Next Article in Journal
Safety, Tolerability, and Pharmacokinetics of β-Cryptoxanthin Supplementation in Healthy Women: A Double-Blind, Randomized, Placebo-Controlled Clinical Trial
Previous Article in Journal
Impacts of High Fructose Diet and Chronic Exercise on Nitric Oxide Synthase and Oxidative Stress in Rat Kidney
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dysregulation of Lipid and Glucose Metabolism in Nonalcoholic Fatty Liver Disease

Cardiovascular Research Center, Department of Internal Medicine, Yale School of Medicine, New Haven, CT 06511, USA
*
Authors to whom correspondence should be addressed.
Current address: Janssen Pharmaceutical Companies of Johnson and Johnson, 1400 McKean Road, Springhouse, PA 19002, USA.
Nutrients 2023, 15(10), 2323; https://doi.org/10.3390/nu15102323
Submission received: 3 April 2023 / Revised: 8 May 2023 / Accepted: 12 May 2023 / Published: 16 May 2023
(This article belongs to the Special Issue Lipid Metabolism in Diabetes and Atherosclerosis)

Abstract

:
Non-Alcoholic Fatty Liver Disease (NAFLD) is a highly prevalent condition affecting approximately a quarter of the global population. It is associated with increased morbidity, mortality, economic burden, and healthcare costs. The disease is characterized by the accumulation of lipids in the liver, known as steatosis, which can progress to more severe stages such as steatohepatitis, fibrosis, cirrhosis, and even hepatocellular carcinoma (HCC). This review focuses on the mechanisms that contribute to the development of diet-induced steatosis in an insulin-resistant liver. Specifically, it discusses the existing literature on carbon flux through glycolysis, ketogenesis, TCA (Tricarboxylic Acid Cycle), and fatty acid synthesis pathways in NAFLD, as well as the altered canonical insulin signaling and genetic predispositions that lead to the accumulation of diet-induced hepatic fat. Finally, the review discusses the current therapeutic efforts that aim to ameliorate various pathologies associated with NAFLD.

1. Introduction

Non-Alcoholic Fatty Liver Disease (NAFLD) currently affects over 25% of the adult population globally [1,2]. If left untreated, the fatty liver can progress to more severe stages such as steatohepatitis (NASH), fibrosis, cirrhosis, and hepatocellular carcinoma, and an increased risk of developing type II diabetes and cardiovascular diseases. This is primarily due to the malfunctioning of the liver affecting flux through glucose and lipid metabolic pathways, thereby affecting the regulation of lipid and glucose levels in the body.
Recent studies have shown that patients with NAFLD have a greater than two-fold risk of developing type II diabetes [3], and those with type II diabetes have a 75% prevalence of NAFLD [1,2,4]. Both NAFLD and type II diabetes are associated with an increased risk of cardiovascular diseases [5]. These findings suggest the need for increased surveillance of NAFLD in patients with cardiovascular diseases and type II diabetes, and vice versa. Therapies developed for treating type II diabetes or cardiovascular diseases may also help reduce cardiovascular complications in patients with NAFLD, and vice versa where patients with NAFLD could benefit from therapies against type II diabetes. Moreover, NAFLD has also become one of the most common liver diseases in the pediatric population, with a 5–10% prevalence globally [6,7,8,9]. These patients are also susceptible to developing cardiovascular diseases and type II diabetes, making it important to carry out prospective follow-ups to prevent serious outcomes. Hence, there is an urgent need to develop novel therapies and investigate their efficacies in ongoing and prospective clinical trials.

1.1. Altered Carbon Flux in NAFLD

The liver plays a crucial role in regulating lipid homeostasis. After digestion, triglycerides (TGs) are transported in lipoprotein particles called chylomicrons. These TGs are broken down by lipoprotein lipase in the capillaries, releasing fatty acids that are taken up by adipose and skeletal muscle tissues. The remaining chylomicron remnants are then taken up by the liver. The liver also synthesizes lipids from dietary sugars through de novo lipogenesis (DNL). Hepatic TGs are packaged and secreted along with free and esterified cholesterol as very low-density lipoprotein particles (VLDL). Through lipolysis of triglycerides and exchange of apolipoproteins, VLDL transforms into intermediate-density lipoprotein particles (IDL) and low-density lipoprotein particles (LDL) [10]. Furthermore, the liver receives free fatty acids through lipolysis of TGs in the adipose tissue, particularly during fasting or insulin resistance states [11]. In this review, we examine how carbon from lipids and carbohydrates flows through various metabolic pathways in the liver and how this flux becomes disrupted in cases of over-nutrition. This is a complex condition since energy flux through multiple pathways is intertwined with signaling from endocrine hormones such as insulin, glucagon, adipokines, cytokines, and myokines.
The pathological response to an excess of carbon flux from energy-rich nutrients, such as sugars and lipids, suggests that metabolic pathways are vulnerable to managing excess nutrients. This vulnerability largely arises due to systemic insulin resistance, which is one of the first predictors of dysregulated lipid and glucose metabolism [1,2,12,13,14,15,16]. Insulin resistance, along with visceral adiposity, elevated triglycerides, and reduced high-density lipoproteins (HDL), are common clinical characteristics of NASH patients, including women with gestational diabetes [1,2,12,13,14,15,16].
Insulin resistance leads to dysregulation of both anabolic (de novo synthesis and lipid accumulation) and catabolic (oxidation and secretion) processes of lipid and glucose metabolism (Figure 1). One of the immediate effects of insulin resistance is impaired glucose disposal in peripheral skeletal muscle [17,18,19]. Normally, insulin binding and phosphorylation of insulin receptors activate a downstream cascade of reactions that culminate in the translocation of glucose transporter Glut4 to the plasma membrane, facilitating glucose uptake by the skeletal muscle [17,18,19,20]. In an insulin-resistant state, the Glut4 receptor fails to translocate to the membrane in the skeletal muscles, hindering the uptake of plasma glucose. As a consequence, the glycogen stores of the muscle are depleted, which is one of the first manifestations of insulin resistance [21,22,23]. The lack of energy stores in the skeletal muscle leads to wasting and sarcopenia in type II diabetic patients [24]. Persistently higher levels of plasma glucose cause increased insulin secretion from the beta cells of the pancreas [13,19,25,26,27,28,29]. The resulting hyperinsulinemia is one of the defining features of dysregulated glucose metabolism and NAFLD. Continued hyperinsulinemia appears to desensitize insulin signaling in skeletal muscle, hepatocytes, and adipocytes, further exacerbating systemic insulin resistance [17,18,19]. The mechanisms causing this desensitization are only beginning to be understood (see below, structure of insulin receptor) [11,30,31,32,33,34].
Under normal conditions, insulin suppresses lipolysis of TGs from the adipose tissue and suppresses glucose production from the liver. Insulin resistance causes increased lipolysis of TGs from adipose tissue and increased glucose production in the liver. The increased lipolysis of TGs in the adipose tissue increases plasma levels of non-esterified fatty acids (NEFA) and glycerol [11,30,31,32,33,34]. Upon entry into hepatocytes, NEFA is esterified into TGs by the action of Glycerol 3-phosphate acyltransferase3 (GPAT) and Diacylglycerol O-transferase1 (DGAT). The other product of adipose tissue lipolysis is Glycerol, which is converted into glycerol-3-phosphate (G3P) in the liver by Glycerol-3 kinase (G3K). The increased flux of glycerol to the liver increases gluconeogenic flux, leading to hyperglycemia [32,35,36] (Figure 2). In NAFLD, G3P is increasingly diverted toward the glycolysis/TCA cycle [37] and gets oxidized to generate oxaloacetate (OAA), which is then converted to phosphoenolpyruvate (PEP) by mitochondrial phosphoenolpyruvate carboxykinase (PEPCK), the rate-limiting enzyme in the gluconeogenesis pathway. This results in increased glucose production from the liver (Figure 2). Moreover, the reductive equivalents generated in the TCA cycle and their subsequent oxidation lead to endoplasmic reticulum (ER) stress [31,38] and reactive oxygen species production, resulting in steatohepatitis [39]. In addition to providing carbons for gluconeogenesis, G3P also serves as a carbon backbone for the esterification of acyl chains by the actions of GPAT and DGAT, thereby contributing to lipid synthesis [37,40].
Despite resistance to insulin action in adipose tissue, de-novo lipogenesis (DNL), an insulin-dependent process, is stimulated during insulin resistance in the liver. During DNL, excess sugars are diverted towards fatty acid synthesis pathways in hepatocytes as demonstrated by stable isotope studies [11]. While glucose and fructose-rich diets stimulate DNL in the liver [11,41,42], labeled tracers of these sugars do not proportionally convert to fatty acids, indicating that they do not directly convert to fatty acids [43]. The gluconeogenic precursors such as alanine, lactate, and glutamine were found to be direct contributors to carbon in fatty acid synthesis through the DNL pathway [44,45]. Furthermore, gut microbiota-derived acetate has recently been shown to increase DNL in response to dietary fructose [46]. The acetyl-CoA generated by lipogenic substrates is released from the mitochondria as citrate, which is converted to acetyl-CoA and oxaloacetate (OAA) in the cytoplasm by citrate lyase [47]. The cytoplasmic acetyl-CoA is directed toward fatty acid synthesis via the DNL pathway, while OAA is converted to phosphoenolpyruvate (PEP) by cytoplasmic PEPCK for gluconeogenesis. The fatty acids generated by the DNL pathway (Figure 2) are esterified by glycerol released from adipose tissue. Therefore, the pathways of gluconeogenesis and lipogenesis are intricately linked and are simultaneously fueled by carbons from glycerol and NEFA released by adipose tissue, acetate from gut microbiota, and circulating acetate, lactate, and alanine.
The contribution of altered beta-oxidation towards the development of NAFLD is likely context-dependent. While some studies report decreased mitochondrial oxidation of fatty acids in NAFLD [48], others indicate a compensatory increase in beta-oxidation [31,38,39,41,49]. A recent report suggested the presence of increased flux through the TCA cycle and reduced ketogenesis without any change in beta-oxidation per se in patients with NAFLD [49]. These discrepancies are likely due to differences in underlying disease mechanisms and/or adaptive mechanisms that counter disease states [48,50]. In conclusion, the altered flux of glycerol, NEFA, and fatty acids increases lipid synthesis and glucose production in the liver, leading to a vicious cycle of hyperlipidemia and hyperglycemia and fatty liver disease.

