Next Article in Journal
Association of Zinc Deficiency with Development of CVD Events in Patients with CKD
Previous Article in Journal
Effects of Three-Month Administration of High-Saturated Fat Diet and High-Polyunsaturated Fat Diets with Different Linoleic Acid (LA, C18:2n–6) to α-Linolenic Acid (ALA, C18:3n–3) Ratio on the Mouse Liver Proteome
Previous Article in Special Issue
The Role of Fructose in Non-Alcoholic Steatohepatitis: Old Relationship and New Insights
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nutrigenomics and Nutrigenetics in Metabolic- (Dysfunction) Associated Fatty Liver Disease: Novel Insights and Future Perspectives

by
Marcello Dallio
1,*,
Mario Romeo
1,
Antonietta Gerarda Gravina
1,
Mario Masarone
2,
Tiziana Larussa
3,
Ludovico Abenavoli
3,
Marcello Persico
2,
Carmelina Loguercio
1 and
Alessandro Federico
1
1
Department of Precision Medicine, University of Campania Luigi Vanvitelli, Via S. Pansini 5, 80131 Naples, Italy
2
Department of Medicine and Surgery, University of Salerno, Via Allende, 84081 Baronissi, Italy
3
Department of Health Sciences, University Magna Graecia, viale Europa, 88100 Catanzaro, Italy
*
Author to whom correspondence should be addressed.
Nutrients 2021, 13(5), 1679; https://doi.org/10.3390/nu13051679
Submission received: 22 March 2021 / Revised: 12 May 2021 / Accepted: 13 May 2021 / Published: 15 May 2021
(This article belongs to the Special Issue Nutrition and Non-alcoholic Fatty Liver Disease: Gene Storm)

Abstract

:
Metabolic- (dysfunction) associated fatty liver disease (MAFLD) represents the predominant hepatopathy and one of the most important systemic, metabolic-related disorders all over the world associated with severe medical and socio-economic repercussions due to its growing prevalence, clinical course (steatohepatitis and/or hepatocellular-carcinoma), and related extra-hepatic comorbidities. To date, no specific medications for the treatment of this condition exist, and the most valid recommendation for patients remains lifestyle change. MAFLD has been associated with metabolic syndrome; its development and progression are widely influenced by the interplay between genetic, environmental, and nutritional factors. Nutrigenetics and nutrigenomics findings suggest nutrition’s capability, by acting on the individual genetic background and modifying the specific epigenetic expression as well, to influence patients’ clinical outcome. Besides, immunity response is emerging as pivotal in this multifactorial scenario, suggesting the interaction between diet, genetics, and immunity as another tangled network that needs to be explored. The present review describes the genetic background contribution to MAFLD onset and worsening, its possibility to be influenced by nutritional habits, and the interplay between nutrients and immunity as one of the most promising research fields of the future in this context.

Graphical Abstract

1. Introduction

Non-alcoholic fatty liver disease (NAFLD) represents one of the most important metabolic-related disorders of the 21st century and the leading cause of chronic liver disease and liver transplantation worldwide [1]. It includes a wide spectrum of conditions ranging from simple steatosis to steatohepatitis (NASH), characterized by the histologic appearance of inflammation and fibrosis, which act as driving factors to fuel the disease progression and complications onset [2].
Currently, NAFLD is considered the hepatic manifestation of metabolic syndrome (MS) [3] and recently, an expert consensus established that metabolic- (dysfunction) associated fatty liver disease (MAFLD) represents the more adequate denomination, revealing its larger and deeper nature as systemic disorder [4] (Figure 1).
Alarmingly, as the diffusion of MS-related conditions continues to increase and the obesity pandemic spreads irrepressibly, MAFLD prevalence, in turn, seems to rise exponentially [1]. The worldwide prevalence ranges from 6% to 35%, with higher levels in the industrialized countries (the Middle East 32%, South America 31%, United States 24%, and Europe 23%) and lower levels in the underdeveloped ones (14%) [5]. Moreover, future perspectives appear decidedly not encouraging, expecting prevalence to reach about 100 million in the United States alone by 2030 [1].
The numbers of the MAFLD spread worldwide raise several important concerns principally due to the very complex and still not fully understood pathogenesis. It gives to the disease an inherent difficulty to support the burden of its optimal medical and social management.
Pathogenetically MAFLD represents a multifactorial disease in which various elements can simultaneously contribute to the genesis and affect natural history, contributing to make its manifestation from patient to patient hugely different [6]. The genome-wide association studies (GWAS) shed light on the genetic susceptibility of the population to MAFLD onset and evolution, identifying several single nucleotide polymorphisms (SNPs), like phospholipase domain-containing protein-3 (PNPLA3) rs738409, the transmembrane 6 superfamily member 2 protein (TM6SF2) rs58542926, and membrane-bound O-acyltransferase domain containing 7 (MBOAT7) rs641738, as crucial in this scenario [7,8]. At the same time, the contemporary involvement of several environmental factors like sedentary life, unhealthy diet regimens promoting insulin resistance (IR), and gut dysbiosis, together with some immune disturbances, have been accepted by the scientific community as “tiles of MAFLD pathogenetic mosaic” [9,10,11,12] (Figure 1).
To date, no specific medications for MAFLD treatment exist [5,13]; hence, given its demonstrated capability to induce huge improvement in IR and liver damage, lifestyle modification remains the most valid and accepted recommendation [14]. This approach classically rests on three enchained pivots: regular physical exercise, weight loss, and healthy diet [15]. Regarding the latter, the scientific evidence suggests the possibility that a regimen characterized by the main consumption of plant-based food, fish, and white meat could contribute significantly to the reduction of several chronic diseases’ occurrence, including MS components and thus MAFLD [16,17,18]. These principles constitute the paradigm of the well-known Mediterranean diet (MD), which, based on the antioxidant and anti-inflammatory properties of the recommended food, currently represents the nutritional gold standard of preventive medicine [19].
However, the different kinds of therapeutic outcomes potentially obtained from this dietary regimen induced the scientific community to focus attention on the identification of the factors involved in determining the dietary therapeutic effect, hypothesizing a mutual relationship between diet and genetics [15]. Recent findings suggested nutrition’s capability, by acting on the individual genetic background and modifying the specific epigenetic expression as well, to influence NAFLD patients’ clinical severity and response to the treatment [20]. The study of the interaction between nutrients and inherited factors constitutes the central aim of emerging and promising lines of research known as nutrigenetics and nutrigenomics [21].
Globally, the main scope of these research fields is strictly linked in a very complex scientific network with other disciplines, such as genomics, transcriptomics, proteomics, metabolomics, and system biology, configuring the “Omics-approach” to the disease that seems to be the key for the full comprehension of MAFLD pathogenesis and evolution [21,22] (Figure 2).
However, further efforts will have to be spent to apply concretely the research findings in routine clinical practice and also to design, in the next future, a tailored therapeutic approach fit with the era of precision medicine.
This review describes the genetic background contribution to MAFLD onset and worsening, its possibility to be influenced by nutritional habits, and the effect of genetics in determining certain biologic responses to nutrients based on the current nutrigenetics and nutrigenomics scientific knowledge concerning metabolic-associated disorders.

2. Genetics and Epigenetics: An Overview on MAFLD Genetic Background

In MAFLD context, heritability seems to play a pivotal role as suggested clearly just considering the higher disease risk of offspring in case of positive family history for MAFLD, particularly when both parents are affected [23,24]. Furthermore, the ethnic susceptibility represents another overwhelming proof. Different ethnicities, in fact, show different propensity towards disease development and progression [25,26]. Interestingly, besides the inter-ethnic one, growing evidence reveals also an inter-individual variability. Therefore, people belonging to the same ethnicity have different possibilities to develop MAFLD, and this appears strictly linked to certain genetic variants as well as genetic expressions able to influence its onset and evolution [27,28,29].

2.1. Most Common Genetic Determinants of MAFLD

A long list of genetic inter-individual variants has been provided by GWAS in this context; it embraces several genes known as genetic determinants of MAFLD, whose expression is involved in the formation of lipid droplets and regulation of oxidative stress, inflammation, fibrogenesis, and many other metabolic pathways [7,29,30] (Table 1).
However, more robust and reproducible associations seem to exist, above all, for PNPLA3, MBOAT7, and TM6SF2 variants that currently are considered the main genetic MAFLD determinants [8].
The rs738409 SNP of the PNPLA3 gene (or Adiponutrin), encoding for the I148M protein variant, showed heavy impact on disease susceptibility. This variant has been associated with higher hepatic fat content in an IR-independent manner, as demonstrated by Romeo et al. [31]. In details, the role of IR in the light of the current scientific knowledge regarding this purpose seems to be extremely complex and, from a certain point of view, still not fully understood. In fact, the PNPLA3 gene encodes for the homonymous transmembrane protein localized on the endoplasmic reticulum [32], mainly expressed in hepatocytes and adipose tissue and exhibiting triglycerides (TG) hydrolase activity regulated by glucose and insulin [32,33]. In case of IR and obesity, both conditions featured by high insulin levels, PNPLA3 expression is induced, and the protein is placed and accumulated on the surface of lipid droplets where it is not catalytically active, disrupting TG and phospholipids turnover and remodelling [34].
PNPLA3 is also involved in the release of the storage form of retinol, known as retinyl-palmitate, in hepatic stellate cells (HSCs) [35,36]. Due to the involvement of retinyl-palmitate in the regulation of cellular fat metabolism and HSCs activation, its retention caused by the I148M genetic variant determines, consequently, the activation of pro-inflammatory and pro-fibrotic responses [35,37]. About that, growing evidence in animal and clinical experimental models suggests the I148M genetic variant relationship with all the steps of the diseases’ natural history, from simple steatosis to NASH, cirrhosis, and hepatocellular carcinoma (HCC) development [38,39]. Moreover, this SNP appears to be involved as the key driver in chronic kidney disease (CKD) development and is currently recognized as an established marker of higher cardiovascular risk in MAFLD [40].
MBOAT7 is a gene encoding for a membrane-bound enzyme whose main function is the incorporation of arachidonic acid (AA) and other unsaturated fatty acids (UFAs) in the phosphatidylinositol (PI) molecule by the Lands cycle (a series of phospholipid-remodelling reactions by which acyl-chains become trans acylated) [41]. In the presence of the rs641738 variant, the expression of the encoded enzyme decreases together with the hepatic levels of PI containing AA and causes a disruption of several cellular pathways regulating TG metabolism, inflammation, fibrosis, and cellular proliferation as well [42,43,44].
The TM6SF2 gene is responsible for the production of a protein mainly localized in the endoplasmic reticulum and Golgi apparatus, involved in the regulation of the hepatic triglyceride secretion [45,46]. The rs58542926 C>T encoding for the E167K variant of this gene is involved in MAFLD development and worsening of histological picture, impairing inflammation, ballooning, and fibrosis, as demonstrated by Liu et al. [42,47]. On the other side, this SNP seems to confer protection from the cardiovascular disease risk through the reduction of lipid secretion and very-low-density lipoproteins (VLDL) synthesis, designing a very complex biologic role that still remains to be fully clarified [48,49,50].
Besides PNPLA3, MBOAT7, and TM6SF2 SNPs, other genetic variants appear potentially involved in this scenario. Strong scientific evidence exists for the glucokinase regulator (GCKR) common missense variants rs1260326 encoding for the P446L protein [51]. The common missense variant rs1260326 disrupts GCKR function, making it unable to inhibit the glucokinase and, consequently, activating hepatic glucose uptake and glycolysis. These phenomena lead to the generation of acetyl-CoA cellular overload and de novo lipogenesis (DNL) [51]. Contrariwise, the protein phosphatase 1 regulatory subunit 3B (PPP1R3B) rs4841132 variant, through the reduction of lipogenesis and the increase of glycogen synthesis, seems to protect against hepatic fat accumulation [52]. Besides the aforementioned one involved in steatosis development, other genetic polymorphisms have been associated with the disease progression.
As known, reactive oxygen species (ROS) overproduction derived from free fatty acids (FFAs) overload in mitochondria secondary to IR and consequent mitochondrial dysfunction, represent critical events in MAFLD worsening [53]. In this regard, some studies highlighted the association of polymorphism rs4880 of Superoxide dismutase 2 (SOD 2) with higher oxidative stress levels, inflammation, and more advanced fibrosis [54,55]. In line with this, the uncoupling protein 3 (UCP3) rs1800849, uncoupling protein 2 (UCP2) rs695366, homeostatic iron regulator (HFE) C282Y rs1800562, and a rare missense variant (A165T) in mitochondrial amidoxime reducing component 1 (MARC1) represent other new interesting topics [7,56,57]. UCP3 is a mitochondrial anion carrier selectively expressed in skeletal muscle involved in the modulation of energy, lipid homeostasis, and thermogenesis by facilitating the proton leak of the mitochondrial inner membrane and uncoupling the oxidative phosphorylation [58]. Interestingly, a Spanish study conducted on a cohort of overweight patients revealed the association of UCP3 rs1800849 variant with IR, increased adiponectin levels, and NASH [59]. On the contrary, the presence of the UCP2 rs695366 variant has been linked to higher gene expression, insulin sensitivity, and protection from liver damage [60].
Other rare genetic variants were demonstrated related to MAFLD clinical context as mainly involved in the disease progression instead of its development: the rs236918 genetic variant in proprotein convertase subtilisin/Kexin type 7 (PCSK7) [61], or protecting from the disease and complication onset: the loss of function in HSD17B13 gene due to rs72613567 variant [62,63].
However, one of the most intriguing questions is how far we are to apply in the routine clinical practice the knowledge derived from GWAS studies. This question could be interpreted from some different point of views on MAFLD clinical picture: predictive, prognostic, and therapeutic ones [28]. On this line, it is necessary to highlight the lack of scientific agreement regarding the best methodological choice to evaluate the utility of genetic variant risk estimates [64]. Considering the PNPLA3 I148M variant, it has great power in the prediction of the disease, which was highlighted by several clinical trials in which its role in disease appearance was totally demonstrated independent from the other classical risk factors [24,65]. However, its pertinence as a heritable factor for NAFLD development was not demonstrated as brilliant as the latter purpose and, for this reason, the actual clinical management guidelines of the European Association for the Study of the Liver do not recommend its routine assessment for NAFLD-related liver damage evaluation [66]. In this regard, because of the lack of sufficient scientific proof to support the use of a single SNP in the prediction of the disease risk, the recent scientific literature has focused attention on the possible applicability of polygenic risk scores (PRSs) for this purpose [64]. This acquires even more relevance considering the accuracy of PRS based on well-established SNP and commonly recognized risk factors for the disease development in NAFLD risk prediction [67,68]. The use of PRS in a cross-sectional study on NAFLD was demonstrated able to induce an improvement of risk prediction in about 20% of patients. Moreover, recently, PNPLA3-TM6SF2-GCKR-MBOAT7 variants combined in a hepatic fat PRS (PRS-HFC) and adjusted for HSD17B13 (PRS-5) were demonstrated as able to predict HCC more efficiently than single variants assessment and that the association between PRS and HCC, mediated by severe fibrosis, was independent from the latter in clinically relevant subgroups and in those without advanced stages fibrosis [69]. Despite that the emerging research topic could have huge scientific impact on future clinical management, the amount of scientific evidence currently remains the most important concern to recommend their routine use. Moreover, regarding the prediction of long-term outcome, independently from the baseline staging of the disease, nowadays, it is not possible to state scientifically coherent conclusions, at least until the publication of data from long-term prospective studies.
The possibility of MAFLD therapeutic outcome prediction in the era of the patients-tailored approach assumes a fascinating aura in the light of some research trials on new therapeutic agents for disease treatment [70]. This novel research line could give the possibility to interpret better the usefulness of a wide range of already known drugs and newly developed ones as well. The WELCOME trial evaluated the response to 4 × 1000-mg capsules of 460 mg eicosapentaenoic acid and 380 mg docosahexaenoic acid administration for 15–18 months on liver fat content and fibrosis in NAFLD patients [71]. At the end of treatment, those patients carrying the 148I/I and 148I/M genotype showed a decrease of liver fat percentage (148I/I: −7.05%, 148I/M: −7.30%) whereas the 148M/M group showed a moderate increase (2.75%) [71].
Even if for some of the identified genes more powerful scientific evidences are still missing and the results of the different studies appear sometimes controversial as well, there is no doubt in considering this field one of the most promising topics of the recent MAFLD research.

