [212].
Table 3. Targeted drugs for the treatment of adult NAFLD being evaluated in phase III clinical trials [212].
MechanismDrugIdentifierInterventionTitleStatus
Lipid metabolismLanifibranorNCT04849728Drug: Lanifibranor|Drug: PlaceboA Phase 3 Study Evaluating Efficacy and Safety of Lanifibranor Followed by an Active Treatment Extension in Adult Patients With (NASH) and Fibrosis Stages F2 and F3 (NATiV3)Recruiting
RosuvastatinNCT05731596Drug: Rosuvastatin 20 mg Oral Tablet|Drug: Coenzyme Q10 100 mg Oral CapsuleComparative Clinical Study to Evaluate the Efficacy and Safety of Rosuvastatin vs. CoQ10 on Non-alcoholic SteatohepatitisNot yet recruiting
OltiprazNCT04142749Drug: Oltipraz|Drug: PlacebosOltipraz for Liver Fat Reduction in Patients with Non-alcoholic Fatty Liver Disease Except for Liver CirrhosisCompleted
NCT02068339Drug: Oltipraz 1 (90 mg)|Drug: Placebo|Drug: Oltipraz 2 (120 mg)Efficacy and Safety of Oltipraz for Liver Fat Reduction in Patients with Non-alcoholic Fatty Liver Disease Except for Liver CirrhosisCompleted
ResmetiromNCT03900429Drug: Resmetirom|Drug: PlaceboA Phase 3 Study to Evaluate the Efficacy and Safety of MGL-3196 (Resmetirom) in Patients with NASH and FibrosisRecruiting
NCT04197479Drug: Placebo|Drug: ResmetiromA Phase 3 Study to Evaluate the Safety and Biomarkers of Resmetirom (MGL-3196) in Non-Alcoholic Fatty Liver Disease (NAFLD) PatientsActive, not recruiting
Glucose metabolismPioglitazoneNCT05521633Drug: Metformin and PioglitazoneComparison of the Effects of Metformin and Pioglitazone on Liver Enzymes and Ultrasound Changes in Non-Diabetic Non-alcoholic Fatty LiverCompleted
NCT05605158Drug: Pioglitazone 30 mg|Drug: Empagliflozin 10 mgComparative Clinical Study Between Empagliflozin Versus Pioglitazone in Non-diabetic Patients with Non-alcoholic SteatohepatitisNot yet recruiting
NCT00063622Drug: Pioglitazone|Dietary Supplement: Vitamin E|Drug: Matching placeboPioglitazone vs. Vitamin E vs. Placebo for Treatment of Non-Diabetic Patients with Non-alcoholic Steatohepatitis (PIVENS)Completed
SemaglutideNCT05067621Drug: Semaglutide Pen Injector|Drug: PlaceboSemaglutide Effects in Obese Youth with Prediabetes/New Onset Type 2 Diabetes and Non-alcoholic Fatty Liver DiseaseNot yet recruiting
NCT03919929Drug: Semaglutide 3 mg and 7 mg [Rybelsus]|Other: Weight loss dietTreating PCOS With Semaglutide vs. Active Lifestyle InterventionRecruiting
NCT04822181Drug: Semaglutide|Drug: PlaceboResearch Study on Whether Semaglutide Works in People with Non-alcoholic Steatohepatitis (NASH)Recruiting
ExenatideNCT00650546Drug: ExenatideRole of Exenatide in NASH-a Pilot StudyCompleted
CotadutideNCT05364931Drug: Cotadutide|Drug: PlaceboA Study to Evaluate the Safety and Efficacy of Cotadutide Given by Subcutaneous Injection in Adult Participants with Non-cirrhotic Non-alcoholic Steatohepatitis With Fibrosis.Active, not recruiting
VildagliptinNCT03925701Drug: Vildagliptin|Drug: vildagliptin\metforminClinical Study Evaluating Vildagliptin Versus Vildagliptin/Metformin on NAFLD With DMRecruiting
DapagliflozinNCT05308160Drug: Dapagliflozin 10 mg Tab|Drug: PlaceboA Single Center, Randomized, Open Label, Parallel Group, Phase 3 Study to Evaluate the Efficacy of Dapagliflozin in Subjects with Non-alcoholic Fatty Liver DiseaseRecruiting
NCT03723252Drug: Dapagliflozin|Drug: PlaceboDapagliflozin Efficacy and Action in NASHRecruiting
EmpagliflozinNCT05605158Drug: Pioglitazone 30 mg|Drug: Empagliflozin 10 mgComparative Clinical Study Between Empagliflozin Versus Pioglitazone in Non-diabetic Patients with Non-alcoholic SteatohepatitisNot yet recruiting
Bile acid metabolismObeticholic acidNCT02548351Drug: Obeticholic Acid|Drug: PlaceboRandomized Global Phase 3 Study to Evaluate the Impact on NASH With Fibrosis of Obeticholic Acid TreatmentActive, not recruiting
NCT03439254Drug: Obeticholic acid (10 mg)|Drug: Obeticholic acid (10 mg to 25 mg)|Drug: PlaceboStudy Evaluating the Efficacy and Safety of Obeticholic Acid in Subjects with Compensated Cirrhosis Due to Non-alcoholic SteatohepatitisCompleted
AramcholNCT04104321Drug: Aramchol|Drug: PlaceboA Clinical Study to Evaluate the Efficacy and Safety of Aramchol in Subjects with NASH (ARMOR) (ARMOR)Suspended
Oxidative stress, inflammation and fibrosisN-acetylcysteineNCT05589584Drug: N acetyl cysteine with weight reductionN-acetyl Cysteine and Patients with Non-alcoholic Fatty Liver DiseaseRecruiting
PentoxifyllineNCT05284448Drug: pentoxifylline (Trental SR®)Pentoxifylline in Treatment of Patients with Non-alcoholic SteatohepatitisActive, not recruiting
NCT00267670Drug: Pentoxifylline|Drug: PlaceboPentoxifylline/Non-alcoholic Steatohepatitis (NASH) Study: The Effect of Pentoxifylline on NASHCompleted
SecukinumabNCT04237116Biological: Investigational Arm—secukinumab|Biological: Control Arm—placeboA Study of Secukinumab Treatment in Patients with Plaque Psoriasis and Coexisting Non-alcoholic Fatty Liver Disease (NAFLD)Completed
LubiprostoneNCT05768334Drug: Lubiprostone 24 Mcg Oral CapEfficacy and Tolerability of Lubiprostone in Patients with Non-alcoholic Fatty Liver DiseaseRecruiting

3.2.3. Targeting Bile Acid Metabolism

Obeticholic acid (OCA) is a potent FXR agonist evaluated for the treatment of NASH-mediated fibrosis thanks to its ability to reduce liver fat and fibrosis in animal models of NAFLD. Currently being tested in phase III clinical trials [213], phase II studies demonstrated that OCA treatment improved multiple histological NASH features [214].
Aramchol is a fatty acid–BA conjugate that has demonstrated an ability to reduce liver fat and inflammation in NAFLD patients. Results from the phase IIb trial showed that aramchol was safe and well tolerated [215]. Although the primary end point of a reduction in liver fat did not meet the pre-specified significance level, the observed safety and changes in liver histology and enzymes encouraged the initiation of phase III trials [215], whose interim analysis revealed that the open-label part met its objectives (Table 3).
These treatments may be especially useful in the management of lean NAFLD as these patients have shown specific alterations of BA metabolism [55].

3.2.4. Targeting Oxidative Stress, Inflammation and Fibrosis

N-acetylcysteine is frequently used where intracellular oxidant–antioxidant balance is concerned and it has protective effects against liver injury [216]. Its potential as antioxidant treatment in NAFLD has been demonstrated in animal models [217,218], whereas information in NAFLD patients is scarce but promising [216]. Thus, a phase III clinical trial evaluating the effect of N-acetylcysteine on markers of oxidative stress and insulin resistance in patients with NAFLD is currently recruiting participants (Table 3).
Pentoxifylline is a methylxanthine derivative with a variety of physiological effects at the cellular level, which include decreases in TNF-α gene transcription, affecting multiple steps in the cytokine/chemokine pathway that has been implicated in NAFLD pathogenesis. Thus, it has been evaluated in several clinical trials mostly showing beneficial effects in weight loss, improved liver function and histological changes in patients with NAFLD/NASH [219]. However, other studies have failed in demonstrating pentoxifylline’s effectiveness in reducing transaminases compared to placebos, and it did not positively affect any of the metabolic markers postulated to contribute to NASH [220].
Secukinumab is a monoclonal antibody against IL-17 used in the treatment of psoriasis. It has been shown to have neutral effects on fasting plasma glucose, lipid parameters and liver enzymes, while reducing levels of CRP, a marker for systemic inflammation, and markers of oxidative stress. Secukinumab produced improvements in arterial elasticity, coronary artery function and myocardial deformation indices, thus protecting from CVR [221]. However, publication of phase III clinical trial results evaluating liver function is pending (Table 3).
Finally, lubiprostone is a laxative drug that improves intestinal permeability. It was reported to ameliorate increases in intestinal permeability induced by a high-fat and high-cholesterol diet in an atherosclerosis mouse model, while in humans it improved the increased intestinal permeability induced by non-steroidal anti-inflammatory drugs. Thus, lubiprostone might prevent the excessive inflammation and fibrosis induced by gut-derived endotoxin in NAFLD patients [222]. Results from the phase IIa study have shown that lubiprostone was well tolerated and reduced the levels of liver enzymes in patients with NAFLD and constipation [222]. Therefore, recruitment for the phase III clinical trial is already open.
Treatments targeting inflammation and fibrosis might be eligible for patients with more advanced disease or those with enhanced inflammation due to co-morbidities such as IMIDs, whereas targeting intestinal permeability could be indicated for those with dysbiosis or IBD.

4. Concluding Remarks

The hallmark of NAFLD is the accumulation of lipids in the liver that results from deranged lipid metabolism. Consequently, NAFLD is strongly associated with obesity, insulin resistance and dyslipidemia. However, inflammation may precede steatosis as inflammatory events may lead to lipid accumulation. Therefore, there are many factors influencing NAFLD initiation and progression, such as environmental exposure, lifestyle, genetic susceptibility, metabolic status and the microbiome. The phenotypic manifestation of fatty liver diseases likely reflects the sum of the dynamic and complex systems-level interactions of these drivers; it follows that effective treatment requires that they be targeted with precision and based on a person’s phenotype [223]. Importantly, morbimortality in patients with NAFLD involves extra-hepatic organs as it is considered a mediator of systemic diseases including CVD. This further contributes to NAFLD’s heterogeneity, representing a major challenge in discovering highly effective therapies. Obtaining a comprehensive landscape of the main NAFLD drivers and patient outcome determinants should facilitate patient stratification and identification of disease subtypes with different natural history and liver disease courses. Therefore, a multi-omic and clinical data integration approach of NAFLD patients could help us to properly subphenotype and stratify patients, paving the way for precision medicine in NAFLD. On the other hand, implementing optimal strategies to promote physical activity, prescribing an appropriate diet and changing the model of care through the use of digital tools such as telemedicine are crucial elements that will help in maintaining healthy habits in patients with NAFLD, as well as being the current curative and preventive options. Prioritizing research in these areas and developing innovative strategies to address this growing public health concern is essential.

Author Contributions

Conceptualization and supervision: M.A.-P., M.T.A.-L., P.I. and J.C. Searching through the literature, interpreting the data and writing the paper: M.A.-P., M.D.B., A.P.-V. and C.J.-G. Figures: A.P.-V. Editing and critical revision of the entire manuscript: A.S.-L., M.A.-P., M.T.A.-L., P.I. and J.C. 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.

Data Availability Statement

Not applicable.