1.2. What Condition Comes First: Insulin Resistance or NAFLD?

The question of whether insulin resistance or NAFLD comes first has remained unanswered. Individuals with metabolic disorders, manifested as NAFLD, type II diabetes, and metabolic syndrome, display a range of heterogenous traits: visceral adiposity, hyperlipidemia, hypertension, atherogenic dyslipidemia, glucose intolerance, prediabetes, and insulin resistance. Several investigators have made attempts to simplify this heterogeneity by classifying patients into different phenotypic classes [51,52,53]. This entails that the disease is highly heterogenous, there is no single cure for all the metabolic defects, and hence, therapeutic strategies have to become increasingly personalized. One factor that can advance precision medicine is a comprehensive knowledge of the underlying genetic background of the disease. While genome-wide association studies have linked many genetic loci to various metabolic traits, common variants that underlie the association of these traits have remained vastly unidentified [51]. Apart from genetic background, the composition of consumed nutrients by individuals can influence how metabolic dysfunction manifests. A diet rich in high fructose corn syrup, used as an additive to processed foods, has been shown to stimulate lipogenesis in the liver [46,54,55]. Similarly, a diet disproportionally enriched in saturated fats may increase the atherogenic lipid particles in the plasma. It is likely that there is not a single linear relationship between the development of insulin resistance and hepatic steatosis, and the mechanistic studies detailed below further show this.
The mechanistic studies in rodents show that hepatic lipid accumulation suppresses insulin signaling and, conversely, insulin resistance stimulates hepatic lipid accumulation. This mutually stimulating cycle causes both hyperglycemia and hyperlipidemia, culminating in T2D and NAFLD. In support of hepatic lipid accumulation suppressing insulin signaling, investigators have shown that lipid accumulation in the liver increases the pool of sn-1,2-diacylglycerol (DAG) in the membrane, which recruits protein kinase C (PKC) epsilon to the membrane. PKCepsilon then suppresses insulin-induced phosphorylation of insulin receptor beta at Y1162 and promotes inhibitory phosphorylation of threonine at residue 1160, suggesting that lipid accumulation inhibits insulin signaling [34,56,57,58,59]. As discussed previously, insulin resistance increases hepatic lipid accumulation by reducing peripheral lipid disposal, increasing adipose tissue lipolysis, and stimulating DNL in the liver [1,11,15,18,19,20,27,41,42,60,61,62,63]. Another example showing the complex relationship between hepatic lipogenesis and insulin resistance is illustrated by insulin-mediated activation of the mammalian target of rapamycin-1 (mTORC1). The activation of canonical insulin signaling results in a cascade of substrate phosphorylations, which inhibits Tumor sclerosis complex 1 and 2 (TSC1/2). TSC1/2 inhibition leads to the activation of mTORC1 through the G-protein Rheb complex [64]. Previously, mTORC1 activation was believed to be necessary for lipogenesis [65,66]. However, soon it was realized that the prolonged activation of mTORC1 negatively feedbacks and dampens insulin signaling [67] by its downstream targets such as S6K1 (Ribosomal Protein S6 Kinase beta 1). S6K1 promotes the inhibitory serine phosphorylation of insulin receptor substrate 1 (IRS1) [68,69]. Therefore, mTORC1 promotes lipogenesis in the liver in response to insulin, but then inhibits insulin signaling upon prolonged activation.
The autonomous insulin signaling activation of hepatocytes is further augmented by the resident macrophages in the liver known as Kupffer cells, which greatly influence overall glucose and lipid homeostasis [70,71]. From rodent studies, it is evident that Kupffer cells are activated into a pro-inflammatory state when fed a high-fat diet. These activated macrophages secrete chemokines that recruit monocytes from the plasma called monocyte-derived macrophages (MoMF) [72,73] and together secrete pro-inflammatory cytokines such as TNFalpha, IL-6, IL1-beta, which suppresses insulin signaling in the hepatocytes [70,71,74,75,76,77] and activate hepatic stellate cells leading to fibrosis [78]. Conversely, other studies suggest that Kupffer cells modulate hepatic steatosis by secreting anti-inflammatory cytokines that stimulate insulin signaling in the hepatocytes [74,79,80,81]. Taken together, these studies indicate a regulatory function of Kupffer cells on hepatic insulin sensitivity and hepatic steatosis. One consistent observation in these different studies is that a pro-inflammatory milieu worsens the phenotype of hepatic steatosis and hence, the diets that reduce inflammation should be recommended for patients suffering from metabolic dysfunction. Altogether, lipid accumulation and insulin resistance mutually stimulate each other in the liver through autonomous and non-autonomous pathways and cascade into progressively worsening conditions such as steatohepatitis and type II diabetes. The key question is: “what is the leading condition in any given individual”. Accordingly, the focus should be on the identification of unique biomarkers that can classify patients based on these conditions and treating them in a personalized fashion.

1.3. The Intersection of Metabolic Flux with the Activities and Transcript Levels of Metabolic Enzymes

The metabolic flux through different pathways is tightly regulated by the expression and/or activity of key rate-limiting enzymes. These rate-limiting enzymes control the flux through different pathways. The substrate flux in each pathway varies depending on the metabolic needs (energy generation or energy conservation) and the metabolic state (insulin-sensitive or insulin-resistant). In a metabolic pathway, there are critical flux-altering nodes that determine whether a cell will be engaged in anabolic or catabolic processes. An example of such a flux-altering node is when the cell decides whether to engage the substrate pyruvate in a decarboxylation reaction towards acetyl-CoA by pyruvate dehydrogenase and promote glucose oxidation or towards carboxylation by pyruvate carboxylase to generate OAA. The latter is then converted to the gluconeogenic precursor PEP (Figure 2). Likewise, acetyl-CoA can be directed towards oxidation in the TCA cycle or exported out of the mitochondria as citrate for lipid synthesis depending upon the ATP levels (Figure 2). Another example of a flux-altering critical node is whether acetyl-CoA is diverted towards ketogenesis, which is a catabolic reaction, or towards cholesterol synthesis, which is anabolic. Depending on the metabolic state, the HMG-CoA generated from acetyl-CoA, can be diverted towards ketogenesis by 3-hydroxy-3-methylglutaryl-CoA lyase (HMG-CoA lyase) or towards cholesterol biosynthesis by 3-hydroxy-3-methylglutaryl-CoA reductase (HMGCR). In sum, whether a given substrate is used for an anabolic or catabolic reaction is determined by the expression levels of key metabolic enzymes and the metabolic state of the cells such as ATP content, nutrient availability, and insulin sensitivity. In NAFLD, the flux is directed toward anabolic pathways. For instance, pyruvate is directed toward gluconeogenesis [32] and acetyl-CoA is directed toward lipid [11] and cholesterol synthesis [11,49].
The endocrine hormones insulin and glucagon exert a potent influence in determining the choice between anabolic versus catabolic pathways. Insulin transcriptionally and post-translationally activates sterol regulatory binding protein 1c (Srebp1-c), which upregulates hepatic lipogenesis [82,83,84,85,86,87,88,89]. Insulin activates the mammalian target of rapamycin complex2 (mTORC2), protein kinase B (PKB), and liver X receptor (LXRa), which transcriptionally activates the expression of Srebp1c [89,90,91]. The genetic mouse models reveal that Srebp-1c, ChRebp1a, LXRa, and mTORC2, increase the expression of enzymes in the glycolytic pathway concomitant with increasing the enzyme levels of the DNL pathway [89,90,91]. Increased levels of glycolytic intermediates such as glucose 6-phosphate, fructose, and the pentose phosphate pathway intermediate xylulose-5-phosphate activate the transcription factors carbohydrate response element binding protein 1a (ChRebp1a) and ChRebp1b [85,90,92,93,94]. ChRebp1a and ChRebp1b are phosphorylated and translocate to the nucleus where they transcriptionally activate genes that promote lipogenesis. The increased glycolysis likely diverts the carbon flux towards an increased lipogenesis [95,96]. However, detailed in vivo flux studies using tracers are essential to confirm this hypothesis and how these processes are disrupted in insulin-resistant conditions. This is especially important in light of observations that DNL, an insulin-dependent process, remains insulin sensitive despite systemic insulin resistance [63]. In contrast to insulin signaling, glucagon–PKA signaling inactivates Srebp-1c by inducing inhibitory phosphorylation [97,98], as well as blocking the nuclear translocation of ChRebp1 by the glucagon-cAMP-PKA signaling pathway [94,99,100].
Previous studies from our group revealed that impaired Wnt signaling in a mouse model of human LRP6R611C mutation results in the activation of the mTORC1-SREBP1c axis and the development of NAFLD [101,102]. The genetic gain and loss of TCF4 (Transcription Factor 4), an effector of Wnt signaling, further confirmed the protective action of Wnt signaling against NAFLD [103]. TCF4 transcriptionally activates Fgf19 in the intestinal epithelium, which suppresses bile synthesis in the liver, prevents dietary lipid uptake from the intestine, and protects against diet-induced fatty liver [103]. Accordingly, an independent study by Novartis reported that the loss of hepatic Wnt/β-catenin activity by Lgr4/5 deletion led to impaired secretion of bile acids, cholestasis, and altered lipid homeostasis, leading to the development of NAFLD [104]. Our group recently discovered another mutation, R102C, in dual specificity tyrosine phosphorylation regulated kinase 1b (Dyrk1b), that is strongly linked with metabolic syndrome in the carriers [105]. Further mechanistic studies revealed that Dyrk1b spontaneously causes fatty liver in rodents by increasing de novo lipogenesis and hepatic fatty acid uptake [56,105,106,107,108], despite suppressing the [104] canonical insulin signaling by PKC epsilon-mediated insulin receptor inactivation [57,59]. Our model recapitulates the selective insulin resistance model previously suggested by other investigators [63]. The detailed mechanistic studies revealed that Dyrk1b increases the activity of the mammalian target of rapamycin complex2 (mTORC2), a stimulator of DNL in the liver [91], in a kinase-independent manner and stimulates the autophosphorylation of mTOR. Importantly, the Dyrk1b mRNA and protein levels were elevated in mouse liver with diet-induced NAFLD and in human NASH samples. Since Dyrk1b is activated by auto-phosphorylation during its translation, transcriptional and post-transcriptional regulation seems to be the most probable mechanism to regulate its activity in the cells and [109,110,111] future studies are required to clarify these findings.

1.4. Selective Insulin Resistance in Hepatocytes: A Perspective from the Structure of Insulin Receptors

The concept of selective insulin resistance was proposed by Brown and Goldstein about 15 years ago when they made a striking observation that systemic insulin resistance impairs the insulin-dependent lowering of plasma glucose, but the insulin-dependent lipogenesis in the liver remains increased [63]. One contributor to NAFLD is the unrestrained supply of fatty acids from the lipolysis of adipose tissue in an insulin-resistant state [33]. While this may explain the 60% contribution to hepatic fat in NAFLD patients, the increase of fatty acid synthesis by the DNL pathway still remains unexplained [11,112]. Insulin stimulates DNL by transcriptional activation of key lipogenic enzymes such as sterol regulatory binding protein 1c (Srebp1c), carbohydrate response element binding protein (ChRebp1), and upstream transcription factor 1 (USF1), and by post-transcriptional regulation of key lipogenic enzymes such as Srebp1c [83]. A potential explanation for this “pathogenic paradox” may be offered by the structural changes in the insulin receptor in response to hyperinsulinemia, which is commonly associated with metabolic syndrome [13]. In a normal unstimulated state, the heterodimeric insulin receptor assumes a symmetrical inverted “V”-shape where the N-terminal end of one alpha subunit interacts with the juxta membrane domains on the other alpha subunit (See references [113,114,115,116] for details). The insulin receptor can bind maximally four insulin molecules [113,114,117,118,119] but the binding of even one insulin molecule to the high-affinity site1 on the insulin receptor is sufficient for the activation of the receptor and its dramatic conformational change from an inverted “V” to an asymmetrical “T” structure [113,114,117,118,119], which is sufficient for downstream signal transduction. These morphological alterations lead to autophosphorylation activation of the receptor, although the precise mechanisms by which insulin induces these conformational changes in the insulin receptor are unclear [114]. It has been observed that a fully occupied (all four insulin binding sites) insulin receptor causes the distance between trans-activation domains to increase as opposed to an asymmetrical partially occupied receptor [113]. It is plausible that a fully occupied receptor, as might be present in hyperinsulinemia, is competent to stimulate the lipogenic pathway but is unable to suppress glycogenolysis and gluconeogenesis. Hyperinsulinemia may desensitize the insulin receptor to suppress glycogenolysis and gluconeogenesis but not DNL. Alternatively, hyperinsulinemia [20,63,120,121,122] may differentially activate insulin receptor substrate-1 (IRS-1) versus IRS-2 in NAFLD [123,124]. Further, more clarity is needed to define the structural changes in the IR in an insulin-resistant state and to determine if activation of the insulin receptor is necessary for lipogenesis in NAFLD after the onset of fatty liver disease. An alternative explanation, independent of canonical insulin signaling, could be that transcriptomic and proteomic changes induced in the liver by the nutritional overflow may increase the expression of factors that stimulate lipogenesis even in the absence of active insulin signaling. One such factor that was recently discovered is Dyrk1b [105,108], which is increased transcriptionally in fatty liver disease. Dyrk1b then activates the central regulator of lipogenesis, mTORC2 [91], in the absence of the canonical insulin signaling and in a manner independent of Dyrk1b’s kinase activity. Dyrk1b increases the flux and expression of enzymes in the DNL pathway [56] while canonical insulin signaling is inhibited.