2.2. Main Epigenetic Mechanisms of MAFLD

As part of the complex genetic background sustaining MAFLD, several epigenetic phenomena, influencing different levels of gene expression regulation, seem to be potentially involved in pathogenesis and clinical history. The mechanisms described encompass DNA methylation, histone modifications, and microRNAs (miRNAs) activity on specific targets [72]. In this context, the development of a characteristic methylation pattern could be critical and fuel the disease progression [73]. Kitamoto et al. compared the levels of DNA methylation of certain CpG islands as CpG99 (which resides in the regulatory region of PNPLA3) and CpG26 (which resides in that of PARVB variant 1) in the livers of patients with mild (fibrosis stages 0 and 1) or advanced (fibrosis stages 2 to 4) steatosis by performing targeted-bisulfite sequencing [74]. Relevantly, in the livers of patients with advanced disease, CpG26 resulted markedly hypomethylated while CpG99 was substantially hypermethylated, suggesting the hypomethylation of CpG26 and the hypermethylation of CpG99 as potential contributors to the severity of fibrosis in patients with MAFLD [74]. Moreover, in individuals affected by severe steatosis compared to mild ones, lower DNA methylation levels characterized specific CpG islands of noted pro-fibrotic genes, such as transforming growth factor- β(TGF-β), collagen 1A1, platelet-derived growth factor- α(PDGF-α), and others [75,76]. On the contrary, hypermethylation status occurred for certain CpG islands in various anti-steatotic and anti-fibrotic genes, such as ApoB and peroxisome proliferator-activated receptor (PPAR)-α [75,77,78]. The loss of PPAR-α functioning seems extremely important because it regulates cytokine production, reducing the expression of pro-inflammatory ones, and, on the other hand, it modulates the proteins involved in the fatty acid binding activity, taking part in the regulation of lipogenesis during the oxidation [77,79].
In this sense, a large number of liver DNA methylation alterations widely affecting disease onset and worsening by promoting inflammation and fibrosis has been related to certain metabolic features in terms of insulin, amino acids, and lipids serum levels. An interesting analysis of the DNA methylation pattern performed on the liver biopsies obtained from 95 obese individuals (34 individuals showing normal liver phenotype, 35 simple steatosis, and 26 steatohepatitis), identified 1292 CpG sites representing 677 unique genes differentially methylated in the livers of individuals with advanced disease (i.e., steatohepatitis) [80]. Focusing on the top-ranking 30 and another 37 CpG sites mapped to genes enriched in pathways of metabolism and cancer all together, the authors revealed 59 steatohepatitis-associated CpG sites correlating with fasting insulin levels independently of age, fasting glucose, or diabetes mellitus type 2 [80].
In addition, in a recent study on 194 obese patients (79 with liver histology classifiable as normal liver, 40 as simple steatosis, and 45 as steatohepatitis) aiming to assess serum aromatic and branched-chain amino acids levels association with steatohepatitis, the tryptophan resulted significantly higher in those with advanced disease compared to those with simple steatosis [81]. Relevantly, these amino acid serum levels result were associated with liver DNA methylation of CpG sites known to be differentially methylated in individuals with steatohepatitis and were correlated positively with serum total and low-density lipoprotein (LDL) cholesterol and, accordingly, with liver low-density lipoprotein receptor (LDL-R) at mRNA-expression level [81].
Histone modifications represent another epigenetic phenomenon potentially related to different metabolic dysfunctions [72]. The disruption of circadian rhythm of histone acetylation regulated by histone deacetylases (HDCA3), implicated in the regulation of the circadian rhythm of hepatic lipogenesis [82], may alter hepatic lipid metabolism leading to IR and obesity [83,84]. In particular, a group of deacetylates, known as silent information regulator 2 proteins (Sirtuins), may exert a key role in MAFLD development [84], according to growing evidence revealing their downregulation in animals and in vitro models [84,85]. In mice, in fact, liver-specific deletion of SIRT1 by impairing of PPAR-α signalling and decreasing fatty acids (FAs) beta-oxidation, leads to IR and related inflammation [85]. Relevantly, both the DNA methylation and histone modifications may also occur in mitochondrial DNA (mt-DNA) [86].
This feature appears in line with the aforementioned mitochondrial dysfunction implicated in MAFLD pathogenesis. Hypermethylation of mitochondrially encoded NADH ubiquinone oxidoreductase core subunit 6 (MT-ND6) was demonstrated on liver biopsy samples from NASH patients compared with subjects affected by liver steatosis [87]. In addition, the authors highlighted the association of MT-ND6 methylated/unmethylated DNA ratio with the NAFLD activity score (NAS) [87]. In keeping, a more recent study by Pirola et al. revealed higher levels of MT-Cytochrome B variance and mt-DNA damage in NASH patients compared to simple steatosis ones [86].
Regarding mt-DNA histone modifications, genetic polymorphism of SIRT3, a mitochondrial sirtuin pivotal for guaranteeing mitochondrial integrity and metabolism during oxidative stress [88], has been associated with the development of MS in mice and humans as well as contributes to downregulate autophagy, leading to lipotoxicity in hepatocytes and thus MAFLD worsening [89,90,91].
The role of epigenetics is not only limited to the DNA access variations being expressed also in post-transcriptional steps acting by miRNAs, small non-coding single strand RNAs (ssRNAs) able to repress or degrade specific mRNAs target, regulate several biological and pathological processes, and intervene in lipid metabolism and inflammatory processes [92,93]. In this context, one of the most relevant studied miRNAs is miRNA-122, identified as a promising biomarker and drug target for MAFLD [94]. Esau et al. firstly highlighted the key role of this miRNA in the liver of high-fat diet (HFD) fed mice; after treatment with antisense oligonucleotides (ASO) inhibitors of miRNA-122, a significant reduction of hepatic biosynthesis of FAs, an increased beta-oxidation, and a reduction of TG accumulation were observed [95]. Recent scientific evidence demonstrated through in vitro and animal models the miRNA-122 suppressing role of SIRT1 expression via binding its 3’-untranslated region (UTR), enhancing lipogenesis, and contributing, thus, to lipid accumulation [94].
In contrast, miRNA-122 knockdown mitigated this consequence through the upregulation of SIRT1 and the activation of the liver kinase B1 (LKB1)/adenosine monophosphate-activated protein kinase (AMPK) signalling pathway [94]. Besides miRNA-122, other miRNAs could interfere with these processes [96]: for instance, miRNA-33a/b action on the target AMPK, by reducing AMPK expression, increases the levels of intrahepatic TG [97]; contrariwise, miRNA-33a/b inhibition promotes beta-oxidation of FAs and insulin sensitivity [97,98]. However, the role of miRNAs was not only related to the induction of steatosis because of the proved effects on the progression to inflammation and fibrosis [99]. In this sense, miRNA-34a results highly expressed in MAFLD patients in a stage-dependent manner [100,101,102]. On the contrary, the expression levels of miRNA-451, able to downregulate nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) and tumour necrosis factor–α(TNF-α) in vitro, resulted significantly decreased in NASH patients [101].
Altogether, these epigenetic mechanisms contribute to MAFLD inter-individual variability in terms of susceptibility towards disease and its progression, influencing the clinical history of each patient. To improve MAFLD individual prognosis, the future challenge is certainly represented by the application of these findings to the routine clinical practice, through the identification of biomarkers and therapeutic targets usable for early diagnosis and personalized therapies. In these terms, the above-presented findings highlight a suggestive relationship between the occurrence of epigenetic phenomena modifying gene expression and metabolic dysfunctions (including, among others, IR and high LDL serum levels). This feature appears widely relevant, representing an important food for thought regarding potential nutrigenomics approaches considering certain nutrients’ capability to influence gene expression, discussed as well in the next paragraph.

3. Nutrigenomics in MAFLD: The Impact of Nutrients on Epigenetic Regulation

The integration of the Omics approach and system biology for MAFLD study certainly represents one of the most important novelties of the recent scientific literature. Thanks to an increasing number of high-throughput Omics investigations published in the last few years, the molecular pathobiology of MAFLD, considered as a systemic disorder, has been better clarified. This fact represented an important starting point to properly debate regarding the appropriacy of the holistic view on MAFLD pathophysiology, potentially leading to some important diagnostic and therapeutic advantages as well [103]. As part of this tangled biologic network, nutrigenomics appears to be crucial in this setting, particularly considering the potential benefit that could derive from the dietary therapeutic application and tailored clinical management. In this sense, nutrigenomics represents a part of the well-known environment exposition/disease development relationship. The origin of this research field is based on the observation of a gene-diet effect due to a drastic change in the exposition to specific macronutrients, such as lipids, carbohydrates, or proteins, as well as considering the availability of micronutrients, involved as cofactors in DNA restructuring. In fact, it seems to be able to influence the expression and thus the activity of several enzymes involved in nutrients’ metabolism conducing, potentially, to a metabolic imbalance [104].

3.1. Lipids

The food enrichment with omega 6 (n-6) 18-carbon polyunsaturated fatty acids (PUFAs), in particular linoleic acid (LA), linked to the alteration of LA/alpha-linolenic acid ratio, represents a classic example of the latter concept, conducing to an altered omega 3 (n-3) and n-6 metabolism of inflammatory PUFAs and enhancing, then, the production of inflammatory mediators involved in several disorders [105,106]. The DNL represents, together with the adipose tissue fat release and the block of beta-oxidation, a crucial mechanism able to trigger the hepatocyte TG overload induced by high short-chain fatty acids (SCFAs) consumption. They have been hypothesized to exert this effect, inducing the activation of the peroxisome proliferator-activated receptor (PPAR)-coactivator-1 beta (PGC-1) and enhancing the expression of stearoyl-CoA desaturase-1, fatty acid synthase, diacylglycerol acyltransferase genes involved in lipid metabolism and homeostasis [107]. The clinical effect of high n-6/n-3 PUFAs ratio and SCFAs-rich food consumption is represented by both hepatocyte and lipocyte dysfunction that could lead to cellular fat accumulation, inflammation, and fibrosis in a typical NASH histologic picture [19,53].
Contrariwise, the increase of n-3 PUFAs in the daily diet was demonstrated in preclinical models as related to suppression of diet-induced steatosis and improvement of IR, mainly reducing the activity of sterol regulatory element-binding proteins (SREBP)-1 transcription factor, the expression of DNL, and fat uptake genes, as well as by increasing the expression of genes involved in FAs oxidation and export thought VLDL secretion [108,109]. These gene-diet molecular mechanisms of interaction lead also to important consequences on the major cause of death in this context: the cardiovascular events. In fact, it was well established that n-3-enriched food consumption or a low n-6/n-3 ratio is linked to an improvement of the blood lipid profile, lowering the LDL level up to 20 % along with glycemia and IR and reducing the risk of death for coronary heart diseases [110,111]. However, as recently highlighted in the FADSDIET and FADSDIET2 trials, an interconnection between the fatty acid desaturase (FADS) rs174550 gene cluster and dietary linoleic and alpha-linolenic acid intake on plasma lipids composition, fasting glucose, and high-sensitivity C-reactive protein (hsCRP) exists, in particular for TT homozygous carriers in which plasma eicosanoid concentrations correlated with the arachidonic acid proportion in plasma and with hsCRP [112,113].
The effect of diet in the explored context embraces also the miRNA pathobiology and, on this purpose, Long et al. highlighted a SIRT1-suppressing effect exerted by FFAs and a high-fat diet able to promote lipogenesis in HepG2, HuH7 cell lines, and mice thought miR-122 upregulation [94]. MiR-181b and miR34a have been recognized to play crucial roles in steatosis development, acting as SIRT1 suppressors in case of palmitic acid exposure and both SIRT1 and PPAR suppressors in case of HFD exposure [114,115], as well as inhibiting β-Klotho, a co-receptor of the fibroblast growth factor 19 (FGF19), mainly involved in glucose metabolism, whose reduction was demonstrated related to inflammation, ballooning, and fibrosis [116].

3.2. Carbohydrates

Regarding carbohydrates, powerful evidence for fructose exposition exists. In fact, fructose-enriched food has been recognized as one of the most important categories of food directly involved in MAFLD onset and evolution [117]. The pleiotropic role of fructose in this setting seems to be exerted though a direct effect as substrate for DNL, increasing the acetil-CoA cellular level and acting, in a relative IR independent-manner, as an enhancer for the expression of several enzymes involved in DNL via SREBP-1c activation [118,119]. The described molecular mechanisms are involved in determining clinical outcomes for MAFLD patients and, on this purpose, the fructose-enriched drink consumption was directly related to the risk of hepatic steatosis, inflammation, and fibrosis development as well as related to extrahepatic complications onset [120,121] even if it has to be note that the highlighted effects appear to have a powerful scientific background only in case of concomitant hypercaloric dietary regimens in clinical settings [122,123].

3.3. Proteins

Differently to the previous categories of macronutrients, the role of dietary proteins in this context appears still obscure and, for certain points of view, controversial due to the low availability of data coming from well-constructed controlled trials that, thus, make it impossible to deduce scientifically coherent conclusions. In order to address the main scope of this review, we believe it correct to pay attention to the association between low dietary protein intake and the risk of sarcopenia development. In fact, sarcopenia and sarcopenic obesity properly represent well known risk factors for MAFLD progression, fueling IR, hepatic inflammation, and fibrosis as well as atherogenesis [124]. Even if the molecular mechanisms that lead to the development of sarcopenia and sarcopenic obesity are not fully clarified yet, it is not questionable that diet, physical exercise, and aging represent the most important factors of the pathogenetic cascade. If then, sarcopenia is able to have a worsened effect on the MAFLD clinical picture, is it possible to hypothesize an effect exerted by the muscle loss on the expression of genes involved in metabolic homeostasis? In this line, the role of autophagy in determining the metabolic imbalance able to trigger fat deposition and fueling muscle loss in a vicious circle seems to be crucial [125]. The generation of the autophagosome for the degradation of misfolded proteins and cellular organelles represents an efficient strategy of energy recovery and re-allocation to other biologic pathways [126]. However, the hyperactivation of this process, in particular when it occurs in mitochondria, determine their dysfunction, impairing the metabolism of the whole cell and then the organ physiology [127]. Higher levels of autophagy and mitophagy have been hypothesized to be important determinants for sarcopenia and reduced physical performance [128,129]. Sarcopenia, in turn, was related to increase in the expression of several genes involved in autophagy regulation, such as p62, autophagy related 7 (ATG7), beclin-1, and p53 [125]. Moreover, a crucial role of sarcopenia in inducing the upregulation of PPAR-α, SREBP-1, and acetyl-CoA carboxylase (ACC) was also demonstrated and related to the increase of fatty acid synthesis responsible for muscle cellular fat accumulation [125].

3.4. Micronutrients

As previously mentioned, the availability of several micronutrients could influence the DNA methylation state, regulating, then, the expression of crucial genes involved in metabolism control. It is necessary to highlight that the scientific knowledge produced about this topic are mainly based on animal experiments, and due to the lack of well-designed clinical trials, some controversies about human results exist. However, it has been shown that betaine, choline, vitamin B12, and folate could be related to MAFLD progression and their supplementation in a high-fat-sucrose- (HFS) fed mice model, acting on fatty acid synthetase (FASN) hypermethylation, improved the liver TG overload [130]. Moreover, betaine has been correlated with hypomethylation of CpG clusters of microsomal triglyceride transfer protein (MTTP), enhancing VLDL excretion from the liver; conversely, vitamin B12 and folate deficiency was demonstrated as related to mitochondrial dysfunction, impaired beta-oxidation, and TG accumulation based on peroxisome proliferator-activated receptor gamma coactivator 1-alpha hypomethylation [131].

3.5. On the Way of a Genetic Based Dietary Approach

Thanks to the actual scientific preclinical evidence, it could be reasonable to hypothesize the possibility of a genetic-based nutritional recommendation for a tailored MAFLD therapeutic management. The improvement of fatty acids oxidation, glutathione, and NAD+ availability may provide interesting strategies in this context [132].
Moreover, in a murine model, the supplementation with betaine, choline, cobalamin (vitamin B12), and folic acid in a high-fat sucrose diet regimen determined the prevention of fat accumulation through the epigenetic modulation of several genes involved in insulin sensitivity and globally lipid metabolism homeostasis [133].
Previously, highlighted in 154 NAFLD patients was also a PNPLA3 involvement in the regulation of the therapeutic response to 30 min per day of physical exercise and an individual-designed diet regimen based on the consumption of fruit, vegetables, moderate-carbohydrate, low-fat, low-glycaemic index, and low-caloric products [134].
G-allele carriers demonstrated a greater reduction of intrahepatic triglyceride content compared to those with the C-allele after 12 months of treatment together with an improvement of body weight, WHtR, total cholesterol, and LDL [134].
Due to the novelty of the topic and the weak, even if hugely promising, scientific evidence, it is not possible to transfer the results of the studies in the routine clinical practice as well as compare the therapeutic outcome of conventional general healthy diet recommendation to a genetic-based approach. Nevertheless, the charm and the power of the latter purpose represents the reason why we observe an exponential rise in the number of scientific studies on that topic and the increase of direct-to-consumer, genetically-based nutritional testing and advice by nutritional companies [135]. Globally, these signals have to be recognized as a measure of the scientific interest and expectation that society shows regarding this topic that may derive from the comprehension of the MAFLD problem worldwide not only among specific categories of healthcare professionals but also in the general population.