Acknowledgments

The English revision of the manuscript by Paula Echevarria and Dermot Erskine is sincerely appreciated. In addition, we want to thank the reviewers for their critical revision of our manuscript to improve its quality.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mota, M.; Banini, B.A.; Cazanave, S.C.; Sanyal, A.J. Molecular mechanisms of lipotoxicity and glucotoxicity in nonalcoholic fatty liver disease. Metabolism 2016, 65, 1049–1061. [Google Scholar] [CrossRef] [Green Version]
  2. Kleiner, D.E.; Brunt, E.M.; Van Natta, M.; Behling, C.; Contos, M.J.; Cummings, O.W.; Ferrell, L.D.; Liu, Y.C.; Torbenson, M.S.; Unalp-Arida, A.; et al. Design and validation of a histological scoring system for nonalcoholic fatty liver disease. Hepatology 2005, 41, 1313–1321. [Google Scholar] [CrossRef]
  3. Nasr, P.; Ignatova, S.; Kechagias, S.; Ekstedt, M. Natural history of nonalcoholic fatty liver disease: A prospective follow-up study with serial biopsies. Hepatol. Commun. 2018, 2, 199–210. [Google Scholar] [CrossRef] [Green Version]
  4. Sanyal, A.J.; Van Natta, M.L.; Clark, J.; Neuschwander-Tetri, B.A.; Diehl, A.; Dasarathy, S.; Loomba, R.; Chalasani, N.; Kowdley, K.; Hameed, B.; et al. Prospective Study of Outcomes in Adults with Nonalcoholic Fatty Liver Disease. N. Engl. J. Med. 2021, 385, 1559–1569. [Google Scholar] [CrossRef]
  5. Paik, J.M.; Henry, L.; De Avila, L.; Younossi, E.; Racila, A.; Younossi, Z.M. Mortality Related to Nonalcoholic Fatty Liver Disease Is Increasing in the United States. Hepatol. Commun. 2019, 3, 1459–1471. [Google Scholar] [CrossRef] [Green Version]
  6. Rosato, V.; Masarone, M.; Dallio, M.; Federico, A.; Aglitti, A.; Persico, M. NAFLD and Extra-Hepatic Comorbidities: Current Evidence on a Multi-Organ Metabolic Syndrome. Int. J. Environ. Res. Public Health 2019, 16, 3415. [Google Scholar] [CrossRef] [Green Version]
  7. Lim, G.E.H.; Tang, A.; Ng, C.H.; Chin, Y.H.; Lim, W.H.; Tan, D.J.H.; Yong, J.N.; Xiao, J.; Lee, C.W.; Chan, M.; et al. An Observational Data Meta-analysis on the Differences in Prevalence and Risk Factors Between MAFLD vs NAFLD. Clin. Gastroenterol. Hepatol. 2023, 21, 619–629.e7. [Google Scholar] [CrossRef]
  8. Eslam, M.; Newsome, P.N.; Sarin, S.K.; Anstee, Q.M.; Targher, G.; Romero-Gomez, M.; Zelber-Sagi, S.; Wai-Sun Wong, V.; Dufour, J.F.; Schattenberg, J.M.; et al. A new definition for metabolic dysfunction-associated fatty liver disease: An international expert consensus statement. J. Hepatol. 2020, 73, 202–209. [Google Scholar] [CrossRef]
  9. Dallio, M.; Sangineto, M.; Romeo, M.; Villani, R.; Romano, A.D.; Loguercio, C.; Serviddio, G.; Federico, A. Immunity as Cornerstone of Non-Alcoholic Fatty Liver Disease: The Contribution of Oxidative Stress in the Disease Progression. Int. J. Mol. Sci. 2021, 22, 436. [Google Scholar] [CrossRef]
  10. Tilg, H.; Moschen, A.R. Evolution of inflammation in nonalcoholic fatty liver disease: The multiple parallel hits hypothesis. Hepatology 2010, 52, 1836–1846. [Google Scholar] [CrossRef]
  11. Juanola, O.; Martínez-López, S.; Francés, R.; Gómez-Hurtado, I. Non-Alcoholic Fatty Liver Disease: Metabolic, Genetic, Epigenetic and Environmental Risk Factors. Int. J. Environ. Res. Public Health 2021, 18, 5227. [Google Scholar] [CrossRef]
  12. Fryk, E.; Olausson, J.; Mossberg, K.; Strindberg, L.; Schmelz, M.; Brogren, H.; Gan, L.M.; Piazza, S.; Provenzani, A.; Becattini, B.; et al. Hyperinsulinemia and insulin resistance in the obese may develop as part of a homeostatic response to elevated free fatty acids: A mechanistic case-control and a population-based cohort study. eBioMedicine 2021, 65, 103264. [Google Scholar] [CrossRef]
  13. Kucuk, S.; Niven, J.; Caamano, J.; Jones, S.W.; Camacho-Muñoz, D.; Nicolaou, A.; Mauro, C. Unwrapping the mechanisms of ceramide and fatty acid-initiated signals leading to immune-inflammatory responses in obesity. Int. J. Biochem. Cell Biol. 2021, 135, 105972. [Google Scholar] [CrossRef]
  14. Ampuero, J.; Romero-Gomez, M. Stratification of patients in NASH clinical trials: A pitfall for trial success. JHEP Rep. 2020, 2, 100148. [Google Scholar] [CrossRef]
  15. Alonso, C.; Fernández-Ramos, D.; Varela-Rey, M.; Martínez-Arranz, I.; Navasa, N.; Van Liempd, S.M.; Lavín Trueba, J.L.; Mayo, R.; Ilisso, C.P.; de Juan, V.G.; et al. Metabolomic Identification of Subtypes of Nonalcoholic Steatohepatitis. Gastroenterology 2017, 152, 1449–1461.e7. [Google Scholar] [CrossRef] [Green Version]
  16. Iruarrizaga-Lejarreta, M.; Varela-Rey, M.; Fernández-Ramos, D.; Martínez-Arranz, I.; Delgado, T.C.; Simon, J.; Juan, V.G.; delaCruz-Villar, L.; Azkargorta, M.; Lavin, J.L.; et al. Role of Aramchol in steatohepatitis and fibrosis in mice. Hepatol. Commun. 2017, 1, 911–927. [Google Scholar] [CrossRef] [Green Version]
  17. Son, J.W.; Shoaie, S.; Lee, S. Systems Biology: A Multi-Omics Integration Approach to Metabolism and the Microbiome. Endocrinol. Metab. 2020, 35, 507–514. [Google Scholar] [CrossRef]
  18. Powell, E.E.; Wong, V.W.; Rinella, M. Non-alcoholic fatty liver disease. Lancet 2021, 397, 2212–2224. [Google Scholar] [CrossRef]
  19. Mózes, F.E.; Lee, J.A.; Selvaraj, E.A.; Jayaswal, A.N.A.; Trauner, M.; Boursier, J.; Fournier, C.; Staufer, K.; Stauber, R.E.; Bugianesi, E.; et al. Diagnostic accuracy of non-invasive tests for advanced fibrosis in patients with NAFLD: An individual patient data meta-analysis. Gut 2022, 71, 1006–1019. [Google Scholar] [CrossRef]
  20. Dufour, J.F.; Anstee, Q.M.; Bugianesi, E.; Harrison, S.; Loomba, R.; Paradis, V.; Tilg, H.; Wong, V.W.; Zelber-Sagi, S. Current therapies and new developments in NASH. Gut 2022, 71, 2123–2134. [Google Scholar] [CrossRef]
  21. Hjelkrem, M.; Stauch, C.; Shaw, J.; Harrison, S.A. Validation of the non-alcoholic fatty liver disease activity score. Aliment. Pharmacol. Ther. 2011, 34, 214–218. [Google Scholar] [CrossRef]
  22. Brunt, E.M.; Clouston, A.D.; Goodman, Z.; Guy, C.; Kleiner, D.E.; Lackner, C.; Tiniakos, D.G.; Wee, A.; Yeh, M.; Leow, W.Q.; et al. Complexity of ballooned hepatocyte feature recognition: Defining a training atlas for artificial intelligence-based imaging in NAFLD. J. Hepatol. 2022, 76, 1030–1041. [Google Scholar] [CrossRef]
  23. Bohte, A.E.; Nederveen, A.J.; Stoker, J. Hepatic fat-content assessment using magnetic resonance-based methods. Imaging Med. 2011, 3, 193–206. [Google Scholar] [CrossRef]
  24. Lee, L.W.; Yen, J.B.; Lu, H.K.; Liao, Y.S. Prediction of Nonalcoholic Fatty Liver Disease by Anthropometric Indices and Bioelectrical Impedance Analysis in Children. Child. Obes. 2021, 17, 551–558. [Google Scholar] [CrossRef]
  25. Vilar-Gomez, E.; Martinez-Perez, Y.; Calzadilla-Bertot, L.; Torres-Gonzalez, A.; Gra-Oramas, B.; Gonzalez-Fabian, L.; Friedman, S.L.; Diago, M.; Romero-Gomez, M. Weight Loss Through Lifestyle Modification Significantly Reduces Features of Nonalcoholic Steatohepatitis. Gastroenterology 2015, 149, 367–378.e5; quiz e314–e365. [Google Scholar] [CrossRef]
  26. Adams, L.A.; Anstee, Q.M.; Tilg, H.; Targher, G. Non-alcoholic fatty liver disease and its relationship with cardiovascular disease and other extrahepatic diseases. Gut 2017, 66, 1138–1153. [Google Scholar] [CrossRef] [Green Version]
  27. Alberti, K.G.; Eckel, R.H.; Grundy, S.M.; Zimmet, P.Z.; Cleeman, J.I.; Donato, K.A.; Fruchart, J.C.; James, W.P.; Loria, C.M.; Smith, S.C.; et al. Harmonizing the metabolic syndrome: A joint interim statement of the International Diabetes Federation Task Force on Epidemiology and Prevention; National Heart, Lung, and Blood Institute; American Heart Association; World Heart Federation; International Atherosclerosis Society; and International Association for the Study of Obesity. Circulation 2009, 120, 1640–1645. [Google Scholar] [CrossRef] [Green Version]
  28. Jarvis, H.; Craig, D.; Barker, R.; Spiers, G.; Stow, D.; Anstee, Q.M.; Hanratty, B. Metabolic risk factors and incident advanced liver disease in non-alcoholic fatty liver disease (NAFLD): A systematic review and meta-analysis of population-based observational studies. PLoS Med. 2020, 17, e1003100. [Google Scholar] [CrossRef]
  29. Ajmera, V.; Cepin, S.; Tesfai, K.; Hofflich, H.; Cadman, K.; Lopez, S.; Madamba, E.; Bettencourt, R.; Richards, L.; Behling, C.; et al. A prospective study on the prevalence of NAFLD, advanced fibrosis, cirrhosis and hepatocellular carcinoma in people with type 2 diabetes. J. Hepatol. 2023, 78, 471–478. [Google Scholar] [CrossRef]
  30. Stefan, N.; Cusi, K. A global view of the interplay between non-alcoholic fatty liver disease and diabetes. Lancet Diabetes Endocrinol. 2022, 10, 284–296. [Google Scholar] [CrossRef]
  31. Targher, G.; Byrne, C.D.; Lonardo, A.; Zoppini, G.; Barbui, C. Non-alcoholic fatty liver disease and risk of incident cardiovascular disease: A meta-analysis. J. Hepatol. 2016, 65, 589–600. [Google Scholar] [CrossRef] [Green Version]
  32. Genua, I.; Iruzubieta, P.; Rodríguez-Duque, J.C.; Pérez, A.; Crespo, J. NAFLD and type 2 diabetes: A practical guide for the joint management. Gastroenterol. Hepatol. 2022. Online ahead of print. [Google Scholar] [CrossRef]
  33. Targher, G.; Corey, K.E.; Byrne, C.D.; Roden, M. The complex link between NAFLD and type 2 diabetes mellitus—Mechanisms and treatments. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 599–612. [Google Scholar] [CrossRef]
  34. Muzica, C.M.; Sfarti, C.; Trifan, A.; Zenovia, S.; Cuciureanu, T.; Nastasa, R.; Huiban, L.; Cojocariu, C.; Singeap, A.M.; Girleanu, I.; et al. Nonalcoholic Fatty Liver Disease and Type 2 Diabetes Mellitus: A Bidirectional Relationship. Can. J. Gastroenterol. Hepatol. 2020, 2020, 6638306. [Google Scholar] [CrossRef]