1.5. A Discrepancy in the Association of Insulin Resistance with Hepatic Fat Content

A strong positive correlation exists between liver triglyceride content and hepatic insulin resistance in the diet-induced models of NAFLD in rodents and humans [2,13,31,63,120,125,126]. As mentioned previously, increased levels of diacylglycerol (DAG), a precursor to triacylglycerol, correlates with insulin resistance, as DAG has been shown to increase PKCepsilon translocation to the plasma membrane, which causes inhibitory phosphorylation in the beta chain of the insulin receptor [34,56,57,59,127]. However, in several rodent genetic models and human genetic models of NAFLD, the association between hepatic fat and insulin resistance does not hold true. The genetic rodent models such as hepatic Akt1/2 and the hepatic knockout of mTORC2 function show reduced lipogenesis but increased insulin resistance [91,122,128,129]. Downstream of Akt, mTORC1 activation promotes lipogenesis in the liver, as revealed by pharmacological inhibition with rapamycin [66,121]. One caveat to the studies examining the function of rapamycin is that prolonged rapamycin treatment inhibits both mTORC1 and mTORC2 [130]. Notably, the loss of mTORC2 has a potent effect on the prevention of hepatic steatosis and increasing hepatic glucose production [91]. The regulation of hepatic lipogenesis by mTORC1 is more complicated. Both the activation of mTORC1 by disruption of TSC1 and loss of Raptor, a mTORC1 specific subunit, protect against NAFLD–NASH [25,131,132,133], suggesting complex regulation of hepatic lipogenesis by mTORC1. This can be explained by a negative feedback mechanism by which the overactivated mTORC1 turns off its own activation [67]. To delineate mechanisms that selectively activate lipogenesis and avoid activation of negative feedback pathways by mTOR complex1, a recent study by Gosis et al. found that mTORC1 activates lipogenesis by phosphorylating transcription factor E3/B (TFE3/B) which then suppresses fatty-acid oxidation and lysosomal and mitochondrial biogenesis [134]. These findings, however, do not recapitulate the diet-induced models of metabolic syndrome in which lipogenesis and glucose production are both elevated. Another condition in which hepatic steatosis exhibits no correlation with diet-induced insulin resistance is the increased hepatic triglyceride content in homozygous patatin-like phospholipase domain containing 3 (PNPLA3) pI148M mutation carriers, which do not develop hepatic insulin resistance [135,136,137,138]. These studies indicate that despite the common phenotype of hepatic lipid accumulation, the heterogeneity of the underlying mechanisms results in different metabolic states. In-depth analyses of dysregulated signaling pathways in NAFLD are necessary to identify novel drug targets and pursue the management of patients with NAFLD in a personalized fashion [63,120,121].

1.6. Therapeutic Interventions to Treat NAFLD

NAFLD begins with a benign fatty acid accumulation but can progress into a pathologically and morphologically complex disease, often associated with CVD and type II diabetes. Therapeutic interventions for preventing type II diabetes are also being investigated for their impacts on alleviating steatosis in the liver (Table 1). These include insulin-sensitizing PPAR agonists, satiety-promoting GLP-1 agonists, and inhibition of the glucose-absorbing Sglt2 cotransporter in the renal tubules. However, as steatosis progresses into inflammation, fibrosis, and HCC, it does not remain merely a metabolic disease. Therefore, other therapies have to be developed to cure advanced disease states such as inflammation, fibrosis, and hepatocellular carcinoma. To that end, Cenicriviroc (CVC), an oral dual CCR2/CCR5 antagonist was studied in the Centaur trial, a randomized, double-blind, placebo-controlled, multinational study of 289 adults with histological evidence of NASH and liver fibrosis. Although the drug failed to demonstrate a statistically significant improvement in the primary endpoint of NASH improvement, defined as ≥2-point improvement in NAS, it showed improvement in fibrosis, especially in subjects with higher disease activity [139]. Nevertheless, the most successful clinical trials have been limited to those that target primary metabolic defects (Table 1). The FDA has yet to approve a therapy for NAFLD.

