Next Article in Journal
Metformin Restrains the Proliferation of CD4+ T Lymphocytes by Inducing Cell Cycle Arrest in Normo- and Hyperglycemic Conditions
Previous Article in Journal
Novel Plasma Biomarkers Associated with Future Peripheral Atherosclerotic Disease and Abdominal Aortic Aneurysm—Insights from Contemporary Prospective Studies from the Malmö Diet and Cancer Study
Previous Article in Special Issue
Establishment of a Rat Model of Alcoholic Liver Fibrosis with Simulated Human Drinking Patterns and Low-Dose Chemical Stimulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulation of Fructose Metabolism in Nonalcoholic Fatty Liver Disease

Department of Molecular and Structural Biochemistry, North Carolina State University, 128 Polk Hall Campus, Box 7622, Raleigh, NC 27695, USA
*
Author to whom correspondence should be addressed.
Biomolecules 2024, 14(7), 845; https://doi.org/10.3390/biom14070845
Submission received: 17 June 2024 / Revised: 2 July 2024 / Accepted: 8 July 2024 / Published: 13 July 2024
(This article belongs to the Special Issue Targets and Treatment Approaches for End-Stage Liver Diseases)

Abstract

:
Elevations in fructose consumption have been reported to contribute significantly to an increased incidence of obesity and metabolic diseases in industrial countries. Mechanistically, a high fructose intake leads to the dysregulation of glucose, triglyceride, and cholesterol metabolism in the liver, and causes elevations in inflammation and drives the progression of nonalcoholic fatty liver disease (NAFLD). A high fructose consumption is considered to be toxic to the body, and there are ongoing measures to develop pharmaceutical therapies targeting fructose metabolism. Although a large amount of work has summarized the effects fructose exposure within the intestine, liver, and kidney, there remains a gap in our knowledge regarding how fructose both indirectly and directly influences immune cell recruitment, activation, and function in metabolic tissues, which are essential to tissue and systemic inflammation. The most recent literature demonstrates that direct fructose exposure regulates oxidative metabolism in macrophages, leading to inflammation. The present review highlights (1) the mechanisms by which fructose metabolism impacts crosstalk between tissues, nonparenchymal cells, microbes, and immune cells; (2) the direct impact of fructose on immune cell metabolism and function; and (3) therapeutic targets of fructose metabolism to treat NAFLD. In addition, the review highlights how fructose disrupts liver tissue homeostasis and identifies new therapeutic targets for treating NAFLD and obesity.

1. Introduction

Worldwide, approximately 46.9 cases of nonalcoholic fatty liver disease (NAFLD) occur among 1000 persons per year [1]. In the U.S., 47.8% of the population is diagnosed with NAFLD, with an average of 30% developing nonalcoholic steatohepatitis (NASH) [1,2,3,4,5]. NAFLD encompasses many pathologies that consist of hepatic lipid accumulation, known as hepatic steatosis, independent of alcohol consumption or other secondary causes of fatty lipid accumulation in the liver [6,7,8,9]. Hepatic steatosis can progress to NASH, defined by inflammation and histological evidence of liver injury [10,11]. NASH is the third most common reason for liver transplants and 5–25% of patients will progress to cirrhosis, with 13% progressing to hepatocellular carcinoma, one of the most common causes of cancer-related deaths [10,12,13,14,15,16,17].
The incidence of NAFLD is increasing each year and is related to several metabolic and lifestyle factors. The increase in individuals with NAFLD is partly due to a rise in the number of patients with Type 2 diabetes mellitus (T2DM), with more than 50–70% of those with T2DM diagnosed with NAFLD [18,19,20,21]. NAFLD is also strongly associated with metabolic syndrome (MetS), as obesity, insulin resistance, and hyperlipidemia are common risk factors [22,23,24,25,26,27,28]. Resmetirom is the only therapy for NASH approved by the U.S. Food and Drug Administration (FDA), but it remains to be seen if this therapy will lead to long-term complications or side effects [29]. Therefore, it is necessary to identify new molecular mechanisms driving the progression of NAFLD to develop effective therapies with minimal complications.
Diet plays a key role in the development of NAFLD. The typical Western diet, characterized by excess fat and carbohydrates, mainly from sugars such as fructose, has been shown to drive obesity and MetS [30,31,32]. Fructose consumption has increased over the past 40 years among adults and children [33,34,35]. Dietary fructose is often accompanied by other sugars which are consumed in excess, such as glucose, sucrose, or high-fructose corn syrup (HFCS) [36]. Sucrose, commonly known as table sugar, consists of a 50:50 ratio of fructose to glucose, while HFCS consists of ~50–55% fructose and the remaining glucose. These sugars are present in yogurts, fast food, and highly consumed sugar-sweetened beverages. While fructose is often consumed in the form of sucrose, both sucrose and HFCS can elevate levels of triglycerides, free fatty acids, and insulin release after consumption [37,38,39].
Fructose metabolism is dependent on the breakdown of adenosine triphosphate (ATP), generating adenosine monophosphate (AMP) which leads to increased uric acid production [40,41]. Zheng et al. recently reported that elevated fructose-to-vitamin C ratios are positively associated with increased serum uric acid levels among African American men [42]. Hyperuricemia and high consumption of fructose also are strongly correlated with the incidence of NAFLD and cardiovascular disease [43,44,45,46,47]. Elevated serum levels of uric acid are strong predictors of NAFLD among obese adults [48]. Due to the harmful effects of increased consumption of sucrose or HFCS on human health, it is imperative to fully distinguish the unique effects of fructose on the development of metabolic disease.
The current dogma regarding fructose metabolism and metabolic disease suggests hepatocytes and enterocytes metabolize fructose, altering glucose and lipid metabolism. In addition, intestinal microbiomes exposed to high concentrations of fructose produce metabolites that contribute to progression of NAFLD. However, there are a number of other cell types exposed to high concentrations of fructose, including immune cells, that interact with parenchymal cells in metabolic tissues to control homeostasis. Myeloid cells and lymphocytes are specifically known to contribute to inflammation and fibrosis in metabolic tissues. However, a knowledge gap remains in our understanding of the role of fructose metabolism in immune cells when exposed to an excess of fructose and the pathogenesis of metabolic diseases, such as NAFLD. The present review highlights the current literature regarding how fructose regulates inflammation in metabolic tissues, the implications of fructose metabolism on immune cell phenotypes, and therapeutic targets of fructose metabolism.

2. Fructose Metabolism and Metabolic Tissues

2.1. Overview of Fructose Metabolism

In healthy adults, normal resting levels of fructose within peripheral blood are 0.04 mM. However, following fructose consumption, peripheral blood levels can increase to 0.4 mM [49,50]. In the small intestine, polysaccharides are broken down into monomeric forms of glucose, fructose, and galactose. Fructose is taken up into the intestine by glucose transporters (GLUT) 2 and 5. The process of fructose transport is passive and insulin-independent, unlike glucose. The knockout of GLUT5 in mouse models leads to a 75% reduction in fructose absorption in the intestine, highlighting its importance in fructose metabolism [51]. GLUT5 is developmentally regulated, with GLUT5 gene expression significantly lower in fetal intestines than in adults [52]. The consumption of fructose by toddlers leads to cases of malabsorption and early-onset weight gain [53,54]. GLUT5 also is expressed in tissues such as brain, kidney, and skeletal muscle and certain types of cancers [55,56,57,58]. Human monocytes express GLUT5, which increases during macrophage differentiation and foam cell development [59,60]. It is possible that with increased levels of fructose consumption, traditionally non-metabolizing fructose cells become subject to fructose-mediated cell plasticity.
Once taken up into the cells, ketohexokinase C (KHK) phosphorylates fructose. The KHK gene encodes two isotopes, KHKC and KHKA. KHKC is primarily expressed in humans and mice in the liver, kidney, and jejunum, with low expression levels in the pancreas and white adipose tissue [61]. KHKA is ubiquitously expressed and has a slower enzymatic rate compared with KHKC, with the Km values being 0.8 mmol/L and 7 mmol/L for fructose, respectively. Although both isoforms can phosphorylate fructose, the majority of fructose uptake and metabolism is mediated by KHKC in the intestine and liver [62]. Pharmacological inhibition and global knockdown of both isoforms of KHK reduce fructose caloric intake, leading to reductions in body weight, adipose tissue expansion, hepatic lipid accumulation, liver injury, and fibrosis [63,64]. Knockdown of KHK in epithelial cells of the small and large intestines in mice leads to reductions in fructose caloric intake, but does not affect chronic fructose-induced lipid accumulation in adipose tissues or the liver [62]. In contrast, hepatocytes were found to be the primary cells responsible for driving fructose-induced MetS. Interestingly, during the development of hepatocellular carcinoma and breast cancer, there is a switch from KHKC to KHKA activity, leading to the phosphorylation of phosphoribosyl pyrophosphate synthetase 1, driving nucleotide synthesis [65,66].
Transcriptional regulation of KHK modulates fructose metabolism and mitigates its effects on intestinal and hepatic lipogenesis and inflammation. Carbohydrate regulatory element binding protein (ChREBP) is the master transcription factor of KHK. Recently, hypoxia inducible-factors alpha-2a (HIF-2α) was discovered to inhibit fructose metabolism by reducing KHK isoforms and aldolase B expression in hepatocytes. In addition, peroxisomal biogenesis factor 2 (PEX2), a peroxisomal membrane protein necessary for peroxisome biogenesis, was also found to control KHK expression in the liver [67]. Fructose metabolism also involves the enzyme hexokinase (HK) [61]. HK phosphorylates fructose to form fructose-6-phosphate, using ATP as a phosphate donor. The HK family is a group of enzymes, including HK I, II, III, and IV which are expressed in liver, while HK I and II are present in adipose tissue [61]. The interplay between fructose and HK is complex, and alterations in fructose metabolism can have implications for metabolic health, including conditions such as NAFLD and MetS.
Unlike HK, the metabolism of fructose by KHK has no regulating feedback loop, thus allowing for the rapid conversion of fructose. Within the intestine, fructose is converted to glucose, glycerate, and tricarboxylic acid cycle (TCA) metabolites, which are then released into the blood. Although the intestine can metabolize low levels of fructose, it can become overly saturated, and excess fructose is transported through the portal vein and metabolized in the liver [68]. GLUT8 is responsible for fructose uptake in hepatocytes and stellate cells, de novo lipogenesis (DNL), and hepatic inflammation [69,70]. Knockdown of GLUT8 in hepatocytes attenuates DNL and inflammation while increasing fatty acid oxidation, lipolysis, and inflammation in stellate cells. GLUT2 also is expressed on hepatocytes and is capable of fructose uptake with a Km ~67 mM [71]. Following entry into a cell, fructose is further metabolized by KHK and aldolase B [72]. KHK phosphorylates fructose to fructose-1-phosphate. Fructose-1-phosphate is then converted to dihydroxyacetone phosphate and glyceraldehyde by aldolase B, which is further converted to glyceraldehyde-3-phosphate [73]. Triose kinase expressed in hepatocytes is the enzyme responsible for converting glyceraldehyde to glyceraldehyde-3-phosphate and directs fructose metabolites into lipogenesis, gluconeogenesis, and triglyceride accumulation in hepatocytes [73]. Although fructose is primarily shuttled through the portal vein, it is projected that 20–30% of ingested fructose may enter systemic circulation, where it can be deposited into other tissues, including the proximal tubes within the kidney [74]. In addition, fructose is also metabolized by immune cell populations in tissues and peripheral circulation [75,76]. Microglia, the resident macrophages in the brain, also express GLUT5 and KHK and metabolize fructose to drive cognitive dysfunction and influence the progression of metabolic diseases [77]. Ultimately, the concentration and frequency of consumption of dietary fructose determines the metabolic fate of these carbohydrates in metabolic tissues [68].
Endogenous fructose production depends on the isomerization of glucose, mannose, or xylose in bacteria and yeast [62,78]. In humans and other mammals, fructose can be endogenously produced by the polyol pathway, which converts glucose into sorbitol using aldose reductase and fructose using sorbitol dehydrogenase [79]. Under hyperglycemic conditions, fructose is generated through the polyol pathway to increase KHKA activity in cancer cells, promoting metastasis [80]. Interestingly, the activation of the polyol pathway in the liver and the intestine are positively associated with complications of diabetes and the progression of NAFLD [81,82,83].