4. Nutrigenetics in MAFLD: Genetics Influence the Response to the Nutrients

A plethora of genetic polymorphisms contribute to the determination of a certain background influencing MAFLD by interfering with different mechanisms involved in several above-mentioned metabolic pathways [7]. In addition to these features, the genetics involvement in affecting the human biologic response to the nutrients (in terms of metabolism, requirement, and food tolerances) has been recently emerging [136]. This latter concept constitutes the study field of nutrigenetics and appears crucial in the genesis and progression of several diseases [21,137]. Consistently with the definition of MAFLD as a multifactorial disorder, the exposition to certain nutrients represents a recognized environmental factor whose interaction with a specific genetic background may lead to different clinical consequences. Paradoxically, in this scenario, a specific kind of food may be hazardous for some MAFLD patients and protective for others, depending on the presence of specific genetic variants of PNPLA3, MBOAT7, TM6SF2, and GKCR (Figure 3) [21,137].
Figure 3 proposes the mechanisms by which the genetic determinants of MAFLD influence the response to nutrients. The PNPLA3 I148M protein variant reduces the hydrolysing of n-9 PUFAs. As a consequence, the intake of foods containing n-3 PUFAs (omega-3) (above all, fishes such as salmon and tuna) ameliorates the hepatic steatosis; conversely, the assumption of n-6 (omega-6) PUFAs (present in palm oil) or n-9 derived PUFAs (contained among others in nutrients as red meat, olive oil, and almonds), through the n-6 overloading, increases the TG synthesis and promotes the production of omega-6 pro-inflammatory-derived species (as AA). Moreover, the omega-3 may downregulate SREBP1 activity while the carbohydrate feeding (particularly sugar) may up-regulate it.
The presence of TM6SF2 E167K protein variant has two apparently opposite effects: on one side, it entails lipid retention (both in hepatocytes and in enterocytes); on the other, by worsening insulin resistance and reducing VLDL secretion from liver cells, it increases hepatic steatosis. These latter effects are highlighted after high carbohydrate assumption or in case of high-fat diet regimen. Besides, in the presence of GCKR P446L protein variant, a high carbohydrate assumption may increase hepatic FAs accumulation and thus steatosis. This latter, in the presence of MBOAT7 rs641738 C>T variant, is most favoured by hyperinsulinemia, as well as the assumption of saturated FAs.

4.1. PNPLA3 rs738409 Variant’s Influence

The rs738409 C>G variant in PNPLA3 influences the response to the dietary introduction of PUFAs, particularly omega 9 (n-9) [138,139]. The n-9 PUFAs derive from n-6 PUFAs or directly from the diet (being present in meat, almonds, and several other components of the MD, such as olive oil) [140]. In case of I148M protein variant, the reduction of PNPLA3 hydrolytic efficiency decreases the hydrolysing of n-9 fatty PUFAs, leading to an n-6 overload and two critical consequences: the TG synthesis with increased hepatocyte overload and the production of n-6 pro-inflammatory-derived species (e.g., AA), worsening the steatosis and promoting NASH [141,142] (Figure 3). In contrast, in a randomized controlled trial (NCT01038102) comparing the effects of n-6 PUFAs with short FAs on liver fat and inflammation, high n-6 PUFAs intake after 10 weeks did not increase inflammation or oxidative stress [143]. Therefore, contrary to the better-known consequences on cardiovascular diseases [144], n-6 PUFAs’ effects on the liver fat content and their capability to promote inflammation still remain unclear [142,143]. Due to the controversy regarding this purpose, the construction of large clinical trials focusing attention on the genetic background based on the actual nutrigenetics and nutrigenomics knowledge could provide to the scientific community the missing key of lecture to interpret in deep the obtained results. In presence of the I149M protein variant, the consumption of n-3 PUFAs, largely contained in food as salmon and tuna, reduced TG accumulation in liver cells, thus ameliorating hepatic steatosis [21,141,145] (Figure 3). Regarding that, a previous cohort study on 127 NAFLD-obese youths’ homozygote for the G allele of the rs738409 revealed that n-3 consumption ameliorated steatosis and alanine aminotransferase (ALT) levels [141].
Moreover, in a recent randomized controlled trial (NCT01556113) on a group of twenty obese adolescents, a 12 weeks’ low n-6:n-3 PUFAs ratio (4:1) norm caloric diet mitigated fatty liver determining the reduction of ALT levels and hepatic fat fraction percentage (HFF%) assessed by multiple abdominal slices with magnetic resonance imaging [146]. Relevantly, even if the HFF% declination was observed in both CC/CG and GG PNPLA3 rs738409 genotype patients groups, it resulted significant in the GG one [146]. These findings suggest the PNPLA3 rs738409 variant capability to affect the response to this dietary intervention.
Besides PUFAs, the presence of the PNPLA3 rs738409 variant has been shown associated also with the response to carbohydrates [147]. In particular, as highlighted in the trial NCT00697580, based on a cohort of 153 Hispanic children, the carriers of the GG genotype were more susceptible to the hepatic fat overload (expressed as HFF%) when exposed to high dietary carbohydrate intake, specifically sugar [147]. Consistently, an intriguing study by Nobili et al. revealed a significant interaction between PNPLA3 I148M and the intake of sweetened beverages, showing increased hepatic fat deposition and steatosis [148]. The above-presented results appear linked to the following contrast effect: on one side, n-3 PUFAs feeding downregulates SREBP1c expression [149]; on the contrary, the carbohydrate feeding upregulates it [150]. The activation of SREBP1c/liver X receptor (LXR) pathway represents a critical moment in MAFLD development since it promotes PNPLA3 expression [150]. Relevantly, in the presence of its rs738409, C>G variant, a n-3 PUFAs-poor and carbohydrate-rich diet may increase I148M protein expression levels, thus promoting FAs overload and inhibiting, in turn, PNPLA3 degradation in a vicious circle (Figure 3).
In addition to PNPLA3, other genetic variants have been associated with a specific response to certain nutrients and dietary habits, showing important repercussions on MAFLD clinical history.

4.2. MBOAT7 rs641738 Variant’s Influence

MBOAT7’s main function is the conjugation of an acyl-CoA to the second acyl-chain of lysophospholipids, using as preferential substrate the arachidonoyl-CoA, regulating, then, the desaturation of phospholipids and the amount of free AA [151]. The rs641738 C>T variant of MBOAT7 has been associated with the development of steatosis and its complications [44,48].
In the presence of the T allele, as well as in the case of hyperinsulinemia, MBOAT7 expression (and function) results as impaired, promoting the formation of saturated phospholipids [152]. The accumulation of these latter compounds favours the synthesis of saturated and mono-unsaturated TG, promoting fat deposition in liver cells. As a consequence, in C>T variant carriers, the high intake of hyperglycaemic nutrients and food rich in saturated fatty acid may unmask MBOAT7 deficiency by increasing the predisposition to fatty liver disease and its advanced stages (Figure 3).

4.3. TM6SF2 rs58542926 and GCKR rs1260326 Variants’ Influence

The rs58542926 C>T TM6SF2 variant, encoding for the E167K protein, has been previously linked with reduced circulating lipid levels and increased hepatic steatosis due to the lipids retention (in enterocytes and hepatocytes) and the reduction of VLDL release [48,49,50]. The presence of this variant has been associated with lower post-prandial lipemia levels even after a high-fat challenge (Figure 3), as shown in a cohort of 3556 individuals enrolled in the Amish Complex Disease Research Program (ACDRP) [153]. However, besides the effects on lipid metabolism, the E167K protein variant seems to also influence the response to carbohydrate assumption and thus glucose tolerance. In line with this, Musso et al. revealed the association of TM6SF2 C>T polymorphism with increased hepatic (and adipose) IR as well as impaired pancreatic β-cell function [154]. These mechanisms, together with the abovementioned lipid retention and reduced VLDL liver secretion, may explain the TM6SF2 C>T carriers’ susceptibility towards hepatic steatosis in case of high-carbohydrate dietary regimens (Figure 3).
Finally, the common missense variant rs1260326 of GCKR gene, encoding for the P446L protein, has been shown to disrupt the regulation of glucose homeostasis by increasing glycolysis and hepatic DNL [51]. Growing evidence suggests the GCKR variant involvement to influence the response to different nutrients [155]. In this sense, Nettleton et al. revealed, in patients carrying the GCKR variant, the association between high whole-grain intake and the reduction of fasting insulin concentration [156]. Moreover, after the consumption of carbohydrate enriched (CHO) drink (75 g glucose and 25 g fructose), a group of 14 obese adolescents carriers of homozygote rs1260326 variant showed increased hepatic lipid synthesis due to the enhanced glycolytic carbon flux to TG formation [157] (Figure 3).

4.4. Personalized Nutritional Strategies Targeting Specific Genetic Backgrounds

Basing on the all above presented observations, a possible targeted therapy for the homozygous patient carriers of the PNPLA3 rs738409 variant may be based on the following three recommendations: (1) to increase the dietary intake of foods rich in n-3 PUFAs (such as salmon, tuna), (2) to follow a diet with a high n-3 PUFAs/ n-6 PUFAs ratio, and (3) to decrease the intake of sweetened food [148]. Differently, in the case of TM6SF2 rs58542926 C>T variant, a reduction of carbohydrate intake rather than lipids may result in more useful outcome [154]. In keeping, a diet based on low carbohydrate and saturated fatty acid consumption may affect the genetic susceptibility to develop fatty liver in carriers of the MBOAT7 rs641738 variant [21]. However, these promising objectives still need to be addressed via large randomized clinical trials from which it might be possible to look at the MAFLD clinical picture from another point of view, based, moreover, on the possibility to corroborate the unquestionable potential advantage that could derive from the use in the routine clinical practice of prognostic/predictive genetic scores as previously mentioned.

5. Future Perspectives: Diet, Immunity, and Genes Interaction, a Very Complex Network That Needs to Be Explored

The tangled network sustaining MAFLD pathobiology seems not only limited to the role of the caloric overload due to the dietary habit, showing extraordinary and deep interconnections among genetics, metabolism homeostasis, gut microbiome modification, and immune response modulation. In this setting, the development of westernized societies as well as low-income countries is linked to deep changes in lifestyle, characterized by sedentary life and western-type dietary regimens, which determines the irrepressible spread of immune-metabolic associated disorders, forcing governments all over the world to spend strong efforts to support the social and economic burden of disease [158,159].
Novel evidence of the recent scientific literature has been identified in the innate and adaptive immunity response triggered by and linked to FFAs-associated mitochondrial dysfunction, a crucial role in the promotion of oxidative stress, IR, and inflammation that represents the pathogenetic triad for MAFLD progression [160,161].
The close interconnection between immune reaction and cell metabolism could be considered the cornerstone for the comprehension of several immune-mediated metabolic disorders, and, in this scenario, the immunometabolic dependence of the MAFLD clinical picture seems to be compliant with the actual knowledge on this topic [10].
We are what we eat, and it is particularly true considering the interplay between dietary habits and the immune system reaction. The effect of diet composition in this context seems very wide due to its influence from cytokine production up to bone marrow immune cells generation or maturation in a specific subtype [162,163,164]. Characterized by the richness in refined sugars, salt, white flour, processed meats, purified animal fats, and food additives and the lack of fibres, vitamins, minerals, and other plant-derived molecules such as antioxidants, western diet regimens were largely explored for their possibility in influencing immune activity [165,166]. It has been well established in animal and clinical models the influence of these dietary patterns with elevated serum markers of inflammation, confirming a kind of mutual relationship between the quantity and quality of daily calories, the nutritional and adipose tissue endocrine functioning, and the immune system activity [167]. In this context, the role of cholesterol as an atherosclerosis trigger was largely investigated, acting as a key factor in macrophages’ activation and systemic inflammation induction [168,169]. Similar to cholesterol, dietary palmitate can activate the murine NLRP3 inflammasome, leading to mitochondrial dysfunction and oxidative stress generation via 5’ AMPK inhibition [170]. However, the effect of diet on immune system functioning is not only limited to a direct activity on immune cells modulating the gut microbiota composition that, in turn, influences the immune homeostasis. The diet-induced modification of gut microbiota and microbiota metabolites’ production represent the crucial events for the intestinal permeability increase, metabolic endotoxemia induction, and systemic inflammation establishment via immune scavenger receptors’ activation [171].
Conversely, dietary patterns rich in fibres, polyphenols, and natural and unprocessed food, such as in the Mediterranean dietary pattern, has been related in some clinical trials with anti-inflammatory properties, confirmed by the lowering of serum inflammatory markers, cytokines, and associated as well with reduced metabolic and cardiovascular risk, including myocardial infarction or stroke, by up to 30% [172,173]. On this line, oleate was related to the dropping of palmitate-related ER stress and apoptosis in THP-1 cells and human macrophages [174].
Omega-3 fatty acids as well can directly modulate inflammatory responses in a GPR120-dependent and -independent manner by lowering the activation of scavenger receptors like TLR2, TLR4, and tumour necrosis factor receptor [175].
Some nondigestible fibres, particularly abundant in a typical Mediterranean diet regimen, are fermented to short-chain fatty acids (SCFA): acetate, propionate, and butyrate that may act locally and systemically as immune modulators by both activating G protein-coupled receptor (GPR) 41, 43, and 109a expressed on murine intestinal cells, limiting in turn dendritic cells’ proinflammatory activity and inhibiting the differentiation of T helper 1 cells [176,177,178]. Moreover, they act directly as inducers for regulatory T cells’ differentiation and B cells’ homeostasis factors [179].
It is reasonable, thus, to hypothesize an effect derived from different dietary habits on the regulation of the immune-related genes’ expression that could fuel, together with the highlighted previously mentioned alterations of metabolic pathways, MAFLD evolution. In this context, inflammation and oxidative stress occur simultaneously, influencing each other and determining the development of the main pathogenetic determinants for the MAFLD-associated extrahepatic complications onset [180].
It has to be noticed that currently, oxidative stress represents the main determinant for adaptive immunity recruitment in MAFLD evolution, mainly triggered by oxidized phospholipids and aldehydes production linked to oxidation-specific epitopes (OSEs) generation [181,182]. It has been demonstrated a high prevalence of specific anti-OSE immunoglobulin G (IgG) production in NAFLD and NASH patients, highlighting also the direct correlation with hepatic lobular inflammation and B- and T-cell tissue infiltration [183,184].
In line with this, in preclinical models of NAFLD and NASH, it was already demonstrated the association between anti-OSE IgG titres and B-cell maturation inhibited by antioxidant treatment [183,185]. Moreover, pre-immunized malondialdehyde-acetaldehyde-adducts mice showed M1 Kupffer cells’ activation sustained by liver Th1 polarization of CD4+ T lymphocytes [183]. It was demonstrated, accordingly, a huge reduction of hepatic CD4+ regulatory T cells (Tregs) lymphocytes [186], mainly due to the induction of apoptosis, hence facilitating Th1 and Th17 proinflammatory activity [187].
Normally, the regulation of cellular cholesterol income is based on LDL-R expression on cellular surface that, in turn, undergoes an intracellular, cholesterol-level negative feedback. However, unhealthy dietary regimens characterized by high fat/sugar consumption or lifestyles such as smoking habits and low physical activity are able to determine oxidative stress imbalance. The oxidized LDL (oxLDL) are recognized by several macrophages’ scavenger receptors (CD36, scavenger receptor A- SR-A, lectin-like oxidized LDL receptor-1- LOX-1, toll-like receptors-TLRs) whose expression is not affected by intracellular cholesterol level [169,188].
OxLDL and the accumulation of crystalized cholesterol lead to macrophages’ activation, NLR family pyrin domain containing 3 (NLRP3) induction, and inflammatory cytokines genes overexpression through several different and not fully clarified mechanisms involving, in part, the recent fascinating concept of the discussed-below trained immunity (TI) pathobiology [170,188,189]. As essential component of the cellular wall, cholesterol seems to exert a dose-dependent immunomodulatory effect via LXR induction. This phenomenon, insured by a normal mitochondria biological activity, is mainly based on the generation of 25 and 27 hydroxycholesterols, physiological ligands for LXR whose activation has been correlated with anti-inflammatory effects [190]. Based on this interesting scientific evidence and due to the well-known mitochondrial dysfunction as well as the metabolic Krebs cycle MAFLD-related imbalance, the dose-dependent anti-inflammatory effects of cholesterol disappear, acquiring, then, a potentially important role in fueling hepatic inflammation [191]. SCFAs are involved in the link between dietary habits and inflammatory genetic reprogramming, acting directly on TLR4 and, consequently, triggering NF-kB or inducing intestinal permeability increase [192]. The FAs overload and adipose tissue accumulation could lead to two major consequences: altered IR-related beta-oxidation and modification of resident immune cells repertoire [193]. Normally the tissue-resident macrophages regulate the adipocytes function, maintaining the anti-inflammatory state, enhancing the production of several cytokines such as Interleukin (IL)-10, and favouring the beta-oxidation genes’ transcription [194]. In the case of chronic low-grade inflammation, a change in the specific macrophages’ subtype and their activity has been demonstrated, correlated to IL-1β, TNF-α, IL-6 genes’ transcriptions and IR worsening, generating a vicious circle able to affect hugely MAFLD [195]. Moreover, this resident immune dysfunction seems to be directly related also to the lowering of tissue-specific IL-10 producer B lymphocytes, a condition that was directly held responsible for the worsening of the main driver of liver steatosis: the IR [196]. As above described, the TI pathobiology could be another interesting and potentially useful therapeutic research field in this context [10]. OxLDL, high amounts of glucose, glycation end products, FAs, or microbiome-associated metabolites can induce an epigenetic reprogramming typically associated with a kind of innate cells’ immunological memory by which especially macrophages show a powerful inflammatory reaction potentially obtained from second aspecific antigenic contact [197]. The mammalian target of rapamycin (mTOR) pathway genes are transcriptionally upregulated in case of trained response that lasts from a few days to several months, thanks to the crosstalk between peripheral macrophages and the new cellular clone in the bone marrow that preserve the typical TI reaction despite the lack of a first antigenic contact [198]. It is, on this purpose, extremely interesting to analyse the concept of obesogenic immunological memory, hypothesized on the basis of the recent TI knowledge, to explain the lack of metabolism homeostasis recovering in case of weight loss, demonstrating also a stable, high cardiovascular risk [199,200]. Western diet, then, represents one of the most important promoters of sterile inflammation acting via NLRP3 and inducing the transcriptional rewriting of innate immune training also at the level of myeloid progenitors in bone marrow [201,202,203]. Notably, it enhances the HSCs activation in the liver, triggering inflammation and fibrogenesis as well as ensuring the progression of the disease and potentially complications onset [202,203,204]. However, even if the highlighted evidence seems to be promising because of the possible consequences in the comprehension of mechanisms sustaining MAFLD and the renewal of clinical and therapeutic management, it has to be noticed that the greater part of the actual knowledge are based on animal studies or referred to other inflammatory dysmetabolic contexts similar to but not as complex as MAFLD. Furthermore, the lack of large cohort studies or clinical trials on these topics still remains the Achilles’ heel that needs to be addressed.