Author Contributions

N.B. and A.M. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The work is supported by grant # 5R35HL135767-05 awarded to A.M.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Anand Narayanan for reading and suggesting edits on the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Loomba, R.; Abraham, M.; Unalp, A.; Wilson, L.; Lavine, J.; Doo, E.; Bass, N.M. Association between diabetes, family history of diabetes, and risk of nonalcoholic steatohepatitis and fibrosis. Hepatology 2012, 56, 943–951. [Google Scholar] [CrossRef] [PubMed]
  2. Younossi, Z.M.; Golabi, P.; de Avila, L.; Paik, J.M.; Srishord, M.; Fukui, N.; Qiu, Y.; Burns, L.; Afendy, A.; Nader, F. The global epidemiology of NAFLD and NASH in patients with type 2 diabetes: A systematic review and meta-analysis. J. Hepatol. 2019, 71, 793–801. [Google Scholar] [CrossRef]
  3. Mantovani, A.; Petracca, G.; Beatrice, G.; Tilg, H.; Byrne, C.D.; Targher, G. Non-alcoholic fatty liver disease and risk of incident diabetes mellitus: An updated meta-analysis of 501 022 adult individuals. Gut 2021, 70, 962–969. [Google Scholar] [CrossRef] [PubMed]
  4. Targher, G.; Corey, K.E.; Byrne, C.D.; Roden, M. The complex link between NAFLD and type 2 diabetes mellitus–mechanisms and treatments. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 599–612. [Google Scholar] [CrossRef] [PubMed]
  5. Targher, G.; Byrne, C.D.; Lonardo, A.; Zoppini, G.; Barbui, C. Non-alcoholic fatty liver disease and risk of incident cardiovascular disease: A meta-analysis. J. Hepatol. 2016, 65, 589–600. [Google Scholar] [CrossRef] [PubMed]
  6. Mann, J.P.; Valenti, L.; Scorletti, E.; Byrne, C.D.; Nobili, V. Nonalcoholic Fatty Liver Disease in Children. Semin. Liver Dis. 2018, 38, 1–13. [Google Scholar] [CrossRef]
  7. Molleston, J.P.; Schwimmer, J.B.; Yates, K.P.; Murray, K.F.; Cummings, O.W.; Lavine, J.E.; Brunt, E.M.; Scheimann, A.O.; Unalp-Arida, A.; Network, N.C.R. Histological abnormalities in children with nonalcoholic fatty liver disease and normal or mildly elevated alanine aminotransferase levels. J. Pediatr. 2014, 164, 707–713.e703. [Google Scholar] [CrossRef]
  8. Yu, E.L.; Schwimmer, J.B. Epidemiology of Pediatric Nonalcoholic Fatty Liver Disease. Clin. Liver Dis. 2021, 17, 196–199. [Google Scholar] [CrossRef]
  9. Lavine, J.E.; Schwimmer, J.B.; Molleston, J.P.; Scheimann, A.O.; Murray, K.F.; Abrams, S.H.; Rosenthal, P.; Sanyal, A.J.; Robuck, P.R.; Brunt, E.M.; et al. Treatment of nonalcoholic fatty liver disease in children: TONIC trial design. Contemp. Clin. Trials 2010, 31, 62–70. [Google Scholar] [CrossRef]
  10. Feingold, K.R. Introduction to Lipids and Lipoproteins. In Endotext; Feingold, K.R., Anawalt, B., Blackman, M.R., Boyce, A., Chrousos, G., Corpas, E., de Herder, W.W., Dhatariya, K., Dungan, K., Hofland, J., et al., Eds.; MDText.com, Inc.: South Dartmouth, MA, USA, 2000. [Google Scholar]
  11. Donnelly, K.L.; Smith, C.I.; Schwarzenberg, S.J.; Jessurun, J.; Boldt, M.D.; Parks, E.J. Sources of fatty acids stored in liver and secreted via lipoproteins in patients with nonalcoholic fatty liver disease. J. Clin. Investig. 2005, 115, 1343–1351. [Google Scholar] [CrossRef]
  12. Tiikkainen, M.; Tamminen, M.; Hakkinen, A.M.; Bergholm, R.; Vehkavaara, S.; Halavaara, J.; Teramo, K.; Rissanen, A.; Yki-Jarvinen, H. Liver-fat accumulation and insulin resistance in obese women with previous gestational diabetes. Obes. Res. 2002, 10, 859–867. [Google Scholar] [CrossRef] [PubMed]
  13. Bril, F.; Lomonaco, R.; Orsak, B.; Ortiz-Lopez, C.; Webb, A.; Tio, F.; Hecht, J.; Cusi, K. Relationship between disease severity, hyperinsulinemia, and impaired insulin clearance in patients with nonalcoholic steatohepatitis. Hepatology 2014, 59, 2178–2187. [Google Scholar] [CrossRef] [PubMed]
  14. Chalasani, N.; Deeg, M.A.; Persohn, S.; Crabb, D.W. Metabolic and anthropometric evaluation of insulin resistance in nondiabetic patients with nonalcoholic steatohepatitis. Am. J. Gastroenterol. 2003, 98, 1849–1855. [Google Scholar] [CrossRef] [PubMed]
  15. Chitturi, S.; Abeygunasekera, S.; Farrell, G.C.; Holmes-Walker, J.; Hui, J.M.; Fung, C.; Karim, R.; Lin, R.; Samarasinghe, D.; Liddle, C.; et al. NASH and insulin resistance: Insulin hypersecretion and specific association with the insulin resistance syndrome. Hepatology 2002, 35, 373–379. [Google Scholar] [CrossRef] [PubMed]
  16. Cassader, M.; Gambino, R.; Musso, G.; Depetris, N.; Mecca, F.; Cavallo-Perin, P.; Pacini, G.; Rizzetto, M.; Pagano, G. Postprandial triglyceride-rich lipoprotein metabolism and insulin sensitivity in nonalcoholic steatohepatitis patients. Lipids 2001, 36, 1117–1124. [Google Scholar] [CrossRef]
  17. DeFronzo, R.A.; Tripathy, D. Skeletal muscle insulin resistance is the primary defect in type 2 diabetes. Diabetes Care 2009, 32 (Suppl. S2), S157–S163. [Google Scholar] [CrossRef]
  18. DeFronzo, R.A.; Ferrannini, E.; Groop, L.; Henry, R.R.; Herman, W.H.; Holst, J.J.; Hu, F.B.; Kahn, C.R.; Raz, I.; Shulman, G.I.; et al. Type 2 diabetes mellitus. Nat. Rev. Dis. Prim. 2015, 1, 15019. [Google Scholar] [CrossRef]
  19. Shulman, G.I. Ectopic fat in insulin resistance, dyslipidemia, and cardiometabolic disease. New Engl. J. Med. 2014, 371, 2237–2238. [Google Scholar] [CrossRef]
  20. Boucher, J.; Kleinridders, A.; Kahn, C.R. Insulin receptor signaling in normal and insulin-resistant states. Cold Spring Harb. Perspect. Biol. 2014, 6, a009191. [Google Scholar] [CrossRef]
  21. Kashyap, S.R.; Belfort, R.; Berria, R.; Suraamornkul, S.; Pratipranawatr, T.; Finlayson, J.; Barrentine, A.; Bajaj, M.; Mandarino, L.; DeFronzo, R.; et al. Discordant effects of a chronic physiological increase in plasma FFA on insulin signaling in healthy subjects with or without a family history of type 2 diabetes. Am. J. Physiol. Endocrinol. Metab. 2004, 287, E537–E546. [Google Scholar] [CrossRef]
  22. Miyake, K.; Ogawa, W.; Matsumoto, M.; Nakamura, T.; Sakaue, H.; Kasuga, M. Hyperinsulinemia, glucose intolerance, and dyslipidemia induced by acute inhibition of phosphoinositide 3-kinase signaling in the liver. J. Clin. Investig. 2002, 110, 1483–1491. [Google Scholar] [CrossRef] [PubMed]
  23. Vaag, A.; Henriksen, J.E.; Beck-Nielsen, H. Decreased insulin activation of glycogen synthase in skeletal muscles in young nonobese Caucasian first-degree relatives of patients with non-insulin-dependent diabetes mellitus. J. Clin. Investig. 1992, 89, 782–788. [Google Scholar] [CrossRef] [PubMed]
  24. Mesinovic, J.; Zengin, A.; De Courten, B.; Ebeling, P.R.; Scott, D. Sarcopenia and type 2 diabetes mellitus: A bidirectional relationship. Diabetes Metab. Syndr. Obes. 2019, 12, 1057–1072. [Google Scholar] [CrossRef] [PubMed]
  25. Uehara, K.; Sostre-Colón, J.; Gavin, M.; Santoleri, D.; Leonard, K.A.; Jacobs, R.L.; Titchenell, P.M. Activation of Liver mTORC1 Protects Against NASH via Dual Regulation of VLDL-TAG Secretion and De Novo Lipogenesis. Cell. Mol. Gastroenterol. Hepatol. 2022, 13, 1625–1647. [Google Scholar] [CrossRef] [PubMed]
  26. El Ouaamari, A.; Kawamori, D.; Dirice, E.; Liew, C.W.; Shadrach, J.L.; Hu, J.; Katsuta, H.; Hollister-Lock, J.; Qian, W.J.; Wagers, A.J.; et al. Liver-derived systemic factors drive beta cell hyperplasia in insulin-resistant states. Cell Rep. 2013, 3, 401–410. [Google Scholar] [CrossRef]
  27. Rhee, E.J.; Lee, W.Y.; Cho, Y.K.; Kim, B.I.; Sung, K.C. Hyperinsulinemia and the development of nonalcoholic Fatty liver disease in nondiabetic adults. Am. J. Med. 2011, 124, 69–76. [Google Scholar] [CrossRef]
  28. Golson, M.L.; Misfeldt, A.A.; Kopsombut, U.G.; Petersen, C.P.; Gannon, M. High Fat Diet Regulation of beta-Cell Proliferation and beta-Cell Mass. Open Endocrinol. J. 2010, 4, 66–77. [Google Scholar] [CrossRef]
  29. Okada, T.; Liew, C.W.; Hu, J.; Hinault, C.; Michael, M.D.; Krtzfeldt, J.; Yin, C.; Holzenberger, M.; Stoffel, M.; Kulkarni, R.N. Insulin receptors in beta-cells are critical for islet compensatory growth response to insulin resistance. Proc. Natl. Acad. Sci. USA 2007, 104, 8977–8982. [Google Scholar] [CrossRef]
  30. Chen, Y.D.; Golay, A.; Swislocki, A.L.; Reaven, G.M. Resistance to insulin suppression of plasma free fatty acid concentrations and insulin stimulation of glucose uptake in noninsulin-dependent diabetes mellitus. J. Clin. Endocrinol. Metab. 1987, 64, 17–21. [Google Scholar] [CrossRef]
  31. Sanyal, A.J.; Campbell-Sargent, C.; Mirshahi, F.; Rizzo, W.B.; Contos, M.J.; Sterling, R.K.; Luketic, V.A.; Shiffman, M.L.; Clore, J.N. Nonalcoholic steatohepatitis: Association of insulin resistance and mitochondrial abnormalities. Gastroenterology 2001, 120, 1183–1192. [Google Scholar] [CrossRef]
  32. Perry, R.J.; Camporez, J.G.; Kursawe, R.; Titchenell, P.M.; Zhang, D.; Perry, C.J.; Jurczak, M.J.; Abudukadier, A.; Han, M.S.; Zhang, X.M.; et al. Hepatic acetyl CoA links adipose tissue inflammation to hepatic insulin resistance and type 2 diabetes. Cell 2015, 160, 745–758. [Google Scholar] [CrossRef] [PubMed]
  33. Vatner, D.F.; Majumdar, S.K.; Kumashiro, N.; Petersen, M.C.; Rahimi, Y.; Gattu, A.K.; Bears, M.; Camporez, J.P.; Cline, G.W.; Jurczak, M.J.; et al. Insulin-independent regulation of hepatic triglyceride synthesis by fatty acids. Proc. Natl. Acad. Sci. USA 2015, 112, 1143–1148. [Google Scholar] [CrossRef] [PubMed]
  34. Petersen, M.C.; Madiraju, A.K.; Gassaway, B.M.; Marcel, M.; Nasiri, A.R.; Butrico, G.; Marcucci, M.J.; Zhang, D.; Abulizi, A.; Zhang, X.M.; et al. Insulin receptor Thr1160 phosphorylation mediates lipid-induced hepatic insulin resistance. J. Clin. Investig. 2016, 126, 4361–4371. [Google Scholar] [CrossRef] [PubMed]
  35. Kalemba, K.M.; Wang, Y.; Xu, H.; Chiles, E.; McMillin, S.M.; Kwon, H.; Su, X.; Wondisford, F.E. Glycerol induces G6pc in primary mouse hepatocytes and is the preferred substrate for gluconeogenesis both in vitro and in vivo. J. Biol. Chem. 2019, 294, 18017–18028. [Google Scholar] [CrossRef]
  36. Nurjhan, N.; Consoli, A.; Gerich, J. Increased lipolysis and its consequences on gluconeogenesis in non-insulin-dependent diabetes mellitus. J. Clin. Investig. 1992, 89, 169–175. [Google Scholar] [CrossRef]