2.2. Fructose and Intestine

Under normal intestinal conditions, fructose has the same metabolic fate in the form of sucrose or fructose alone [13,68]. Fructose carbons are rapidly converted to glucose, glycerate, and TCA metabolites within enterocytes, the primary cell type within the intestine. These byproducts are typically shuttled through the portal vein to the liver for further processing and metabolism. In enterocytes, fructose is converted to fructose-1-phosphate by KHK (Figure 1). Elevated intracellular levels of fructose-1-phosphate can be toxic, leading to endoplasmic reticulum stress and inflammation, which is reported to induce barrier deterioration and shorten colon length [84]. Taylor et al. discovered that fructose promotes hypoxic cell survival via extended villi length, partly through fructose-1-phosphate inhibition of the hypoxic adaptation protein pyruvate kinase isozyme M2 (PKM2) [85]. The extension of the length of the villi increased weight gain, nutritional absorption, and fat accumulation. High levels of fructose lead to a misshapen colon and cecum as well as infiltration of inflammatory macrophages, neutrophils, dendritic cells, and natural killer T cells into the lamina propria. These cells express the inflammatory cytokines interleukin 1 beta (IL-1β), interleukin 2 (IL-2), and interleukin 6 (IL-6) within the small intestine of mice [86,87]. In addition, cecal metabolites such as prostaglandin B1 and I2, which are important for regulating blood pressure, inflammation, and fibrosis, are elevated in male mice on a fructose diet [88]. Although fructose metabolism begins within the small intestine at low doses, it can bypass initial metabolism at higher doses and be shuttled through the portal vein for further processing [89].
High concentrations of fructose reduce the ratio of Bacteroidetes/Firmicutes while increasing the abundance of the bacterial genera Bifidobacterium, Enterococcus, Lactobacilus, Romboutsia, and Turicibater [86,90]. These microbes metabolize fructose through fructokinase, the phosphoenolpyruvate-dependent sugar phosphotransferase system, and other hexose metabolism pathways [91,92,93,94,95]. Additionally, the pro-inflammatory metabolites cresol, arachidonic acid, stearic acid, palmitic acid, and indole sulfuric acid were demonstrated to be elevated, while anti-inflammatory short chain fatty acids (SCFAs) such as acetic acid, propionic acid, butyric acid, and pentanoic acid were reduced in male mice and humans consuming high-fructose diets [86,96]. Among obese and overweight teenagers, Eubacterium and Streptococcus were found to be negatively correlated with fructose consumption [97]. The gut bacteria of obese youth also are reported to have a higher capability to ferment fructose compared with microbes in the guts of lean subjects [98]. The abundance of Barnesiella is elevated, while the abundance of Ruminococcus, Faecalibacterium, and Erysipelatoclostridium are reduced among healthy adult women consuming a HFCS diet [96]. Fructose malabsorption can occur when dietary concentrations exceed enterocyte absorption capacity [99]. In addition to increased bacterial growth, Kawabata et al. found that a high fructose consumption induced a leaky gut by decreasing protein expression of the tight junction protein occludin in male mice [100]. Ultimately, a leaky gut may lead to the increased permeability of the epithelial cells lining the digestive tract and induce inflammatory cell infiltration.

2.3. Fructose and Liver

Over the last decade, studies have linked monosaccharide fructose to the progression of NAFLD, as fructose consumption has increased from the recommended 5% of daily energy intake to 17% within sugar-sweetened beverages alone, correlating with the rise in obesity-related diseases [101]. As high doses of fructose saturate intestinal metabolism, the remaining fructose is metabolized in the liver (Figure 1). The metabolism of fructose produces triose phosphates that form acetyl-CoA, the essential building block for fatty acid synthesis. Fructose metabolism stimulates DNL through triose phosphate-activated fatty acid synthesis and transcriptional regulation of enzymes involved in DNL. Softic et al. demonstrate that male mice consuming fructose-sweetened water in the presence of a high-fat diet exhibited increased fatty acid synthesis and hepatic insulin resistance compared with glucose-sweetened water in a process dependent on KHK [102]. Fructose supplementation increases the levels of the transcription factors sterol regulatory element binding protein 1 (SREBP-1) and ChREBP, as well as the downstream targets acetyl-CoA carboxylase (ACC), fatty acid synthase (FAS), and the fatty acid transporter CD36 [103]. Fructose metabolism also reduces hepatic ATP levels and the activation of AMP-activated protein kinase (AMPK), which impairs insulin signaling and increases DNL [104,105,106].
Multiple studies in male mice and rats have demonstrated that a high-fat fructose diet leads to increased infiltration of macrophages within hepatic tissues [107,108,109]. Furthermore, emerging evidence indicates the recruitment of myeloid cells from the bone marrow and mature macrophages originating from surrounding tissues, such as peritoneal and intestinal macrophages [110]. The emergence of single-cell RNA sequencing has allowed for the identification of residential and recruited macrophage subsets based on gene signatures [111]. In addition, Kupffer cells, residential macrophages of the liver, or recruited macrophages may express a fluid phenotype, displaying pro- and anti-inflammatory properties. External stimuli such as lipopolysaccharide (LPS) or interferon-gamma (IFNγ) activate Toll-like receptors (TLR) of macrophages to secrete cytokines such as IL-6 and tumor necrosis factor alpha (TNFα) to aid in-wound repair through the increased flux of glycolysis [112]. During prolonged inflammation, macrophages can take on an anti-inflammatory phenotype, releasing extracellular matrix proteins to signal wound healing. A chronic fructose diet increases hepatic transitioning monocytes and reduces Kupffer cells in male mice [75]. In vitro studies demonstrate that Kupffer cells metabolize fructose through glycolysis and the pentose phosphate pathway. In addition to increased monocyte infiltration, Kupffer cells express a wound healing and anti-inflammatory phenotype, indicating fructose regulates Kupffer cell phenotype and viability (Figure 2). This prolonged, chronic inflammation can lead to tissue damage and, ultimately, result in the development of NASH.

2.4. Fructose and Adipose Tissue

Adipose tissue is dynamic in cell populations and crosstalk fluidity, thus multiple cell types are exposed to dietary carbohydrates. Fructose also increases adipogenesis [113] (Figure 1). Adipocytes depend on GLUT5 to transport fructose and HK for phosphorylation [114]. Once isomerized to glucose-6-phosphate (G6P), the fate of fructose diverges through glycolysis. Using C13-labeled fructose, fructose has been demonstrated to be shuttled into acetyl-CoA and incorporated into the TCA cycle, causing an increase in the release of fatty acid and palmitate [115]. In the presence of glucose, fructose increases the production of nicotinamide adenine dinucleotide phosphate (NADP) and carbon dioxide, which can be utilized for lipid storage. Fructose increases gene expression of peroxisome proliferator-activated receptor γ (PPARγ), CCAAT/enhancer-binding protein alpha (C/EBPα), GLUT4, adiponectin, and adipocyte fatty acid binding protein 2 (AP2) in 3T3-L1 preadipocytes [116]. GLUT5 was determined to be essential for adipocyte differentiation, as knockdown of this transporter reduced the expression of PPARy and AP2, as well as the ability of adipocytes to store lipids. Fructose also reduces beta-oxidation by decreasing PPAR beta and delta, leading to increased inflammation in adipose tissue and the liver [117]. Adipocytes maintain homeostasis by secreting cytokines to signal for immune cell recruitment. Fructose increases the secretion of the pro-inflammatory cytokines interleukin 18 (IL-18), IL-1β, and TNFα in the adipose tissue of male and female rats [118]. Interestingly, the anti-inflammatory proteins interleukin 10 (IL-10) and nuclear receptor factor 2 (Nrf2) also were elevated, potentially to reduce the pro-inflammatory response [118]. Taken together, these studies indicate a direct effect of fructose exposure on energy production and maturation of adipocytes.