  35. Lomonaco, R.; Godinez Leiva, E.; Bril, F.; Shrestha, S.; Mansour, L.; Budd, J.; Portillo Romero, J.; Schmidt, S.; Chang, K.L.; Samraj, G.; et al. Advanced Liver Fibrosis Is Common in Patients with Type 2 Diabetes Followed in the Outpatient Setting: The Need for Systematic Screening. Diabetes Care 2021, 44, 399–406. [Google Scholar] [CrossRef]
  36. Younossi, Z.M.; Golabi, P.; de Avila, L.; Paik, J.M.; Srishord, M.; Fukui, N.; Qiu, Y.; Burns, L.; Afendy, A.; Nader, F. The global epidemiology of NAFLD and NASH in patients with type 2 diabetes: A systematic review and meta-analysis. J. Hepatol. 2019, 71, 793–801. [Google Scholar] [CrossRef]
  37. Cusi, K.; Isaacs, S.; Barb, D.; Basu, R.; Caprio, S.; Garvey, W.T.; Kashyap, S.; Mechanick, J.I.; Mouzaki, M.; Nadolsky, K.; et al. American Association of Clinical Endocrinology Clinical Practice Guideline for the Diagnosis and Management of Nonalcoholic Fatty Liver Disease in Primary Care and Endocrinology Clinical Settings: Co-Sponsored by the American Association for the Study of Liver Diseases (AASLD). Endocr. Pract. 2022, 28, 528–562. [Google Scholar] [CrossRef]
  38. Younossi, Z.M.; Koenig, A.B.; Abdelatif, D.; Fazel, Y.; Henry, L.; Wymer, M. Global epidemiology of nonalcoholic fatty liver disease-Meta-analytic assessment of prevalence, incidence, and outcomes. Hepatology 2016, 64, 73–84. [Google Scholar] [CrossRef] [Green Version]
  39. Le, M.H.; Yeo, Y.H.; Li, X.; Li, J.; Zou, B.; Wu, Y.; Ye, Q.; Huang, D.Q.; Zhao, C.; Zhang, J.; et al. 2019 Global NAFLD Prevalence: A Systematic Review and Meta-analysis. Clin. Gastroenterol. Hepatol. 2022, 20, 2809–2817.e28. [Google Scholar] [CrossRef]
  40. Liu, J.; Mu, C.; Li, K.; Luo, H.; Liu, Y.; Li, Z. Estimating Global Prevalence of Metabolic Dysfunction-Associated Fatty Liver Disease in Overweight or Obese Children and Adolescents: Systematic Review and Meta-Analysis. Int. J. Public Health 2021, 66, 1604371. [Google Scholar] [CrossRef]
  41. Polyzos, S.A.; Kountouras, J.; Mantzoros, C.S. Obesity and nonalcoholic fatty liver disease: From pathophysiology to therapeutics. Metabolism 2019, 92, 82–97. [Google Scholar] [CrossRef]
  42. Lu, F.B.; Hu, E.D.; Xu, L.M.; Chen, L.; Wu, J.L.; Li, H.; Chen, D.Z.; Chen, Y.P. The relationship between obesity and the severity of non-alcoholic fatty liver disease: Systematic review and meta-analysis. Expert Rev. Gastroenterol. Hepatol. 2018, 12, 491–502. [Google Scholar] [CrossRef]
  43. Younossi, Z.M.; Paik, J.M.; Al Shabeeb, R.; Golabi, P.; Younossi, I.; Henry, L. Are there outcome differences between NAFLD and metabolic-associated fatty liver disease? Hepatology 2022, 76, 1423–1437. [Google Scholar] [CrossRef]
  44. Aller, R.; Fernández-Rodríguez, C.; Lo Iacono, O.; Bañares, R.; Abad, J.; Carrión, J.A.; García-Monzón, C.; Caballería, J.; Berenguer, M.; Rodríguez-Perálvarez, M.; et al. Consensus document. Management of non-alcoholic fatty liver disease (NAFLD). Clinical practice guideline. Gastroenterol. Hepatol. 2018, 41, 328–349. [Google Scholar] [CrossRef]
  45. Younossi, Z.M.; Stepanova, M.; Negro, F.; Hallaji, S.; Younossi, Y.; Lam, B.; Srishord, M. Nonalcoholic fatty liver disease in lean individuals in the United States. Medicine 2012, 91, 319–327. [Google Scholar] [CrossRef]
  46. Ye, Q.; Zou, B.; Yeo, Y.H.; Li, J.; Huang, D.Q.; Wu, Y.; Yang, H.; Liu, C.; Kam, L.Y.; Tan, X.X.E.; et al. Global prevalence, incidence, and outcomes of non-obese or lean non-alcoholic fatty liver disease: A systematic review and meta-analysis. Lancet Gastroenterol. Hepatol. 2020, 5, 739–752. [Google Scholar] [CrossRef]
  47. Cheng, Y.M.; Kao, J.H.; Wang, C.C. The metabolic profiles and body composition of lean metabolic associated fatty liver disease. Hepatol. Int. 2021, 15, 405–412. [Google Scholar] [CrossRef]
  48. Leung, J.C.; Loong, T.C.; Wei, J.L.; Wong, G.L.; Chan, A.W.; Choi, P.C.; Shu, S.S.; Chim, A.M.; Chan, H.L.; Wong, V.W. Histological severity and clinical outcomes of nonalcoholic fatty liver disease in nonobese patients. Hepatology 2017, 65, 54–64. [Google Scholar] [CrossRef]
  49. Fracanzani, A.L.; Petta, S.; Lombardi, R.; Pisano, G.; Russello, M.; Consonni, D.; Di Marco, V.; Cammà, C.; Mensi, L.; Dongiovanni, P.; et al. Liver and Cardiovascular Damage in Patients with Lean Nonalcoholic Fatty Liver Disease, and Association with Visceral Obesity. Clin. Gastroenterol. Hepatol. 2017, 15, 1604–1611.e1. [Google Scholar] [CrossRef]
  50. Kim, D.; Kim, W.; Joo, S.K.; Han, J.; Kim, J.H.; Harrison, S.A.; Younossi, Z.M.; Ahmed, A. Association between body size-metabolic phenotype and nonalcoholic steatohepatitis and significant fibrosis. J. Gastroenterol. 2020, 55, 330–341. [Google Scholar] [CrossRef]
  51. Tan, E.X.; Lee, J.W.; Jumat, N.H.; Chan, W.K.; Treeprasertsuk, S.; Goh, G.B.; Fan, J.G.; Song, M.J.; Charatcharoenwitthaya, P.; Duseja, A.; et al. Non-obese non-alcoholic fatty liver disease (NAFLD) in Asia: An international registry study. Metabolism 2022, 126, 154911. [Google Scholar] [CrossRef]
  52. Kim, Y.; Han, E.; Lee, J.S.; Lee, H.W.; Kim, B.K.; Kim, M.K.; Kim, H.S.; Park, J.Y.; Kim, D.Y.; Ahn, S.H.; et al. Cardiovascular Risk Is Elevated in Lean Subjects with Nonalcoholic Fatty Liver Disease. Gut Liver 2022, 16, 290–299. [Google Scholar] [CrossRef]
  53. Ahadi, M.; Molooghi, K.; Masoudifar, N.; Namdar, A.B.; Vossoughinia, H.; Farzanehfar, M. A review of non-alcoholic fatty liver disease in non-obese and lean individuals. J. Gastroenterol. Hepatol. 2021, 36, 1497–1507. [Google Scholar] [CrossRef]
  54. Feldman, A.; Eder, S.K.; Felder, T.K.; Kedenko, L.; Paulweber, B.; Stadlmayr, A.; Huber-Schönauer, U.; Niederseer, D.; Stickel, F.; Auer, S.; et al. Clinical and Metabolic Characterization of Lean Caucasian Subjects with Non-alcoholic Fatty Liver. Am. J. Gastroenterol. 2017, 112, 102–110. [Google Scholar] [CrossRef]
  55. Chen, F.; Esmaili, S.; Rogers, G.B.; Bugianesi, E.; Petta, S.; Marchesini, G.; Bayoumi, A.; Metwally, M.; Azardaryany, M.K.; Coulter, S.; et al. Lean NAFLD: A Distinct Entity Shaped by Differential Metabolic Adaptation. Hepatology 2020, 71, 1213–1227. [Google Scholar] [CrossRef]
  56. Long, M.T.; Noureddin, M.; Lim, J.K. AGA Clinical Practice Update: Diagnosis and Management of Nonalcoholic Fatty Liver Disease in Lean Individuals: Expert Review. Gastroenterology 2022, 163, 764–774.e1. [Google Scholar] [CrossRef]
  57. Byrne, C.D.; Targher, G. Non-alcoholic fatty liver disease-related risk of cardiovascular disease and other cardiac complications. Diabetes Obes. Metab. 2022, 24 (Suppl. S2), 28–43. [Google Scholar] [CrossRef]
  58. Kim, D.; Konyn, P.; Sandhu, K.K.; Dennis, B.B.; Cheung, A.C.; Ahmed, A. Metabolic dysfunction-associated fatty liver disease is associated with increased all-cause mortality in the United States. J. Hepatol. 2021, 75, 1284–1291. [Google Scholar] [CrossRef]
  59. Martínez-Arranz, I.; Bruzzone, C.; Noureddin, M.; Gil-Redondo, R.; Mincholé, I.; Bizkarguenaga, M.; Arretxe, E.; Iruarrizaga-Lejarreta, M.; Fernández-Ramos, D.; Lopitz-Otsoa, F.; et al. Metabolic subtypes of patients with NAFLD exhibit distinctive cardiovascular risk profiles. Hepatology 2022, 76, 1121–1134. [Google Scholar] [CrossRef]
  60. Golabi, P.; Fukui, N.; Paik, J.; Sayiner, M.; Mishra, A.; Younossi, Z.M. Mortality Risk Detected by Atherosclerotic Cardiovascular Disease Score in Patients with Nonalcoholic Fatty Liver Disease. Hepatol. Commun. 2019, 3, 1050–1060. [Google Scholar] [CrossRef] [Green Version]
  61. Paik, J.M.; Deshpande, R.; Golabi, P.; Younossi, I.; Henry, L.; Younossi, Z.M. The impact of modifiable risk factors on the long-term outcomes of non-alcoholic fatty liver disease. Aliment. Pharmacol. Ther. 2020, 51, 291–304. [Google Scholar] [CrossRef]
  62. Yongpisarn, T.; Namasondhi, A.; Iamsumang, W.; Rattanakaemakorn, P.; Suchonwanit, P. Liver fibrosis prevalence and risk factors in patients with psoriasis: A systematic review and meta-analysis. Front. Med. 2022, 9, 1068157. [Google Scholar] [CrossRef]
  63. Rodriguez-Duque, J.C.; Calleja, J.L.; Iruzubieta, P.; Hernández-Conde, M.; Rivas-Rivas, C.; Vera, M.I.; Garcia, M.J.; Pascual, M.; Castro, B.; García-Blanco, A.; et al. Increased risk of MAFLD and Liver Fibrosis in Inflammatory Bowel Disease Independent of Classic Metabolic Risk Factors. Clin. Gastroenterol. Hepatol. 2023, 21, 406–414.e407. [Google Scholar] [CrossRef]
  64. Durán-Vian, C.; Arias-Loste, M.T.; Hernández, J.L.; Fernández, V.; González, M.; Iruzubieta, P.; Rasines, L.; González-Vela, C.; Vaqué, J.P.; Blanco, R.; et al. High prevalence of non-alcoholic fatty liver disease among hidradenitis suppurativa patients independent of classic metabolic risk factors. J. Eur. Acad. Dermatol. Venereol. 2019, 33, 2131–2136. [Google Scholar] [CrossRef]
  65. Romero-Gómez, M.; Aller, R.; Ampuero, J.; Fernández Rodríguez, C.; Augustín, S.; Latorre, R.; Rivera-Esteban, J.; Martínez Urroz, B.; Gutiérrez García, M.L.; López, S.A.; et al. AEEH «Consensus about detection and referral of hidden prevalent liver diseases». Gastroenterol. Hepatol. 2023, 46, 236–247. [Google Scholar] [CrossRef]
  66. Rinella, M.E.; Neuschwander-Tetri, B.A.; Siddiqui, M.S.; Abdelmalek, M.F.; Caldwell, S.; Barb, D.; Kleiner, D.E.; Loomba, R. AASLD Practice Guidance on the clinical assessment and management of nonalcoholic fatty liver disease. Hepatology 2023, 77, 1797–1835. [Google Scholar] [CrossRef]
  67. Schwimmer, J.B.; Celedon, M.A.; Lavine, J.E.; Salem, R.; Campbell, N.; Schork, N.J.; Shiehmorteza, M.; Yokoo, T.; Chavez, A.; Middleton, M.S.; et al. Heritability of nonalcoholic fatty liver disease. Gastroenterology 2009, 136, 1585–1592. [Google Scholar] [CrossRef] [Green Version]