  37. Jin, E.S.; Browning, J.D.; Murphy, R.E.; Malloy, C.R. Fatty liver disrupts glycerol metabolism in gluconeogenic and lipogenic pathways in humans. J. Lipid Res. 2018, 59, 1685–1694. [Google Scholar] [CrossRef]
  38. Miele, L.; Grieco, A.; Armuzzi, A.; Candelli, M.; Forgione, A.; Gasbarrini, A.; Gasbarrini, G. Hepatic mitochondrial beta-oxidation in patients with nonalcoholic steatohepatitis assessed by 13C-octanoate breath test. Am. J. Gastroenterol. 2003, 98, 2335–2336. [Google Scholar] [CrossRef]
  39. Koliaki, C.; Szendroedi, J.; Kaul, K.; Jelenik, T.; Nowotny, P.; Jankowiak, F.; Herder, C.; Carstensen, M.; Krausch, M.; Knoefel, W.T.; et al. Adaptation of hepatic mitochondrial function in humans with non-alcoholic fatty liver is lost in steatohepatitis. Cell Metab. 2015, 21, 739–746. [Google Scholar] [CrossRef]
  40. Yu, J.; Loh, K.; Song, Z.Y.; Yang, H.Q.; Zhang, Y.; Lin, S. Update on glycerol-3-phosphate acyltransferases: The roles in the development of insulin resistance. Nutr. Diabetes 2018, 8, 34. [Google Scholar] [CrossRef]
  41. Parks, E.J.; Krauss, R.M.; Christiansen, M.P.; Neese, R.A.; Hellerstein, M.K. Effects of a low-fat, high-carbohydrate diet on VLDL-triglyceride assembly, production, and clearance. J. Clin. Investig. 1999, 104, 1087–1096. [Google Scholar] [CrossRef]
  42. Parks, E.J.; Skokan, L.E.; Timlin, M.T.; Dingfelder, C.S. Dietary sugars stimulate fatty acid synthesis in adults. J. Nutr. 2008, 138, 1039–1046. [Google Scholar] [CrossRef] [PubMed]
  43. Sun, S.Z.; Empie, M.W. Fructose metabolism in humans–what isotopic tracer studies tell us. Nutr. Metab. 2012, 9, 89. [Google Scholar] [CrossRef] [PubMed]
  44. Domènech, M.; López-Soriano, F.J.; Argilés, J.M. Alanine as a lipogenic precursor in isolated hepatocytes from obese Zucker rats. Cell Mol. Biol. 1993, 39, 693–699. [Google Scholar]
  45. Palacín, M.; Lasunción, M.A.; Herrera, E. Utilization of glucose, alanine, lactate, and glycerol as lipogenic substrates by periuterine adipose tissue in situ in fed and starved rats. J. Lipid Res. 1988, 29, 26–32. [Google Scholar] [CrossRef] [PubMed]
  46. Zhao, S.; Jang, C.; Liu, J.; Uehara, K.; Gilbert, M.; Izzo, L.; Zeng, X.; Trefely, S.; Fernandez, S.; Carrer, A.; et al. Dietary fructose feeds hepatic lipogenesis via microbiota-derived acetate. Nature 2020, 579, 586–591. [Google Scholar] [CrossRef] [PubMed]
  47. Zhang, Z.; TeSlaa, T.; Xu, X.; Zeng, X.; Yang, L.; Xing, G.; Tesz, G.J.; Clasquin, M.F.; Rabinowitz, J.D. Serine catabolism generates liver NADPH and supports hepatic lipogenesis. Nat. Metab. 2021, 3, 1608–1620. [Google Scholar] [CrossRef]
  48. Ipsen, D.H.; Lykkesfeldt, J.; Tveden-Nyborg, P. Molecular mechanisms of hepatic lipid accumulation in non-alcoholic fatty liver disease. Cell. Mol. Life Sci. CMLS 2018, 75, 3313–3327. [Google Scholar] [CrossRef]
  49. Fletcher, J.A.; Deja, S.; Satapati, S.; Fu, X.; Burgess, S.C.; Browning, J.D. Impaired ketogenesis and increased acetyl-CoA oxidation promote hyperglycemia in human fatty liver. JCI Insight 2019, 5, e127737. [Google Scholar] [CrossRef]
  50. Rector, R.S.; Thyfault, J.P.; Uptergrove, G.M.; Morris, E.M.; Naples, S.P.; Borengasser, S.J.; Mikus, C.R.; Laye, M.J.; Laughlin, M.H.; Booth, F.W.; et al. Mitochondrial dysfunction precedes insulin resistance and hepatic steatosis and contributes to the natural history of non-alcoholic fatty liver disease in an obese rodent model. J. Hepatol. 2010, 52, 727–736. [Google Scholar] [CrossRef]
  51. Abou Ziki, M.D.; Mani, A. Metabolic syndrome: Genetic insights into disease pathogenesis. Curr. Opin. Lipidol. 2016, 27, 162–171. [Google Scholar] [CrossRef]
  52. Hulman, A.; Witte, D.R.; Vistisen, D.; Balkau, B.; Dekker, J.M.; Herder, C.; Hatunic, M.; Konrad, T.; Faerch, K.; Manco, M. Pathophysiological Characteristics Underlying Different Glucose Response Curves: A Latent Class Trajectory Analysis From the Prospective EGIR-RISC Study. Diabetes Care 2018, 41, 1740–1748. [Google Scholar] [CrossRef] [PubMed]
  53. Elksnis, A.; Martinell, M.; Eriksson, O.; Espes, D. Heterogeneity of Metabolic Defects in Type 2 Diabetes and Its Relation to Reactive Oxygen Species and Alterations in Beta-Cell Mass. Front. Physiol. 2019, 10, 107. [Google Scholar] [CrossRef] [PubMed]
  54. Ishimoto, T.; Lanaspa, M.A.; Rivard, C.J.; Roncal-Jimenez, C.A.; Orlicky, D.J.; Cicerchi, C.; McMahan, R.H.; Abdelmalek, M.F.; Rosen, H.R.; Jackman, M.R.; et al. High-fat and high-sucrose (western) diet induces steatohepatitis that is dependent on fructokinase. Hepatology 2013, 58, 1632–1643. [Google Scholar] [CrossRef] [PubMed]
  55. Jürgens, H.; Haass, W.; Castañeda, T.R.; Schürmann, A.; Koebnick, C.; Dombrowski, F.; Otto, B.; Nawrocki, A.R.; Scherer, P.E.; Spranger, J.; et al. Consuming fructose-sweetened beverages increases body adiposity in mice. Obes. Res. 2005, 13, 1146–1156. [Google Scholar] [CrossRef]
  56. Bhat, N.; Narayanan, A.; Fathzadeh, M.; Kahn, M.; Zhang, D.; Goedeke, L.; Neogi, A.; Cardone, R.L.; Kibbey, R.G.; Fernandez-Hernando, C.; et al. Dyrk1b promotes hepatic lipogenesis by bypassing canonical insulin signaling and directly activating mTORC2 in mice. J. Clin. Investig. 2021, 132, e153724. [Google Scholar] [CrossRef]
  57. Lyu, K.; Zhang, Y.; Zhang, D.; Kahn, M.; Ter Horst, K.W.; Rodrigues, M.R.S.; Gaspar, R.C.; Hirabara, S.M.; Luukkonen, P.K.; Lee, S.; et al. A Membrane-Bound Diacylglycerol Species Induces PKC-Mediated Hepatic Insulin Resistance. Cell Metab. 2020, 32, 654–664.e655. [Google Scholar] [CrossRef]
  58. Aryal, B.; Singh, A.K.; Zhang, X.; Varela, L.; Rotllan, N.; Goedeke, L.; Chaube, B.; Camporez, J.P.; Vatner, D.F.; Horvath, T.L.; et al. Absence of ANGPTL4 in adipose tissue improves glucose tolerance and attenuates atherogenesis. JCI Insight 2018, 3, e97918. [Google Scholar] [CrossRef]
  59. Samuel, V.T.; Liu, Z.X.; Wang, A.; Beddow, S.A.; Geisler, J.G.; Kahn, M.; Zhang, X.M.; Monia, B.P.; Bhanot, S.; Shulman, G.I. Inhibition of protein kinase Cepsilon prevents hepatic insulin resistance in nonalcoholic fatty liver disease. J. Clin. Investig. 2007, 117, 739–745. [Google Scholar] [CrossRef]
  60. Rinella, M.E. Nonalcoholic fatty liver disease: A systematic review. Jama 2015, 313, 2263–2273. [Google Scholar] [CrossRef]
  61. Sun, Z.; Lazar, M.A. Dissociating fatty liver and diabetes. Trends Endocrinol. Metab. 2013, 24, 4–12. [Google Scholar] [CrossRef]
  62. Shanik, M.H.; Xu, Y.; Skrha, J.; Dankner, R.; Zick, Y.; Roth, J. Insulin resistance and hyperinsulinemia: Is hyperinsulinemia the cart or the horse? Diabetes Care 2008, 31 (Suppl. S2), S262–S268. [Google Scholar] [CrossRef] [PubMed]
  63. Brown, M.S.; Goldstein, J.L. Selective versus total insulin resistance: A pathogenic paradox. Cell Metab. 2008, 7, 95–96. [Google Scholar] [CrossRef] [PubMed]
  64. Han, J.; Wang, Y. mTORC1 signaling in hepatic lipid metabolism. Protein Cell 2018, 9, 145–151. [Google Scholar] [CrossRef] [PubMed]
  65. Saxton, R.A.; Sabatini, D.M. mTOR Signaling in Growth, Metabolism, and Disease. Cell 2017, 168, 960–976. [Google Scholar] [CrossRef] [PubMed]
  66. Peterson, T.R.; Sengupta, S.S.; Harris, T.E.; Carmack, A.E.; Kang, S.A.; Balderas, E.; Guertin, D.A.; Madden, K.L.; Carpenter, A.E.; Finck, B.N.; et al. mTOR complex 1 regulates lipin 1 localization to control the SREBP pathway. Cell 2011, 146, 408–420. [Google Scholar] [CrossRef] [PubMed]
  67. Ardestani, A.; Lupse, B.; Kido, Y.; Leibowitz, G.; Maedler, K. mTORC1 Signaling: A Double-Edged Sword in Diabetic beta Cells. Cell Metab. 2018, 27, 314–331. [Google Scholar] [CrossRef]
  68. Aguirre, V.; Werner, E.D.; Giraud, J.; Lee, Y.H.; Shoelson, S.E.; White, M.F. Phosphorylation of Ser307 in insulin receptor substrate-1 blocks interactions with the insulin receptor and inhibits insulin action. J. Biol. Chem. 2002, 277, 1531–1537. [Google Scholar] [CrossRef]
  69. Um, S.H.; Frigerio, F.; Watanabe, M.; Picard, F.; Joaquin, M.; Sticker, M.; Fumagalli, S.; Allegrini, P.R.; Kozma, S.C.; Auwerx, J.; et al. Absence of S6K1 protects against age- and diet-induced obesity while enhancing insulin sensitivity. Nature 2004, 431, 200–205. [Google Scholar] [CrossRef]
  70. Nguyen-Lefebvre, A.T.; Horuzsko, A. Kupffer Cell Metabolism and Function. J. Enzymol. Metab. 2015, 1, 101. [Google Scholar]
  71. Mayoral Monibas, R.; Johnson, A.M.; Osborn, O.; Traves, P.G.; Mahata, S.K. Distinct Hepatic Macrophage Populations in Lean and Obese Mice. Front. Endocrinol. 2016, 7, 152. [Google Scholar] [CrossRef]
  72. Morinaga, H.; Mayoral, R.; Heinrichsdorff, J.; Osborn, O.; Franck, N.; Hah, N.; Walenta, E.; Bandyopadhyay, G.; Pessentheiner, A.R.; Chi, T.J.; et al. Characterization of distinct subpopulations of hepatic macrophages in HFD/obese mice. Diabetes 2015, 64, 1120–1130. [Google Scholar] [CrossRef] [PubMed]
  73. Zigmond, E.; Samia-Grinberg, S.; Pasmanik-Chor, M.; Brazowski, E.; Shibolet, O.; Halpern, Z.; Varol, C. Infiltrating monocyte-derived macrophages and resident kupffer cells display different ontogeny and functions in acute liver injury. J. Immunol. 2014, 193, 344–353. [Google Scholar] [CrossRef] [PubMed]
  74. Tacke, F. Targeting hepatic macrophages to treat liver diseases. J. Hepatol. 2017, 66, 1300–1312. [Google Scholar] [CrossRef] [PubMed]
  75. Tencerova, M.; Aouadi, M.; Vangala, P.; Nicoloro, S.M.; Yawe, J.C.; Cohen, J.L.; Shen, Y.; Garcia-Menendez, L.; Pedersen, D.J.; Gallagher-Dorval, K.; et al. Activated Kupffer cells inhibit insulin sensitivity in obese mice. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2015, 29, 2959–2969. [Google Scholar] [CrossRef]
  76. Huang, W.; Metlakunta, A.; Dedousis, N.; Zhang, P.; Sipula, I.; Dube, J.J.; Scott, D.K.; O’Doherty, R.M. Depletion of liver Kupffer cells prevents the development of diet-induced hepatic steatosis and insulin resistance. Diabetes 2010, 59, 347–357. [Google Scholar] [CrossRef]
  77. Lanthier, N.; Molendi-Coste, O.; Horsmans, Y.; van Rooijen, N.; Cani, P.D.; Leclercq, I.A. Kupffer cell activation is a causal factor for hepatic insulin resistance. Am. J. Physiol. Gastrointest Liver Physiol. 2010, 298, G107–G116. [Google Scholar] [CrossRef]
  78. Tsuchida, T.; Friedman, S.L. Mechanisms of hepatic stellate cell activation. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 397–411. [Google Scholar] [CrossRef]