2.5. Fructose Metabolism Regulation of Immune Cell Function

2.5.1. Myeloid Cells

Many studies have examined fructose metabolism in the parenchymal cells of metabolic tissues, such as the intestine, liver, and adipose tissue. However, macrophages can also metabolize fructose to regulate their function and phenotype (Figure 2) [59,60,77]. Glucose is known as the primary fuel source for pro-inflammatory macrophages, with glycolysis providing energy for the production and secretion of cytokines. Peripheral blood monocytes metabolize fructose and shuttle carbons through oxidative phosphorylation (OXPHOS) compared with the lactate production that occurs with glucose metabolism [76]. TCA metabolites such as succinate, malate, and fumarate were also elevated to a greater extent than glucose, most likely to produce energy and reducing agents to maintain increased OXPHOS. The inflammatory cytokines IL-6, IL-1β, IL-18, and TNFα are detected at higher concentrations in fructose-treated monocytes compared with those treated with glucose [76]. This response was, in part, mediated by upregulated signaling of mTORC1. Glutamine carbons mainly contributed to TCA cycle intermediates and cytokine secretion within fructose treatments. These findings demonstrate that fructose directly regulates inflammation in monocytes.
Although macrophages regulate inflammation and wound repair, the liver relies on other immune cells to patrol the environment for damage and infection. Dendritic cells (DCs) can process and present antigens to surrounding lymphocytes to commence and regulate an adaptive immune response [119]. In human dendritic cells, 15mM fructose treatment induces a chronic inflammatory protein secretion of IL-1β, TNFα, and IL-6 [120]. Fructose-treated DCs stimulate T cell IFNγ secretion more significantly than glucose. This response was initiated and controlled by the secretion of TNFα from fructose exposed DCs. The DCs had elevated extracellular acidification rate (ECAR) levels, indicating a metabolic flux through glycolysis to keep up with the demands of cytokine secretion. This process is most likely regulated by the activation of nuclear factor kappa b subunit (NFκB) via advanced glycation end products (AGE) as these proteins were upregulated with fructose exposure [120].

2.5.2. B Cells

B cells play a critical role in the progression of NAFLD via the regulation of T cells and fibrosis [121]. However, specific subsets of B cells are known to drive the progression of NAFLD, while others prevent progression [122,123]. Pathogenic B2 cells produce pro-the inflammatory cytokines TNFα, IL-6, and IL-1β, which attract other immune cells to the liver [124]. Additionally, B cells act as antigen-presenting cells to activate T cells and amplify hepatic inflammation and fibrosis. The antibodies produced by B cells can trigger liver injury and fibrosis in NASH. Using C13 labeling nuclear magnetic resonance (NMR), carbohydrate metabolism differs based on the carbon source in B cells isolated from blood [125]. Fructose at 15mM is phosphorylated by HK in B cells (Figure 2). Interestingly, when fructose and glucose are supplied together, HK-mediated metabolism is constrained by glucose affinity, leaving fructose metabolism dependent upon the KHK pathway. This switch in metabolism causes DHAP and lactate buildup from methylglyoxal partitioning. Fructose supplementation reduces B cell lymphopoiesis in bone marrow and splenic marginal zone B cells compared with glucose in male mice [126]. Tan et al. demonstrated that glycolysis and oxidative phosphorylation are the primary metabolic pathways for the generation of B cells [126]. Currently, only one study has demonstrated that a chronic fructose diet decreases hepatic B cells in male mice compared with glucose or control diets [75].

3. Regulation of Fructose Intake and Metabolism to Prevent Metabolic Disease

3.1. Dietary Modifications and Physical Activity

Reducing fructose intake and metabolism can effectively treat obesity and NAFLD in men and women [127]. Patients with NAFLD are informed to limit soda, juice, and sports drinks that are high in sugar, which are associated with higher pro-inflammatory markers, triglyceride/high-density lipoprotein cholesterol ratios, and intrahepatic lipids [128,129,130]. In addition, selecting low-fructose-containing fruits and vegetables and avoiding foods with high concentrations of fructose and HFCS is recommended. Consuming high-fiber foods, such as vegetables, legumes, and whole grains, slows down the absorption of carbohydrates and promotes excretion. Soluble fibers, such as psyllium husk, oats, legumes, and insoluble fibers, reduce absorption of fructose and improve liver enzymes, intrahepatic lipid content, and liver fibrosis [131,132,133,134]. Maintaining a balanced gut microbiota through the consumption of probiotics or prebiotics also may help modulate fructose metabolism and reduce its detrimental effects on the liver. The fructose and water-soluble polymer inulin, fermented in the small intestine and colon to SCFAs, provides an energy source for the growth of protective Bifidobacterium [135,136,137]. Regular exercise improves insulin sensitivity and metabolic health and attenuates the fructose-mediated glycemic response in healthy adults, which can mitigate the negative impact of fructose on liver metabolism [138,139,140,141,142]. While effective, these interventions can be challenging for patients to adhere to over time, requiring other therapies to help reduce the harmful effects of increased fructose consumption.

3.2. Nutritional Supplements

Antioxidants may help to alleviate fructose-induced liver damage, oxidative stress, and inflammation to improve overall health (Figure 3A). Both vitamin E and resveratrol attenuate fructose-induced hepatic steatosis and liver damage by reducing lipid peroxidation and inflammation in mice and rats [143,144,145,146]. Fisetin, a naturally occurring flavonoid, has been reported to reduce fructose-induced liver damage in animal models by suppressing NFkB and inducing the transcription factor Nrf2, which regulates antioxidant enzymes [147]. However, limited studies demonstrate that these antioxidants reduce levels of liver enzymes and steatosis among patients with NAFLD [148,149,150,151].
L-carnitine is critical in transporting fatty acids into the mitochondria, the cellular organelles responsible for energy production by facilitating beta-oxidation. Studies exploring the potential benefits of L-carnitine supplementation for NAFLD suggest that L-carnitine supplementation improves the levels of liver enzymes, reduces the accumulation of liver fat, and even ameliorates insulin sensitivity in hepatocyte, animal, and human studies [152]. L-carnitine could enhance the transport of fatty acids into the mitochondria, preventing the excessive accumulation of lipids in liver cells induced by a high fructose intake. Hepatic long-chain acylcarnitines (LACs) with 18 and 20 carbons are increased in fructose-fed livers [75]. Serum LACs are positively correlated with the progression of fibrosis and hepatocellular carcinoma and are elevated with fructose supplementation in male mice [102,153]. LACs are elevated when triglyceride synthesis is inhibited and may act in an autocrine or paracrine mechanism to induce inflammation and cell death in macrophages and hepatocytes. Dietary fructose inhibits fatty acid oxidation in the liver and impairs mitochondrial function, which correlates with elevations in long-chain acylcarnitines [102]. In addition, L-carnitine exhibits antioxidant properties by reducing reactive oxygen species (ROS) and inducing the expression of the antioxidant enzyme superoxide dismutase 2 to reduce oxidative stress and the associated inflammation in fructose-treated hepatocytes [154].

3.3. Pharmacological Interventions

Targeting fructose metabolism for potential therapeutic interventions in biomolecules involves various strategies (Figure 3B). The design and development of small molecules or compounds that selectively inhibit KHK or HK, which could reduce the metabolism of fructose and its downstream effects on DNL and insulin resistance, are at the forefront of the field. Targeting fructose metabolism with KHK inhibitors has several potential therapeutic benefits, including treating obesity, type 2 diabetes, and fatty liver disease [64,155]. In one study, patients with T2DM treated with KHK inhibitors experienced significant reductions in blood sugar levels. KHK inhibitors also improve insulin sensitivity and reduce inflammation [64]. Bempedoic acid, an FDA-approved drug for hypercholesterolemia, reduces high fat and fructose diet-induced hepatic KHK protein expression, fructose uptake, and hepatic beta-oxidation in female rats [156]. Numerous clinical trials on KHK inhibitors have yielded opposing results. One of the most promising clinical trials was a Phase 2 trial of PF-06835919, a reversible KHKA/C inhibitor developed by Pfizer [157,158]. In this trial with 164 participants, PF-06835919 effectively reduced liver fat and inflammation in patients with NAFLD at 16 weeks. In addition, patients experienced significant reductions in insulin concentrations. However, PF-06835919 at 300 mg or 150 mg did not change fasting plasma cholesterol or low-density lipoprotein cholesterol levels. The KHK inhibitor had minimum safety concerns, with both placebo group and groups receiving the high and low concentration of the inhibitor reporting similar treatment-emergent adverse events ranging from diarrhea to hypoglycemia.
Other enzymes involved in fructose metabolism, such as aldolase B, triokinase, aldose reductase, and sorbitol dehydrogenase, can be targeted to modulate the overall flux of fructose metabolism and its impact on liver function. While targeting these enzymes with inhibitors holds promise as a potential therapeutic agent for targeting fructose metabolism, there also are limitations and challenges. For example, aldolase B, the enzyme responsible for breaking down fructose-1-phosphate into glyceraldehyde and dihydroxyacetone phosphate, was initially considered a potential therapeutic target for colon cancer as it mediates the metabolic reprogramming necessary for tumor metastasis [159]. However, patients with a mutation in the aldolase B gene developed hereditary fructose intolerance (HFI), leading to liver dysfunction and even failure, making this treatment toxic unless administered alongside a fructose-restricted diet [160]. Ultimately, due to these potential harmful off-target effects, therapeutically targeting aldolase B to regulate fructose metabolism was deemed undesirable. High selectivity is crucial to avoid off-target effects on other enzymes or metabolic pathways. Complete inhibition of these enzymes may disrupt glucose metabolism and have systemic metabolic consequences beyond the liver. Therefore, careful assessment of tissue and cell-specific expression patterns and potential off-target effects is necessary.