6. Conclusions

Due to the worldwide spread of the metabolic pandemic and the epidemiological change of the modern hepatology, the need to pay attention to the mechanisms sustaining MAFLD and related disorders have become a diktat for future research perspectives. In light of recent scientific knowledge, diet and genetics establish the cornerstone of this complex pathologic picture for their involvement in all the different steps of the disease: development, worsening, complication onset, and therapy as well. The key for the deep and complete comprehension of the mechanisms underlying MAFLD could be referred to as diet-genes mutual relationship also in order to take advantage of designing tailored management approaches for the future of medicine. To achieve this goal, future, strong scientific efforts in designing randomized controlled trials aimed at the study of the interaction between diet and genetic factors, enrolling large cohorts of patients with long-term follow up, have to be spent.

Author Contributions

Conceptualization, M.D., M.R., and A.F.; resources, C.L., M.P., A.G.G., T.L., and L.A.; writing-original draft preparation, M.D. and M.R.; visualization, C.L., M.M., and A.F.; supervision; L.A. and M.P.; project administration, A.F. and M.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The representation of cellular structures, molecular pathways, and other elements in the figures was created using the graphic support and the available icons offered by the BioRender software (biorender.com accessed on 14 May 2021).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AAArachidonic acid
ACCAcetyl-CoA carboxylase
ACDRPAmish Complex Disease Research Program
AMPKAdenosine monophosphate-activated protein kinase
ASOantisense oligonucleotides
ATG7Autophagy related 7
CHOCarbohydrate enriched
CKDChronic kidney disease
DNLDe novo lipogenesis
FAsFatty acids
FASNFatty acid synthetase
FFAsFree fatty acids
GCKRGlucokinase regulator
GWASGenome-wide association studies
GPRG protein-coupled receptor
HCCHepatocellular carcinoma
HDCA3Histone deacetylases
HFDHigh-fat diet
HFEHomeostatic iron regulator
HFF%Hepatic fat fraction percentage
HFSHigh-fat-sucrose
HSCsHepatic stellate cells
ILInterleukin
IRInsulin resistance
LALinoleic acid
LDLLow-density lipoproteins
LDL-RLow density lipoprotein receptor
LKB1Liver kinase B1 (LKB1)
LOX-1Lectin-like oxidized LDL receptor-1
LXRLiver X receptor
MAFLDMetabolic (dysfunction) associated fatty liver disease
MARC1Mitochondrial amidoxime reducing component 1
MBOAT7Membrane-bound O-acyltransferase domain containing 7
MDMediterranean diet
MiRNAsMicroRNAs
MSMetabolic syndrome
Mt-DNAMitochondrial DNA
MT-ND6Mitochondrially encoded NADH ubiquinone oxidoreductase core subunit 6
mTORmammalian target of rapamycin
MTTPMicrosomal triglyceride transfer protein
n−3Omega 3
n−6Omega 6
n−9Omega 9
NAFLDNon-alcoholic fatty liver disease
NASNAFLD activity score
NASHNon-alcoholic steatohepatitis
NF-kBNuclear factor kappa-light-chain-enhancer of activated B cells
NLRP3NLR family pyrin domain containing 3
OxLDLoxidized low-density lipoproteins
PARVB variant 1Homo sapiens parvin beta transcript variant 1
PCSK7Proprotein convertase subtilisin/Kexin type 7
PDGF-αPlatelet derived growth factor-alpha
PGC1peroxisome proliferator-activated receptor -coactivator-1beta
PIPhosphatidylinositol
PNPLA3Phospholipase domain-containing protein-3
PPARPeroxisome proliferator-activated receptor
PPP1R3BProtein phosphatase 1 regulatory subunit 3B
PUFAsPolyunsaturated fatty acids
ROSReactive oxygen species
SCFAsshort chain fatty acids
SIRTSirtuin
SirtuinsSilent information regulator 2 proteins
SNPSingle nucleotide polymorphism
SOD 2Superoxide dismutase 2
SR-AScavenger receptor A
SREBP1Sterol regulatory element-binding proteins
SsRNAsSmall non-coding single strand RNAs
TGTriglycerides
TGF-βTransforming growth factor-beta
TITrained immunity
TLRsToll like receptors
TM6SF2Transmembrane 6 superfamily member 2 protein
TNF-αTumour necrosis factor-α
UCP2Uncoupling protein 2
UCP3Uncoupling Protein 3
UFAsUnsaturated fatty acids
UTRUntranslated region
VLDLVery-low-density lipoproteins