  79. Papackova, Z.; Palenickova, E.; Dankova, H.; Zdychova, J.; Skop, V.; Kazdova, L.; Cahova, M. Kupffer cells ameliorate hepatic insulin resistance induced by high-fat diet rich in monounsaturated fatty acids: The evidence for the involvement of alternatively activated macrophages. Nutr. Metab. 2012, 9, 22. [Google Scholar] [CrossRef]
  80. Clementi, A.H.; Gaudy, A.M.; van Rooijen, N.; Pierce, R.H.; Mooney, R.A. Loss of Kupffer cells in diet-induced obesity is associated with increased hepatic steatosis, STAT3 signaling, and further decreases in insulin signaling. Biochim. Biophys. Acta 2009, 1792, 1062–1072. [Google Scholar] [CrossRef]
  81. Odegaard, J.I.; Ricardo-Gonzalez, R.R.; Red Eagle, A.; Vats, D.; Morel, C.R.; Goforth, M.H.; Subramanian, V.; Mukundan, L.; Ferrante, A.W.; Chawla, A. Alternative M2 activation of Kupffer cells by PPARdelta ameliorates obesity-induced insulin resistance. Cell Metab. 2008, 7, 496–507. [Google Scholar] [CrossRef]
  82. Foretz, M.; Guichard, C.; Ferré, P.; Foufelle, F. Sterol regulatory element binding protein-1c is a major mediator of insulin action on the hepatic expression of glucokinase and lipogenesis-related genes. Proc. Natl. Acad. Sci. USA 1999, 96, 12737–12742. [Google Scholar] [CrossRef] [PubMed]
  83. Horton, J.D.; Goldstein, J.L.; Brown, M.S. SREBPs: Activators of the complete program of cholesterol and fatty acid synthesis in the liver. J. Clin. Investig. 2002, 109, 1125–1131. [Google Scholar] [CrossRef] [PubMed]
  84. Liang, G.; Yang, J.; Horton, J.D.; Hammer, R.E.; Goldstein, J.L.; Brown, M.S. Diminished hepatic response to fasting/refeeding and liver X receptor agonists in mice with selective deficiency of sterol regulatory element-binding protein-1c. J. Biol. Chem. 2002, 277, 9520–9528. [Google Scholar] [CrossRef] [PubMed]
  85. Linden, A.G.; Li, S.; Choi, H.Y.; Fang, F.; Fukasawa, M.; Uyeda, K.; Hammer, R.E.; Horton, J.D.; Engelking, L.J.; Liang, G. Interplay between ChREBP and SREBP-1c coordinates postprandial glycolysis and lipogenesis in livers of mice. J. Lipid Res. 2018, 59, 475–487. [Google Scholar] [CrossRef] [PubMed]
  86. Owen, J.L.; Zhang, Y.; Bae, S.H.; Farooqi, M.S.; Liang, G.; Hammer, R.E.; Goldstein, J.L.; Brown, M.S. Insulin stimulation of SREBP-1c processing in transgenic rat hepatocytes requires p70 S6-kinase. Proc. Natl. Acad. Sci. USA 2012, 109, 16184–16189. [Google Scholar] [CrossRef]
  87. Shimomura, I.; Bashmakov, Y.; Ikemoto, S.; Horton, J.D.; Brown, M.S.; Goldstein, J.L. Insulin selectively increases SREBP-1c mRNA in the livers of rats with streptozotocin-induced diabetes. Proc. Natl. Acad. Sci. USA 1999, 96, 13656–13661. [Google Scholar] [CrossRef]
  88. Shimomura, I.; Matsuda, M.; Hammer, R.E.; Bashmakov, Y.; Brown, M.S.; Goldstein, J.L. Decreased IRS-2 and increased SREBP-1c lead to mixed insulin resistance and sensitivity in livers of lipodystrophic and ob/ob mice. Mol. Cell 2000, 6, 77–86. [Google Scholar] [CrossRef]
  89. Yahagi, N.; Shimano, H.; Hasty, A.H.; Matsuzaka, T.; Ide, T.; Yoshikawa, T.; Amemiya-Kudo, M.; Tomita, S.; Okazaki, H.; Tamura, Y.; et al. Absence of sterol regulatory element-binding protein-1 (SREBP-1) ameliorates fatty livers but not obesity or insulin resistance in Lep(ob)/Lep(ob) mice. J. Biol. Chem. 2002, 277, 19353–19357. [Google Scholar] [CrossRef]
  90. Benhamed, F.; Denechaud, P.D.; Lemoine, M.; Robichon, C.; Moldes, M.; Bertrand-Michel, J.; Ratziu, V.; Serfaty, L.; Housset, C.; Capeau, J.; et al. The lipogenic transcription factor ChREBP dissociates hepatic steatosis from insulin resistance in mice and humans. J. Clin. Investig. 2012, 122, 2176–2194. [Google Scholar] [CrossRef]
  91. Hagiwara, A.; Cornu, M.; Cybulski, N.; Polak, P.; Betz, C.; Trapani, F.; Terracciano, L.; Heim, M.H.; Ruegg, M.A.; Hall, M.N. Hepatic mTORC2 activates glycolysis and lipogenesis through Akt, glucokinase, and SREBP1c. Cell Metab. 2012, 15, 725–738. [Google Scholar] [CrossRef]
  92. Kim, M.S.; Krawczyk, S.A.; Doridot, L.; Fowler, A.J.; Wang, J.X.; Trauger, S.A.; Noh, H.L.; Kang, H.J.; Meissen, J.K.; Blatnik, M.; et al. ChREBP regulates fructose-induced glucose production independently of insulin signaling. J. Clin. Investig. 2016, 126, 4372–4386. [Google Scholar] [CrossRef] [PubMed]
  93. Ma, L.; Tsatsos, N.G.; Towle, H.C. Direct role of ChREBP.Mlx in regulating hepatic glucose-responsive genes. J. Biol. Chem. 2005, 280, 12019–12027. [Google Scholar] [CrossRef] [PubMed]
  94. Uyeda, K.; Repa, J.J. Carbohydrate response element binding protein, ChREBP, a transcription factor coupling hepatic glucose utilization and lipid synthesis. Cell Metab. 2006, 4, 107–110. [Google Scholar] [CrossRef] [PubMed]
  95. Dentin, R.; Pégorier, J.P.; Benhamed, F.; Foufelle, F.; Ferré, P.; Fauveau, V.; Magnuson, M.A.; Girard, J.; Postic, C. Hepatic glucokinase is required for the synergistic action of ChREBP and SREBP-1c on glycolytic and lipogenic gene expression. J. Biol. Chem. 2004, 279, 20314–20326. [Google Scholar] [CrossRef]
  96. Matsuda, T.; Noguchi, T.; Yamada, K.; Takenaka, M.; Tanaka, T. Regulation of the gene expression of glucokinase and L-type pyruvate kinase in primary cultures of rat hepatocytes by hormones and carbohydrates. J. Biochem. 1990, 108, 778–784. [Google Scholar] [CrossRef]
  97. Dong, Q.; Giorgianni, F.; Deng, X.; Beranova-Giorgianni, S.; Bridges, D.; Park, E.A.; Raghow, R.; Elam, M.B. Phosphorylation of sterol regulatory element binding protein-1a by protein kinase A (PKA) regulates transcriptional activity. Biochem. Biophys. Res. Commun. 2014, 449, 449–454. [Google Scholar] [CrossRef]
  98. Lu, M.; Shyy, J.Y. Sterol regulatory element-binding protein 1 is negatively modulated by PKA phosphorylation. Am. J. Physiol. Cell Physiol. 2006, 290, C1477–C1486. [Google Scholar] [CrossRef]
  99. Katz, L.S.; Baumel-Alterzon, S.; Scott, D.K.; Herman, M.A. Adaptive and maladaptive roles for ChREBP in the liver and pancreatic islets. J. Biol. Chem. 2021, 296, 100623. [Google Scholar] [CrossRef]
  100. Iizuka, K.; Takao, K.; Yabe, D. ChREBP-Mediated Regulation of Lipid Metabolism: Involvement of the Gut Microbiota Liver, and Adipose Tissue. Front. Endocrinol. 2020, 11, 587189. [Google Scholar] [CrossRef]
  101. Wang, S.; Song, K.; Srivastava, R.; Dong, C.; Go, G.W.; Li, N.; Iwakiri, Y.; Mani, A. Nonalcoholic fatty liver disease induced by noncanonical Wnt and its rescue by Wnt3a. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2015, 29, 3436–3445. [Google Scholar] [CrossRef]
  102. Go, G.W.; Srivastava, R.; Hernandez-Ono, A.; Gang, G.; Smith, S.B.; Booth, C.J.; Ginsberg, H.N.; Mani, A. The combined hyperlipidemia caused by impaired Wnt-LRP6 signaling is reversed by Wnt3a rescue. Cell Metab. 2014, 19, 209–220. [Google Scholar] [CrossRef] [PubMed]
  103. Bhat, N.; Esteghamat, F.; Chaube, B.K.; Gunawardhana, K.; Mani, M.; Thames, C.; Jain, D.; Ginsberg, H.N.; Fernandes-Hernando, C.; Mani, A. TCF7L2 transcriptionally regulates Fgf15 to maintain bile acid and lipid homeostasis through gut-liver crosstalk. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2022, 36, e22185. [Google Scholar] [CrossRef] [PubMed]
  104. Saponara, E.; Penno, C.; Orsini, V.; Wang, Z.Y.; Fischer, A.; Aebi, A.; Matadamas-Guzman, M.L.; Brun, V.; Fischer, B.; Brousseau, M.; et al. Loss of Hepatic Leucine-Rich Repeat-Containing G-Protein Coupled Receptors 4 and 5 Promotes Nonalcoholic Fatty Liver Disease. Am. J. Pathol. 2023, 193, 161–181. [Google Scholar] [CrossRef] [PubMed]
  105. Keramati, A.R.; Fathzadeh, M.; Go, G.W.; Singh, R.; Choi, M.; Faramarzi, S.; Mane, S.; Kasaei, M.; Sarajzadeh-Fard, K.; Hwa, J.; et al. A form of the metabolic syndrome associated with mutations in DYRK1B. New Engl. J. Med. 2014, 370, 1909–1919. [Google Scholar] [CrossRef] [PubMed]
  106. Bhat, N.; Narayanan, A.; Fathzadeh, M.; Shah, K.; Dianatpour, M.; Abou Ziki, M.D.; Mani, A. Dyrk1b promotes autophagy during skeletal muscle differentiation by upregulating 4e-bp1. Cell. Signal. 2021, 90, 110186. [Google Scholar] [CrossRef]
  107. Leder, S.; Czajkowska, H.; Maenz, B.; De Graaf, K.; Barthel, A.; Joost, H.G.; Becker, W. Alternative splicing variants of dual specificity tyrosine phosphorylated and regulated kinase 1B exhibit distinct patterns of expression and functional properties. Biochem. J. 2003, 372, 881–888. [Google Scholar] [CrossRef]
  108. Leder, S.; Weber, Y.; Altafaj, X.; Estivill, X.; Joost, H.G.; Becker, W. Cloning and characterization of DYRK1B, a novel member of the DYRK family of protein kinases. Biochem. Biophys. Res. Commun. 1999, 254, 474–479. [Google Scholar] [CrossRef]
  109. Ashford, A.L.; Dunkley, T.P.; Cockerill, M.; Rowlinson, R.A.; Baak, L.M.; Gallo, R.; Balmanno, K.; Goodwin, L.M.; Ward, R.A.; Lochhead, P.A.; et al. Identification of DYRK1B as a substrate of ERK1/2 and characterisation of the kinase activity of DYRK1B mutants from cancer and metabolic syndrome. Cell. Mol. Life Sci. CMLS 2016, 73, 883–900. [Google Scholar] [CrossRef]
  110. Deng, X.; Ewton, D.Z.; Pawlikowski, B.; Maimone, M.; Friedman, E. Mirk/dyrk1B is a Rho-induced kinase active in skeletal muscle differentiation. J. Biol. Chem. 2003, 278, 41347–41354. [Google Scholar] [CrossRef]
  111. Deng, X.; Hu, J.; Ewton, D.Z.; Friedman, E. Mirk/dyrk1B kinase is upregulated following inhibition of mTOR. Carcinogenesis 2014, 35, 1968–1976. [Google Scholar] [CrossRef]
  112. Smith, G.I.; Shankaran, M.; Yoshino, M.; Schweitzer, G.G.; Chondronikola, M.; Beals, J.W.; Okunade, A.L.; Patterson, B.W.; Nyangau, E.; Field, T.; et al. Insulin resistance drives hepatic de novo lipogenesis in nonalcoholic fatty liver disease. J. Clin. Investig. 2020, 130, 1453–1460. [Google Scholar] [CrossRef] [PubMed]
  113. Nielsen, J.; Brandt, J.; Boesen, T.; Hummelshøj, T.; Slaaby, R.; Schluckebier, G.; Nissen, P. Structural Investigations of Full-Length Insulin Receptor Dynamics and Signalling. J. Mol. Biol. 2022, 434, 167458. [Google Scholar] [CrossRef] [PubMed]
  114. De Meyts, P. The insulin receptor: A prototype for dimeric, allosteric membrane receptors? Trends Biochem. Sci. 2008, 33, 376–384. [Google Scholar] [CrossRef]
  115. Ebina, Y.; Ellis, L.; Jarnagin, K.; Edery, M.; Graf, L.; Clauser, E.; Ou, J.H.; Masiarz, F.; Kan, Y.W.; Goldfine, I.D.; et al. The human insulin receptor cDNA: The structural basis for hormone-activated transmembrane signalling. Cell 1985, 40, 747–758. [Google Scholar] [CrossRef] [PubMed]