3.4. Fructose Transporter Modulators

Fructose transporter inhibitors may selectively target fructose metabolism by reducing fructose absorption and metabolism, improving hepatic insulin sensitivity. Inhibitors specifically targeting GLUT5 or GLUT8 could limit fructose availability for metabolism in the liver. GLUT5 is primarily expressed by intestinal epithelial cells and adipose tissue. However, under NASH conditions GLUT5 gene expression is increased in patients with NASH compared to those with normal and steatotic livers [63,161]. The increased expression may be due to infiltrating peripheral monocytes, which are reported to express GLUT5 [60]. Tripp et al. developed a yeast-based screening system to identify potential GLUT5 inhibitors [162]. The inhibitors N-[4-(methylsulfonyl)-2-nitrophenyl]- 1,3-benzodioxol-5-amine (MSNBA) and (−)-epicatechin-gallate (ECG) selectively inhibited GLUT5 without affecting other glucose transporters. Treatment with these inhibitors reduced the growth of mutant yeast cells in fructose-rich media [162,163]. ECG attenuates high-fat-diet-induced hepatic steatosis, ferroptosis, and oxidative stress [164]. MSNBA prevents the fructose-induced proliferation of colon cancer cells (HT-29) with little impact on the neighboring, healthy intestinal mucosa [57]. ECG, a flavonoid belonging to the group of catechins found in tea, cocoa, and fruits, reduces fructose uptake in Caco-2 cells and high-fat and high-fructose-induced obesity and NAFLD in male mice [165,166]. SW157765 is the only inhibitor reported to bind to GLUT8 and reduces glucose uptake in GLUT8 expressing cancer cells [167].

3.5. Insulin-Sensitizing Agents

The anti-hyperglycemia drug metformin is reported to reduce levels of liver aspartate aminotransferase (AST) and alanine aminotransferase (ALT) in patients with T2DM and NAFLD [168,169,170,171,172,173,174]. In a study conducted by Spruss et al., mice supplemented with metformin showed significantly lower hepatic steatosis, liver injury, and levels of plasma leptin and retinol-binding protein 4 (RBP-4) [174]. Additionally, metformin reduced the hepatic expression of TNFα [175]. The direct target of metformin remains to be determined. Metformin indirectly activates AMP-activated kinase and reduces mitochondrial function to reduce hepatic glucose production and lipid metabolism [176,177]. In male mice on a 47% fructose diet for ten weeks, metformin increased the expression of genes regulating mitochondria biogenesis and antioxidants and reduced genes associated with DNL [178]. Metformin also prevents LPS-induced M1 macrophage polarization by activating AMPK reducing ROS production [179,180,181]. Taken together, these data suggest that metformin protects the liver from the onset of fructose-induced NAFLD by altering intestinal permeability and activation of hepatic Kupffer cells.
Glucagon-like-peptide-1-receptor agonists (GLP-1-RA) are another class of drugs developed to treat type 2 diabetes. GLP-1, the peptide hormone that binds the GLP-1 receptor, is secreted by the L cells of the jejunum and ileum [182]. Fatty acids and sucrose are known to increase GLP-1 secretion, and GLP-1 causes the postprandial secretion of insulin in the pancreas [183,184]. In adult males, acute sucrose consumption causes a smaller increase in circulating GLP-1 levels compared with females [185]. The GLP-1 receptor is found in the kidneys, lungs, and hypothalamus [183]. Boland et al. demonstrated that the GLP-1 receptors also are expressed on mouse and human monocytes and macrophages [186]. GLP-1 decreases hunger and food intake [187,188,189]. GLP-1-RA suppresses appetite through AMPK inhibition in the hypothalamus [187,188]. Burmeister et al. demonstrated that intracerebroventricular injections of fructose attenuated the anorectic response induced by the GLP-1-RA, Exendin-4 (Ex-4) [188]. Fructose can activate AMPK and block the decrease in appetite and satiety associated with GLP-1 [190]. Likewise, Ghidewon et al. demonstrated that GLP-1-RA semaglutide decreased the intake of sucrose-enriched pellets in male mice [191]. GLP-1-RA combats metabolic disease through several mechanisms. For example, these drugs also increase the expression of GLUT4 receptors in adipose and muscle tissue, mediate lipid metabolism, and reduce the levels of ROS, leading to inflammation [192]. GLP-1-RA also increase glucose-dependent insulin secretion by pancreatic β-cells [182].
GLP-1-RAs have shown significant promise in treating NASH. In male rats administered fructose, GLP-1-RA significantly reduced hepatic steatosis by increasing b-catenin nuclear translocation, leading to a reduction in lipid synthesis enzymes, ACC, FAS, and stearoyl-coenzyme A desaturase-1 (SCD-1) [193]. Liraglutide, a GLP-1RA, reduced serum AST and ALT levels in hepatic ischemia–reperfusion injury [194]. GLP-1 receptor signaling also prevents M1 macrophage polarization [194]. The GLP-1 and glucagon receptor dual agonist cotadutide significantly reduced hepatic steatosis, inflammation, and fibrosis in NASH male mice on a high fructose, trans-fat, and cholesterol diet [195]. Ultimately, more research is needed to elucidate the effects of GLP-1-RAs on fructose absorption and metabolism in metabolic diseases. However, these drugs are still promising therapeutic targets for fructose-associated NAFLD.

4. Limitations and Future Directions

It is important to note that targeting fructose metabolism alone may not be sufficient for addressing the complexities of NAFLD, and a comprehensive approach involving dietary modifications, lifestyle changes, and management of metabolic risk factors is required. Notably, current research on fructose-induced disease progression and cellular metabolism is limited by the exaggerated concentrations of fructose used in studies. Depending on the location, specific cells and tissues would not be exposed to such high concentrations of circulating fructose. Interestingly, a high fructose consumption may be protective during acute liver injury. In male mice, fructose reduces acetaminophen-induced liver injury by increasing ChREBP and serum levels of fibroblast growth factor 21 [196]. Although we gain insight into fructose metabolism within these models, the current gap in our understanding regarding dietitian-recommended levels of fructose remains large. It is also necessary to note that human studies have demonstrated no correlation between fructose consumption and NAFLD development. Among older Finnish older, a high fructose intake was associated with a lower risk of NAFLD [197]. Similarly, total fructose intake was not different between patients with steatosis, NASH, or cirrhosis. Both studies were dependent on food frequency questionnaires and dietary recalls [198], which may lead to underreporting dietary intake of fructose. Thus, there is a need for metabolic tracing studies and biomarkers to track fructose consumption.
In addition, very few groups have identified the effects of fructose on immune cell populations other than the plasticity of fructose-induced macrophage and B cell populations. Collectively, there are differences in fructose metabolism depending on tissue and cell type, mainly through the actions of KHK or HK. The availability and concentrations of fructose and glucose limit the initial steps in fructose metabolism. For example, residential intestine and liver cells utilize KHK-mediated metabolism, which shuttles primarily through glycolysis conversion intermediates. However, bone marrow-derived macrophages depend on oxidative phosphorylation and the TCA cycle to generate intermediates for cytokine secretion. Because of the differences in metabolism, it is unsurprising that fructose impacts parenchymal and nonparenchymal cellular function through different mechanisms. As research on fructose exposure and immune cells in disease increases, we may find novel metabolic pathways.
The impact of sex on fructose metabolism and metabolic outcomes remains limited. Sex differences are reported for the effects of fructose metabolism on kidney function with fructose increasing GLUT5 protein expression in males compared with females [199]. Isotope tracing of fructose in healthy young males and females demonstrated that fructose is metabolized less efficiently in females, leading to reduced plasma insulin, lactate, and uric acid levels [200]. In animal models, fructose supplementation in the drinking water increased the body weight of male rats compared with female rats [201]. In addition, hepatic TNFα levels are reduced in female rats compared with male rats at 10 weeks of high-fructose feeding [202]. From the United Kingdom Biobank, fructose consumption from sugary drinks is associated with a 2% higher risk of hyperandrogenism among middle-aged women [130]. These findings suggest menopause or age in women alters fructose metabolism, leading to a possibly higher risk of metabolic diseases such as NAFLD. Elevations in androgens leads to increased plasma lipids, hepatic steatosis, and insulin resistance due to mitochondrial dysfunction in female rats [203]. Likewise, supplementation of 30% fructose in a high-fat diet increases liver weight, inflammation, and injury compared with high-fat-diet-fed ovariectomized female mice [204]. In males, fructose consumption causes metabolic disease but reduces testosterone and luteinizing hormone release in the anterior pituitary [205]. Studies linking mechanisms by which fructose regulates sex hormones and alters tissue specific fructose metabolism in males and females remain to be investigated.

5. Conclusions

NAFLD remains a significant health concern and managing its effects on hepatic lipid accumulation and insulin resistance is critical. To treat this condition effectively and prevent its progression to severe liver disease, we need a comprehensive approach that combines dietary modification, weight management, and pharmacological interventions. The literature discussed in this review defines the cellular and molecular mechanisms regulating fructose metabolism and the development of NAFLD. Fructose regulates lipid metabolism, inflammation, and immune cell function in the intestine, adipose tissue, and liver. Several studies show varying results depending on the concentrations of fructose supplied and the duration of the animal and clinical studies. These inconsistencies raise questions about how fructose affects individual cells, tissues, and biological systems. For example, how does fructose control subsets and functions of T cells, and how does it affect NAFLD? Also, understanding how fructose modulates intestinal function and identifying critical cellular and molecular signaling pathways are essential. Therefore, more research is needed to understand the critical pathways regulated by fructose in metabolic systems and the pharmacological interventions targeting various steps in fructose metabolism. By addressing fructose metabolism, we can effectively manage NAFLD and prevent its progression to more severe liver disease.

Author Contributions

Design and conception, M.L. and A.K. drafting and revising manuscript. M.L., R.D. and A.K. contributed to the article and approved the submitted version. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Institute of Diabetes and Digestive and Kidney Diseases [R21 DK128678].