  68. Anstee, Q.M.; Day, C.P. The genetics of NAFLD. Nat. Rev. Gastroenterol. Hepatol. 2013, 10, 645–655. [Google Scholar] [CrossRef]
  69. Eslam, M.; Valenti, L.; Romeo, S. Genetics and epigenetics of NAFLD and NASH: Clinical impact. J. Hepatol. 2018, 68, 268–279. [Google Scholar] [CrossRef]
  70. Romeo, S.; Kozlitina, J.; Xing, C.; Pertsemlidis, A.; Cox, D.; Pennacchio, L.A.; Boerwinkle, E.; Cohen, J.C.; Hobbs, H.H. Genetic variation in PNPLA3 confers susceptibility to nonalcoholic fatty liver disease. Nat. Genet. 2008, 40, 1461–1465. [Google Scholar] [CrossRef] [Green Version]
  71. Sookoian, S.; Pirola, C.J. Genetics in non-alcoholic fatty liver disease: The role of risk alleles through the lens of immune response. Clin. Mol. Hepatol. 2023, 29, S184–S195. [Google Scholar] [CrossRef]
  72. Buch, S.; Stickel, F.; Trépo, E.; Way, M.; Herrmann, A.; Nischalke, H.D.; Brosch, M.; Rosendahl, J.; Berg, T.; Ridinger, M.; et al. 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]
  73. Zhang, H.B.; Su, W.; Xu, H.; Zhang, X.Y.; Guan, Y.F. HSD17B13: A Potential Therapeutic Target for NAFLD. Front. Mol. Biosci. 2021, 8, 824776. [Google Scholar] [CrossRef]
  74. Claussnitzer, M.; Dankel, S.N.; Kim, K.H.; Quon, G.; Meuleman, W.; Haugen, C.; Glunk, V.; Sousa, I.S.; Beaudry, J.L.; Puviindran, V.; et al. FTO Obesity Variant Circuitry and Adipocyte Browning in Humans. N. Engl. J. Med. 2015, 373, 895–907. [Google Scholar] [CrossRef] [Green Version]
  75. Reue, K.; Zhang, P. The lipin protein family: Dual roles in lipid biosynthesis and gene expression. FEBS Lett. 2008, 582, 90–96. [Google Scholar] [CrossRef] [Green Version]
  76. Mahdessian, H.; Taxiarchis, A.; Popov, S.; Silveira, A.; Franco-Cereceda, A.; Hamsten, A.; Eriksson, P.; van’t Hooft, F. TM6SF2 is a regulator of liver fat metabolism influencing triglyceride secretion and hepatic lipid droplet content. Proc. Natl. Acad. Sci. USA 2014, 111, 8913–8918. [Google Scholar] [CrossRef] [Green Version]
  77. Sliz, E.; Sebert, S.; Würtz, P.; Kangas, A.J.; Soininen, P.; Lehtimäki, T.; Kähönen, M.; Viikari, J.; Männikkö, M.; Ala-Korpela, M.; et al. NAFLD risk alleles in PNPLA3, TM6SF2, GCKR and LYPLAL1 show divergent metabolic effects. Hum. Mol. Genet. 2018, 27, 2214–2223. [Google Scholar] [CrossRef] [Green Version]
  78. Speliotes, E.K.; Yerges-Armstrong, L.M.; Wu, J.; Hernaez, R.; Kim, L.J.; Palmer, C.D.; Gudnason, V.; Eiriksdottir, G.; Garcia, M.E.; Launer, L.J.; et al. Genome-wide association analysis identifies variants associated with nonalcoholic fatty liver disease that have distinct effects on metabolic traits. PLoS Genet. 2011, 7, e1001324. [Google Scholar] [CrossRef] [Green Version]
  79. Onuma, H.; Tabara, Y.; Kawamoto, R.; Shimizu, I.; Kawamura, R.; Takata, Y.; Nishida, W.; Ohashi, J.; Miki, T.; Kohara, K.; et al. The GCKR rs780094 polymorphism is associated with susceptibility of type 2 diabetes, reduced fasting plasma glucose levels, increased triglycerides levels and lower HOMA-IR in Japanese population. J. Hum. Genet. 2010, 55, 600–604. [Google Scholar] [CrossRef] [Green Version]
  80. Seko, Y.; Yamaguchi, K.; Itoh, Y. The genetic backgrounds in nonalcoholic fatty liver disease. Clin. J. Gastroenterol. 2018, 11, 97–102. [Google Scholar] [CrossRef] [Green Version]
  81. Männistö, V.; Kaminska, D.; Käkelä, P.; Neuvonen, M.; Niemi, M.; Alvarez, M.; Pajukanta, P.; Romeo, S.; Nieuwdorp, M.; Groen, A.K.; et al. Protein Phosphatase 1 Regulatory Subunit 3B Genotype at rs4240624 Has a Major Effect on Gallbladder Bile Composition. Hepatol. Commun. 2021, 5, 244–257. [Google Scholar] [CrossRef]
  82. Al-Serri, A.; Anstee, Q.M.; Valenti, L.; Nobili, V.; Leathart, J.B.; Dongiovanni, P.; Patch, J.; Fracanzani, A.; Fargion, S.; Day, C.P.; et al. The SOD2 C47T polymorphism influences NAFLD fibrosis severity: Evidence from case-control and intra-familial allele association studies. J. Hepatol. 2012, 56, 448–454. [Google Scholar] [CrossRef] [Green Version]
  83. Petta, S.; Valenti, L.; Marra, F.; Grimaudo, S.; Tripodo, C.; Bugianesi, E.; Cammà, C.; Cappon, A.; Di Marco, V.; Di Maira, G.; et al. MERTK rs4374383 polymorphism affects the severity of fibrosis in non-alcoholic fatty liver disease. J. Hepatol. 2016, 64, 682–690. [Google Scholar] [CrossRef]
  84. Petta, S.; Valenti, L.; Svegliati-Baroni, G.; Ruscica, M.; Pipitone, R.M.; Dongiovanni, P.; Rychlicki, C.; Ferri, N.; Cammà, C.; Fracanzani, A.L.; et al. Fibronectin Type III Domain-Containing Protein 5 rs3480 A>G Polymorphism, Irisin, and Liver Fibrosis in Patients with Nonalcoholic Fatty Liver Disease. J. Clin. Endocrinol. Metab. 2017, 102, 2660–2669. [Google Scholar] [CrossRef]
  85. Miele, L.; Beale, G.; Patman, G.; Nobili, V.; Leathart, J.; Grieco, A.; Abate, M.; Friedman, S.L.; Narla, G.; Bugianesi, E.; et al. The Kruppel-like factor 6 genotype is associated with fibrosis in nonalcoholic fatty liver disease. Gastroenterology 2008, 135, 282–291.e1. [Google Scholar] [CrossRef] [Green Version]
  86. Aravinthan, A.; Mells, G.; Allison, M.; Leathart, J.; Kotronen, A.; Yki-Jarvinen, H.; Daly, A.K.; Day, C.P.; Anstee, Q.M.; Alexander, G. Gene polymorphisms of cellular senescence marker p21 and disease progression in non-alcohol-related fatty liver disease. Cell Cycle 2014, 13, 1489–1494. [Google Scholar] [CrossRef] [Green Version]
  87. Petta, S.; Grimaudo, S.; Cammà, C.; Cabibi, D.; Di Marco, V.; Licata, G.; Pipitone, R.M.; Craxì, A. IL28B and PNPLA3 polymorphisms affect histological liver damage in patients with non-alcoholic fatty liver disease. J. Hepatol. 2012, 56, 1356–1362. [Google Scholar] [CrossRef]
  88. Luukkonen, P.K.; Nick, A.; Hölttä-Vuori, M.; Thiele, C.; Isokuortti, E.; Lallukka-Brück, S.; Zhou, Y.; Hakkarainen, A.; Lundbom, N.; Peltonen, M.; et al. Human PNPLA3-I148M variant increases hepatic retention of polyunsaturated fatty acids. JCI Insight 2019, 4, e127902. [Google Scholar] [CrossRef] [Green Version]
  89. Yuan, L.; Terrrault, N.A. PNPLA3 and nonalcoholic fatty liver disease: Towards personalized medicine for fatty liver. Hepatobiliary Surg. Nutr. 2020, 9, 353–356. [Google Scholar] [CrossRef]
  90. Eslam, M.; George, J. Genetic contributions to NAFLD: Leveraging shared genetics to uncover systems biology. Nat. Rev. Gastroenterol. Hepatol. 2020, 17, 40–52. [Google Scholar] [CrossRef]
  91. Dongiovanni, P.; Petta, S.; Maglio, C.; Fracanzani, A.L.; Pipitone, R.; Mozzi, E.; Motta, B.M.; Kaminska, D.; Rametta, R.; Grimaudo, S.; et al. Transmembrane 6 superfamily member 2 gene variant disentangles nonalcoholic steatohepatitis from cardiovascular disease. Hepatology 2015, 61, 506–514. [Google Scholar] [CrossRef] [Green Version]
  92. Valenti, L.; Alisi, A.; Nobili, V. Unraveling the genetics of fatty liver in obese children: Additive effect of P446L GCKR and I148M PNPLA3 polymorphisms. Hepatology 2012, 55, 661–663. [Google Scholar] [CrossRef]
  93. 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.e1216. [Google Scholar] [CrossRef] [Green Version]
  94. Wang, P.; Wu, C.X.; Li, Y.; Shen, N. HSD17B13 rs72613567 protects against liver diseases and histological progression of nonalcoholic fatty liver disease: A systematic review and meta-analysis. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 8997–9007. [Google Scholar]
  95. 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]
  96. Gilbert, J.A.; Blaser, M.J.; Caporaso, J.G.; Jansson, J.K.; Lynch, S.V.; Knight, R. Current understanding of the human microbiome. Nat. Med. 2018, 24, 392–400. [Google Scholar] [CrossRef]
  97. Clemente, J.C.; Ursell, L.K.; Parfrey, L.W.; Knight, R. The impact of the gut microbiota on human health: An integrative view. Cell 2012, 148, 1258–1270. [Google Scholar] [CrossRef] [Green Version]
  98. Bombin, A.; Yan, S.; Bombin, S.; Mosley, J.D.; Ferguson, J.F. Obesity influences composition of salivary and fecal microbiota and impacts the interactions between bacterial taxa. Physiol. Rep. 2022, 10, e15254. [Google Scholar] [CrossRef]
  99. Turnbaugh, P.J.; Hamady, M.; Yatsunenko, T.; Cantarel, B.L.; Duncan, A.; Ley, R.E.; Sogin, M.L.; Jones, W.J.; Roe, B.A.; Affourtit, J.P.; et al. A core gut microbiome in obese and lean twins. Nature 2009, 457, 480–484. [Google Scholar] [CrossRef] [Green Version]
  100. Loomba, R.; Seguritan, V.; Li, W.; Long, T.; Klitgord, N.; Bhatt, A.; Dulai, P.S.; Caussy, C.; Bettencourt, R.; Highlander, S.K.; et al. Gut Microbiome-Based Metagenomic Signature for Non-invasive Detection of Advanced Fibrosis in Human Nonalcoholic Fatty Liver Disease. Cell Metab. 2017, 25, 1054–1062.e5. [Google Scholar] [CrossRef]
  101. Khan, A.; Ding, Z.; Ishaq, M.; Bacha, A.S.; Khan, I.; Hanif, A.; Li, W.; Guo, X. Understanding the Effects of Gut Microbiota Dysbiosis on Nonalcoholic Fatty Liver Disease and the Possible Probiotics Role: Recent Updates. Int. J. Biol. Sci. 2021, 17, 818–833. [Google Scholar] [CrossRef]
  102. Kolodziejczyk, A.A.; Zheng, D.; Shibolet, O.; Elinav, E. The role of the microbiome in NAFLD and NASH. EMBO Mol. Med. 2019, 11, e9302. [Google Scholar] [CrossRef]
  103. Albillos, A.; de Gottardi, A.; Rescigno, M. The gut-liver axis in liver disease: Pathophysiological basis for therapy. J. Hepatol. 2020, 72, 558–577. [Google Scholar] [CrossRef] [Green Version]
  104. Bauer, K.C.; Littlejohn, P.T.; Ayala, V.; Creus-Cuadros, A.; Finlay, B.B. Nonalcoholic Fatty Liver Disease and the Gut-Liver Axis: Exploring an Undernutrition Perspective. Gastroenterology 2022, 162, 1858–1875.e2. [Google Scholar] [CrossRef]