  116. Ullrich, A.; Bell, J.R.; Chen, E.Y.; Herrera, R.; Petruzzelli, L.M.; Dull, T.J.; Gray, A.; Coussens, L.; Liao, Y.C.; Tsubokawa, M.; et al. Human insulin receptor and its relationship to the tyrosine kinase family of oncogenes. Nature 1985, 313, 756–761. [Google Scholar] [CrossRef] [PubMed]
  117. McKern, N.M.; Lawrence, M.C.; Streltsov, V.A.; Lou, M.Z.; Adams, T.E.; Lovrecz, G.O.; Elleman, T.C.; Richards, K.M.; Bentley, J.D.; Pilling, P.A.; et al. Structure of the insulin receptor ectodomain reveals a folded-over conformation. Nature 2006, 443, 218–221. [Google Scholar] [CrossRef]
  118. Menting, J.G.; Whittaker, J.; Margetts, M.B.; Whittaker, L.J.; Kong, G.K.; Smith, B.J.; Watson, C.J.; Záková, L.; Kletvíková, E.; Jiráček, J.; et al. How insulin engages its primary binding site on the insulin receptor. Nature 2013, 493, 241–245. [Google Scholar] [CrossRef]
  119. Scapin, G.; Dandey, V.P.; Zhang, Z.; Prosise, W.; Hruza, A.; Kelly, T.; Mayhood, T.; Strickland, C.; Potter, C.S.; Carragher, B. Structure of the insulin receptor-insulin complex by single-particle cryo-EM analysis. Nature 2018, 556, 122–125. [Google Scholar] [CrossRef]
  120. Cook, J.R.; Langlet, F.; Kido, Y.; Accili, D. Pathogenesis of selective insulin resistance in isolated hepatocytes. J. Biol. Chem. 2015, 290, 13972–13980. [Google Scholar] [CrossRef]
  121. Li, S.; Brown, M.S.; Goldstein, J.L. Bifurcation of insulin signaling pathway in rat liver: mTORC1 required for stimulation of lipogenesis, but not inhibition of gluconeogenesis. Proc. Natl. Acad. Sci. USA 2010, 107, 3441–3446. [Google Scholar] [CrossRef]
  122. Titchenell, P.M.; Quinn, W.J.; Lu, M.; Chu, Q.; Lu, W.; Li, C.; Chen, H.; Monks, B.R.; Chen, J.; Rabinowitz, J.D.; et al. Direct Hepatocyte Insulin Signaling Is Required for Lipogenesis but Is Dispensable for the Suppression of Glucose Production. Cell Metab. 2016, 23, 1154–1166. [Google Scholar] [CrossRef] [PubMed]
  123. Honma, M.; Sawada, S.; Ueno, Y.; Murakami, K.; Yamada, T.; Gao, J.; Kodama, S.; Izumi, T.; Takahashi, K.; Tsukita, S.; et al. Selective insulin resistance with differential expressions of IRS-1 and IRS-2 in human NAFLD livers. Int. J. Obes. 2018, 42, 1544–1555. [Google Scholar] [CrossRef] [PubMed]
  124. Bouzakri, K.; Zachrisson, A.; Al-Khalili, L.; Zhang, B.B.; Koistinen, H.A.; Krook, A.; Zierath, J.R. siRNA-based gene silencing reveals specialized roles of IRS-1/Akt2 and IRS-2/Akt1 in glucose and lipid metabolism in human skeletal muscle. Cell Metab. 2006, 4, 89–96. [Google Scholar] [CrossRef] [PubMed]
  125. Croci, I.; Byrne, N.M.; Choquette, S.; Hills, A.P.; Chachay, V.S.; Clouston, A.D.; O’Moore-Sullivan, T.M.; Macdonald, G.A.; Prins, J.B.; Hickman, I.J. Whole-body substrate metabolism is associated with disease severity in patients with non-alcoholic fatty liver disease. Gut 2013, 62, 1625–1633. [Google Scholar] [CrossRef]
  126. Czech, M.P.; Tencerova, M.; Pedersen, D.J.; Aouadi, M. Insulin signalling mechanisms for triacylglycerol storage. Diabetologia 2013, 56, 949–964. [Google Scholar] [CrossRef]
  127. Gassaway, B.M.; Petersen, M.C.; Surovtseva, Y.V.; Barber, K.W.; Sheetz, J.B.; Aerni, H.R.; Merkel, J.S.; Samuel, V.T.; Shulman, G.I.; Rinehart, J. PKCepsilon contributes to lipid-induced insulin resistance through cross talk with p70S6K and through previously unknown regulators of insulin signaling. Proc. Natl. Acad. Sci. USA 2018, 115, E8996–E9005. [Google Scholar] [CrossRef]
  128. Dong, X.C.; Copps, K.D.; Guo, S.; Li, Y.; Kollipara, R.; DePinho, R.A.; White, M.F. Inactivation of hepatic Foxo1 by insulin signaling is required for adaptive nutrient homeostasis and endocrine growth regulation. Cell Metab. 2008, 8, 65–76. [Google Scholar] [CrossRef]
  129. Yuan, M.; Pino, E.; Wu, L.; Kacergis, M.; Soukas, A.A. Identification of Akt-independent regulation of hepatic lipogenesis by mammalian target of rapamycin (mTOR) complex 2. J. Biol. Chem. 2012, 287, 29579–29588. [Google Scholar] [CrossRef]
  130. Lamming, D.W.; Ye, L.; Katajisto, P.; Goncalves, M.D.; Saitoh, M.; Stevens, D.M.; Davis, J.G.; Salmon, A.B.; Richardson, A.; Ahima, R.S.; et al. Rapamycin-induced insulin resistance is mediated by mTORC2 loss and uncoupled from longevity. Science 2012, 335, 1638–1643. [Google Scholar] [CrossRef]
  131. Quinn, W.J., 3rd; Wan, M.; Shewale, S.V.; Gelfer, R.; Rader, D.J.; Birnbaum, M.J.; Titchenell, P.M. mTORC1 stimulates phosphatidylcholine synthesis to promote triglyceride secretion. J. Clin. Investig. 2017, 127, 4207–4215. [Google Scholar] [CrossRef]
  132. Yecies, J.L.; Zhang, H.H.; Menon, S.; Liu, S.; Yecies, D.; Lipovsky, A.I.; Gorgun, C.; Kwiatkowski, D.J.; Hotamisligil, G.S.; Lee, C.H.; et al. Akt stimulates hepatic SREBP1c and lipogenesis through parallel mTORC1-dependent and independent pathways. Cell Metab. 2011, 14, 21–32. [Google Scholar] [CrossRef] [PubMed]
  133. Kenerson, H.L.; Yeh, M.M.; Yeung, R.S. Tuberous sclerosis complex-1 deficiency attenuates diet-induced hepatic lipid accumulation. PLoS ONE 2011, 6, e18075. [Google Scholar] [CrossRef] [PubMed]
  134. Gosis, B.S.; Wada, S.; Thorsheim, C.; Li, K.; Jung, S.; Rhoades, J.H.; Yang, Y.; Brandimarto, J.; Li, L.; Uehara, K.; et al. Inhibition of nonalcoholic fatty liver disease in mice by selective inhibition of mTORC1. Science 2022, 376, eabf8271. [Google Scholar] [CrossRef]
  135. 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]
  136. Kantartzis, K.; Peter, A.; Machicao, F.; Machann, J.; Wagner, S.; Königsrainer, I.; Königsrainer, A.; Schick, F.; Fritsche, A.; Häring, H.U.; et al. Dissociation between fatty liver and insulin resistance in humans carrying a variant of the patatin-like phospholipase 3 gene. Diabetes 2009, 58, 2616–2623. [Google Scholar] [CrossRef]
  137. Romeo, S.; Kozlitina, J.; Xing, C.; Pertsemlidis, A.; Cox, D.; Pennacchio, L.A.; Boerwinkle, E.; Cohen, J.C.; Hobbs, H.H. Genetic variation in PNPLA3 confers susceptibility to nonalcoholic fatty liver disease. Nat. Genet. 2008, 40, 1461–1465. [Google Scholar] [CrossRef] [PubMed]
  138. Sevastianova, K.; Kotronen, A.; Gastaldelli, A.; Perttilä, J.; Hakkarainen, A.; Lundbom, J.; Suojanen, L.; Orho-Melander, M.; Lundbom, N.; Ferrannini, E.; et al. Genetic variation in PNPLA3 (adiponutrin) confers sensitivity to weight loss-induced decrease in liver fat in humans. Am. J. Clin. Nutr. 2011, 94, 104–111. [Google Scholar] [CrossRef]
  139. Friedman, S.L.; Ratziu, V.; Harrison, S.A.; Abdelmalek, M.F.; Aithal, G.P.; Caballeria, J.; Francque, S.; Farrell, G.; Kowdley, K.V.; Craxi, A.; et al. A randomized, placebo-controlled trial of cenicriviroc for treatment of nonalcoholic steatohepatitis with fibrosis. Hepatology 2018, 67, 1754–1767. [Google Scholar] [CrossRef]
  140. Boeckmans, J.; Natale, A.; Rombaut, M.; Buyl, K.; Rogiers, V.; De Kock, J.; Vanhaecke, T.; Rodrigues, M.R. Anti-NASH Drug Development Hitches a Lift on PPAR Agonism. Cells 2019, 9, 37. [Google Scholar] [CrossRef]
  141. Francque, S.M.; Bedossa, P.; Ratziu, V.; Anstee, Q.M.; Bugianesi, E.; Sanyal, A.J.; Loomba, R.; Harrison, S.A.; Balabanska, R.; Mateva, L.; et al. A Randomized, Controlled Trial of the Pan-PPAR Agonist Lanifibranor in NASH. New Engl. J. Med. 2021, 385, 1547–1558. [Google Scholar] [CrossRef]
  142. Gawrieh, S.; Noureddin, M.; Loo, N.; Mohseni, R.; Awasty, V.; Cusi, K.; Kowdley, K.V.; Lai, M.; Schiff, E.; Parmar, D.; et al. Saroglitazar, a PPAR-α/γ Agonist, for Treatment of NAFLD: A Randomized Controlled Double-Blind Phase 2 Trial. Hepatology 2021, 74, 1809–1824. [Google Scholar] [CrossRef] [PubMed]
  143. Armstrong, M.J.; Gaunt, P.; Aithal, G.P.; Barton, D.; Hull, D.; Parker, R.; Hazlehurst, J.M.; Guo, K.; Abouda, G.; Aldersley, M.A.; et al. Liraglutide safety and efficacy in patients with non-alcoholic steatohepatitis (LEAN): A multicentre, double-blind, randomised, placebo-controlled phase 2 study. Lancet 2016, 387, 679–690. [Google Scholar] [CrossRef] [PubMed]
  144. Newsome, P.N.; Buchholtz, K.; Cusi, K.; Linder, M.; Okanoue, T.; Ratziu, V.; Sanyal, A.J.; Sejling, A.S.; Harrison, S.A. A Placebo-Controlled Trial of Subcutaneous Semaglutide in Nonalcoholic Steatohepatitis. New Engl. J. Med. 2021, 384, 1113–1124. [Google Scholar] [CrossRef] [PubMed]
  145. Flint, A.; Andersen, G.; Hockings, P.; Johansson, L.; Morsing, A.; Sundby Palle, M.; Vogl, T.; Loomba, R.; Plum-Mörschel, L. Randomised clinical trial: Semaglutide versus placebo reduced liver steatosis but not liver stiffness in subjects with non-alcoholic fatty liver disease assessed by magnetic resonance imaging. Aliment. Pharm. 2021, 54, 1150–1161. [Google Scholar] [CrossRef]
  146. Saponaro, F.; Sestito, S.; Runfola, M.; Rapposelli, S.; Chiellini, G. Selective Thyroid Hormone Receptor-Beta (TRβ) Agonists: New Perspectives for the Treatment of Metabolic and Neurodegenerative Disorders. Front. Med. 2020, 7, 331. [Google Scholar] [CrossRef]
  147. Harrison, S.A.; Bashir, M.; Moussa, S.E.; McCarty, K.; Pablo Frias, J.; Taub, R.; Alkhouri, N. Effects of Resmetirom on Noninvasive Endpoints in a 36-Week Phase 2 Active Treatment Extension Study in Patients With NASH. Hepatol. Commun. 2021, 5, 573–588. [Google Scholar] [CrossRef]
  148. Harrison, S.A.; Bashir, M.R.; Guy, C.D.; Zhou, R.; Moylan, C.A.; Frias, J.P.; Alkhouri, N.; Bansal, M.B.; Baum, S.; Neuschwander-Tetri, B.A.; et al. Resmetirom (MGL-3196) for the treatment of non-alcoholic steatohepatitis: A multicentre, randomised, double-blind, placebo-controlled, phase 2 trial. Lancet 2019, 394, 2012–2024. [Google Scholar] [CrossRef]
  149. Charles, E.D.; Neuschwander-Tetri, B.A.; Pablo Frias, J.; Kundu, S.; Luo, Y.; Tirucherai, G.S.; Christian, R. Pegbelfermin (BMS-986036), PEGylated FGF21, in Patients with Obesity and Type 2 Diabetes: Results from a Randomized Phase 2 Study. Obesity 2019, 27, 41–49. [Google Scholar] [CrossRef]
  150. Harrison, S.A.; Ruane, P.J.; Freilich, B.L.; Neff, G.; Patil, R.; Behling, C.A.; Hu, C.; Fong, E.; de Temple, B.; Tillman, E.J.; et al. Efruxifermin in non-alcoholic steatohepatitis: A randomized, double-blind, placebo-controlled, phase 2a trial. Nat. Med. 2021, 27, 1262–1271. [Google Scholar] [CrossRef]