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Riazi, K.; Azhari, H.; Charette, J.H.; Underwood, F.E.; King, J.A.; Afshar, E.E.; Swain, M.G.; Congly, S.E.; Kaplan, G.G.; Shaheen, A.A. The prevalence and incidence of NAFLD worldwide: A systematic review and meta-analysis. Lancet Gastroenterol. Hepatol. 2022, 7, 851–861. [Google Scholar] [CrossRef] [PubMed]
  2. Vernon, G.; Baranova, A.; Younossi, Z.M. Systematic review: The epidemiology and natural history of non-alcoholic fatty liver disease and non-alcoholic steatohepatitis in adults. Aliment. Pharmacol. Ther. 2011, 34, 274–285. [Google Scholar] [CrossRef] [PubMed]
  3. Tilg, H.; Moschen, A.R. Evolution of inflammation in nonalcoholic fatty liver disease: The multiple parallel hits hypothesis. Hepatology 2010, 52, 1836–1846. [Google Scholar] [CrossRef] [PubMed]
  4. Wong, V.W.; Wong, G.L.; Choi, P.C.; Chan, A.W.; Li, M.K.; Chan, H.Y.; Chim, A.M.; Yu, J.; Sung, J.J.; Chan, H.L. Disease progression of non-alcoholic fatty liver disease: A prospective study with paired liver biopsies at 3 years. Gut 2010, 59, 969–974. [Google Scholar] [CrossRef] [PubMed]
  5. Masoodi, M.; Gastaldelli, A.; Hyötyläinen, T.; Arretxe, E.; Alonso, C.; Gaggini, M.; Brosnan, J.; Anstee, Q.M.; Millet, O.; Ortiz, P.; et al. Metabolomics and lipidomics in NAFLD: Biomarkers and non-invasive diagnostic tests. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 835–856. [Google Scholar] [CrossRef] [PubMed]
  6. Benedict, M.; Zhang, X. Non-alcoholic fatty liver disease: An expanded review. World J. Hepatol. 2017, 9, 715–732. [Google Scholar] [CrossRef] [PubMed]
  7. Ludwig, J.; Viggiano, T.R.; McGill, D.B.; Oh, B.J. Nonalcoholic steatohepatitis: Mayo Clinic experiences with a hitherto unnamed disease. Mayo Clin. Proc. 1980, 55, 434–438. [Google Scholar] [PubMed]
  8. Chalasani, N.; Younossi, Z.; Lavine, J.E.; Diehl, A.M.; Brunt, E.M.; Cusi, K.; Charlton, M.; Sanyal, A.J.; Association, A.G. The diagnosis and management of non-alcoholic fatty liver disease: Practice guideline by the American Gastroenterological Association, American Association for the Study of Liver Diseases, and American College of Gastroenterology. Gastroenterology 2012, 142, 1592–1609. [Google Scholar] [CrossRef] [PubMed]
  9. Kneeman, J.M.; Misdraji, J.; Corey, K.E. Secondary causes of nonalcoholic fatty liver disease. Therap Adv. Gastroenterol. 2012, 5, 199–207. [Google Scholar] [CrossRef] [PubMed]
  10. Stephenson, K.; Kennedy, L.; Hargrove, L.; Demieville, J.; Thomson, J.; Alpini, G.; Francis, H. Updates on Dietary Models of Nonalcoholic Fatty Liver Disease: Current Studies and Insights. Gene Expr. 2018, 18, 5–17. [Google Scholar] [CrossRef] [PubMed]
  11. Popescu, M.; Popescu, I.A.; Stanciu, M.; Cazacu, S.M.; Ianosi, N.G.; Comanescu, M.V.; Singer, C.E.; Neagoe, C.D. Non-alcoholic fatty liver disease—Clinical and histopathological aspects. Rom. J. Morphol. Embryol. 2016, 57, 1295–1302. [Google Scholar] [PubMed]
  12. Ekstedt, M.; Franzén, L.E.; Mathiesen, U.L.; Thorelius, L.; Holmqvist, M.; Bodemar, G.; Kechagias, S. Long-term follow-up of patients with NAFLD and elevated liver enzymes. Hepatology 2006, 44, 865–873. [Google Scholar] [CrossRef] [PubMed]
  13. White, D.L.; Kanwal, F.; El-Serag, H.B. Association between nonalcoholic fatty liver disease and risk for hepatocellular cancer, based on systematic review. Clin. Gastroenterol. Hepatol. 2012, 10, 1342–1359.e2. [Google Scholar] [CrossRef] [PubMed]
  14. Önnerhag, K.; Nilsson, P.M.; Lindgren, S. Increased risk of cirrhosis and hepatocellular cancer during long-term follow-up of patients with biopsy-proven NAFLD. Scand. J. Gastroenterol. 2014, 49, 1111–1118. [Google Scholar] [CrossRef] [PubMed]
  15. Powell, E.E.; Cooksley, W.G.; Hanson, R.; Searle, J.; Halliday, J.W.; Powell, L.W. The natural history of nonalcoholic steatohepatitis: A follow-up study of forty-two patients for up to 21 years. Hepatology 1990, 11, 74–80. [Google Scholar] [CrossRef] [PubMed]
  16. Caldwell, S.H.; Oelsner, D.H.; Iezzoni, J.C.; Hespenheide, E.E.; Battle, E.H.; Driscoll, C.J. Cryptogenic cirrhosis: Clinical characterization and risk factors for underlying disease. Hepatology 1999, 29, 664–669. [Google Scholar] [CrossRef] [PubMed]
  17. Calzadilla Bertot, L.; Adams, L.A. The Natural Course of Non-Alcoholic Fatty Liver Disease. Int. J. Mol. Sci. 2016, 17, 774. [Google Scholar] [CrossRef] [PubMed]
  18. Williams, C.D.; Stengel, J.; Asike, M.I.; Torres, D.M.; Shaw, J.; Contreras, M.; Landt, C.L.; Harrison, S.A. Prevalence of nonalcoholic fatty liver disease and nonalcoholic steatohepatitis among a largely middle-aged population utilizing ultrasound and liver biopsy: A prospective study. Gastroenterology 2011, 140, 124–131. [Google Scholar] [CrossRef] [PubMed]
  19. Younossi, Z.M.; Stepanova, M.; Afendy, M.; Fang, Y.; Younossi, Y.; Mir, H.; Srishord, M. Changes in the prevalence of the most common causes of chronic liver diseases in the United States from 1988 to 2008. Clin. Gastroenterol. Hepatol 2011, 9, 524–530.e1. quiz e560. [Google Scholar] [CrossRef] [PubMed]
  20. Adams, L.A.; Sanderson, S.; Lindor, K.D.; Angulo, P. The histological course of nonalcoholic fatty liver disease: A longitudinal study of 103 patients with sequential liver biopsies. J. Hepatol. 2005, 42, 132–138. [Google Scholar] [CrossRef] [PubMed]
  21. Williamson, R.M.; Price, J.F.; Glancy, S.; Perry, E.; Nee, L.D.; Hayes, P.C.; Frier, B.M.; Van Look, L.A.; Johnston, G.I.; Reynolds, R.M.; et al. Prevalence of and risk factors for hepatic steatosis and nonalcoholic Fatty liver disease in people with type 2 diabetes: The Edinburgh Type 2 Diabetes Study. Diabetes Care 2011, 34, 1139–1144. [Google Scholar] [CrossRef] [PubMed]
  22. Muzurovic, E.; Mikhailidis, D.P.; Mantzoros, C. Non-alcoholic fatty liver disease, insulin resistance, metabolic syndrome and their association with vascular risk. Metabolism 2021, 119, 154770. [Google Scholar] [CrossRef] [PubMed]
  23. Lee, J.H.; Kwon, Y.J.; Park, K.; Lee, H.S.; Park, H.K.; Han, J.H.; Ahn, S.B. Metabolic Score for Insulin Resistance Is Inversely Related to Incident Advanced Liver Fibrosis in Patients with Non-Alcoholic Fatty Liver Disease. Nutrients 2022, 14, 3039. [Google Scholar] [CrossRef] [PubMed]
  24. Chew, N.W.S.; Hui Pan, X.; Chong, B.; Chandramouli, C.; Muthiah, M.; Lam, C.S.P. Type 2 diabetes mellitus and cardiometabolic outcomes in metabolic dysfunction-associated steatotic liver disease population. Diabetes Res. Clin. Pract. 2024, 211, 111652. [Google Scholar] [CrossRef] [PubMed]
  25. Hamaguchi, M.; Kojima, T.; Takeda, N.; Nakagawa, T.; Taniguchi, H.; Fujii, K.; Omatsu, T.; Nakajima, T.; Sarui, H.; Shimazaki, M.; et al. The metabolic syndrome as a predictor of nonalcoholic fatty liver disease. Ann. Intern. Med. 2005, 143, 722–728. [Google Scholar] [CrossRef] [PubMed]
  26. Marchesini, G.; Brizi, M.; Bianchi, G.; Tomassetti, S.; Bugianesi, E.; Lenzi, M.; McCullough, A.J.; Natale, S.; Forlani, G.; Melchionda, N. Nonalcoholic fatty liver disease: A feature of the metabolic syndrome. Diabetes 2001, 50, 1844–1850. [Google Scholar] [CrossRef] [PubMed]
  27. Ratziu, V.; Bellentani, S.; Cortez-Pinto, H.; Day, C.; Marchesini, G. A position statement on NAFLD/NASH based on the EASL 2009 special conference. J. Hepatol. 2010, 53, 372–384. [Google Scholar] [CrossRef] [PubMed]