  105. Huang, C.; Zhou, Y.; Cheng, J.; Guo, X.; Shou, D.; Quan, Y.; Chen, H. Pattern recognition receptors in the development of nonalcoholic fatty liver disease and progression to hepatocellular carcinoma: An emerging therapeutic strategy. Front. Endocrinol. 2023, 14, 1145392. [Google Scholar] [CrossRef]
  106. Iruzubieta, P.; Medina, J.M.; Fernández-López, R.; Crespo, J.; de la Cruz, F. A Role for Gut Microbiome Fermentative Pathways in Fatty Liver Disease Progression. J. Clin. Med. 2020, 9, 1369. [Google Scholar] [CrossRef]
  107. De Minicis, S.; Rychlicki, C.; Agostinelli, L.; Saccomanno, S.; Candelaresi, C.; Trozzi, L.; Mingarelli, E.; Facinelli, B.; Magi, G.; Palmieri, C.; et al. Dysbiosis contributes to fibrogenesis in the course of chronic liver injury in mice. Hepatology 2014, 59, 1738–1749. [Google Scholar] [CrossRef]
  108. Yuan, J.; Chen, C.; Cui, J.; Lu, J.; Yan, C.; Wei, X.; Zhao, X.; Li, N.; Li, S.; Xue, G.; et al. Fatty Liver Disease Caused by High-Alcohol-Producing Klebsiella pneumoniae. Cell Metab. 2019, 30, 675–688.e7. [Google Scholar] [CrossRef]
  109. Mir, H.; Meena, A.S.; Chaudhry, K.K.; Shukla, P.K.; Gangwar, R.; Manda, B.; Padala, M.K.; Shen, L.; Turner, J.R.; Dietrich, P.; et al. Occludin deficiency promotes ethanol-induced disruption of colonic epithelial junctions, gut barrier dysfunction and liver damage in mice. Biochim. Biophys. Acta 2016, 1860, 765–774. [Google Scholar] [CrossRef] [Green Version]
  110. Zhu, R.; Baker, S.S.; Moylan, C.A.; Abdelmalek, M.F.; Guy, C.D.; Zamboni, F.; Wu, D.; Lin, W.; Liu, W.; Baker, R.D.; et al. Systematic transcriptome analysis reveals elevated expression of alcohol-metabolizing genes in NAFLD livers. J. Pathol. 2016, 238, 531–542. [Google Scholar] [CrossRef]
  111. Koh, A.; De Vadder, F.; Kovatcheva-Datchary, P.; Bäckhed, F. From Dietary Fiber to Host Physiology: Short-Chain Fatty Acids as Key Bacterial Metabolites. Cell 2016, 165, 1332–1345. [Google Scholar] [CrossRef] [Green Version]
  112. Chu, H.; Duan, Y.; Yang, L.; Schnabl, B. Small metabolites, possible big changes: A microbiota-centered view of non-alcoholic fatty liver disease. Gut 2019, 68, 359–370. [Google Scholar] [CrossRef]
  113. Cresci, G.A.; Glueck, B.; McMullen, M.R.; Xin, W.; Allende, D.; Nagy, L.E. Prophylactic tributyrin treatment mitigates chronic-binge ethanol-induced intestinal barrier and liver injury. J. Gastroenterol. Hepatol. 2017, 32, 1587–1597. [Google Scholar] [CrossRef]
  114. Ji, Y.; Yin, Y.; Li, Z.; Zhang, W. Gut Microbiota-Derived Components and Metabolites in the Progression of Non-Alcoholic Fatty Liver Disease (NAFLD). Nutrients 2019, 11, 1712. [Google Scholar] [CrossRef] [Green Version]
  115. Zhang, L.S.; Davies, S.S. Microbial metabolism of dietary components to bioactive metabolites: Opportunities for new therapeutic interventions. Genome Med. 2016, 8, 46. [Google Scholar] [CrossRef] [Green Version]
  116. Del Barrio, M.; Lavín, L.; Santos-Laso, Á.; Arias-Loste, M.T.; Odriozola, A.; Rodriguez-Duque, J.C.; Rivas, C.; Iruzubieta, P.; Crespo, J. Faecal Microbiota Transplantation, Paving the Way to Treat Non-Alcoholic Fatty Liver Disease. Int. J. Mol. Sci. 2023, 24, 6123. [Google Scholar] [CrossRef]
  117. European Association for the Study of the Liver (EASL); European Association for the Study of Diabetes (EASD); European Association for the Study of Obesity (EASO). EASL-EASD-EASO Clinical Practice Guidelines for the management of non-alcoholic fatty liver disease. J. Hepatol. 2016, 64, 1388–1402. [Google Scholar] [CrossRef]
  118. Chalasani, N.; Younossi, Z.; Lavine, J.E.; Charlton, M.; Cusi, K.; Rinella, M.; Harrison, S.A.; Brunt, E.M.; Sanyal, A.J. The diagnosis and management of nonalcoholic fatty liver disease: Practice guidance from the American Association for the Study of Liver Diseases. Hepatology 2018, 67, 328–357. [Google Scholar] [CrossRef] [Green Version]
  119. WHO. Global Status Report on Alcohol and Health 2018; WHO: Geneva, Switzerland, 2018. [Google Scholar]
  120. Staufer, K.; Huber-Schönauer, U.; Strebinger, G.; Pimingstorfer, P.; Suesse, S.; Scherzer, T.M.; Paulweber, B.; Ferenci, P.; Stimpfl, T.; Yegles, M.; et al. Ethyl glucuronide in hair detects a high rate of harmful alcohol consumption in presumed non-alcoholic fatty liver disease. J. Hepatol. 2022, 77, 918–930. [Google Scholar] [CrossRef]
  121. Long, M.T.; Massaro, J.M.; Hoffmann, U.; Benjamin, E.J.; Naimi, T.S. Alcohol Use Is Associated with Hepatic Steatosis Among Persons with Presumed Nonalcoholic Fatty Liver Disease. Clin. Gastroenterol. Hepatol. 2020, 18, 1831–1841.e5. [Google Scholar] [CrossRef]
  122. Petroni, M.L.; Brodosi, L.; Marchignoli, F.; Musio, A.; Marchesini, G. Moderate Alcohol Intake in Non-Alcoholic Fatty Liver Disease: To Drink or Not to Drink? Nutrients 2019, 11, 3048. [Google Scholar] [CrossRef] [Green Version]
  123. Cabezas, J.; Lucey, M.R.; Bataller, R. Biomarkers for monitoring alcohol use. Clin. Liver Dis. 2016, 8, 59–63. [Google Scholar] [CrossRef]
  124. Farrell, G.C. Drugs and Steatohepatitis. Semin. Liver Dis. 2002, 22, 185–194. [Google Scholar] [CrossRef]
  125. Dash, A.; Figler, R.A.; Sanyal, A.J.; Wamhoff, B.R. Drug-induced steatohepatitis. Expert Opin. Drug Metab. Toxicol. 2017, 13, 193–204. [Google Scholar] [CrossRef]
  126. Guigui, B.; Perrot, S.; Berry, J.P.; Fleury-Feith, J.; Martin, N.; Métreau, J.M.; Dhumeaux, D.; Zafrani, E.S. Amiodarone-induced hepatic phospholipidosis: A morphological alteration independent of pseudoalcoholic liver disease. Hepatology 1988, 8, 1063–1068. [Google Scholar] [CrossRef]
  127. Schultz, J.C.; Adamson, J.S.; Workman, W.W.; Norman, T.D. Fatal liver disease after intravenous administration of tetracycline in high dosage. N. Engl. J. Med. 1963, 269, 999–1004. [Google Scholar] [CrossRef]
  128. Fréneaux, E.; Labbe, G.; Letteron, P.; Dinh, T.L.; Degott, C.; Genève, J.; Larrey, D.; Pessayre, D. Inhibition of the mitochondrial oxidation of fatty acids by tetracycline in mice and in man: Possible role in microvesicular steatosis induced by this antibiotic. Hepatology 1988, 8, 1056–1062. [Google Scholar] [CrossRef]
  129. Caldwell, S.H.; Hespenheide, E.E.; von Borstel, R.W. Myositis, microvesicular hepatitis, and progression to cirrhosis from troglitazone added to simvastatin. Dig. Dis. Sci. 2001, 46, 376–378. [Google Scholar] [CrossRef]
  130. Fukano, M.; Amano, S.; Sato, J.; Yamamoto, K.; Adachi, H.; Okabe, H.; Fujiyama, Y.; Bamba, T. Subacute hepatic failure associated with a new antidiabetic agent, troglitazone: A case report with autopsy examination. Hum. Pathol. 2000, 31, 250–253. [Google Scholar] [CrossRef]
  131. Kohlroser, J.; Mathai, J.; Reichheld, J.; Banner, B.F.; Bonkovsky, H.L. Hepatotoxicity due to troglitazone: Report of two cases and review of adverse events reported to the United States Food and Drug Administration. Am. J. Gastroenterol. 2000, 95, 272–276. [Google Scholar] [CrossRef]
  132. Luef, G.; Rauchenzauner, M.; Waldmann, M.; Sturm, W.; Sandhofer, A.; Seppi, K.; Trinka, E.; Unterberger, I.; Ebenbichler, C.F.; Joannidis, M.; et al. Non-alcoholic fatty liver disease (NAFLD), insulin resistance and lipid profile in antiepileptic drug treatment. Epilepsy Res. 2009, 86, 42–47. [Google Scholar] [CrossRef]
  133. Luef, G.J.; Waldmann, M.; Sturm, W.; Naser, A.; Trinka, E.; Unterberger, I.; Bauer, G.; Lechleitner, M. Valproate therapy and nonalcoholic fatty liver disease. Ann. Neurol. 2004, 55, 729–732. [Google Scholar] [CrossRef]
  134. Hamed, S.A.; Fathy, R.A.; Radwan, M.E.; Abdellah, M.M. Fatty liver in adults receiving antiepileptic medications: Relationship to the metabolic risks. Expert Rev. Clin. Pharmacol. 2016, 9, 617–624. [Google Scholar] [CrossRef]
  135. Lettéron, P.; Brahimi-Bourouina, N.; Robin, M.A.; Moreau, A.; Feldmann, G.; Pessayre, D. Glucocorticoids inhibit mitochondrial matrix acyl-CoA dehydrogenases and fatty acid beta-oxidation. Am. J. Physiol. 1997, 272, G1141–G1150. [Google Scholar] [CrossRef]
  136. Miyake, K.; Hayakawa, K.; Nishino, M.; Morimoto, T.; Mukaihara, S. Effects of oral 5-fluorouracil drugs on hepatic fat content in patients with colon cancer. Acad. Radiol. 2005, 12, 722–727. [Google Scholar] [CrossRef]
  137. Pilgrim, C.H.; Satgunaseelan, L.; Pham, A.; Murray, W.; Link, E.; Smith, M.; Usatoff, V.; Evans, P.M.; Banting, S.; Thomson, B.N.; et al. Correlations between histopathological diagnosis of chemotherapy-induced hepatic injury, clinical features, and perioperative morbidity. HPB 2012, 14, 333–340. [Google Scholar] [CrossRef] [Green Version]
  138. Mani, T.R.; Reuben, R.; Akiyama, J. Field efficacy of “Mosbar” mosquito repellent soap against vectors of bancroftian filariasis and Japanese encephalitis in southern India. J. Am. Mosq. Control Assoc. 1991, 7, 565–568. [Google Scholar]
  139. Pawlik, T.M.; Olino, K.; Gleisner, A.L.; Torbenson, M.; Schulick, R.; Choti, M.A. Preoperative chemotherapy for colorectal liver metastases: Impact on hepatic histology and postoperative outcome. J. Gastrointest. Surg. 2007, 11, 860–868. [Google Scholar] [CrossRef]
  140. Wolf, P.S.; Park, J.O.; Bao, F.; Allen, P.J.; DeMatteo, R.P.; Fong, Y.; Jarnagin, W.R.; Kingham, T.P.; Gönen, M.; Kemeny, N.; et al. Preoperative chemotherapy and the risk of hepatotoxicity and morbidity after liver resection for metastatic colorectal cancer: A single institution experience. J. Am. Coll. Surg. 2013, 216, 41–49. [Google Scholar] [CrossRef]
  141. Gabbi, C.; Carubbi, F.; Losi, L.; Loria, P.; Costantini, M.; Bertolotti, M.; Carulli, N. Nonalcoholic fatty liver disease induced by leuprorelin acetate. J. Clin. Gastroenterol. 2008, 42, 107–110. [Google Scholar] [CrossRef]
  142. Langman, G.; Hall, P.M.; Todd, G. Role of non-alcoholic steatohepatitis in methotrexate-induced liver injury. J. Gastroenterol. Hepatol. 2001, 16, 1395–1401. [Google Scholar] [CrossRef]