  151. Sanyal, A.; Charles, E.D.; Neuschwander-Tetri, B.A.; Loomba, R.; Harrison, S.A.; Abdelmalek, M.F.; Lawitz, E.J.; Halegoua-DeMarzio, D.; Kundu, S.; Noviello, S.; et al. Pegbelfermin (BMS-986036), a PEGylated fibroblast growth factor 21 analogue, in patients with non-alcoholic steatohepatitis: A randomised, double-blind, placebo-controlled, phase 2a trial. Lancet 2019, 392, 2705–2717. [Google Scholar] [CrossRef]
  152. Younossi, Z.M.; Ratziu, V.; Loomba, R.; Rinella, M.; Anstee, Q.M.; Goodman, Z.; Bedossa, P.; Geier, A.; Beckebaum, S.; Newsome, P.N.; et al. Obeticholic acid for the treatment of non-alcoholic steatohepatitis: Interim analysis from a multicentre, randomised, placebo-controlled phase 3 trial. Lancet 2019, 394, 2184–2196. [Google Scholar] [CrossRef] [PubMed]
  153. Han, E.; Lee, Y.H.; Lee, B.W.; Kang, E.S.; Cha, B.S. Ipragliflozin Additively Ameliorates Non-Alcoholic Fatty Liver Disease in Patients with Type 2 Diabetes Controlled with Metformin and Pioglitazone: A 24-Week Randomized Controlled Trial. J. Clin. Med. 2020, 9, 259. [Google Scholar] [CrossRef] [PubMed]
  154. Harrison, S.A.; Manghi, F.P.; Smith, W.B.; Alpenidze, D.; Aizenberg, D.; Klarenbeek, N.; Chen, C.Y.; Zuckerman, E.; Ravussin, E.; Charatcharoenwitthaya, P.; et al. Licogliflozin for nonalcoholic steatohepatitis: A randomized, double-blind, placebo-controlled, phase 2a study. Nat. Med. 2022, 28, 1432–1438. [Google Scholar] [CrossRef]
  155. Inoue, M.; Hayashi, A.; Taguchi, T.; Arai, R.; Sasaki, S.; Takano, K.; Inoue, Y.; Shichiri, M. Effects of canagliflozin on body composition and hepatic fat content in type 2 diabetes patients with non-alcoholic fatty liver disease. J. Diabetes Investig. 2019, 10, 1004–1011. [Google Scholar] [CrossRef] [PubMed]
  156. Ribeiro Dos Santos, L.; Baer Filho, R. Treatment of nonalcoholic fatty liver disease with dapagliflozin in non-diabetic patients. Metabol. Open 2020, 5, 100028. [Google Scholar] [CrossRef] [PubMed]
  157. Shimizu, M.; Suzuki, K.; Kato, K.; Jojima, T.; Iijima, T.; Murohisa, T.; Iijima, M.; Takekawa, H.; Usui, I.; Hiraishi, H.; et al. Evaluation of the effects of dapagliflozin, a sodium-glucose co-transporter-2 inhibitor, on hepatic steatosis and fibrosis using transient elastography in patients with type 2 diabetes and non-alcoholic fatty liver disease. Diabetes Obes. Metab. 2019, 21, 285–292. [Google Scholar] [CrossRef]
  158. Calle, R.A.; Amin, N.B.; Carvajal-Gonzalez, S.; Ross, T.T.; Bergman, A.; Aggarwal, S.; Crowley, C.; Rinaldi, A.; Mancuso, J.; Aggarwal, N.; et al. ACC inhibitor alone or co-administered with a DGAT2 inhibitor in patients with non-alcoholic fatty liver disease: Two parallel, placebo-controlled, randomized phase 2a trials. Nat. Med. 2021, 27, 1836–1848. [Google Scholar] [CrossRef]
  159. Loomba, R.; Morgan, E.; Watts, L.; Xia, S.; Hannan, L.A.; Geary, R.S.; Baker, B.F.; Bhanot, S. Novel antisense inhibition of diacylglycerol O-acyltransferase 2 for treatment of non-alcoholic fatty liver disease: A multicentre, double-blind, randomised, placebo-controlled phase 2 trial. Lancet Gastroenterol. Hepatol. 2020, 5, 829–838. [Google Scholar] [CrossRef]
  160. Ratziu, V.; de Guevara, L.; Safadi, R.; Poordad, F.; Fuster, F.; Flores-Figueroa, J.; Arrese, M.; Fracanzani, A.L.; Ben Bashat, D.; Lackner, K.; et al. Aramchol in patients with nonalcoholic steatohepatitis: A randomized, double-blind, placebo-controlled phase 2b trial. Nat. Med. 2021, 27, 1825–1835. [Google Scholar] [CrossRef]
Figure 1. The flow of lipids from the intestine to the adipose tissue and liver, and from the liver into systemic circulation.
Figure 1. The flow of lipids from the intestine to the adipose tissue and liver, and from the liver into systemic circulation.
Nutrients 15 02323 g001
Figure 2. A review of carbon flux through glycolysis, the Krebs cycle, and lipid, amino, and nucleic acid synthesis pathways in NAFLD. The glucose enters the hepatocytes through Glut2 transporters independently of insulin and undergoes phosphorylation by hexokinases, trapping glucose intracellularly. Glucose 6-phosphate can be metabolized in three ways: by being stored as glycogen typically after a meal, by being metabolized into nucleic acids, amino acids, and NADPH through the pentose phosphate pathway, or by being broken down into glyceraldehyde 3-phosphate (GAP) and DHAP (di-hydroxy acetone phosphate) in glycolysis. Other contributors to the GAP pool are fructose and glycerol, the latter being released as a byproduct of lipolysis from adipose tissue. Fructose is readily taken up by the liver and broken down by fructose 1-P aldolase into GAP. GAP is a major contributor to the pyruvate pool and serves as a backbone for triglycerides. The pyruvate pool receives major contributions from lactate, alanine, and glutamine released by the muscle. The pyruvate pool, in equilibrium with lactate, can be converted to acetate or contribute to acetyl coA or oxaloacetate (OAA). Acetyl CoA contributes to ketogenesis and acetate production or is cycled through the Kreb’s cycle. The acetyl-CoA combines with OAA to produce citrate, which can be shuttled out of mitochondria and reconverted to acetyl CoA and OAA. Acetyl CoA can generate ketones, especially during fasting, and/or serve as a precursor for triglycerides and cholesterol synthesis. The OAA is exported out through a malate shuttle and is converted to phosphoenol pyruvate (PEP) through the action of the gluconeogenic enzyme PEPCK (phosphor enol pyruvate carboxy kinase). Carbon flux through the various anabolic (lipid, glucose, nucleic, and protein synthesis) and catabolic (glycolysis, beta-oxidation, and Krebs–Electron transport chain) pathways is influenced by the overall nutritional load and the status of insulin sensitivity. Insulin sensitivity regulates the flux through lipid synthesis and glucose-producing pathways in the hepatocytes. The pathways highlighted in red are increased in NAFLD.
Figure 2. A review of carbon flux through glycolysis, the Krebs cycle, and lipid, amino, and nucleic acid synthesis pathways in NAFLD. The glucose enters the hepatocytes through Glut2 transporters independently of insulin and undergoes phosphorylation by hexokinases, trapping glucose intracellularly. Glucose 6-phosphate can be metabolized in three ways: by being stored as glycogen typically after a meal, by being metabolized into nucleic acids, amino acids, and NADPH through the pentose phosphate pathway, or by being broken down into glyceraldehyde 3-phosphate (GAP) and DHAP (di-hydroxy acetone phosphate) in glycolysis. Other contributors to the GAP pool are fructose and glycerol, the latter being released as a byproduct of lipolysis from adipose tissue. Fructose is readily taken up by the liver and broken down by fructose 1-P aldolase into GAP. GAP is a major contributor to the pyruvate pool and serves as a backbone for triglycerides. The pyruvate pool receives major contributions from lactate, alanine, and glutamine released by the muscle. The pyruvate pool, in equilibrium with lactate, can be converted to acetate or contribute to acetyl coA or oxaloacetate (OAA). Acetyl CoA contributes to ketogenesis and acetate production or is cycled through the Kreb’s cycle. The acetyl-CoA combines with OAA to produce citrate, which can be shuttled out of mitochondria and reconverted to acetyl CoA and OAA. Acetyl CoA can generate ketones, especially during fasting, and/or serve as a precursor for triglycerides and cholesterol synthesis. The OAA is exported out through a malate shuttle and is converted to phosphoenol pyruvate (PEP) through the action of the gluconeogenic enzyme PEPCK (phosphor enol pyruvate carboxy kinase). Carbon flux through the various anabolic (lipid, glucose, nucleic, and protein synthesis) and catabolic (glycolysis, beta-oxidation, and Krebs–Electron transport chain) pathways is influenced by the overall nutritional load and the status of insulin sensitivity. Insulin sensitivity regulates the flux through lipid synthesis and glucose-producing pathways in the hepatocytes. The pathways highlighted in red are increased in NAFLD.
Nutrients 15 02323 g002
Table 1. Therapeutics for NAFLD and their mechanism of action.
Table 1. Therapeutics for NAFLD and their mechanism of action.
Mechanism of ActionT2DSteatosisFibrosis Other IndicationsReferences
PPAR agonistInsulin sensitizing, transcriptional controlyesYes; trials
ongoing
Yes; trials
ongoing
Weight gain
Bone fractures
[140,141,142]
GLP-1 agonistIncreases insulin secretion, enhances satietyyesYes; trials
ongoing
Not testedWeight loss[143,144,145]
Thyroid receptor beta-agonistIncreases mitochondrial respiration and breakdown of lipids. Not
tested
Yes; trials ongoingYes; trials ongoing Liver and isoform-specific [146,147,148]
Fgf21Increases energy expenditure; improves glucose and lipid balanceNot
tested
Yes; trials ongoing Yes; trials ongoing Some Fgf21 analogues were discontinued [149,150,151]
Obeticholic acidActivator of FXR. Suppresses liver fat content by increasing Fgf19 signaling from gut to the liver. Not
tested
Yes; trials ongoing Yes; trials ongoingPromising results for NASH patients. Approved drug for cholangitis. [152]
Sglt2 inhibitor Inhibits tubular absorption of sugars in the kidneyyesYes; trials ongoingYes; trials ongoingBone fractures,
frequent urination, urinary tract infections.
[153,154,155,156,157]
ACC + DGAT
Inhibitor
Inhibits the activity of enzymes that stimulate lipogenesis in the liverNot
tested
Yes; trials
ongoing
Not assessed [158,159]
SCD1 inhibitor Suppresses synthesis of saturated fats in the liverNot
tested
Yes, trials ongoingYes; trials ongoing [160]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bhat, N.; Mani, A. Dysregulation of Lipid and Glucose Metabolism in Nonalcoholic Fatty Liver Disease. Nutrients 2023, 15, 2323. https://doi.org/10.3390/nu15102323

AMA Style

Bhat N, Mani A. Dysregulation of Lipid and Glucose Metabolism in Nonalcoholic Fatty Liver Disease. Nutrients. 2023; 15(10):2323. https://doi.org/10.3390/nu15102323

Chicago/Turabian Style

Bhat, Neha, and Arya Mani. 2023. "Dysregulation of Lipid and Glucose Metabolism in Nonalcoholic Fatty Liver Disease" Nutrients 15, no. 10: 2323. https://doi.org/10.3390/nu15102323

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