  28. Hillenbrand, A.; Kiebler, B.; Schwab, C.; Scheja, L.; Xu, P.; Henne-Bruns, D.; Wolf, A.M.; Knippschild, U. Prevalence of non-alcoholic fatty liver disease in four different weight related patient groups: Association with small bowel length and risk factors. BMC Res. Notes 2015, 8, 290. [Google Scholar] [CrossRef] [PubMed]
  29. Harrison, S.A.; Bedossa, P.; Guy, C.D.; Schattenberg, J.M.; Loomba, R.; Taub, R.; Labriola, D.; Moussa, S.E.; Neff, G.W.; Rinella, M.E.; et al. A Phase 3, Randomized, Controlled Trial of Resmetirom in NASH with Liver Fibrosis. N. Engl. J. Med. 2024, 390, 497–509. [Google Scholar] [CrossRef] [PubMed]
  30. Dharmalingam, M.; Yamasandhi, P.G. Nonalcoholic Fatty Liver Disease and Type 2 Diabetes Mellitus. Indian J. Endocrinol. Metab. 2018, 22, 421–428. [Google Scholar] [CrossRef] [PubMed]
  31. 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] [PubMed]
  32. Lonardo, A.; Bellentani, S.; Argo, C.K.; Ballestri, S.; Byrne, C.D.; Caldwell, S.H.; Cortez-Pinto, H.; Grieco, A.; Machado, M.V.; Miele, L.; et al. Epidemiological modifiers of non-alcoholic fatty liver disease: Focus on high-risk groups. Dig. Liver Dis. 2015, 47, 997–1006. [Google Scholar] [CrossRef] [PubMed]
  33. Marriott, B.P.; Cole, N.; Lee, E. National estimates of dietary fructose intake increased from 1977 to 2004 in the United States. J. Nutr. 2009, 139, 1228S–1235S. [Google Scholar] [CrossRef] [PubMed]
  34. Lin, W.T.; Chan, T.F.; Huang, H.L.; Lee, C.Y.; Tsai, S.; Wu, P.W.; Yang, Y.C.; Wang, T.N.; Lee, C.H. Fructose-Rich Beverage Intake and Central Adiposity, Uric Acid, and Pediatric Insulin Resistance. J. Pediatr. 2016, 171, 90–96.e1. [Google Scholar] [CrossRef] [PubMed]
  35. DeChristopher, L.R. 40 years of adding more fructose to high fructose corn syrup than is safe, through the lens of malabsorption and altered gut health-gateways to chronic disease. Nutr. J. 2024, 23, 16. [Google Scholar] [CrossRef] [PubMed]
  36. Li, X.; Luan, Y.; Li, Y.; Ye, S.; Wang, G.; Cai, X.; Liang, Y.; Kord Varkaneh, H.; Luan, Y. The effect of high-fructose corn syrup vs. sucrose on anthropometric and metabolic parameters: A systematic review and meta-analysis. Front. Nutr. 2022, 9, 1013310. [Google Scholar] [CrossRef] [PubMed]
  37. Rippe, J.M.; Angelopoulos, T.J. Sucrose, high-fructose corn syrup, and fructose, their metabolism and potential health effects: What do we really know? Adv. Nutr. 2013, 4, 236–245. [Google Scholar] [CrossRef] [PubMed]
  38. Bray, G.A.; Nielsen, S.J.; Popkin, B.M. Consumption of high-fructose corn syrup in beverages may play a role in the epidemic of obesity. Am. J. Clin. Nutr. 2004, 79, 537–543. [Google Scholar] [CrossRef] [PubMed]
  39. Ter Horst, K.W.; Serlie, M.J. Fructose Consumption, Lipogenesis, and Non-Alcoholic Fatty Liver Disease. Nutrients 2017, 9, 981. [Google Scholar] [CrossRef] [PubMed]
  40. Faller, J.; Fox, I.H. Ethanol-induced hyperuricemia: Evidence for increased urate production by activation of adenine nucleotide turnover. N. Engl. J. Med. 1982, 307, 1598–1602. [Google Scholar] [CrossRef] [PubMed]
  41. Zhang, P.; Sun, H.; Cheng, X.; Li, Y.; Zhao, Y.; Mei, W.; Wei, X.; Zhou, H.; Du, Y.; Zeng, C. Dietary intake of fructose increases purine de novo synthesis: A crucial mechanism for hyperuricemia. Front. Nutr. 2022, 9, 1045805. [Google Scholar] [CrossRef] [PubMed]
  42. Zheng, Z.; Harman, J.L.; Coresh, J.; Kottgen, A.; McAdams-DeMarco, M.A.; Correa, A.; Young, B.A.; Katz, R.; Rebholz, C.M. The Dietary Fructose:Vitamin C Intake Ratio Is Associated with Hyperuricemia in African-American Adults. J. Nutr. 2018, 148, 419–426. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, C.; Yang, S.; Xu, W.; Zhang, J.; Fu, W.; Feng, C. Association between the hyperuricemia and nonalcoholic fatty liver disease risk in a Chinese population: A retrospective cohort study. PLoS ONE 2017, 12, e0177249. [Google Scholar] [CrossRef] [PubMed]
  44. Zhou, M.; Yang, N.; Xing, X.; Chang, D.; Li, J.; Deng, J.; Chen, Y.; Hu, C.; Zhang, R.; Lu, X.; et al. Obesity interacts with hyperuricemia on the severity of non-alcoholic fatty liver disease. BMC Gastroenterol. 2021, 21, 43. [Google Scholar] [CrossRef] [PubMed]
  45. Huang, C.; Liang, Z.; Ma, J.; Hu, D.; Yao, F.; Qin, P. Total sugar, added sugar, fructose, and sucrose intake and all-cause, cardiovascular, and cancer mortality: A systematic review and dose-response meta-analysis of prospective cohort studies. Nutrition 2023, 111, 112032. [Google Scholar] [CrossRef] [PubMed]
  46. Schaefer, S.M.; Kaiser, A.; Eichner, G.; Fasshauer, M. Association of sugar intake from different sources with cardiovascular disease incidence in the prospective cohort of UK Biobank participants. Nutr. J. 2024, 23, 22. [Google Scholar] [CrossRef] [PubMed]
  47. Khan, T.A.; Tayyiba, M.; Agarwal, A.; Mejia, S.B.; de Souza, R.J.; Wolever, T.M.S.; Leiter, L.A.; Kendall, C.W.C.; Jenkins, D.J.A.; Sievenpiper, J.L. Relation of Total Sugars, Sucrose, Fructose, and Added Sugars with the Risk of Cardiovascular Disease: A Systematic Review and Dose-Response Meta-analysis of Prospective Cohort Studies. Mayo Clin. Proc. 2019, 94, 2399–2414. [Google Scholar] [CrossRef] [PubMed]
  48. Kilonzo, S.B.; Kamala, E.; Jaka, H.; Ngoya, P. Non-alcoholic fatty liver disease in Tanzania: Prevalence, determinants, and diagnostic performance of triglycerides-glucose index and triglycerides-glucose inde -body mass index compared to the hepatic ultrasound in overweight and obese individuals. BMC Gastroenterol. 2024, 24, 96. [Google Scholar] [CrossRef] [PubMed]
  49. Wahjudi, P.N.; Patterson, M.E.; Lim, S.; Yee, J.K.; Mao, C.S.; Lee, W.N. Measurement of glucose and fructose in clinical samples using gas chromatography/mass spectrometry. Clin. Biochem. 2010, 43, 198–207. [Google Scholar] [CrossRef] [PubMed]
  50. Preston, G.M.; Calle, R.A. Elevated Serum Sorbitol and not Fructose in Type 2 Diabetic Patients. Biomark. Insights 2010, 5, 33–38. [Google Scholar] [CrossRef] [PubMed]
  51. Barone, S.; Fussell, S.L.; Singh, A.K.; Lucas, F.; Xu, J.; Kim, C.; Wu, X.; Yu, Y.; Amlal, H.; Seidler, U.; et al. Slc2a5 (Glut5) is essential for the absorption of fructose in the intestine and generation of fructose-induced hypertension. J. Biol. Chem. 2009, 284, 5056–5066. [Google Scholar] [CrossRef] [PubMed]
  52. Davidson, N.O.; Hausman, A.M.; Ifkovits, C.A.; Buse, J.B.; Gould, G.W.; Burant, C.F.; Bell, G.I. Human intestinal glucose transporter expression and localization of GLUT5. Am. J. Physiol. 1992, 262, C795–C800. [Google Scholar] [CrossRef] [PubMed]
  53. Ferraris, R.P. Dietary and developmental regulation of intestinal sugar transport. Biochem. J. 2001, 360, 265–276. [Google Scholar] [CrossRef] [PubMed]
  54. Kretowicz, M.; Johnson, R.J.; Ishimoto, T.; Nakagawa, T.; Manitius, J. The impact of fructose on renal function and blood pressure. Int. J. Nephrol. 2011, 2011, 315879. [Google Scholar] [CrossRef] [PubMed]
  55. Dyer, J.; Wood, I.S.; Palejwala, A.; Ellis, A.; Shirazi-Beechey, S.P. Expression of monosaccharide transporters in intestine of diabetic humans. Am. J. Physiol. Gastrointest. Liver Physiol. 2002, 282, G241–G248. [Google Scholar] [CrossRef] [PubMed]
  56. Stuart, C.A.; Howell, M.E.; Yin, D. Overexpression of GLUT5 in diabetic muscle is reversed by pioglitazone. Diabetes Care 2007, 30, 925–931. [Google Scholar] [CrossRef] [PubMed]
  57. Wlodarczyk, J.; Wlodarczyk, M.; Zielinska, M.; Jedrzejczak, B.; Dziki, L.; Fichna, J. Blockade of fructose transporter protein GLUT5 inhibits proliferation of colon cancer cells: Proof of concept for a new class of anti-tumor therapeutics. Pharmacol. Rep. 2021, 73, 939–945. [Google Scholar] [CrossRef] [PubMed]