References

  1. Cotter, T.G.; Rinella, M. Nonalcoholic Fatty Liver Disease 2020: The State of the Disease. Gastroenterology 2020, 158, 1851–1864. [Google Scholar] [CrossRef]
  2. Lindenmeyer, C.C.; McCullough, A.J. The Natural History of Nonalcoholic Fatty Liver Disease—An Evolving View. Clin. Liver Dis. 2018, 22, 11–21. [Google Scholar] [CrossRef]
  3. 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]
  4. Eslam, M.; Sanyal, A.J.; George, J. International Consensus Panel. MAFLD: A Consensus-Driven Proposed Nomenclature for Metabolic Associated Fatty Liver Disease. Gastroenterology 2020, 158, 1999–2014. [Google Scholar] [CrossRef]
  5. Mundi, M.S.; Velapati, S.; Patel, J.; Kellogg, T.A.; Abu Dayyeh, B.K.; Hurt, R.T. Evolution of NAFLD and Its Management. Nutr. Clin. Pr. 2020, 35, 72–84. [Google Scholar] [CrossRef] [PubMed]
  6. Fang, Y.-L.; Chen, H.; Wang, C.-L.; Liang, L. Pathogenesis of non-alcoholic fatty liver disease in children and adolescence: From “two hit theory” to “multiple hit model. ” World J. Gastroenterol. 2018, 24, 2974–2983. [Google Scholar] [CrossRef]
  7. Taliento, A.E.; Dallio, M.; Federico, A.; Prati, D.; Valenti, L. Novel Insights into the Genetic Landscape of Nonalcoholic Fatty Liver Disease. Int. J. Environ. Res. Public Health 2019, 16, 2755. [Google Scholar] [CrossRef] [Green Version]
  8. Krawczyk, M.; Rau, M.; Schattenberg, J.M.; Bantel, H.; Pathil, A.; Demir, M.; Kluwe, J.; Boettler, T.; Lammert, F.; Geier, A. Combined effects of the PNPLA3 rs738409, TM6SF2 rs58542926, and MBOAT7 rs641738 variants on NAFLD severity: A multicenter biopsy-based study. J. Lipid Res. 2017, 58, 247–255. [Google Scholar] [CrossRef] [Green Version]
  9. Dallio, M.; Masarone, M.; Errico, S.; Gravina, A.G.; Nicolucci, C.; Di Sarno, R.; Gionti, L.; Tuccillo, C.; Persico, M.; Stiuso, P.; et al. Role of bisphenol A as environmental factor in the promotion of non-alcoholic fatty liver disease: In vitro and clinical study. Aliment. Pharmacol. Ther. 2018, 47, 826–837. [Google Scholar] [CrossRef] [Green Version]
  10. 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]
  11. Bibbò, S.; Ianiro, G.; Dore, M.P.; Simonelli, C.; Newton, E.E.; Cammarota, G. Gut Microbiota as a Driver of Inflammation in Nonalcoholic Fatty Liver Disease. Mediat. Inflamm. 2018, 2018, 1–7. [Google Scholar] [CrossRef] [Green Version]
  12. Nicolucci, C.; Errico, S.; Federico, A.; Dallio, M.; Loguercio, C.; Diano, N. Human exposure to Bisphenol A and liver health status: Quantification of urinary and circulating levels by LC–MS/MS. J. Pharm. Biomed. Anal. 2017, 140, 105–112. [Google Scholar] [CrossRef] [PubMed]
  13. Lama, S.; Vanacore, D.; Diano, N.; Nicolucci, C.; Errico, S.; Dallio, M.; Federico, A.; Loguercio, C.; Stiuso, P. Ameliorative effect of Silybin on bisphenol A induced oxidative stress, cell proliferation and steroid hormones oxidation in HepG2 cell cultures. Sci. Rep. 2019, 9, 3228. [Google Scholar] [CrossRef]
  14. Eslam, M.; Sarin, S.K.; Wong, V.W.-S.; Fan, J.-G.; Kawaguchi, T.; Ahn, S.H.; Zheng, M.-H.; Shiha, G.; Yilmaz, Y.; Gani, R.; et al. The Asian Pacific Association for the Study of the Liver clinical practice guidelines for the diagnosis and management of metabolic associated fatty liver disease. Hepatol. Int. 2020, 14, 889–919. [Google Scholar] [CrossRef]
  15. Dongiovanni, P.; Valenti, L. A Nutrigenomic Approach to Non-Alcoholic Fatty Liver Disease. Int. J. Mol. Sci. 2017, 18, 1534. [Google Scholar] [CrossRef] [Green Version]
  16. De Lorenzo, A.; Noce, A.; Bigioni, M.; Calabrese, V.; Della Rocca, D.; Daniele, N.; Tozzo, C.; Renzo, L. The Effects of Italian Mediterranean Organic Diet (IMOD) on Health Status. Curr. Pharm. Des. 2010, 16, 814–824. [Google Scholar] [CrossRef]
  17. Davis, C.R.; Bryan, J.; Hodgson, J.M.; Murphy, K.J. Definition of the Mediterranean Diet: A Literature Review. Nutrients 2015, 7, 9139–9153. [Google Scholar] [CrossRef]
  18. Kim, K.; Hyeon, J.; Lee, S.A.; Kwon, S.O.; Lee, H.; Keum, N.; Lee, J.; Park, S.M. Role of Total, Red, Processed, and White Meat Consumption in Stroke Incidence and Mortality: A Systematic Review and Meta-Analysis of Prospective Cohort Studies. J. Am. Hear. Assoc. 2017, 6, e005983. [Google Scholar] [CrossRef]
  19. Abenavoli, L.; Boccuto, L.; Federico, A.; Dallio, M.; Loguercio, C.; Di Renzo, L.; De Lorenzo, A. Diet and Non-Alcoholic Fatty Liver Disease: The Mediterranean Way. Int. J. Environ. Res. Public Health 2019, 16, 3011. [Google Scholar] [CrossRef] [Green Version]
  20. Dayeh, T.A.; Olsson, A.H.; Volkov, P.; Almgren, P.; Rönn, T.; Ling, C. Identification of CpG-SNPs associated with type 2 diabetes and differential DNA methylation in human pancreatic islets. Diabetologia 2013, 56, 1036–1046. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Meroni, M.; Longo, M.; Rustichelli, A.; Dongiovanni, P. Nutrition and Genetics in NAFLD: The Perfect Binomium. Int. J. Mol. Sci. 2020, 21, 2986. [Google Scholar] [CrossRef] [Green Version]
  22. Ferguson, L.R.; De Caterina, R.; Görman, U.; Allayee, H.; Kohlmeier, M.; Prasad, C.; Choi, M.S.; Curi, R.; De Luis, D.A.; Gil, Á.; et al. Guide and Position of the International Society of Nutrigenetics/Nutrigenomics on Personalised Nutrition: Part 1—Fields of Precision Nutrition. Lifestyle Genom. 2016, 9, 12–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Long, M.T.; Gurary, E.B.; Massaro, J.M.; Ma, J.; Hoffmann, U.; Chung, R.T.; Benjamin, E.J.; Loomba, R. Parental non-alcoholic fatty liver disease increases risk of non-alcoholic fatty liver disease in offspring. Liver Int. 2019, 39, 740–747. [Google Scholar] [CrossRef]
  24. Eslam, M.; Valenti, L.; Romeo, S. Genetics and epigenetics of NAFLD and NASH: Clinical impact. J. Hepatol. 2018, 68, 268–279. [Google Scholar] [CrossRef]
  25. Sánchez-Torrijos, Y.; Herrojo, J.A.; Palacios, D.P.; Gallego-Durán, R.; Romero-Gómez, M.; Ampuero, J. Analysis of the burden and variability in the management of NAFLD patients in the clinical practice: Unifying the required criteria. Rev. Española Enferm. Dig. 2019, 111, 270–274. [Google Scholar] [CrossRef]
  26. Rinella, M.E.; Tacke, F.; Sanyal, A.J.; Anstee, Q.M.; the Participants of the AASLD/EASL Workshop. Report on the AASLD/EASL Joint Workshop on Clinical Trial Endpoints in NAFLD. Hepatology 2019, 70, 1424–1436. [Google Scholar] [CrossRef]
  27. Clarke, J.D.; Cherrington, N.J. Nonalcoholic steatohepatitis in precision medicine: Unraveling the factors that contribute to individual variability. Pharmacol. Ther. 2015, 151, 99–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Trépo, E.; Valenti, L. Update on NAFLD genetics: From new variants to the clinic. J. Hepatol. 2020, 72, 1196–1209. [Google Scholar] [CrossRef]
  29. Choudhary, N.S.; Duseja, A. Genetic and epigenetic disease modifiers: Non-alcoholic fatty liver disease (NAFLD) and alcoholic liver disease (ALD). Transl. Gastroenterol. Hepatol. 2021, 6, 2. [Google Scholar] [CrossRef] [PubMed]
  30. Li, Y.-Y. Genetic and epigenetic variants influencing the development of nonalcoholic fatty liver disease. World J. Gastroenterol. 2012, 18, 6546–6551. [Google Scholar] [CrossRef]
  31. Romeo, S.; Kozlitina, J.; Xing, C.; Pertsemlidis, A.; Cox, D.; A Pennacchio, L.; 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] [PubMed] [Green Version]
  32. Pingitore, P.; Pirazzi, C.; Mancina, R.M.; Motta, B.M.; Indiveri, C.; Pujia, A.; Montalcini, T.; Hedfalk, K.; Romeo, S. Recombinant PNPLA3 protein shows triglyceride hydrolase activity and its I148M mutation results in loss of function. Biochim. Biophys. Acta (BBA) Mol. Cell Biol. Lipids 2014, 1841, 574–580. [Google Scholar] [CrossRef] [Green Version]
  33. Rae-Whitcombe, S.M.; Kennedy, D.; Voyles, M.; Thompson, M.P. Regulation of the promoter region of the human adiponutrin/PNPLA3 gene by glucose and insulin. Biochem. Biophys. Res. Commun. 2010, 402, 767–772. [Google Scholar] [CrossRef]
  34. BasuRay, S.; Wang, Y.; Smagris, E.; Cohen, J.C.; Hobbs, H.H. Accumulation of PNPLA3 on lipid droplets is the basis of associated hepatic steatosis. Proc. Natl. Acad. Sci. USA 2019, 116, 9521–9526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Mondul, A.; Mancina, R.M.; Merlo, A.; Dongiovanni, P.; Rametta, R.; Montalcini, T.; Valenti, L.; Albanes, D.; Romeo, S. PNPLA3 I148M Variant Influences Circulating Retinol in Adults with Nonalcoholic Fatty Liver Disease or Obesity. J. Nutr. 2015, 145, 1687–1691. [Google Scholar] [CrossRef] [Green Version]
  36. Haaker, M.W.; Vaandrager, A.B.; Helms, J.B. Retinoids in health and disease: A role for hepatic stellate cells in affecting retinoid levels. Biochim. Biophys. Acta (BBA) Mol. Cell Biol. Lipids 2020, 1865, 158674. [Google Scholar] [CrossRef] [PubMed]
  37. Higashi, T.; Friedman, S.L.; Hoshida, Y. Hepatic stellate cells as key target in liver fibrosis. Adv. Drug Deliv. Rev. 2017, 121, 27–42. [Google Scholar] [CrossRef] [PubMed]
  38. Sookoian, S.; Pirola, C.J. Meta-analysis of the influence of I148M variant of patatin-like phospholipase domain containing 3 gene (PNPLA3) on the susceptibility and histological severity of nonalcoholic fatty liver disease. Hepatology 2011, 53, 1883–1894. [Google Scholar] [CrossRef] [PubMed]
  39. Valenti, L.; Al-Serri, A.; Daly, A.K.; Galmozzi, E.; Rametta, R.; Dongiovanni, P.; Nobili, V.; Mozzi, E.; Roviaro, G.; Vanni, E.; et al. Homozygosity for the patatin-like phospholipase-3/adiponutrin I148M polymorphism influences liver fibrosis in patients with nonalcoholic fatty liver disease. Hepatology 2010, 51, 1209–1217. [Google Scholar] [CrossRef] [PubMed]
  40. Musso, G.; Cassader, M.; Gambino, R. PNPLA3 rs738409 and TM6SF2 rs58542926 gene variants affect renal disease and function in nonalcoholic fatty liver disease. Hepatology 2015, 62, 658–659. [Google Scholar] [CrossRef]
  41. Murphy, R.C.; Folco, G. Lysophospholipid acyltransferases and leukotriene biosynthesis: Intersection of the Lands cycle and the arachidonate PI cycle. J. Lipid Res. 2019, 60, 219–226. [Google Scholar] [CrossRef] [Green Version]
  42. Mancina, R.M.; Dongiovanni, P.; Petta, S.; Pingitore, P.; Meroni, M.; Rametta, R.; Borén, J.; Montalcini, T.; Pujia, A.; Wiklund, O.; et al. The MBOAT7-TMC4 Variant rs641738 Increases Risk of Nonalcoholic Fatty Liver Disease in Individuals of European Descent. Gastroenterology 2016, 150, 1219–1230. [Google Scholar] [CrossRef] [Green Version]
  43. Luukkonen, P.K.; Zhou, Y.; Hyötyläinen, T.; Leivonen, M.; Arola, J.; Orho-Melander, M.; Orešič, M.; Yki-Järvinen, H. The MBOAT7 variant rs641738 alters hepatic phosphatidylinositols and increases severity of non-alcoholic fatty liver disease in humans. J. Hepatol. 2016, 65, 1263–1265. [Google Scholar] [CrossRef] [Green Version]
  44. Donati, B.; Dongiovanni, P.; Romeo, S.; Meroni, M.; McCain, M.; Miele, L.; Petta, S.; Maier, S.; Rosso, C.; De Luca, L.; et al. MBOAT7 rs641738 variant and hepatocellular carcinoma in non-cirrhotic individuals. Sci. Rep. 2017, 7, 1–10. [Google Scholar] [CrossRef] [Green Version]
  45. Kozlitina, J.; Smagris, E.; Stender, S.; Nordestgaard, B.G.; Zhou, H.H.; Tybjærg-Hansen, A.; Vogt, T.F.; Hobbs, H.H.; Cohen, J.C. Exome-wide association study identifies a TM6SF2 variant that confers susceptibility to nonalcoholic fatty liver disease. Nat. Genet. 2014, 46, 352–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Mahdessian, H.; Taxiarchis, A.; Popov, S.; Silveira, A.; Franco-Cereceda, A.; Hamsten, A.; Eriksson, P.; Hooft, F.V. 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]
  47. Liu, Y.-L.; Reeves, H.L.; Burt, A.D.; Tiniakos, D.G.; McPherson, S.W.; Leathart, J.B.S.; Allison, M.E.D.; Alexander, G.J.; Piguet, A.-C.; Anty, R.; et al. TM6SF2 rs58542926 influences hepatic fibrosis progression in patients with non-alcoholic fatty liver disease. Nat. Commun. 2014, 5, 4309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Luukkonen, P.K.; Zhou, Y.; Haridas, P.N.; Dwivedi, O.P.; Hyötyläinen, T.; Ali, A.; Juuti, A.; Leivonen, M.; Tukiainen, T.; Ahonen, L.; et al. Impaired hepatic lipid synthesis from polyunsaturated fatty acids in TM6SF2 E167K variant carriers with NAFLD. J. Hepatol. 2017, 67, 128–136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Kim, D.S.; Jackson, A.U.; Li, Y.K.; Stringham, H.M.; Kuusisto, J.; Kangas, A.J.; Soininen, P.; Ala-Korpela, M.; Burant, C.F.; Salomaa, V.; et al. Novel association of TM6SF2 rs58542926 genotype with increased serum tyrosine levels and decreased apoB-100 particles in Finns. J. Lipid Res. 2017, 58, 1471–1481. [Google Scholar] [CrossRef] [Green Version]
  50. 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] [PubMed]
  51. Beer, N.L.; Tribble, N.D.; McCulloch, L.J.; Roos, C.; Johnson, P.R.; Orho-Melander, M.; Gloyn, A.L. The P446L variant in GCKR associated with fasting plasma glucose and triglyceride levels exerts its effect through increased glucokinase activity in liver. Hum. Mol. Genet. 2009, 18, 4081–4088. [Google Scholar] [CrossRef] [Green Version]
  52. Dongiovanni, P.; Meroni, M.; Mancina, R.M.; Baselli, G.; Rametta, R.; Pelusi, S.; Männistö, V.; Fracanzani, A.L.; Badiali, S.; Miele, L.; et al. Protein phosphatase 1 regulatory subunit 3B gene variation protects against hepatic fat accumulation and fibrosis in individuals at high risk of nonalcoholic fatty liver disease. Hepatol. Commun. 2018, 2, 666–675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Masarone, M.; Rosato, V.; Dallio, M.; Gravina, A.G.; Aglitti, A.; Loguercio, C.; Federico, A.; Persico, M. Role of Oxidative Stress in Pathophysiology of Nonalcoholic Fatty Liver Disease. Oxidative Med. Cell. Longev. 2018, 2018, 1–14. [Google Scholar] [CrossRef] [PubMed]
  54. 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]
  55. Namikawa, C.; Shu-Ping, Z.; Vyselaar, J.R.; Nozaki, Y.; Nemoto, Y.; Ono, M.; Akisawa, N.; Saibara, T.; Hiroi, M.; Enzan, H.; et al. Polymorphisms of microsomal triglyceride transfer protein gene and manganese superoxide dismutase gene in non-alcoholic steatohepatitis. J. Hepatol. 2004, 40, 781–786. [Google Scholar] [CrossRef]
  56. Macaluso, F.S.; Maida, M.; Petta, S. Genetic background in nonalcoholic fatty liver disease: A comprehensive review. World J. Gastroenterol. 2015, 21, 11088–11111. [Google Scholar] [CrossRef] [PubMed]
  57. Valenti, L.; Fracanzani, A.L.; Bugianesi, E.; Dongiovanni, P.; Galmozzi, E.; Vanni, E.; Canavesi, E.; Lattuada, E.; Roviaro, G.; Marchesini, G.; et al. HFE Genotype, Parenchymal Iron Accumulation, and Liver Fibrosis in Patients With Nonalcoholic Fatty Liver Disease. Gastroenterology 2010, 138, 905–912. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Cadenas, S. Mitochondrial uncoupling, ROS generation and cardioprotection. Biochim. Biophys. Acta (BBA) Bioenerg. 2018, 1859, 940–950. [Google Scholar] [CrossRef] [PubMed]
  59. Aller, R.; A De Luis, D.; Izaola, O.; Sagrado, M.G.; Conde, R.; Alvarez, T.; Pacheco, D.; Velasco, M.C. Role of -55CT polymorphism of UCP3 gene on non alcoholic fatty liver disease and insulin resistance in patients with obesity. Nutr. Hosp. 2010, 25, 572–576. [Google Scholar]
  60. Fares, R.; Petta, S.; Lombardi, R.; Grimaudo, S.; Dongiovanni, P.; Pipitone, R.; Rametta, R.; Fracanzani, A.L.; Mozzi, E.; Craxì, A.; et al. The UCP2 -866 G>A promoter region polymorphism is associated with nonalcoholic steatohepatitis. Liver Int. 2015, 35, 1574–1580. [Google Scholar] [CrossRef]
  61. Dongiovanni, P.; Meroni, M.; Baselli, G.; Mancina, R.M.; Ruscica, M.; Longo, M.; Rametta, R.; Cespiati, A.; Pelusi, S.; Ferri, N.; et al. PCSK7 gene variation bridges atherogenic dyslipidemia with hepatic inflammation in NAFLD patients. J. Lipid Res. 2019, 60, 1144–1153. [Google Scholar] [CrossRef] [PubMed]