  143. Saphner, T.; Triest-Robertson, S.; Li, H.; Holzman, P. The association of nonalcoholic steatohepatitis and tamoxifen in patients with breast cancer. Cancer 2009, 115, 3189–3195. [Google Scholar] [CrossRef]
  144. Nguyen, M.C.; Stewart, R.B.; Banerji, M.A.; Gordon, D.H.; Kral, J.G. Relationships between tamoxifen use, liver fat and body fat distribution in women with breast cancer. Int. J. Obes. Relat. Metab. Disord. 2001, 25, 296–298. [Google Scholar] [CrossRef] [Green Version]
  145. Sulkowski, M.S.; Mehta, S.H.; Torbenson, M.; Afdhal, N.H.; Mirel, L.; Moore, R.D.; Thomas, D.L. Hepatic steatosis and antiretroviral drug use among adults coinfected with HIV and hepatitis C virus. AIDS 2005, 19, 585–592. [Google Scholar] [CrossRef]
  146. Chariot, P.; Drogou, I.; de Lacroix-Szmania, I.; Eliezer-Vanerot, M.C.; Chazaud, B.; Lombès, A.; Schaeffer, A.; Zafrani, E.S. Zidovudine-induced mitochondrial disorder with massive liver steatosis, myopathy, lactic acidosis, and mitochondrial DNA depletion. J. Hepatol. 1999, 30, 156–160. [Google Scholar] [CrossRef]
  147. Riddle, T.M.; Kuhel, D.G.; Woollett, L.A.; Fichtenbaum, C.J.; Hui, D.Y. HIV protease inhibitor induces fatty acid and sterol biosynthesis in liver and adipose tissues due to the accumulation of activated sterol regulatory element-binding proteins in the nucleus. J. Biol. Chem. 2001, 276, 37514–37519. [Google Scholar] [CrossRef] [Green Version]
  148. Yamamoto, A.; Adachi, S.; Ishikawa, K.; Yokomura, T.; Kitani, T. Studies on drug-induced lipidosis. 3. Lipid composition of the liver and some other tissues in clinical cases of “Niemann-Pick-like syndrome” induced by 4,4′-diethylaminoethoxyhexestrol. J. Biochem. 1971, 70, 775–784. [Google Scholar] [CrossRef]
  149. Deschamps, D.; DeBeco, V.; Fisch, C.; Fromenty, B.; Guillouzo, A.; Pessayre, D. Inhibition by perhexiline of oxidative phosphorylation and the beta-oxidation of fatty acids: Possible role in pseudoalcoholic liver lesions. Hepatology 1994, 19, 948–961. [Google Scholar]
  150. Pessayre, D.; Bichara, M.; Degott, C.; Potet, F.; Benhamou, J.P.; Feldmann, G. Perhexiline maleate-induced cirrhosis. Gastroenterology 1979, 76, 170–177. [Google Scholar]
  151. Dulai, P.S.; Sirlin, C.B.; Loomba, R. MRI and MRE for non-invasive quantitative assessment of hepatic steatosis and fibrosis in NAFLD and NASH: Clinical trials to clinical practice. J. Hepatol. 2016, 65, 1006–1016. [Google Scholar] [CrossRef] [Green Version]
  152. Park, C.C.; Nguyen, P.; Hernandez, C.; Bettencourt, R.; Ramirez, K.; Fortney, L.; Hooker, J.; Sy, E.; Savides, M.T.; Alquiraish, M.H.; et al. Magnetic Resonance Elastography vs Transient Elastography in Detection of Fibrosis and Noninvasive Measurement of Steatosis in Patients with Biopsy-Proven Nonalcoholic Fatty Liver Disease. Gastroenterology 2017, 152, 598–607.e2. [Google Scholar] [CrossRef] [Green Version]
  153. Browning, J.D.; Szczepaniak, L.S.; Dobbins, R.; Nuremberg, P.; Horton, J.D.; Cohen, J.C.; Grundy, S.M.; Hobbs, H.H. Prevalence of hepatic steatosis in an urban population in the United States: Impact of ethnicity. Hepatology 2004, 40, 1387–1395. [Google Scholar] [CrossRef]
  154. Mofrad, P.; Contos, M.J.; Haque, M.; Sargeant, C.; Fisher, R.A.; Luketic, V.A.; Sterling, R.K.; Shiffman, M.L.; Stravitz, R.T.; Sanyal, A.J. Clinical and histologic spectrum of nonalcoholic fatty liver disease associated with normal ALT values. Hepatology 2003, 37, 1286–1292. [Google Scholar] [CrossRef]
  155. Fracanzani, A.L.; Valenti, L.; Bugianesi, E.; Andreoletti, M.; Colli, A.; Vanni, E.; Bertelli, C.; Fatta, E.; Bignamini, D.; Marchesini, G.; et al. Risk of severe liver disease in nonalcoholic fatty liver disease with normal aminotransferase levels: A role for insulin resistance and diabetes. Hepatology 2008, 48, 792–798. [Google Scholar] [CrossRef]
  156. Pavlik, L.; Regev, A.; Ardayfio, P.A.; Chalasani, N.P. Drug-Induced Steatosis and Steatohepatitis: The Search for Novel Serum Biomarkers Among Potential Biomarkers for Non-Alcoholic Fatty Liver Disease and Non-Alcoholic Steatohepatitis. Drug Saf. 2019, 42, 701–711. [Google Scholar] [CrossRef]
  157. Lemoine, M.; Barbu, V.; Girard, P.M.; Kim, M.; Bastard, J.P.; Wendum, D.; Paye, F.; Housset, C.; Capeau, J.; Serfaty, L. Altered hepatic expression of SREBP-1 and PPARgamma is associated with liver injury in insulin-resistant lipodystrophic HIV-infected patients. AIDS 2006, 20, 387–395. [Google Scholar] [CrossRef]
  158. Agarwal, N.; Iyer, D.; Gabbi, C.; Saha, P.; Patel, S.G.; Mo, Q.; Chang, B.; Goswami, B.; Schubert, U.; Kopp, J.B.; et al. HIV-1 viral protein R (Vpr) induces fatty liver in mice via LXRα and PPARα dysregulation: Implications for HIV-specific pathogenesis of NAFLD. Sci. Rep. 2017, 7, 13362. [Google Scholar] [CrossRef] [Green Version]
  159. Tuyama, A.C.; Hong, F.; Saiman, Y.; Wang, C.; Ozkok, D.; Mosoian, A.; Chen, P.; Chen, B.K.; Klotman, M.E.; Bansal, M.B. Human immunodeficiency virus (HIV)-1 infects human hepatic stellate cells and promotes collagen I and monocyte chemoattractant protein-1 expression: Implications for the pathogenesis of HIV/hepatitis C virus-induced liver fibrosis. Hepatology 2010, 52, 612–622. [Google Scholar] [CrossRef] [Green Version]
  160. Bruno, R.; Galastri, S.; Sacchi, P.; Cima, S.; Caligiuri, A.; DeFranco, R.; Milani, S.; Gessani, S.; Fantuzzi, L.; Liotta, F.; et al. gp120 modulates the biology of human hepatic stellate cells: A link between HIV infection and liver fibrogenesis. Gut 2010, 59, 513–520. [Google Scholar] [CrossRef]
  161. Verna, E.C. Non-alcoholic fatty liver disease and non-alcoholic steatohepatitis in patients with HIV. Lancet Gastroenterol. Hepatol. 2017, 2, 211–223. [Google Scholar] [CrossRef]
  162. Squillace, N.; Soria, A.; Bozzi, G.; Gori, A.; Bandera, A. Nonalcoholic fatty liver disease and steatohepatitis in people living with HIV. Expert Rev. Gastroenterol. Hepatol. 2019, 13, 643–650. [Google Scholar] [CrossRef]
  163. Morrison, M.; Hughes, H.Y.; Naggie, S.; Syn, W.K. Nonalcoholic Fatty Liver Disease Among Individuals with HIV Mono-infection: A Growing Concern? Dig. Dis. Sci. 2019, 64, 3394–3401. [Google Scholar] [CrossRef]
  164. Guaraldi, G.; Squillace, N.; Stentarelli, C.; Orlando, G.; D’Amico, R.; Ligabue, G.; Fiocchi, F.; Zona, S.; Loria, P.; Esposito, R.; et al. Nonalcoholic fatty liver disease in HIV-infected patients referred to a metabolic clinic: Prevalence, characteristics, and predictors. Clin. Infect. Dis. 2008, 47, 250–257. [Google Scholar] [CrossRef] [Green Version]
  165. Crum-Cianflone, N.; Dilay, A.; Collins, G.; Asher, D.; Campin, R.; Medina, S.; Goodman, Z.; Parker, R.; Lifson, A.; Capozza, T.; et al. Nonalcoholic fatty liver disease among HIV-infected persons. J. Acquir. Immune Defic. Syndr. 2009, 50, 464–473. [Google Scholar] [CrossRef] [Green Version]
  166. Brown, T.T.; Xu, X.; John, M.; Singh, J.; Kingsley, L.A.; Palella, F.J.; Witt, M.D.; Margolick, J.B.; Dobs, A.S. Fat distribution and longitudinal anthropometric changes in HIV-infected men with and without clinical evidence of lipodystrophy and HIV-uninfected controls: A substudy of the Multicenter AIDS Cohort Study. AIDS Res. Ther. 2009, 6, 8. [Google Scholar] [CrossRef] [Green Version]
  167. Joy, T.; Keogh, H.M.; Hadigan, C.; Dolan, S.E.; Fitch, K.; Liebau, J.; Johnsen, S.; Lo, J.; Grinspoon, S.K. Relation of body composition to body mass index in HIV-infected patients with metabolic abnormalities. J. Acquir. Immune Defic. Syndr. 2008, 47, 174–184. [Google Scholar] [CrossRef] [Green Version]
  168. Lemoine, M.; Assoumou, L.; Girard, P.M.; Valantin, M.A.; Katlama, C.; De Wit, S.; Campa, P.; Rougier, H.; Meynard, J.L.; Necsoi, C.; et al. Screening HIV Patients at Risk for NAFLD Using MRI-PDFF and Transient Elastography: A European Multicenter Prospective Study. Clin. Gastroenterol. Hepatol. 2023, 21, 713–722.e3. [Google Scholar] [CrossRef]
  169. Lemoine, M.; Assoumou, L.; De Wit, S.; Girard, P.M.; Valantin, M.A.; Katlama, C.; Necsoi, C.; Campa, P.; Huefner, A.D.; Schulze Zur Wiesch, J.; et al. Diagnostic Accuracy of Noninvasive Markers of Steatosis, NASH, and Liver Fibrosis in HIV-Monoinfected Individuals at Risk of Nonalcoholic Fatty Liver Disease (NAFLD): Results from the ECHAM Study. J. Acquir. Immune Defic. Syndr. 2019, 80, e86–e94. [Google Scholar] [CrossRef]
  170. Lombardi, R.; Sambatakou, H.; Mariolis, I.; Cokkinos, D.; Papatheodoridis, G.V.; Tsochatzis, E.A. Prevalence and predictors of liver steatosis and fibrosis in unselected patients with HIV mono-infection. Dig. Liver Dis. 2016, 48, 1471–1477. [Google Scholar] [CrossRef]
  171. Price, J.C.; Ma, Y.; Kuniholm, M.H.; Adimora, A.A.; Fischl, M.; French, A.L.; Golub, E.T.; Konkle-Parker, D.; Minkoff, H.; Ofotokun, I.; et al. Human Immunodeficiency Virus Is Associated with Elevated FibroScan-Aspartate Aminotransferase (FAST) Score. Clin. Infect. Dis. 2022, 75, 2119–2127. [Google Scholar] [CrossRef]
  172. Busca, C.; Sánchez-Conde, M.; Rico, M.; Rosas, M.; Valencia, E.; Moreno, A.; Moreno, V.; Martín-Carbonero, L.; Moreno, S.; Pérez-Valero, I.; et al. Assessment of Noninvasive Markers of Steatosis and Liver Fibrosis in Human Immunodeficiency Virus-Monoinfected Patients on Stable Antiretroviral Regimens. Open. Forum Infect. Dis. 2022, 9, ofac279. [Google Scholar] [CrossRef]
  173. Vodkin, I.; Valasek, M.A.; Bettencourt, R.; Cachay, E.; Loomba, R. Clinical, biochemical and histological differences between HIV-associated NAFLD and primary NAFLD: A case-control study. Aliment. Pharmacol. Ther. 2015, 41, 368–378. [Google Scholar] [CrossRef]