  58. Wuest, M.; Hamann, I.; Bouvet, V.; Glubrecht, D.; Marshall, A.; Trayner, B.; Soueidan, O.M.; Krys, D.; Wagner, M.; Cheeseman, C.; et al. Molecular Imaging of GLUT1 and GLUT5 in Breast Cancer: A Multitracer Positron Emission Tomography Imaging Study in Mice. Mol. Pharmacol. 2018, 93, 79–89. [Google Scholar] [CrossRef] [PubMed]
  59. Fu, Y.; Maianu, L.; Melbert, B.R.; Garvey, W.T. Facilitative glucose transporter gene expression in human lymphocytes, monocytes, and macrophages: A role for GLUT isoforms 1, 3, and 5 in the immune response and foam cell formation. Blood Cells Mol. Dis. 2004, 32, 182–190. [Google Scholar] [CrossRef] [PubMed]
  60. Malide, D.; Davies-Hill, T.M.; Levine, M.; Simpson, I.A. Distinct localization of GLUT-1, -3, and -5 in human monocyte-derived macrophages: Effects of cell activation. Am. J. Physiol. 1998, 274, E516–E526. [Google Scholar] [CrossRef] [PubMed]
  61. Grossbard, L.; Schimke, R.T. Multiple hexokinases of rat tissues. Purification and comparison of soluble forms. J. Biol. Chem. 1966, 241, 3546–3560. [Google Scholar] [CrossRef] [PubMed]
  62. Andres-Hernando, A.; Orlicky, D.J.; Kuwabara, M.; Ishimoto, T.; Nakagawa, T.; Johnson, R.J.; Lanaspa, M.A. Deletion of Fructokinase in the Liver or in the Intestine Reveals Differential Effects on Sugar-Induced Metabolic Dysfunction. Cell Metab. 2020, 32, 117–127.e3. [Google Scholar] [CrossRef] [PubMed]
  63. Shepherd, E.L.; Saborano, R.; Northall, E.; Matsuda, K.; Ogino, H.; Yashiro, H.; Pickens, J.; Feaver, R.E.; Cole, B.K.; Hoang, S.A.; et al. Ketohexokinase inhibition improves NASH by reducing fructose-induced steatosis and fibrogenesis. JHEP Rep. 2021, 3, 100217. [Google Scholar] [CrossRef] [PubMed]
  64. Gutierrez, J.A.; Liu, W.; Perez, S.; Xing, G.; Sonnenberg, G.; Kou, K.; Blatnik, M.; Allen, R.; Weng, Y.; Vera, N.B.; et al. Pharmacologic inhibition of ketohexokinase prevents fructose-induced metabolic dysfunction. Mol. Metab. 2021, 48, 101196. [Google Scholar] [CrossRef] [PubMed]
  65. Li, X.; Qian, X.; Peng, L.X.; Jiang, Y.; Hawke, D.H.; Zheng, Y.; Xia, Y.; Lee, J.H.; Cote, G.; Wang, H.; et al. A splicing switch from ketohexokinase-C to ketohexokinase-A drives hepatocellular carcinoma formation. Nat. Cell Biol. 2016, 18, 561–571. [Google Scholar] [CrossRef] [PubMed]
  66. Kim, J.; Kang, J.; Kang, Y.L.; Woo, J.; Kim, Y.; Huh, J.; Park, J.W. Ketohexokinase-A acts as a nuclear protein kinase that mediates fructose-induced metastasis in breast cancer. Nat. Commun. 2020, 11, 5436. [Google Scholar] [CrossRef] [PubMed]
  67. Eberhart, T.; Schonenberger, M.J.; Walter, K.M.; Charles, K.N.; Faust, P.L.; Kovacs, W.J. Peroxisome-Deficiency and HIF-2alpha Signaling Are Negative Regulators of Ketohexokinase Expression. Front. Cell Dev. Biol. 2020, 8, 566. [Google Scholar] [CrossRef] [PubMed]
  68. Jang, C.; Hui, S.; Lu, W.; Cowan, A.J.; Morscher, R.J.; Lee, G.; Liu, W.; Tesz, G.J.; Birnbaum, M.J.; Rabinowitz, J.D. The Small Intestine Converts Dietary Fructose into Glucose and Organic Acids. Cell Metab. 2018, 27, 351–361.e3. [Google Scholar] [CrossRef] [PubMed]
  69. Debosch, B.J.; Chen, Z.; Saben, J.L.; Finck, B.N.; Moley, K.H. Glucose transporter 8 (GLUT8) mediates fructose-induced de novo lipogenesis and macrosteatosis. J. Biol. Chem. 2014, 289, 10989–10998. [Google Scholar] [CrossRef] [PubMed]
  70. Novelle, M.G.; Bravo, S.B.; Deshons, M.; Iglesias, C.; García-Vence, M.; Annells, R.; da Silva Lima, N.; Nogueiras, R.; Fernández-Rojo, M.A.; Diéguez, C.; et al. Impact of liver-specific GLUT8 silencing on fructose-induced inflammation and omega oxidation. iScience 2021, 24, 102071. [Google Scholar] [CrossRef] [PubMed]
  71. Colville, C.A.; Seatter, M.J.; Jess, T.J.; Gould, G.W.; Thomas, H.M. Kinetic analysis of the liver-type (GLUT2) and brain-type (GLUT3) glucose transporters in Xenopus oocytes: Substrate specificities and effects of transport inhibitors. Biochem. J. 1993, 290, 701–706. [Google Scholar] [CrossRef] [PubMed]
  72. Oppelt, S.A.; Sennott, E.M.; Tolan, D.R. Aldolase-B knockout in mice phenocopies hereditary fructose intolerance in humans. Mol. Genet. Metab. 2015, 114, 445–450. [Google Scholar] [CrossRef] [PubMed]
  73. Liu, L.; Li, T.; Liao, Y.; Wang, Y.; Gao, Y.; Hu, H.; Huang, H.; Wu, F.; Chen, Y.G.; Xu, S.; et al. Triose Kinase Controls the Lipogenic Potential of Fructose and Dietary Tolerance. Cell Metab. 2020, 32, 605–618.e7. [Google Scholar] [CrossRef] [PubMed]
  74. Bjorkman, O.; Felig, P. Role of the kidney in the metabolism of fructose in 60-hour fasted humans. Diabetes 1982, 31, 516–520. [Google Scholar] [CrossRef] [PubMed]
  75. Lodge, M.; Scheidemantle, G.; Adams, V.R.; Cottam, M.A.; Richard, D.; Breuer, D.; Thompson, P.; Shrestha, K.; Liu, X.; Kennedy, A. Fructose regulates the pentose phosphate pathway and induces an inflammatory and resolution phenotype in Kupffer cells. Sci. Rep. 2024, 14, 4020. [Google Scholar] [CrossRef] [PubMed]
  76. Jones, N.; Blagih, J.; Zani, F.; Rees, A.; Hill, D.G.; Jenkins, B.J.; Bull, C.J.; Moreira, D.; Bantan, A.I.M.; Cronin, J.G.; et al. Fructose reprogrammes glutamine-dependent oxidative metabolism to support LPS-induced inflammation. Nat. Commun. 2021, 12, 1209. [Google Scholar] [CrossRef] [PubMed]
  77. Li, Y.; Jiang, T.; Du, M.; He, S.; Huang, N.; Cheng, B.; Yan, C.; Tang, W.; Gao, W.; Guo, H.; et al. Ketohexokinase-dependent metabolism of cerebral endogenous fructose in microglia drives diabetes-associated cognitive dysfunction. Exp. Mol. Med. 2023, 55, 2417–2432. [Google Scholar] [CrossRef] [PubMed]
  78. Endo, A.; Maeno, S.; Tanizawa, Y.; Kneifel, W.; Arita, M.; Dicks, L.; Salminen, S. Fructophilic Lactic Acid Bacteria, a Unique Group of Fructose-Fermenting Microbes. Appl. Environ. Microbiol. 2018, 84, 19. [Google Scholar] [CrossRef] [PubMed]
  79. Lanaspa, M.A.; Ishimoto, T.; Li, N.; Cicerchi, C.; Orlicky, D.J.; Ruzycki, P.; Rivard, C.; Inaba, S.; Roncal-Jimenez, C.A.; Bales, E.S.; et al. Endogenous fructose production and metabolism in the liver contributes to the development of metabolic syndrome. Nat. Commun. 2013, 4, 2434. [Google Scholar] [CrossRef] [PubMed]
  80. Kang, Y.L.; Kim, J.; Kwak, S.B.; Kim, Y.S.; Huh, J.; Park, J.W. The polyol pathway and nuclear ketohexokinase A signaling drive hyperglycemia-induced metastasis of gastric cancer. Exp. Mol. Med. 2024, 56, 220–234. [Google Scholar] [CrossRef] [PubMed]
  81. Andres-Hernando, A.; Orlicky, D.J.; Kuwabara, M.; Cicerchi, C.; Pedler, M.; Petrash, M.J.; Johnson, R.J.; Tolan, D.R.; Lanaspa, M.A. Endogenous Fructose Production and Metabolism Drive Metabolic Dysregulation and Liver Disease in Mice with Hereditary Fructose Intolerance. Nutrients 2023, 15, 4376. [Google Scholar] [CrossRef] [PubMed]
  82. Jeon, D.; Choi, W.M.; Kim, J.S.; Jung, Y.; Lee, S.Y.; Seo, H.R.; Kim, K.M. Serum Sorbitol Dehydrogenase as a Novel Prognostic Factor for Hepatocellular Carcinoma after Surgical Resection. Cancers 2021, 13, 6143. [Google Scholar] [CrossRef] [PubMed]
  83. Sanchez-Lozada, L.G.; Andres-Hernando, A.; Garcia-Arroyo, F.E.; Cicerchi, C.; Li, N.; Kuwabara, M.; Roncal-Jimenez, C.A.; Johnson, R.J.; Lanaspa, M.A. Uric acid activates aldose reductase and the polyol pathway for endogenous fructose and fat production causing development of fatty liver in rats. J. Biol. Chem. 2019, 294, 4272–4281. [Google Scholar] [CrossRef] [PubMed]
  84. Todoric, J.; Di Caro, G.; Reibe, S.; Henstridge, D.C.; Green, C.R.; Vrbanac, A.; Ceteci, F.; Conche, C.; McNulty, R.; Shalapour, S.; et al. Fructose stimulated de novo lipogenesis is promoted by inflammation. Nat. Metab. 2020, 2, 1034–1045. [Google Scholar] [CrossRef] [PubMed]