  62. 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-TruncatingHSD17B13Variant and Protection from Chronic Liver Disease. N. Engl. J. Med. 2018, 378, 1096–1106. [Google Scholar] [CrossRef] [PubMed]
  63. Su, W.; Mao, Z.; Liu, Y.; Zhang, X.; Zhang, W.; Gustafsson, J.-A.; Guan, Y. Role of HSD17B13 in the liver physiology and pathophysiology. Mol. Cell. Endocrinol. 2019, 489, 119–125. [Google Scholar] [CrossRef] [PubMed]
  64. Torkamani, A.; Wineinger, N.E.; Topol, E.J. The personal and clinical utility of polygenic risk scores. Nat. Rev. Genet. 2018, 19, 581–590. [Google Scholar] [CrossRef] [PubMed]
  65. Trépo, E.; Romeo, S.; Zucman-Rossi, J.; Nahon, P. PNPLA3 gene in liver diseases. J. Hepatol. 2016, 65, 399–412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. European Association for the Study of The Liver; European Association for the Study of Diabetes. EASL-EASD-EASO Clinical Practice Guidelines for the management of non-alcoholic fatty liver disease. J. Hepatol. 2016, 64, 1388–1402. [Google Scholar] [CrossRef]
  67. Lee, A.; Mavaddat, N.; Wilcox, A.N.; Msc, A.P.C.; Carver, T.; Hartley, S.; De Villiers, C.B.; Izquierdo, A.; Simard, J.; Schmidt, M.K.; et al. BOADICEA: A comprehensive breast cancer risk prediction model incorporating genetic and nongenetic risk factors. Genet. Med. 2019, 21, 1708–1718. [Google Scholar] [CrossRef] [Green Version]
  68. Pelusi, S.; Baselli, G.; Pietrelli, A.; Dongiovanni, P.; Donati, B.; McCain, M.V.; Meroni, M.; Fracanzani, A.L.; Romagnoli, R.; Petta, S.; et al. Rare Pathogenic Variants Predispose to Hepatocellular Carcinoma in Nonalcoholic Fatty Liver Disease. Sci. Rep. 2019, 9, 1–10. [Google Scholar] [CrossRef]
  69. Bianco, C.; Jamialahmadi, O.; Pelusi, S.; Baselli, G.; Dongiovanni, P.; Zanoni, I.; Santoro, L.; Maier, S.; Liguori, A.; Meroni, M.; et al. Non-invasive stratification of hepatocellular carcinoma risk in non-alcoholic fatty liver using polygenic risk scores. J. Hepatol. 2021, 74, 775–782. [Google Scholar] [CrossRef]
  70. Guzman, C.B.; Duvvuru, S.; Akkari, A.; Bhatnagar, P.; Battioui, C.; Foster, W.; Zhang, X.M.; Shankar, S.S.; Deeg, M.A.; Chalasani, N.; et al. Coding variants in PNPLA3 and TM6SF2 are risk factors for hepatic steatosis and elevated serum alanine aminotransferases caused by a glucagon receptor antagonist. Hepatol. Commun. 2018, 2, 561–570. [Google Scholar] [CrossRef]
  71. Scorletti, E.; Bhatia, L.; McCormick, K.; Clough, G.; Nash, K.; Calder, P.; Byrne, C.D. Design and rationale of the WELCOME trial: A randomised, placebo controlled study to test the efficacy of purified long chain omega-3 fatty treatment in non-alcoholic fatty liver disease. Contemp. Clin. Trials 2014, 37, 301–311. [Google Scholar] [CrossRef]
  72. Sun, C.; Fan, J.-G.; Qiao, L. Potential Epigenetic Mechanism in Non-Alcoholic Fatty Liver Disease. Int. J. Mol. Sci. 2015, 16, 5161–5179. [Google Scholar] [CrossRef]
  73. Hotta, K.; Kitamoto, A.; Kitamoto, T.; Ogawa, Y.; Honda, Y.; Kessoku, T.; Yoneda, M.; Imajo, K.; Tomeno, W.; Saito, S.; et al. Identification of differentially methylated region (DMR) networks associated with progression of nonalcoholic fatty liver disease. Sci. Rep. 2018, 8, 13567. [Google Scholar] [CrossRef] [PubMed]
  74. Kitamoto, T.; Kitamoto, A.; Ogawa, Y.; Honda, Y.; Imajo, K.; Saito, S.; Yoneda, M.; Nakamura, T.; Nakajima, A.; Hotta, K. Targeted-bisulfite sequence analysis of the methylation of CpG islands in genes encoding PNPLA3, SAMM50, and PARVB of patients with non-alcoholic fatty liver disease. J. Hepatol. 2015, 63, 494–502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Murphy, S.K.; Yang, H.; Moylan, C.A.; Pang, H.; Dellinger, A.; Abdelmalek, M.F.; Garrett, M.E.; Ashley–Koch, A.; Suzuki, A.; Tillmann, H.L.; et al. Relationship between Methylome and Transcriptome in Patients with Nonalcoholic Fatty Liver Disease. Gastroenterology 2013, 145, 1076–1087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Zeybel, M.; Hardy, T.; Robinson, S.M.; Fox, C.; Anstee, Q.M.; Ness, T.; Masson, S.; Mathers, J.C.; French, J.; White, S.; et al. Differential DNA methylation of genes involved in fibrosis progression in non-alcoholic fatty liver disease and alcoholic liver disease. Clin. Epigenetics 2015, 7, 25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Li, Y.Y.; Tang, D.; Du, Y.L.; Cao, C.Y.; Nie, Y.Q.; Cao, J.; Zhou, Y.J. Fatty liver mediated by peroxisome proliferator-activated receptor-α DNA methylation can be reversed by a methylation inhibitor and curcumin. J. Dig. Dis. 2018, 19, 421–430. [Google Scholar] [CrossRef]
  78. Chen, H.-C.; Chen, Y.-Z.; Wang, C.-H.; Lin, F.-J. The nonalcoholic fatty liver disease-like phenotype and lowered serum VLDL are associated with decreased expression and DNA hypermethylation of hepatic ApoB in male offspring of ApoE deficient mothers fed a with Western diet. J. Nutr. Biochem. 2020, 77, 108319. [Google Scholar] [CrossRef]
  79. Tyagi, S.; Sharma, S.; Gupta, P.; Saini, A.S.; Kaushal, C. The peroxisome proliferator-activated receptor: A family of nuclear receptors role in various diseases. J. Adv. Pharm. Technol. Res. 2011, 2, 236–240. [Google Scholar] [CrossRef]
  80. De Mello, V.D.; Matte, A.; Perfilyev, A.; Männistö, V.; Rönn, T.; Nilsson, E.; Käkelä, P.; Ling, C.; Pihlajamäki, J. Human liver epigenetic alterations in non-alcoholic steatohepatitis are related to insulin action. Epigenetics 2017, 12, 287–295. [Google Scholar] [CrossRef] [Green Version]
  81. de Mello, V.D.; Sehgal, R.; Männistö, V.; Klåvus, A.; Nilsson, E.; Perfilyev, A.; Kaminska, D.; Miao, Z.; Pajukanta, P.; Ling, C.; et al. Serum aromatic and branched-chain amino acids associated with NASH demonstrate divergent associations with serum lipids. Liver Int. 2021, 41, 754–763. [Google Scholar] [CrossRef]
  82. Saran, A.R.; Dave, S.; Zarrinpar, A. Circadian Rhythms in the Pathogenesis and Treatment of Fatty Liver Disease. Gastroenterology 2020, 158, 1948–1966.e1. [Google Scholar] [CrossRef] [PubMed]
  83. Mazzoccoli, G.; Vinciguerra, M.; Oben, J.; Tarquini, R.; De Cosmo, S. Non-alcoholic fatty liver disease: The role of nuclear receptors and circadian rhythmicity. Liver Int. 2014, 34, 1133–1152. [Google Scholar] [CrossRef] [PubMed]
  84. Nassir, F.; A Ibdah, J. Sirtuins and nonalcoholic fatty liver disease. World J. Gastroenterol. 2016, 22, 10084–10092. [Google Scholar] [CrossRef]
  85. Purushotham, A.; Schug, T.T.; Xu, Q.; Surapureddi, S.; Guo, X.; Li, X. Hepatocyte-Specific Deletion of SIRT1 Alters Fatty Acid Metabolism and Results in Hepatic Steatosis and Inflammation. Cell Metab. 2009, 9, 327–338. [Google Scholar] [CrossRef] [Green Version]
  86. Pirola, C.J.; Garaycoechea, M.; Flichman, D.; O Castaño, G.; Sookoian, S. Liver mitochondrial DNA damage and genetic variability of Cytochrome b—A key component of the respirasome—Drive the severity of fatty liver disease. J. Intern. Med. 2021, 289, 84–96. [Google Scholar] [CrossRef] [PubMed]
  87. Pirola, C.J.; Gianotti, T.F.; Burgueño, A.L.; Rey-Funes, M.; Loidl, C.F.; Mallardi, P.; Martino, J.S.; Castaño, G.O.; Sookoian, S. Epigenetic modification of liver mitochondrial DNA is associated with histological severity of nonalcoholic fatty liver disease. Gut 2013, 62, 1356–1363. [Google Scholar] [CrossRef] [PubMed]
  88. Kim, H.-S.; Patel, K.; Muldoon-Jacobs, K.; Bisht, K.S.; Aykin-Burns, N.; Pennington, J.D.; van der Meer, R.; Nguyen, P.; Savage, J.; Owens, K.M.; et al. SIRT3 Is a Mitochondria-Localized Tumor Suppressor Required for Maintenance of Mitochondrial Integrity and Metabolism during Stress. Cancer Cell 2010, 17, 41–52. [Google Scholar] [CrossRef] [Green Version]
  89. Hirschey, M.D.; Shimazu, T.; Jing, E.; Grueter, C.A.; Collins, A.M.; Aouizerat, B.; Stančáková, A.; Goetzman, E.; Lam, M.M.; Schwer, B.; et al. SIRT3 Deficiency and Mitochondrial Protein Hyperacetylation Accelerate the Development of the Metabolic Syndrome. Mol. Cell 2011, 44, 177–190. [Google Scholar] [CrossRef] [Green Version]
  90. Yoshino, J.; Imai, S.-I. Mitochondrial SIRT3: A New Potential Therapeutic Target for Metabolic Syndrome. Mol. Cell 2011, 44, 170–171. [Google Scholar] [CrossRef] [Green Version]
  91. Nguyen, T.B.; Olzmann, J.A. Lipid droplets and lipotoxicity during autophagy. Autophagy 2017, 13, 2002–2003. [Google Scholar] [CrossRef] [Green Version]
  92. Finch, M.L.; Marquardt, J.U.; Yeoh, G.C.; Callus, B.A. Regulation of microRNAs and their role in liver development, regeneration and disease. Int. J. Biochem. Cell Biol. 2014, 54, 288–303. [Google Scholar] [CrossRef] [Green Version]
  93. Lu, T.X.; Rothenberg, M.E. MicroRNA. J. Allergy Clin. Immunol. 2018, 141, 1202–1207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Long, J.-K.; Dai, W.; Zheng, Y.-W.; Zhao, S.-P. miR-122 promotes hepatic lipogenesis via inhibiting the LKB1/AMPK pathway by targeting Sirt1 in non-alcoholic fatty liver disease. Mol. Med. 2019, 25, 26. [Google Scholar] [CrossRef] [Green Version]
  95. Esau, C.; Davis, S.; Murray, S.F.; Yu, X.X.; Pandey, S.K.; Pear, M.; Watts, L.; Booten, S.L.; Graham, M.; McKay, R.; et al. miR-122 regulation of lipid metabolism revealed by in vivo antisense targeting. Cell Metab. 2006, 3, 87–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Salvoza, N.C.; Klinzing, D.C.; Gopez-Cervantes, J.; Baclig, M.O. Association of Circulating Serum miR-34a and miR-122 with Dyslipidemia among Patients with Non-Alcoholic Fatty Liver Disease. PLoS ONE 2016, 11, e0153497. [Google Scholar] [CrossRef]
  97. Davalos, A.; Goedeke, L.; Smibert, P.; Ramirez, C.M.; Warrier, N.P.; Andreo, U.; Salinas, D.C.; Rayner, K.; Suresh, U.; Pastor-Pareja, J.C.; et al. miR-33a/b contribute to the regulation of fatty acid metabolism and insulin signaling. Proc. Natl. Acad. Sci. USA 2011, 108, 9232–9237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Erhartova, D.; Cahova, M.; Dankova, H.; Heczkova, M.; Mikova, I.; Sticova, E.; Spicak, J.; Seda, O.; Trunecka, P. Serum miR-33a is associated with steatosis and inflammation in patients with non-alcoholic fatty liver disease after liver transplantation. PLoS ONE 2019, 14, e0224820. [Google Scholar] [CrossRef] [PubMed]
  99. Gjorgjieva, M.; Sobolewski, C.; Dolicka, D.; De Sousa, M.C.; Foti, M. miRNAs and NAFLD: From pathophysiology to therapy. Gut 2019, 68, 2065–2079. [Google Scholar] [CrossRef] [PubMed]
  100. Pillai, S.S.; Lakhani, H.V.; Zehra, M.; Wang, J.; Dilip, A.; Puri, N.; O’Hanlon, K.; Sodhi, K. Predicting Nonalcoholic Fatty Liver Disease through a Panel of Plasma Biomarkers and MicroRNAs in Female West Virginia Population. Int. J. Mol. Sci. 2020, 21, 6698. [Google Scholar] [CrossRef]
  101. Hur, W.; Lee, J.H.; Kim, S.W.; Kim, J.-H.; Bae, S.H.; Kim, M.; Hwang, D.; Kim, Y.S.; Park, T.; Um, S.-J.; et al. Downregulation of microRNA-451 in non-alcoholic steatohepatitis inhibits fatty acid-induced proinflammatory cytokine production through the AMPK/AKT pathway. Int. J. Biochem. Cell Biol. 2015, 64, 265–276. [Google Scholar] [CrossRef] [Green Version]
  102. Liu, X.-L.; Pan, Q.; Zhang, R.-N.; Shen, F.; Yan, S.-Y.; Sun, C.; Xu, Z.-J.; Chen, Y.-W.; Fan, J.-G. Disease-specific miR-34a as diagnostic marker of non-alcoholic steatohepatitis in a Chinese population. World J. Gastroenterol. 2016, 22, 9844–9852. [Google Scholar] [CrossRef]
  103. Perakakis, N.; Stefanakis, K.; Mantzoros, C.S. The role of omics in the pathophysiology, diagnosis and treatment of non-alcoholic fatty liver disease. Metabolism 2020, 111, 154320. [Google Scholar] [CrossRef]
  104. Mullins, V.A.; Bresette, W.; Johnstone, L.; Hallmark, B.; Chilton, F.H. Genomics in Personalized Nutrition: Can You “Eat for Your Genes”? Nutrients 2020, 12, 3118. [Google Scholar] [CrossRef] [PubMed]
  105. Blasbalg, T.L.; Hibbeln, J.; Ramsden, C.E.; Majchrzak, S.F.; Rawlings, R.R. Changes in consumption of omega-3 and omega-6 fatty acids in the United States during the 20th century. Am. J. Clin. Nutr. 2011, 93, 950–962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Serhan, C.N.; Chiang, N.; Van Dyke, T.E. Resolving inflammation: Dual anti-inflammatory and pro-resolution lipid mediators. Nat. Rev. Immunol. 2008, 8, 349–361. [Google Scholar] [CrossRef] [Green Version]
  107. Lottenberg, A.M.; Afonso, M.D.S.; Lavrador, M.S.F.; Machado, R.M.; Nakandakare, E.R. The role of dietary fatty acids in the pathology of metabolic syndrome. J. Nutr. Biochem. 2012, 23, 1027–1040. [Google Scholar] [CrossRef] [PubMed]
  108. Jump, D.B.; Lytle, K.A.; Depner, C.M.; Tripathy, S. Omega-3 polyunsaturated fatty acids as a treatment strategy for nonalcoholic fatty liver disease. Pharmacol. Ther. 2018, 181, 108–125. [Google Scholar] [CrossRef]
  109. Levy, J.R.; Clore, J.N.; Stevens, W. Dietary n-3 polyunsaturated fatty acids decrease hepatic triglycerides in Fischer 344 rats. Hepatology 2004, 39, 608–616. [Google Scholar] [CrossRef]
  110. Campos, H.; Baylin, A.; Willett, W.C. α-Linolenic Acid and Risk of Nonfatal Acute Myocardial Infarction. Circulation 2008, 118, 339–345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Serra, M.C.; Ryan, A.S.; Hafer-Macko, C.E.; Yepes, M.; Nahab, F.B.; Ziegler, T.R. Dietary and Serum Omega-6/Omega-3 Fatty Acids Are Associated with Physical and Metabolic Function in Stroke Survivors. Nutrients 2020, 12, 701. [Google Scholar] [CrossRef] [Green Version]
  112. Lankinen, M.A.; Fauland, A.; Shimizu, B.-I.; Ågren, J.; E Wheelock, C.; Laakso, M.; Schwab, U.; Pihlajamäki, J. Inflammatory response to dietary linoleic acid depends on FADS1 genotype. Am. J. Clin. Nutr. 2019, 109, 165–175. [Google Scholar] [CrossRef]
  113. Lankinen, M.A.; de Mello, V.D.; Meuronen, T.; Sallinen, T.; Ågren, J.; Virtanen, K.A.; Laakso, M.; Pihlajamäki, J.; Schwab, U. The FADS1 Genotype Modifies Metabolic Responses to the Linoleic Acid and Alpha-linolenic Acid Containing Plant Oils–Genotype Based Randomized Trial FADSDIET2. Mol. Nutr. Food Res. 2021, 65, e2001004. [Google Scholar] [CrossRef] [PubMed]
  114. Ding, J.; Li, M.; Wan, X.; Jin, X.; Chen, S.; Yu, C.; Li, Y. Effect of miR-34a in regulating steatosis by targeting PPARα expression in nonalcoholic fatty liver disease. Sci. Rep. 2015, 5, 13729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Wang, Y.; Zhu, K.; Yu, W.; Wang, H.; Liu, L.; Wu, Q.; Li, S.; Guo, J. MiR-181b regulates steatosis in nonalcoholic fatty liver disease via targeting SIRT1. Biochem. Biophys. Res. Commun. 2017, 493, 227–232. [Google Scholar] [CrossRef] [PubMed]
  116. Dongiovanni, P.; Crudele, A.; Panera, N.; Romito, I.; Meroni, M.; De Stefanis, C.; Palma, A.; Comparcola, D.; Fracanzani, A.L.; Miele, L.; et al. β-Klotho gene variation is associated with liver damage in children with NAFLD. J. Hepatol. 2020, 72, 411–419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Schwarz, J.-M.; Noworolski, S.M.; Wen, M.J.; Dyachenko, A.; Prior, J.L.; Weinberg, M.E.; Herraiz, L.A.; Tai, V.W.; Bergeron, N.; Bersot, T.P.; et al. Effect of a High-Fructose Weight-Maintaining Diet on Lipogenesis and Liver Fat. J. Clin. Endocrinol. Metab. 2015, 100, 2434–2442. [Google Scholar] [CrossRef]
  118. Jornayvaz, F.R.; Shulman, G.I. Diacylglycerol Activation of Protein Kinase Cε and Hepatic Insulin Resistance. Cell Metab. 2012, 15, 574–584. [Google Scholar] [CrossRef] [Green Version]
  119. Santos-Marcos, J.A.; Perez-Jimenez, F.; Camargo, A. The role of diet and intestinal microbiota in the development of metabolic syndrome. J. Nutr. Biochem. 2019, 70, 1–27. [Google Scholar] [CrossRef]
  120. Ma, J.; Fox, C.S.; Jacques, P.F.; Speliotes, E.K.; Hoffmann, U.; Smith, C.E.; Saltzman, E.; McKeown, N.M. Sugar-sweetened beverage, diet soda, and fatty liver disease in the Framingham Heart Study cohorts. J. Hepatol. 2015, 63, 462–469. [Google Scholar] [CrossRef] [Green Version]