  174. Chew, K.W.; Wu, K.; Tassiopoulos, K.; Palella, F.J.; Naggie, S.; Utay, N.S.; Overton, E.T.; Sulkowski, M. Liver Inflammation Is Common and Linked to Metabolic Derangements in Persons with Treated Human Immunodeficiency Virus (HIV). Clin. Infect. Dis. 2023, 76, e571–e579. [Google Scholar] [CrossRef]
  175. Sebastiani, G.; Cocciolillo, S.; Mazzola, G.; Malagoli, A.; Falutz, J.; Cervo, A.; Petta, S.; Pembroke, T.; Ghali, P.; Besutti, G.; et al. Application of guidelines for the management of nonalcoholic fatty liver disease in three prospective cohorts of HIV-monoinfected patients. HIV Med. 2020, 21, 96–108. [Google Scholar] [CrossRef]
  176. Iogna Prat, L.; Roccarina, D.; Lever, R.; Lombardi, R.; Rodger, A.; Hall, A.; Luong, T.V.; Bhagani, S.; Tsochatzis, E.A. Etiology and Severity of Liver Disease in HIV-Positive Patients with Suspected NAFLD: Lessons from a Cohort with Available Liver Biopsies. J. Acquir. Immune Defic. Syndr. 2019, 80, 474–480. [Google Scholar] [CrossRef]
  177. Sebastiani, G.; Milic, J.; Gioe, C.; Al Hinai, A.S.; Cervo, A.; Lebouche, B.; Deschenes, M.; Cascio, A.; Mazzola, G.; Guaraldi, G. Diagnosis of liver fibrosis in ageing patients with HIV at risk for non-alcoholic fatty liver disease in Italy and Canada: Assessment of a two-tier pathway. Lancet HIV 2022, 9 (Suppl. S1), S4. [Google Scholar] [CrossRef]
  178. Lombardi, R.; Lever, R.; Smith, C.; Marshall, N.; Rodger, A.; Bhagani, S.; Tsochatzis, E. Liver test abnormalities in patients with HIV mono-infection: Assessment with simple noninvasive fibrosis markers. Ann. Gastroenterol. 2017, 30, 349–356. [Google Scholar] [CrossRef]
  179. Androutsakos, T.; Schina, M.; Pouliakis, A.; Kontos, A.; Sipsas, N.; Hatzis, G. Liver Fibrosis Assessment in a Cohort of Greek HIV Mono-Infected Patients by Non-Invasive Biomarkers. Curr. HIV Res. 2019, 17, 173–182. [Google Scholar] [CrossRef]
  180. Yanavich, C.; Pacheco, A.G.; Cardoso, S.W.; Nunes, E.P.; Chaves, U.; Freitas, G.; Santos, R.; Morata, M.; Veloso, V.G.; Grinsztejn, B.; et al. Diagnostic value of serological biomarkers for detection of non-alcoholic fatty liver disease (NAFLD) and/or advanced liver fibrosis in people living with HIV. HIV Med. 2021, 22, 445–456. [Google Scholar] [CrossRef]
  181. Praktiknjo, M.; Djayadi, N.; Mohr, R.; Schierwagen, R.; Bischoff, J.; Dold, L.; Pohlmann, A.; Schwarze-Zander, C.; Wasmuth, J.C.; Boesecke, C.; et al. Fibroblast growth factor 21 is independently associated with severe hepatic steatosis in non-obese HIV-infected patients. Liver Int. 2019, 39, 1514–1520. [Google Scholar] [CrossRef]
  182. Sim, J.H.; Sherman, J.B.; Stanley, T.L.; Corey, K.E.; Fitch, K.V.; Looby, S.E.; Robinson, J.A.; Lu, M.T.; Burdo, T.H.; Lo, J. Pro-Inflammatory Interleukin-18 is Associated with Hepatic Steatosis and Elevated Liver Enzymes in People with HIV Monoinfection. AIDS Res. Hum. Retrovir. 2021, 37, 385–390. [Google Scholar] [CrossRef]
  183. Benmassaoud, A.; Ghali, P.; Cox, J.; Wong, P.; Szabo, J.; Deschenes, M.; Osikowicz, M.; Lebouche, B.; Klein, M.B.; Sebastiani, G. Screening for nonalcoholic steatohepatitis by using cytokeratin 18 and transient elastography in HIV mono-infection. PLoS ONE 2018, 13, e0191985. [Google Scholar] [CrossRef]
  184. Fourman, L.T.; Stanley, T.L.; Ockene, M.W.; McClure, C.M.; Toribio, M.; Corey, K.E.; Chung, R.T.; Torriani, M.; Kleiner, D.E.; Hadigan, C.M.; et al. Proteomic Analysis of Hepatic Fibrosis in Human Immunodeficiency Virus-Associated Nonalcoholic Fatty Liver Disease Demonstrates Up-regulation of Immune Response and Tissue Repair Pathways. J. Infect. Dis. 2023, 227, 565–576. [Google Scholar] [CrossRef]
  185. Núñez-Torres, R.; Macías, J.; Rivero-Juarez, A.; Neukam, K.; Merino, D.; Téllez, F.; Merchante, N.; Gómez-Mateos, J.; Rivero, A.; Pineda, J.A.; et al. Fat mass and obesity-associated gene variations are related to fatty liver disease in HIV-infected patients. HIV Med. 2017, 18, 546–554. [Google Scholar] [CrossRef] [Green Version]
  186. Martinez, M.A.; Tural, C.; Franco, S. Circulating MicroRNAs as a Tool for Diagnosis of Liver Disease Progression in People Living with HIV-1. Viruses 2022, 14, 1118. [Google Scholar] [CrossRef]
  187. Simpson, S.J.; Raubenheimer, D.; Cogger, V.C.; Macia, L.; Solon-Biet, S.M.; Le Couteur, D.G.; George, J. The nutritional geometry of liver disease including non-alcoholic fatty liver disease. J. Hepatol. 2018, 68, 316–325. [Google Scholar] [CrossRef]
  188. Berná, G.; Romero-Gomez, M. The role of nutrition in non-alcoholic fatty liver disease: Pathophysiology and management. Liver Int. 2020, 40 (Suppl. S1), 102–108. [Google Scholar] [CrossRef] [Green Version]
  189. Iruzubieta, P.; Bataller, R.; Arias-Loste, M.T.; Arrese, M.; Calleja, J.L.; Castro-Narro, G.; Cusi, K.; Dillon, J.F.; Martínez-Chantar, M.L.; Mateo, M.; et al. Research Priorities for Precision Medicine in NAFLD. Clin. Liver Dis. 2023, 27, 535–551. [Google Scholar] [CrossRef]
  190. Lassailly, G.; Caiazzo, R.; Buob, D.; Pigeyre, M.; Verkindt, H.; Labreuche, J.; Raverdy, V.; Leteurtre, E.; Dharancy, S.; Louvet, A.; et al. Bariatric Surgery Reduces Features of Nonalcoholic Steatohepatitis in Morbidly Obese Patients. Gastroenterology 2015, 149, 379–388; quiz e315–376. [Google Scholar] [CrossRef] [Green Version]
  191. Lassailly, G.; Caiazzo, R.; Ntandja-Wandji, L.C.; Gnemmi, V.; Baud, G.; Verkindt, H.; Ningarhari, M.; Louvet, A.; Leteurtre, E.; Raverdy, V.; et al. Bariatric Surgery Provides Long-term Resolution of Nonalcoholic Steatohepatitis and Regression of Fibrosis. Gastroenterology 2020, 159, 1290–1301.e1295. [Google Scholar] [CrossRef]
  192. Salomone, F.; Sharaiha, R.Z.; Boškoski, I. Endoscopic bariatric and metabolic therapies for non-alcoholic fatty liver disease: Evidence and perspectives. Liver Int. 2020, 40, 1262–1268. [Google Scholar] [CrossRef] [Green Version]
  193. Stine, J.G.; Long, M.T.; Corey, K.E.; Sallis, R.E.; Allen, A.M.; Armstrong, M.J.; Conroy, D.E.; Cuthbertson, D.J.; Duarte-Rojo, A.; Hallsworth, K.; et al. American College of Sports Medicine (ACSM) International Multidisciplinary Roundtable report on physical activity and nonalcoholic fatty liver disease. Hepatol. Commun. 2023, 7, e0108. [Google Scholar] [CrossRef]
  194. Thorp, A.; Stine, J.G. Exercise as Medicine: The Impact of Exercise Training on Nonalcoholic Fatty Liver Disease. Curr. Hepatol. Rep. 2020, 19, 402–411. [Google Scholar] [CrossRef]
  195. Stine, J.G.; DiJoseph, K.; Pattison, Z.; Harrington, A.; Chinchilli, V.M.; Schmitz, K.H.; Loomba, R. Exercise Training Is Associated with Treatment Response in Liver Fat Content by Magnetic Resonance Imaging Independent of Clinically Significant Body Weight Loss in Patients with Nonalcoholic Fatty Liver Disease: A Systematic Review and Meta-Analysis. Am. J. Gastroenterol. 2022. [Google Scholar] [CrossRef]
  196. Stine, J.G.; Schreibman, I.R.; Faust, A.J.; Dahmus, J.; Stern, B.; Soriano, C.; Rivas, G.; Hummer, B.; Kimball, S.R.; Geyer, N.R.; et al. NASHFit: A randomized controlled trial of an exercise training program to reduce clotting risk in patients with NASH. Hepatology 2022, 76, 172–185. [Google Scholar] [CrossRef]
  197. Farrugia, M.A.; Le Garf, S.; Chierici, A.; Piche, T.; Gual, P.; Iannelli, A.; Anty, R. Therapeutic Physical Exercise Programs in the Context of NASH Cirrhosis and Liver Transplantation: A Systematic Review. Metabolites 2023, 13, 330. [Google Scholar] [CrossRef]
  198. Wong, V.W.; Wong, G.L.; Chan, R.S.; Shu, S.S.; Cheung, B.H.; Li, L.S.; Chim, A.M.; Chan, C.K.; Leung, J.K.; Chu, W.C.; et al. Beneficial effects of lifestyle intervention in non-obese patients with non-alcoholic fatty liver disease. J. Hepatol. 2018, 69, 1349–1356. [Google Scholar] [CrossRef]
  199. Santos-Laso, A.; Gutiérrez-Larrañaga, M.; Alonso-Peña, M.; Medina, J.M.; Iruzubieta, P.; Arias-Loste, M.T.; López-Hoyos, M.; Crespo, J. Pathophysiological Mechanisms in Non-Alcoholic Fatty Liver Disease: From Drivers to Targets. Biomedicines 2021, 10, 46. [Google Scholar] [CrossRef]
  200. Ng, C.H.; Xiao, J.; Lim, W.H.; Chin, Y.H.; Yong, J.N.; Tan, D.J.H.; Tay, P.; Syn, N.; Foo, R.; Chan, M.; et al. Placebo effect on progression and regression in NASH: Evidence from a meta-analysis. Hepatology 2022, 75, 1647–1661. [Google Scholar] [CrossRef]
  201. Allen, A.M.; Therneau, T.M.; Ahmed, O.T.; Gidener, T.; Mara, K.C.; Larson, J.J.; Canning, R.E.; Benson, J.T.; Kamath, P.S. Clinical course of non-alcoholic fatty liver disease and the implications for clinical trial design. J. Hepatol. 2022, 77, 1237–1245. [Google Scholar] [CrossRef]
  202. Kim, W.; Kim, B.G.; Lee, J.S.; Lee, C.K.; Yeon, J.E.; Chang, M.S.; Kim, J.H.; Kim, H.; Yi, S.; Lee, J.; et al. Randomised clinical trial: The efficacy and safety of oltipraz, a liver X receptor alpha-inhibitory dithiolethione in patients with non-alcoholic fatty liver disease. Aliment. Pharmacol. Ther. 2017, 45, 1073–1083. [Google Scholar] [CrossRef] [Green Version]
  203. Harrison, S.A.; Bashir, M.R.; Guy, C.D.; Zhou, R.; Moylan, C.A.; Frias, J.P.; Alkhouri, N.; Bansal, M.B.; Baum, S.; Neuschwander-Tetri, B.A.; et al. Resmetirom (MGL-3196) for the treatment of non-alcoholic steatohepatitis: A multicentre, randomised, double-blind, placebo-controlled, phase 2 trial. Lancet 2019, 394, 2012–2024. [Google Scholar] [CrossRef]
  204. Ahn, H.Y.; Kim, H.H.; Hwang, J.Y.; Park, C.; Cho, B.Y.; Park, Y.J. Effects of Pioglitazone on Nonalcoholic Fatty Liver Disease in the Absence of Constitutive Androstane Receptor Expression. PPAR Res. 2018, 2018, 9568269. [Google Scholar] [CrossRef] [Green Version]