  121. Abdelmalek, M.F.; Suzuki, A.; Guy, C.D.; Unalp-Arida, A.; Colvin, R.; Johnson, R.J.; Diehl, A.M.; for the Nonalcoholic Steatohepatitis Clinical Research Network. Increased fructose consumption is associated with fibrosis severity in patients with nonalcoholic fatty liver disease. Hepatology 2010, 51, 1961–1971. [Google Scholar] [CrossRef] [Green Version]
  122. Haufe, S.; Engeli, S.; Kast, P.; Böhnke, J.; Utz, W.; Haas, V.; Hermsdorf, M.; Mähler, A.; Wiesner, S.; Birkenfeld, A.L.; et al. Randomized comparison of reduced fat and reduced carbohydrate hypocaloric diets on intrahepatic fat in overweight and obese human subjects. Hepatology 2011, 53, 1504–1514. [Google Scholar] [CrossRef] [PubMed]
  123. Volynets, V.; Machann, J.; Küper, M.A.; Maier, I.B.; Spruss, A.; Königsrainer, A.; Bischoff, S.C.; Bergheim, I. A moderate weight reduction through dietary intervention decreases hepatic fat content in patients with non-alcoholic fatty liver disease (NAFLD): A pilot study. Eur. J. Nutr. 2013, 52, 527–535. [Google Scholar] [CrossRef]
  124. Lee, Y.-H.; Jung, K.S.; Kim, S.U.; Yoon, H.-J.; Yun, Y.J.; Lee, B.-W.; Kang, E.S.; Han, K.-H.; Lee, H.C.; Cha, B.-S. Sarcopaenia is associated with NAFLD independently of obesity and insulin resistance: Nationwide surveys (KNHANES 2008–2011). J. Hepatol. 2015, 63, 486–493. [Google Scholar] [CrossRef]
  125. Wu, H.; Jang, J.; Dridi, S.; Ferrando, A.A.; Wolfe, R.R.; Kim, I.; Baum, J.I. Net protein balance correlates with expression of autophagy, mitochondrial biogenesis, and fat metabolism-related genes in skeletal muscle from older adults. Physiol. Rep. 2020, 8. [Google Scholar] [CrossRef] [PubMed]
  126. Sandri, M. Autophagy in skeletal muscle. FEBS Lett. 2010, 584, 1411–1416. [Google Scholar] [CrossRef] [PubMed]
  127. Kim, I.; Rodriguez-Enriquez, S.; Lemasters, J.J. Selective degradation of mitochondria by mitophagy. Arch. Biochem. Biophys. 2007, 462, 245–253. [Google Scholar] [CrossRef] [Green Version]
  128. Drummond, M.J.; Addison, O.; Brunker, L.; Hopkins, P.N.; McClain, D.A.; LaStayo, P.C.; Marcus, R.L. Downregulation of E3 Ubiquitin Ligases and Mitophagy-Related Genes in Skeletal Muscle of Physically Inactive, Frail Older Women: A Cross-Sectional Comparison. J. Gerontol. Ser. A Boil. Sci. Med. Sci. 2014, 69, 1040–1048. [Google Scholar] [CrossRef] [PubMed]
  129. Carter, H.N.; Kim, Y.; Erlich, A.T.; Zarrin-Khat, D.; Hood, D.A. Autophagy and mitophagy flux in young and aged skeletal muscle following chronic contractile activity. J. Physiol. 2018, 596, 3567–3584. [Google Scholar] [CrossRef]
  130. Cordero, P.; Milagro, F.I.; Campion, J.; Martinez, J.A. Maternal methyl donors supplementation during lactation prevents the hyperhomocysteinemia induced by a high-fat-sucrose intake by dams. Int. J. Mol. Sci. 2013, 14, 24422–24437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Wang, L.-J.; Zhang, H.-W.; Zhou, J.-Y.; Liu, Y.; Yang, Y.; Chen, X.-L.; Zhu, C.-H.; Zheng, R.-D.; Ling, W.-H.; Zhu, H.-L. Betaine attenuates hepatic steatosis by reducing methylation of the MTTP promoter and elevating genomic methylation in mice fed a high-fat diet. J. Nutr. Biochem. 2014, 25, 329–336. [Google Scholar] [CrossRef]
  132. Mardinoglu, A.; Bjornson, E.; Zhang, C.; Klevstig, M.; Söderlund, S.; Ståhlman, M.; Adiels, M.; Hakkarainen, A.; Lundbom, N.; Kilicarslan, M.; et al. Personal model-assisted identification of NAD + and glutathione metabolism as intervention target in NAFLD. Mol. Syst. Biol. 2017, 13, 916. [Google Scholar] [CrossRef]
  133. Wang, C.; Liu, W.; Yao, L.; Zhang, X.; Zhang, X.; Ye, C.; Jiang, H.; He, J.; Zhu, Y.; Ai, D. Hydroxyeicosapentaenoic acids and epoxyeicosatetraenoic acids attenuate early occurrence of nonalcoholic fatty liver disease. Br. J. Pharmacol. 2017, 174, 2358–2372. [Google Scholar] [CrossRef]
  134. Shen, J.; Wong, G.L.-H.; Chan, H.L.-Y.; Chan, R.S.-M.; Chan, H.-Y.; Chu, W.C.-W.; Cheung, B.H.-K.; Yeung, D.K.-W.; Li, L.S.; Sea, M.M.-M.; et al. PNPLA3 gene polymorphism and response to lifestyle modification in patients with nonalcoholic fatty liver disease. J. Gastroenterol. Hepatol. 2015, 30, 139–146. [Google Scholar] [CrossRef]
  135. Caulfield, T.; Ries, N.; Ray, P.; Shuman, C.; Wilson, B.; Ries, N. Direct-to-consumer genetic testing: Good, bad or benign? Clin. Genet. 2010, 77, 101–105. [Google Scholar] [CrossRef] [PubMed]
  136. Heianza, Y.; Qi, L. Gene-Diet Interaction and Precision Nutrition in Obesity. Int. J. Mol. Sci. 2017, 18, 787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Stover, P.J.; Caudill, M.A. Genetic and Epigenetic Contributions to Human Nutrition and Health: Managing Genome–Diet Interactions. J. Am. Diet. Assoc. 2008, 108, 1480–1487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. 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, 4. [Google Scholar] [CrossRef] [PubMed]
  139. Qadri, S.; Lallukka-Brück, S.; Luukkonen, P.K.; Zhou, Y.; Gastaldelli, A.; Orho-Melander, M.; Sammalkorpi, H.; Juuti, A.; Penttilä, A.K.; Perttilä, J.; et al. The PNPLA3-I148M variant increases polyunsaturated triglycerides in human adipose tissue. Liver Int. 2020, 40, 2128–2138. [Google Scholar] [CrossRef] [PubMed]
  140. Simopoulos, A.P. An Increase in the Omega-6/Omega-3 Fatty Acid Ratio Increases the Risk for Obesity. Nutrients 2016, 8, 128. [Google Scholar] [CrossRef] [Green Version]
  141. Santoro, N.; Savoye, M.; Kim, G.; Marotto, K.; Shaw, M.M.; Pierpont, B.; Caprio, S. Hepatic Fat Accumulation Is Modulated by the Interaction between the rs738409 Variant in the PNPLA3 Gene and the Dietary Omega6/Omega3 PUFA Intake. PLoS ONE 2012, 7, e37827. [Google Scholar] [CrossRef] [Green Version]
  142. Innes, J.K.; Calder, P.C. Omega-6 fatty acids and inflammation. Prostaglandins Leukot. Essent. Fat. Acids 2018, 132, 41–48. [Google Scholar] [CrossRef] [Green Version]
  143. Bjermo, H.; Iggman, D.; Kullberg, J.; Dahlman, I.; Johansson, L.; Persson, L.; Berglund, J.; Pulkki, K.; Basu, S.; Uusitupa, M.; et al. Effects of n-6 PUFAs compared with SFAs on liver fat, lipoproteins, and inflammation in abdominal obesity: A randomized controlled trial. Am. J. Clin. Nutr. 2012, 95, 1003–1012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Kain, V.; Ingle, K.A.; Kachman, M.; Baum, H.; Shanmugam, G.; Rajasekaran, N.S.; Young, M.E.; Halade, G.V. Excess ω-6 fatty acids influx in aging drives metabolic dysregulation, electrocardiographic alterations, and low-grade chronic inflammation. Am. J. Physiol. Circ. Physiol. 2018, 314, H160–H169. [Google Scholar] [CrossRef] [PubMed]
  145. Shahidi, F.; Ambigaipalan, P. Omega-3 Polyunsaturated Fatty Acids and Their Health Benefits. Annu. Rev. Food Sci. Technol. 2018, 9, 345–381. [Google Scholar] [CrossRef]
  146. Van Name, M.A.; Savoye, M.; Chick, J.M.; Galuppo, B.T.; Feldstein, A.E.; Pierpont, B.; Johnson, C.; Shabanova, V.; Ekong, U.; Valentino, P.L.; et al. A Low ω-6 to ω-3 PUFA Ratio (n–6:n–3 PUFA) Diet to Treat Fatty Liver Disease in Obese Youth. J. Nutr. 2020, 150, 2314–2321. [Google Scholar] [CrossRef]
  147. Davis, J.N.; Lê, K.-A.; Walker, R.W.; Vikman, S.; Spruijt-Metz, N.; Weigensberg, M.J.; Allayee, H.; Goran, M.I. Increased hepatic fat in overweight Hispanic youth influenced by interaction between genetic variation in PNPLA3 and high dietary carbohydrate and sugar consumption. Am. J. Clin. Nutr. 2010, 92, 1522–1527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Nobili, V.; Liccardo, D.; Bedogni, G.; Salvatori, G.; Gnani, D.; Bersani, I.; Alisi, A.; Valenti, L.; Raponi, M. Influence of dietary pattern, physical activity, and I148M PNPLA3 on steatosis severity in at-risk adolescents. Genes Nutr. 2014, 9, 392. [Google Scholar] [CrossRef] [Green Version]
  149. Mater, M.K.; Thelen, A.P.; Pan, D.A.; Jump, D.B. Sterol Response Element-binding Protein 1c (SREBP1c) Is Involved in the Polyunsaturated Fatty Acid Suppression of Hepatic S14 Gene Transcription. J. Biol. Chem. 1999, 274, 32725–32732. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Huang, Y.; He, S.; Li, J.Z.; Seo, Y.-K.; Osborne, T.F.; Cohen, J.C.; Hobbs, H.H. A feed-forward loop amplifies nutritional regulation of PNPLA3. Proc. Natl. Acad. Sci. USA 2010, 107, 7892–7897. [Google Scholar] [CrossRef] [Green Version]
  151. Ma, D.; Wang, Z.; Merrikh, C.N.; Lang, K.S.; Lu, P.; Li, X.; Merrikh, H.; Rao, Z.; Xu, W. Crystal structure of a membrane-bound O-acyltransferase. Nat. Cell Biol. 2018, 562, 286–290. [Google Scholar] [CrossRef]
  152. Meroni, M.; Dongiovanni, P.; Longo, M.; Carli, F.; Baselli, G.; Rametta, R.; Pelusi, S.; Badiali, S.; Maggioni, M.; Gaggini, M.; et al. Mboat7 down-regulation by hyper-insulinemia induces fat accumulation in hepatocytes. EBioMedicine 2020, 52, 102658. [Google Scholar] [CrossRef] [Green Version]
  153. O’Hare, E.A.; Yang, R.; Yerges-Armstrong, L.M.; Sreenivasan, U.; McFarland, R.; Leitch, C.C.; Wilson, M.H.; Narina, S.; Gorden, A.; Ryan, K.A.; et al. TM6SF2 rs58542926 impacts lipid processing in liver and small intestine. Hepatology 2017, 65, 1526–1542. [Google Scholar] [CrossRef] [Green Version]
  154. Musso, G.; Cipolla, U.; Cassader, M.; Pinach, S.; Saba, F.; De Michieli, F.; Paschetta, E.; Bongiovanni, D.; Framarin, L.; Leone, N.; et al. TM6SF2 rs58542926 variant affects postprandial lipoprotein metabolism and glucose homeostasis in NAFLD. J. Lipid Res. 2017, 58, 1221–1229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Franzago, M.; Fraticelli, F.; Nicolucci, A.; Celentano, C.; Liberati, M.; Stuppia, L.; Vitacolonna, E. Molecular Analysis of a Genetic Variants Panel Related to Nutrients and Metabolism: Association with Susceptibility to Gestational Diabetes and Cardiometabolic Risk in Affected Women. J. Diabetes Res. 2017, 2017, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Nettleton, J.A.; McKeown, N.M.; Kanoni, S.; Lemaitre, R.N.; Hivert, M.-F.; Ngwa, J.; Van Rooij, F.J.A.; Sonestedt, E.; Wojczynski, M.K.; Ye, Z.; et al. Interactions of Dietary Whole-Grain Intake With Fasting Glucose- and Insulin-Related Genetic Loci in Individuals of European Descent: A meta-analysis of 14 cohort studies. Diabetes Care 2010, 33, 2684–2691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Santoro, N.; Caprio, S.; Pierpont, B.; Van Name, M.; Savoye, M.; Parks, E.J. Hepatic De Novo Lipogenesis in Obese Youth Is Modulated by a Common Variant in the GCKR Gene. J. Clin. Endocrinol. Metab. 2015, 100, E1125–E1132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Imamura, F.; Micha, R.; Khatibzadeh, S.; Fahimi, S.; Shi, P.; Powles, J.; Mozaffarian, D. Dietary quality among men and women in 187 countries in 1990 and 2010: A systematic assessment. Lancet Glob. Health 2015, 3, e132–e142. [Google Scholar] [CrossRef] [Green Version]
  159. Monteiro, C.A.; Moubarac, J.-C.; Cannon, G.; Ng, S.W.; Popkin, B. Ultra-processed products are becoming dominant in the global food system. Obes. Rev. 2013, 14, 21–28. [Google Scholar] [CrossRef]
  160. Zhang, L.; Wang, X.; Cueto, R.; Effi, C.; Zhang, Y.; Tan, H.; Qin, X.; Ji, Y.; Yang, X.; Wang, H. Biochemical basis and metabolic interplay of redox regulation. Redox Biol. 2019, 26, 101284. [Google Scholar] [CrossRef]
  161. Bellanti, F.; Villani, R.; Facciorusso, A.; Vendemiale, G.; Serviddio, G. Lipid oxidation products in the pathogenesis of non-alcoholic steatohepatitis. Free. Radic. Biol. Med. 2017, 111, 173–185. [Google Scholar] [CrossRef]
  162. Jacobs, S.R.; Herman, C.E.; Maciver, N.J.; Wofford, J.A.; Wieman, H.L.; Hammen, J.J.; Rathmell, J.C. Glucose Uptake Is Limiting in T Cell Activation and Requires CD28-Mediated Akt-Dependent and Independent Pathways. J. Immunol. 2008, 180, 4476–4486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Michalek, R.D.; Gerriets, V.A.; Jacobs, S.R.; MacIntyre, A.N.; Maciver, N.J.; Mason, E.F.; Sullivan, S.A.; Nichols, A.G.; Rathmell, J.C. Cutting Edge: Distinct Glycolytic and Lipid Oxidative Metabolic Programs Are Essential for Effector and Regulatory CD4+ T Cell Subsets. J. Immunol. 2011, 186, 3299–3303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Pearce, E.L.; Walsh, M.C.; Cejas, P.J.; Harms, G.M.; Shen, H.; Wang, L.-S.; Jones, R.G.; Choi, Y. Enhancing CD8 T-cell memory by modulating fatty acid metabolism. Nature 2009, 460, 103–107. [Google Scholar] [CrossRef] [PubMed]
  165. Mozaffarian, D. Dietary and Policy Priorities for Cardiovascular Disease, Diabetes, and Obesity: A Comprehensive Review. Circulation 2016, 133, 187–225. [Google Scholar] [CrossRef]
  166. Atkinson, F.S.; Foster-Powell, K.; Brand-Miller, J.C. International Tables of Glycemic Index and Glycemic Load Values: 2008. Diabetes Care 2008, 31, 2281–2283. [Google Scholar] [CrossRef] [Green Version]
  167. Hosseini, Z.; Whiting, S.J.; Vatanparast, H. Current evidence on the association of the metabolic syndrome and dietary patterns in a global perspective. Nutr. Res. Rev. 2016, 29, 152–162. [Google Scholar] [CrossRef]
  168. Hansson, G.K.; Hermansson, A. The immune system in atherosclerosis. Nat. Immunol. 2011, 12, 204–212. [Google Scholar] [CrossRef]
  169. Tall, A.R.; Yvan-Charvet, L. Cholesterol, inflammation and innate immunity. Nat. Rev. Immunol. 2015, 15, 104–116. [Google Scholar] [CrossRef] [Green Version]
  170. Wen, H.; Gris, D.; Lei, Y.; Jha, S.; Zhang, L.; Huang, M.T.-H.; Brickey, W.J.; Ting, J.P.-Y. Fatty acid–induced NLRP3-ASC inflammasome activation interferes with insulin signaling. Nat. Immunol. 2011, 12, 408–415. [Google Scholar] [CrossRef] [Green Version]
  171. Lancaster, G.I.; Langley, K.G.; Berglund, N.A.; Kammoun, H.L.; Reibe, S.; Estevez, E.; Weir, J.; Mellett, N.A.; Pernes, G.; Conway, J.R.; et al. Evidence that TLR4 Is Not a Receptor for Saturated Fatty Acids but Mediates Lipid-Induced Inflammation by Reprogramming Macrophage Metabolism. Cell Metab. 2018, 27, 1096–1110.e5. [Google Scholar] [CrossRef] [Green Version]
  172. Schwingshackl, L.; Hoffmann, G. Mediterranean dietary pattern, inflammation and endothelial function: A systematic review and meta-analysis of intervention trials. Nutr. Metab. Cardiovasc. Dis. 2014, 24, 929–939. [Google Scholar] [CrossRef]
  173. Estruch, R.; Ros, E.; Salas-Salvadó, J.; Covas, M.-I.; Corella, D.; Arós, F.; Gómez-Gracia, E.; Ruiz-Gutiérrez, V.; Fiol, M.; Lapetra, J.; et al. Primary Prevention of Cardiovascular Disease with a Mediterranean Diet Supplemented with Extra-Virgin Olive Oil or Nuts. N. Engl. J. Med. 2018, 378, e34. [Google Scholar] [CrossRef]
  174. Ishiyama, J.; Taguchi, R.; Akasaka, Y.; Shibata, S.; Ito, M.; Nagasawa, M.; Murakami, K. Unsaturated FAs prevent palmitate-induced LOX-1 induction via inhibition of ER stress in macrophages. J. Lipid Res. 2011, 52, 299–307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Oh, D.Y.; Talukdar, S.; Bae, E.J.; Imamura, T.; Morinaga, H.; Fan, W.Q.; Li, P.; Lu, W.J.; Watkins, S.M.; Olefsky, J.M. GPR120 Is an Omega-3 Fatty Acid Receptor Mediating Potent Anti-inflammatory and Insulin-Sensitizing Effects. Cell 2010, 142, 687–698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Gurav, A.; Sivaprakasam, S.; Bhutia, Y.D.; Boettger, T.; Singh, N.; Ganapathy, V. Slc5a8, a Na+-coupled high-affinity transporter for short-chain fatty acids, is a conditional tumour suppressor in colon that protects against colitis and colon cancer under low-fibre dietary conditions. Biochem. J. 2015, 469, 267–278. [Google Scholar] [CrossRef] [PubMed]
  177. Kim, M.H.; Kang, S.G.; Park, J.H.; Yanagisawa, M.; Kim, C.H. Short-Chain Fatty Acids Activate GPR41 and GPR43 on Intestinal Epithelial Cells to Promote Inflammatory Responses in Mice. Gastroenterology 2013, 145, 396–406.e10. [Google Scholar] [CrossRef]
  178. Nastasi, C.; Candela, M.; Bonefeld, C.M.; Geisler, C.; Hansen, M.; Krejsgaard, T.; Biagi, E.; Andersen, M.H.; Brigidi, P.; Ødum, N.; et al. The effect of short-chain fatty acids on human monocyte-derived dendritic cells. Sci. Rep. 2015, 5, 16148. [Google Scholar] [CrossRef]
  179. Smith, P.M.; Howitt, M.R.; Panikov, N.; Michaud, M.; Gallini, C.A.; Bohlooly-Y, M.; Glickman, J.N.; Garrett, W.S. The Microbial Metabolites, Short-Chain Fatty Acids, Regulate Colonic Treg Cell Homeostasis. Science 2013, 341, 569–573. [Google Scholar] [CrossRef] [Green Version]
  180. Mittal, M.; Siddiqui, M.R.; Tran, K.; Reddy, S.P.; Malik, A.B. Reactive Oxygen Species in Inflammation and Tissue Injury. Antioxid. Redox Signal. 2014, 20, 1126–1167. [Google Scholar] [CrossRef] [Green Version]
  181. Papac-Milicevic, N.; Busch, C.-L.; Binder, C. Malondialdehyde Epitopes as Targets of Immunity and the Implications for Atherosclerosis. Adv. Immunol. 2016, 131, 1–59. [Google Scholar] [CrossRef] [Green Version]
  182. Chou, M.-Y.; Fogelstrand, L.; Hartvigsen, K.; Hansen, L.F.; Woelkers, D.; Shaw, P.X.; Choi, J.; Perkmann, T.; Bäckhed, F.; Miller, Y.I.; et al. Oxidation-specific epitopes are dominant targets of innate natural antibodies in mice and humans. J. Clin. Investig. 2009, 119, 1335–1349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Sutti, S.; Jindal, A.; Locatelli, I.; Vacchiano, M.; Gigliotti, L.; Bozzola, C.; Albano, E. Adaptive immune responses triggered by oxidative stress contribute to hepatic inflammation in NASH. Hepatology 2014, 59, 886–897. [Google Scholar] [CrossRef] [PubMed]
  184. Albano, E.; Mottaran, E.; Vidali, M.; Reale, E.; Saksena, S.; Occhino, G.; Burt, A.D.; Day, C.P. Immune response towards lipid peroxidation products as a predictor of progression of non-alcoholic fatty liver disease to advanced fibrosis. Gut 2005, 54, 987–993. [Google Scholar] [CrossRef] [PubMed]
  185. Bruzzì, S.; Sutti, S.; Giudici, G.; Burlone, M.E.; Ramavath, N.N.; Toscani, A.; Bozzola, C.; Schneider, P.; Morello, E.; Parola, M.; et al. B2-Lymphocyte responses to oxidative stress-derived antigens contribute to the evolution of nonalcoholic fatty liver disease (NAFLD). Free Radic. Biol. Med. 2018, 124, 249–259. [Google Scholar] [CrossRef]
  186. Ma, X.; Hua, J.; Mohamood, A.R.; Hamad, A.R.A.; Ravi, R.; Li, Z. A high-fat diet and regulatory T cells influence susceptibility to endotoxin-induced liver injury. Hepatology 2007, 46, 1519–1529. [Google Scholar] [CrossRef]
  187. Ma, C.; Kesarwala, A.H.; Eggert, T.; Medina-Echeverz, J.; Kleiner, D.E.; Jin, P.; Stroncek, P.J.D.F.; Terabe, M.; Kapoor, V.; Elgindi, M.; et al. NAFLD causes selective CD4+ T lymphocyte loss and promotes hepatocarcinogenesis. Nat. Cell Biol. 2016, 531, 253–257. [Google Scholar] [CrossRef] [Green Version]
  188. Sheedy, F.J.; Grebe, A.; Rayner, K.J.; Kalantari, P.; Ramkhelawon, B.; Carpenter, S.B.; Becker, C.E.; Ediriweera, H.N.; Mullick, A.E.; Golenbock, D.T.; et al. CD36 coordinates NLRP3 inflammasome activation by facilitating intracellular nucleation of soluble ligands into particulate ligands in sterile inflammation. Nat. Immunol. 2013, 14, 812–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Stewart, C.R.; Stuart, L.M.; Wilkinson, K.; Van Gils, J.M.; Deng, J.; Halle, A.; Rayner, K.J.; Boyer, L.; Zhong, R.; Frazier, W.A.; et al. CD36 ligands promote sterile inflammation through assembly of a Toll-like receptor 4 and 6 heterodimer. Nat. Immunol. 2010, 11, 155–161. [Google Scholar] [CrossRef] [Green Version]
  190. Reboldi, A.; Dang, E.V.; McDonald, J.G.; Liang, G.; Russell, D.; Cyster, J.G. 25-Hydroxycholesterol suppresses interleukin-1–driven inflammation downstream of type I interferon. Science 2014, 345, 679–684. [Google Scholar] [CrossRef] [Green Version]
  191. Meex, R.C.R.; Blaak, E.E. Mitochondrial Dysfunction is a Key Pathway that Links Saturated Fat Intake to the Development and Progression of NAFLD. Mol. Nutr. Food Res. 2021, 65, e1900942. [Google Scholar] [CrossRef]
  192. Federico, A.; Dallio, M.; Godos, J.; Loguercio, C.; Salomone, F.; Loguercio, C. Targeting gut-liver axis for the treatment of nonalcoholic steatohepatitis: Translational and clinical evidence. Transl. Res. 2016, 167, 116–124. [Google Scholar] [CrossRef]
  193. Kosteli, A.; Sugaru, E.; Haemmerle, G.; Martin, J.F.; Lei, J.; Zechner, R.; Ferrante, A.W. Weight loss and lipolysis promote a dynamic immune response in murine adipose tissue. J. Clin. Investig. 2010, 120, 3466–3479. [Google Scholar] [CrossRef] [Green Version]
  194. Molofsky, A.B.; Nussbaum, J.C.; Liang, H.-E.; Van Dyken, S.J.; Cheng, L.E.; Mohapatra, A.; Chawla, A.; Locksley, R.M. Innate lymphoid type 2 cells sustain visceral adipose tissue eosinophils and alternatively activated macrophages. J. Exp. Med. 2013, 210, 535–549. [Google Scholar] [CrossRef] [PubMed]
  195. Lackey, D.E.; Olefsky, J.M. Regulation of metabolism by the innate immune system. Nat. Rev. Endocrinol. 2016, 12, 15–28. [Google Scholar] [CrossRef]
  196. Kosaraju, R.; Guesdon, W.; Crouch, M.J.; Teague, H.L.; Sullivan, E.M.; Karlsson, E.A.; Schultz-Cherry, S.; Gowdy, K.; Bridges, L.C.; Reese, L.R.; et al. B Cell Activity Is Impaired in Human and Mouse Obesity and Is Responsive to an Essential Fatty Acid upon Murine Influenza Infection. J. Immunol. 2017, 198, 4738–4752. [Google Scholar] [CrossRef] [Green Version]
  197. Mulder, W.J.M.; Ochando, J.; Joosten, L.A.B.; Fayad, Z.A.; Netea, M.G. Therapeutic targeting of trained immunity. Nat. Rev. Drug Discov. 2019, 18, 553–566. [Google Scholar] [CrossRef] [PubMed]
  198. Netea, M.G.; Joosten, L.A.B.; Latz, E.; Mills, K.H.G.; Natoli, G.; Stunnenberg, H.G.; O’Neill, L.A.J.; Xavier, R.J. Trained immunity: A program of innate immune memory in health and disease. Science 2016, 352, aaf1098. [Google Scholar] [CrossRef] [Green Version]
  199. Wing, R.R.; Bolin, P.; Brancati, F.L.; Bray, G.A.; Clark, J.M.; Coday, M.; Crow, R.S.; Curtis, J.M.; Egan, C.M.; Espeland, M.A.; et al. Cardiovascular effects of intensive lifestyle intervention in type 2 diabetes. N. Engl. J. Med. 2013, 369, 145–154. [Google Scholar] [CrossRef] [Green Version]
  200. Schmitz, J.; Evers, N.; Awazawa, M.; Nicholls, H.; Brönneke, H.; Dietrich, A.; Mauer, J.; Blüher, M.; Brüning, J. Obesogenic memory can confer long-term increases in adipose tissue but not liver inflammation and insulin resistance after weight loss. Mol. Metab. 2016, 5, 328–339. [Google Scholar] [CrossRef] [PubMed]
  201. Nagareddy, P.R.; Murphy, A.J.; Stirzaker, R.A.; Hu, Y.; Yu, S.; Miller, R.G.; Ramkhelawon, B.; Distel, E.; Westerterp, M.; Huang, L.-S.; et al. Hyperglycemia Promotes Myelopoiesis and Impairs the Resolution of Atherosclerosis. Cell Metab. 2013, 17, 695–708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Seijkens, T.; Hoeksema, M.A.; Beckers, L.; Smeets, E.; Meiler, S.; Levels, J.; Tjwa, M.; De Winther, M.P.J.; Lutgens, E. Hypercholesterolemia-induced priming of hematopoietic stem and progenitor cells aggravates atherosclerosis. FASEB J. 2014, 28, 2202–2213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Van Kampen, E.; Jaminon, A.; Van Berkel, T.J.C.; Van Eck, M. Diet-induced (epigenetic) changes in bone marrow augment atherosclerosis. J. Leukoc. Biol. 2014, 96, 833–841. [Google Scholar] [CrossRef] [Green Version]
  204. Christ, A.; Günther, P.; Lauterbach, M.A.; Duewell, P.; Biswas, D.; Pelka, K.; Scholz, C.J.; Oosting, M.; Haendler, K.; Baßler, K.; et al. Western Diet Triggers NLRP3-Dependent Innate Immune Reprogramming. Cell 2018, 172, 162–175.e14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. NAFLD: a metabolic systemic disease. The modern approach considers NAFLD as a metabolic, systemic disease characterized by several extra-hepatic manifestations mostly linked by a status of insulin resistance (IR). EDCs: Endocrine disrupting compounds; IR: Insulin resistance; MAFLD: metabolic (dysfunction) associated fatty liver disease; NAFLD: Non-alcoholic fatty liver disease; NASH: Non-alcoholic steatohepatitis; HCC: Hepatocellular carcinoma; OSA: Obstructive sleep apnea; PCOS: Polycystic ovarian syndrome; T2DM: Type 2 Diabetes Mellitus.
Figure 1. NAFLD: a metabolic systemic disease. The modern approach considers NAFLD as a metabolic, systemic disease characterized by several extra-hepatic manifestations mostly linked by a status of insulin resistance (IR). EDCs: Endocrine disrupting compounds; IR: Insulin resistance; MAFLD: metabolic (dysfunction) associated fatty liver disease; NAFLD: Non-alcoholic fatty liver disease; NASH: Non-alcoholic steatohepatitis; HCC: Hepatocellular carcinoma; OSA: Obstructive sleep apnea; PCOS: Polycystic ovarian syndrome; T2DM: Type 2 Diabetes Mellitus.
Nutrients 13 01679 g001
Figure 2. Nutrigenomics and nutrigenetics: two sides of the same medal with several linked approaches. Genome influences the responsiveness to nutrients (nutrigenetics’ field); at the same time, nutrition can also modify gene expression involving epigenetic mechanisms (nutrigenomics’ field). Nutrigenomics investigates the effects deriving from the interaction between the nutritional environment and inherited factors. Given the complexity of the scenario, nutrigenomics quests include several approaches involving many other disciplines. Nutritional factors and genetic ones influence each other: on one side, nutrients affect DNA metabolism, gene expression, and genetic variability; on the other side, genetic variants (as SNPs), by determining specific individual genotype, influence dietary habits. In turn, nutrigenetics could also be influenced by epigenetics. Altogether, these mechanisms contribute to determinate the status of health or a condition disease. SNPs, Single Nucleotide Polymorphisms.
Figure 2. Nutrigenomics and nutrigenetics: two sides of the same medal with several linked approaches. Genome influences the responsiveness to nutrients (nutrigenetics’ field); at the same time, nutrition can also modify gene expression involving epigenetic mechanisms (nutrigenomics’ field). Nutrigenomics investigates the effects deriving from the interaction between the nutritional environment and inherited factors. Given the complexity of the scenario, nutrigenomics quests include several approaches involving many other disciplines. Nutritional factors and genetic ones influence each other: on one side, nutrients affect DNA metabolism, gene expression, and genetic variability; on the other side, genetic variants (as SNPs), by determining specific individual genotype, influence dietary habits. In turn, nutrigenetics could also be influenced by epigenetics. Altogether, these mechanisms contribute to determinate the status of health or a condition disease. SNPs, Single Nucleotide Polymorphisms.
Nutrients 13 01679 g002
Figure 3. Genetic determinants of MAFLD influence response to the nutrients. PNPLA3, Phospholipase domain-containing protein-3; MBOAT7, Membrane-bound O-acyltransferase domain containing 7; TM6SF2, Transmembrane 6 superfamily member 2 protein; GCKR, Glucokinase regulator; SREBP1, Sterol regulatory element-binding proteins; VLD, Very-low-density lipoproteins; FAs, Fatty acids.
Figure 3. Genetic determinants of MAFLD influence response to the nutrients. PNPLA3, Phospholipase domain-containing protein-3; MBOAT7, Membrane-bound O-acyltransferase domain containing 7; TM6SF2, Transmembrane 6 superfamily member 2 protein; GCKR, Glucokinase regulator; SREBP1, Sterol regulatory element-binding proteins; VLD, Very-low-density lipoproteins; FAs, Fatty acids.
Nutrients 13 01679 g003
Table 1. Main genetic determinants of NAFLD.
Table 1. Main genetic determinants of NAFLD.
GeneVariants/SNPs/Protein Variants Relative Effects/Association with
Major and most common genetic determinants of MAFLDPNPLA3rs738409
(I148M)
Disruption of triglycerides and phospholipids turnover and remodelling: increased hepatic fat accumulation;
Disruption of retinol storage in HSCs leading to higher risk of inflammation, fibrosis, and HCC progression.
MBOAT 7rs641738Higher risk of MAFLD development, inflammation, fibrosis, and HCC progression.
TM6SF2rs58542926Favouring liver fat accumulation;Protection against the development of cardiovascular diseases.
Other genetic determinants involved in lipid metabolismGCKRrs1260326Increased de novo lipogenesis and worsened hepatic steatosis.
PPP1R3Brs4841132Reduction of de novo lipogenesis and thus protection from hepatic fat accumulation.
APOBSeveral and differentReduced VLDL export from hepatocytes.
Other genetic determinants involved in oxidative stress imbalanceSOD2rs4880Higher oxidative stress and more advanced fibrosis.
UCP3rs1800849IR worsening, increased adiponectin levels, and NASH development.
UCP2rs695366Higher insulin sensitivity and protection against liver damage.
MARC1A165TLower hepatic fat accumulation and decreased levels of several biomarkers of liver disease.
HFErs1800562 (C282Y)Iron overload and related oxidative stress imbalance.
Other genetic determinants involved in inflammation and fibrosisTLR4D299G and T399IProtection against fibrosis (in animal models).
IFNL4rs368234815Induction of severe inflammation.
IFN/IL-28rs12979860Promotes inflammation and fibrosis (it is predictive for advanced stage of the disease).
PCSK7rs236918Liver damage and altered fibrogenesis association.
MERTKrs4374383Protection against fibrosis.
HSD17B13rs72613567Reduced risk of NASH (but not steatosis).
Genetic variants identified as associated with MAFLD (Metabolic (dysfunction) associated fatty liver disease) and NASH (Non-alcoholic steatohepatitis) encodes for genes involved in several metabolism pathways. Some variants seem able to protect from MAFLD onset and progression to NASH; unfortunately, many others promote hepatic steatosis and its worsening in severe inflammation and fibrosis. PNPLA3, Patatin-like phospholipase domain-containing 3; MBOAT 7, Membrane-bound O-acyltransferase domain-containing 7; TM6SF2, Transmembrane 6 superfamily member 2; GCKR, glucokinase regulator; PPP1R3B, protein phosphatase 1 regulatory subunit 3B; APOB, APOB100; SOD2, manganese-dependent superoxide dismutase; UCP2, Uncoupling protein 2; UCP3, Uncoupling protein 3; MARC1, Mitochondrial Amidoxime Reducing Component 1; TLR4, toll-like receptor 4; IFNL4, Interferon Lambda 4; PCSK7, Proprotein convertase subtilisin/Kexin type 7; MERTK, Mer T kinase; HSCs, Hepatic stellate cells; HCC, Hepatocellular carcinoma; VLDL, Very-low-density lipoproteins.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dallio, M.; Romeo, M.; Gravina, A.G.; Masarone, M.; Larussa, T.; Abenavoli, L.; Persico, M.; Loguercio, C.; Federico, A. Nutrigenomics and Nutrigenetics in Metabolic- (Dysfunction) Associated Fatty Liver Disease: Novel Insights and Future Perspectives. Nutrients 2021, 13, 1679. https://doi.org/10.3390/nu13051679

AMA Style

Dallio M, Romeo M, Gravina AG, Masarone M, Larussa T, Abenavoli L, Persico M, Loguercio C, Federico A. Nutrigenomics and Nutrigenetics in Metabolic- (Dysfunction) Associated Fatty Liver Disease: Novel Insights and Future Perspectives. Nutrients. 2021; 13(5):1679. https://doi.org/10.3390/nu13051679

Chicago/Turabian Style

Dallio, Marcello, Mario Romeo, Antonietta Gerarda Gravina, Mario Masarone, Tiziana Larussa, Ludovico Abenavoli, Marcello Persico, Carmelina Loguercio, and Alessandro Federico. 2021. "Nutrigenomics and Nutrigenetics in Metabolic- (Dysfunction) Associated Fatty Liver Disease: Novel Insights and Future Perspectives" Nutrients 13, no. 5: 1679. https://doi.org/10.3390/nu13051679

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