  205. Sanyal, A.J.; Chalasani, N.; Kowdley, K.V.; McCullough, A.; Diehl, A.M.; Bass, N.M.; Neuschwander-Tetri, B.A.; Lavine, J.E.; Tonascia, J.; Unalp, A.; et al. Pioglitazone, vitamin E, or placebo for nonalcoholic steatohepatitis. N. Engl. J. Med. 2010, 362, 1675–1685. [Google Scholar] [CrossRef] [Green Version]
  206. Zhao, X.; Wang, M.; Wen, Z.; Lu, Z.; Cui, L.; Fu, C.; Xue, H.; Liu, Y.; Zhang, Y. GLP-1 Receptor Agonists: Beyond Their Pancreatic Effects. Front. Endocrinol. 2021, 12, 721135. [Google Scholar] [CrossRef]
  207. Kenny, P.R.; Brady, D.E.; Torres, D.M.; Ragozzino, L.; Chalasani, N.; Harrison, S.A. Exenatide in the treatment of diabetic patients with non-alcoholic steatohepatitis: A case series. Am. J. Gastroenterol. 2010, 105, 2707–2709. [Google Scholar] [CrossRef]
  208. Nahra, R.; Wang, T.; Gadde, K.M.; Oscarsson, J.; Stumvoll, M.; Jermutus, L.; Hirshberg, B.; Ambery, P. Effects of Cotadutide on Metabolic and Hepatic Parameters in Adults with Overweight or Obesity and Type 2 Diabetes: A 54-Week Randomized Phase 2b Study. Diabetes Care 2021, 44, 1433–1442. [Google Scholar] [CrossRef]
  209. Subrahmanyan, N.A.; Koshy, R.M.; Jacob, K.; Pappachan, J.M. Efficacy and Cardiovascular Safety of DPP-4 Inhibitors. Curr. Drug Saf. 2021, 16, 154–164. [Google Scholar] [CrossRef]
  210. He, K.; Li, J.; Xi, W.; Ge, J.; Sun, J.; Jing, Z. Dapagliflozin for nonalcoholic fatty liver disease: A systematic review and meta-analysis. Diabetes Res. Clin. Pract. 2022, 185, 109791. [Google Scholar] [CrossRef]
  211. Yu, J.; Lee, S.H.; Kim, M.K. Recent Updates to Clinical Practice Guidelines for Diabetes Mellitus. Endocrinol. Metab. 2022, 37, 26–37. [Google Scholar] [CrossRef]
  212. National Library of Medicine. Available online: https://clinicaltrials.gov/ (accessed on 8 April 2023).
  213. Ratziu, V.; Sanyal, A.J.; Loomba, R.; Rinella, M.; Harrison, S.; Anstee, Q.M.; Goodman, Z.; Bedossa, P.; MacConell, L.; Shringarpure, R.; et al. REGENERATE: Design of a pivotal, randomised, phase 3 study evaluating the safety and efficacy of obeticholic acid in patients with fibrosis due to nonalcoholic steatohepatitis. Contemp. Clin. Trials 2019, 84, 105803. [Google Scholar] [CrossRef] [Green Version]
  214. Neuschwander-Tetri, B.A.; Loomba, R.; Sanyal, A.J.; Lavine, J.E.; Van Natta, M.L.; Abdelmalek, M.F.; Chalasani, N.; Dasarathy, S.; Diehl, A.M.; Hameed, B.; et al. Farnesoid X nuclear receptor ligand obeticholic acid for non-cirrhotic, non-alcoholic steatohepatitis (FLINT): A multicentre, randomised, placebo-controlled trial. Lancet 2015, 385, 956–965. [Google Scholar] [CrossRef] [Green Version]
  215. Ratziu, V.; de Guevara, L.; Safadi, R.; Poordad, F.; Fuster, F.; Flores-Figueroa, J.; Arrese, M.; Fracanzani, A.L.; Ben Bashat, D.; Lackner, K.; et al. Aramchol in patients with nonalcoholic steatohepatitis: A randomized, double-blind, placebo-controlled phase 2b trial. Nat. Med. 2021, 27, 1825–1835. [Google Scholar] [CrossRef]
  216. Khoshbaten, M.; Aliasgarzadeh, A.; Masnadi, K.; Tarzamani, M.K.; Farhang, S.; Babaei, H.; Kiani, J.; Zaare, M.; Najafipoor, F. N-acetylcysteine improves liver function in patients with non-alcoholic Fatty liver disease. Hepat. Mon. 2010, 10, 12–16. [Google Scholar]
  217. Hang, W.; Shu, H.; Wen, Z.; Liu, J.; Jin, Z.; Shi, Z.; Chen, C.; Wang, D.W. N-Acetyl Cysteine Ameliorates High-Fat Diet-Induced Nonalcoholic Fatty Liver Disease and Intracellular Triglyceride Accumulation by Preserving Mitochondrial Function. Front. Pharmacol. 2021, 12, 636204. [Google Scholar] [CrossRef]
  218. Tsai, C.C.; Chen, Y.J.; Yu, H.R.; Huang, L.T.; Tain, Y.L.; Lin, I.C.; Sheen, J.M.; Wang, P.W.; Tiao, M.M. Long term N-acetylcysteine administration rescues liver steatosis via endoplasmic reticulum stress with unfolded protein response in mice. Lipids Health Dis. 2020, 19, 105. [Google Scholar] [CrossRef]
  219. Du, J.; Ma, Y.Y.; Yu, C.H.; Li, Y.M. Effects of pentoxifylline on nonalcoholic fatty liver disease: A meta-analysis. World J. Gastroenterol. 2014, 20, 569–577. [Google Scholar] [CrossRef]
  220. Castro-Narro, G.; Moctezuma-Velázquez, C.; Male-Velázquez, R.; Trejo-Estrada, R.; Bosques, F.J.; Moreno-Alcántar, R.; Rodríguez-Hernández, H.; Bautista-Santos, A.; Córtez-Hernández, C.; Cerda-Reyes, E.; et al. Position statement on the use of albumin in liver cirrhosis. Ann. Hepatol. 2022, 27, 100708. [Google Scholar] [CrossRef]
  221. Balak, D.M.W.; Piaserico, S.; Kasujee, I. Non-Alcoholic Fatty Liver Disease (NAFLD) in Patients with Psoriasis: A Review of the Hepatic Effects of Systemic Therapies. Psoriasis 2021, 11, 151–168. [Google Scholar] [CrossRef]
  222. Kessoku, T.; Imajo, K.; Kobayashi, T.; Ozaki, A.; Iwaki, M.; Honda, Y.; Kato, T.; Ogawa, Y.; Tomeno, W.; Kato, S.; et al. Lubiprostone in patients with non-alcoholic fatty liver disease: A randomised, double-blind, placebo-controlled, phase 2a trial. Lancet Gastroenterol. Hepatol. 2020, 5, 996–1007. [Google Scholar] [CrossRef]
  223. Friedman, S.L.; Neuschwander-Tetri, B.A.; Rinella, M.; Sanyal, A.J. Mechanisms of NAFLD development and therapeutic strategies. Nat. Med. 2018, 24, 908–922. [Google Scholar] [CrossRef]
Figure 1. Non-alcoholic fatty liver disease (NAFLD) includes a spectrum of liver manifestations ranging from steatosis to cirrhosis. Progression of the disease is determined by different pathological pathways, mainly metabolism impairment and liver inflammation, which are influenced by patients’ lifestyles, genetics, microbiomes and environmental factors. Interventions on these factors could promote regression at early stages of the disease.
Figure 1. Non-alcoholic fatty liver disease (NAFLD) includes a spectrum of liver manifestations ranging from steatosis to cirrhosis. Progression of the disease is determined by different pathological pathways, mainly metabolism impairment and liver inflammation, which are influenced by patients’ lifestyles, genetics, microbiomes and environmental factors. Interventions on these factors could promote regression at early stages of the disease.
Ijms 24 10718 g001
Figure 2. Non-alcoholic fatty liver disease (NAFLD) includes a wide spectrum of patients whose phenotype is determined by the different affection of key processes involved in the pathophysiology of the disease, including body weight and composition, the presence of metabolic syndrome and type 2 diabetes mellitus, liver histological features, genetic predisposition, the use of toxic substances such as small amounts of alcohol and steatogenic drugs, infections and alterations in the gut microbiome, development of vascular disease and systemic inflammation.
Figure 2. Non-alcoholic fatty liver disease (NAFLD) includes a wide spectrum of patients whose phenotype is determined by the different affection of key processes involved in the pathophysiology of the disease, including body weight and composition, the presence of metabolic syndrome and type 2 diabetes mellitus, liver histological features, genetic predisposition, the use of toxic substances such as small amounts of alcohol and steatogenic drugs, infections and alterations in the gut microbiome, development of vascular disease and systemic inflammation.
Ijms 24 10718 g002
Table 1. Genetic variants associated with NAFLD.
Table 1. Genetic variants associated with NAFLD.
GeneSNPsFunctionReferences
PNPLA3rs738409Lipid metabolism and inflammatory response[70,71]
rs6006460
MBOAT7rs641738Lipid metabolism[72]
HSD17B13rs72613567Lipid metabolism[73]
FTOrs1421085Lipid metabolism and adipogenesis[74]
LPIN1rs13412852Lipid metabolism and adipogenesis[75]
TM6SF2rs58542926VLDL secretion[76]
LYPLAL1rs12137855Glucose homeostasis[77]
GCKRrs780094Regulation of de novo lipogenesis and insulin resistance[78,79]
ENPP1rs1044498Insulin signaling inhibitor[80]
PPP1R3Brs4240624Glycogen metabolism[81]
SOD2rs4880Fibrosis and oxidative stress[82]
MERTKrs4374383Immune response[83]
FNDC5rs3480Liver fibrogenesis[84]
KLF6rs3750861Liver fibrogenesis[85]
CDKN1Ars762623Cell senescence[86]
IL28Brs12979860Inflammatory response[87]
Table 2. Steatogenic drugs.
Table 2. Steatogenic drugs.
Therapeutic ClassDrug/GroupReferences
AntiarrhythmicsAmiodarone[126]
AntibioticsTetracyclines[127,128]
AntidiabeticsTroglitazone[129,130,131]
AntiepilepticsCarbamazepine[132]
Valproic acid[132,133,134]
Anti-inflammatoriesDexamethasone[135]
Antitumor drugs5-Fluorouracil[136,137]
Irinotecan[137,138,139,140]
Leuprorelin acetate[141]
Methotrexate[142]
Tamoxifen[143,144]
AntiretroviralsNucleoside Reverse Transcriptase Inhibitors[145,146]
Protease Inhibitors[147]
HormonesEstrogens[148]
Vasodilator agentsPerhexiline maleate[149,150]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alonso-Peña, M.; Del Barrio, M.; Peleteiro-Vigil, A.; Jimenez-Gonzalez, C.; Santos-Laso, A.; Arias-Loste, M.T.; Iruzubieta, P.; Crespo, J. Innovative Therapeutic Approaches in Non-Alcoholic Fatty Liver Disease: When Knowing Your Patient Is Key. Int. J. Mol. Sci. 2023, 24, 10718. https://doi.org/10.3390/ijms241310718

AMA Style

Alonso-Peña M, Del Barrio M, Peleteiro-Vigil A, Jimenez-Gonzalez C, Santos-Laso A, Arias-Loste MT, Iruzubieta P, Crespo J. Innovative Therapeutic Approaches in Non-Alcoholic Fatty Liver Disease: When Knowing Your Patient Is Key. International Journal of Molecular Sciences. 2023; 24(13):10718. https://doi.org/10.3390/ijms241310718

Chicago/Turabian Style

Alonso-Peña, Marta, Maria Del Barrio, Ana Peleteiro-Vigil, Carolina Jimenez-Gonzalez, Alvaro Santos-Laso, Maria Teresa Arias-Loste, Paula Iruzubieta, and Javier Crespo. 2023. "Innovative Therapeutic Approaches in Non-Alcoholic Fatty Liver Disease: When Knowing Your Patient Is Key" International Journal of Molecular Sciences 24, no. 13: 10718. https://doi.org/10.3390/ijms241310718

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