  85. Taylor, S.R.; Ramsamooj, S.; Liang, R.J.; Katti, A.; Pozovskiy, R.; Vasan, N.; Hwang, S.K.; Nahiyaan, N.; Francoeur, N.J.; Schatoff, E.M.; et al. Dietary fructose improves intestinal cell survival and nutrient absorption. Nature 2021, 597, 263–267. [Google Scholar] [CrossRef] [PubMed]
  86. Tan, R.; Dong, H.; Chen, Z.; Jin, M.; Yin, J.; Li, H.; Shi, D.; Shao, Y.; Wang, H.; Chen, T.; et al. Intestinal Microbiota Mediates High-Fructose and High-Fat Diets to Induce Chronic Intestinal Inflammation. Front. Cell Infect. Microbiol. 2021, 11, 654074. [Google Scholar] [CrossRef] [PubMed]
  87. Montrose, D.C.; Nishiguchi, R.; Basu, S.; Staab, H.A.; Zhou, X.K.; Wang, H.; Meng, L.; Johncilla, M.; Cubillos-Ruiz, J.R.; Morales, D.K.; et al. Dietary Fructose Alters the Composition, Localization, and Metabolism of Gut Microbiota in Association with Worsening Colitis. Cell Mol. Gastroenterol. Hepatol. 2021, 11, 525–550. [Google Scholar] [CrossRef] [PubMed]
  88. Letsinger, A.C.; Menon, R.; Iyer, A.R.; Vellers, H.L.; Granados, J.Z.; Jayaraman, A.; Lightfoot, J.T. A High Fat/High Sugar Diet Alters the Gastrointestinal Metabolome in a Sex Dependent Manner. Metabolites 2020, 10, 421. [Google Scholar] [CrossRef] [PubMed]
  89. Ham, M.; Lee, J.W.; Choi, A.H.; Jang, H.; Choi, G.; Park, J.; Kozuka, C.; Sears, D.D.; Masuzaki, H.; Kim, J.B. Macrophage glucose-6-phosphate dehydrogenase stimulates proinflammatory responses with oxidative stress. Mol. Cell Biol. 2013, 33, 2425–2435. [Google Scholar] [CrossRef] [PubMed]
  90. Wang, X.; Zhu, L.; Li, X.; Hao, R.; Li, J. Effects of high fructose corn syrup on intestinal microbiota structure and obesity in mice. NPJ Sci. Food 2022, 6, 17. [Google Scholar] [CrossRef] [PubMed]
  91. Oh, J.H.; Alexander, L.M.; Pan, M.; Schueler, K.L.; Keller, M.P.; Attie, A.D.; Walter, J.; van Pijkeren, J.P. Dietary Fructose and Microbiota-Derived Short-Chain Fatty Acids Promote Bacteriophage Production in the Gut Symbiont Lactobacillus reuteri. Cell Host Microbe 2019, 25, 273–284.e6. [Google Scholar] [CrossRef] [PubMed]
  92. Caescu, C.I.; Vidal, O.; Krzewinski, F.; Artenie, V.; Bouquelet, S. Bifidobacterium longum requires a fructokinase (Frk; ATP:D-fructose 6-phosphotransferase, EC 2.7.1.4) for fructose catabolism. J. Bacteriol. 2004, 186, 6515–6525. [Google Scholar] [CrossRef] [PubMed]
  93. Wei, L.; Xia, F.; Wang, J.; Ran, S.; Liang, Y.; Zhou, W.; Huang, Z.; Liang, J. Carbohydrate Metabolism Affects Macrophage-Mediated Killing of Enterococcus faecalis. mSystems 2021, 6, e0043421. [Google Scholar] [CrossRef] [PubMed]
  94. Isaac, S.; Flor-Duro, A.; Carruana, G.; Puchades-Carrasco, L.; Quirant, A.; Lopez-Nogueroles, M.; Pineda-Lucena, A.; Garcia-Garcera, M.; Ubeda, C. Microbiome-mediated fructose depletion restricts murine gut colonization by vancomycin-resistant Enterococcus. Nat. Commun. 2022, 13, 7718. [Google Scholar] [CrossRef] [PubMed]
  95. Gerritsen, J.; Hornung, B.; Renckens, B.; van Hijum, S.A.F.T.; Martins Dos Santos, V.A.P.; Rijkers, G.T.; Schaap, P.J.; de Vos, W.M.; Smidt, H. Genomic and functional analysis of. PeerJ 2017, 5, e3698. [Google Scholar] [CrossRef] [PubMed]
  96. Beisner, J.; Gonzalez-Granda, A.; Basrai, M.; Damms-Machado, A.; Bischoff, S.C. Fructose-Induced Intestinal Microbiota Shift Following Two Types of Short-Term High-Fructose Dietary Phases. Nutrients 2020, 12, 3444. [Google Scholar] [CrossRef] [PubMed]
  97. Jones, R.B.; Alderete, T.L.; Kim, J.S.; Millstein, J.; Gilliland, F.D.; Goran, M.I. High intake of dietary fructose in overweight/obese teenagers associated with depletion of. Gut Microbes 2019, 10, 712–719. [Google Scholar] [CrossRef] [PubMed]
  98. Goffredo, M.; Mass, K.; Parks, E.J.; Wagner, D.A.; McClure, E.A.; Graf, J.; Savoye, M.; Pierpont, B.; Cline, G.; Santoro, N. Role of Gut Microbiota and Short Chain Fatty Acids in Modulating Energy Harvest and Fat Partitioning in Youth. J. Clin. Endocrinol. Metab. 2016, 101, 4367–4376. [Google Scholar] [CrossRef] [PubMed]
  99. Jones, H.F.; Butler, R.N.; Brooks, D.A. Intestinal fructose transport and malabsorption in humans. Am. J. Physiol. Gastrointest. Liver Physiol. 2011, 300, G202–G206. [Google Scholar] [CrossRef] [PubMed]
  100. Kawabata, K.; Kanmura, S.; Morinaga, Y.; Tanaka, A.; Makino, T.; Fujita, T.; Arima, S.; Sasaki, F.; Nasu, Y.; Tanoue, S.; et al. A high-fructose diet induces epithelial barrier dysfunction and exacerbates the severity of dextran sulfate sodium-induced colitis. Int. J. Mol. Med. 2019, 43, 1487–1496. [Google Scholar] [CrossRef] [PubMed]
  101. Taskinen, M.R.; Packard, C.J.; Boren, J. Dietary Fructose and the Metabolic Syndrome. Nutrients 2019, 11, 1987. [Google Scholar] [CrossRef] [PubMed]
  102. Softic, S.; Meyer, J.G.; Wang, G.X.; Gupta, M.K.; Batista, T.M.; Lauritzen, H.P.M.M.; Fujisaka, S.; Serra, D.; Herrero, L.; Willoughby, J.; et al. Dietary Sugars Alter Hepatic Fatty Acid Oxidation via Transcriptional and Post-translational Modifications of Mitochondrial Proteins. Cell Metab. 2019, 30, 735–753.e4. [Google Scholar] [CrossRef] [PubMed]
  103. Gligorovska, L.; Teofilovic, A.; Vojnovic Milutinovic, D.; Miladinovic, N.; Kovacevic, S.; Velickovic, N.; Djordjevic, A. Macrophage migration inhibitory factor deficiency aggravates effects of fructose-enriched diet on lipid metabolism in the mouse liver. Biofactors 2021, 47, 363–375. [Google Scholar] [CrossRef] [PubMed]
  104. Petito, G.; Giacco, A.; Cioffi, F.; Mazzoli, A.; Magnacca, N.; Iossa, S.; Goglia, F.; Senese, R.; Lanni, A. Short-term fructose feeding alters tissue metabolic pathways by modulating microRNAs expression both in young and adult rats. Front. Cell Dev. Biol. 2023, 11, 1101844. [Google Scholar] [CrossRef] [PubMed]
  105. Lanaspa, M.A.; Cicerchi, C.; Garcia, G.; Li, N.; Roncal-Jimenez, C.A.; Rivard, C.J.; Hunter, B.; Andres-Hernando, A.; Ishimoto, T.; Sanchez-Lozada, L.G.; et al. Counteracting roles of AMP deaminase and AMP kinase in the development of fatty liver. PLoS ONE 2012, 7, e48801. [Google Scholar] [CrossRef] [PubMed]
  106. Li, Y.; Wang, B.; Shen, J.; Bai, M.; Xu, E. Berberine attenuates fructose-induced insulin resistance by stimulating the hepatic LKB1/AMPK/PGC1alpha pathway in mice. Pharm. Biol. 2020, 58, 385–392. [Google Scholar] [CrossRef] [PubMed]
  107. Lozano, I.; Van der Werf, R.; Bietiger, W.; Seyfritz, E.; Peronet, C.; Pinget, M.; Jeandidier, N.; Maillard, E.; Marchioni, E.; Sigrist, S.; et al. High-fructose and high-fat diet-induced disorders in rats: Impact on diabetes risk, hepatic and vascular complications. Nutr. Metab. 2016, 13, 15. [Google Scholar] [CrossRef] [PubMed]
  108. Kawanishi, N.; Yano, H.; Mizokami, T.; Takahashi, M.; Oyanagi, E.; Suzuki, K. Exercise training attenuates hepatic inflammation, fibrosis and macrophage infiltration during diet induced-obesity in mice. Brain Behav. Immun. 2012, 26, 931–941. [Google Scholar] [CrossRef] [PubMed]
  109. Kohli, R.; Kirby, M.; Xanthakos, S.A.; Softic, S.; Feldstein, A.E.; Saxena, V.; Tang, P.H.; Miles, L.; Miles, M.V.; Balistreri, W.F.; et al. High-fructose, medium chain trans fat diet induces liver fibrosis and elevates plasma coenzyme Q9 in a novel murine model of obesity and nonalcoholic steatohepatitis. Hepatology 2010, 52, 934–944. [Google Scholar] [CrossRef] [PubMed]
  110. Wen, Y.; Lambrecht, J.; Ju, C.; Tacke, F. Hepatic macrophages in liver homeostasis and diseases-diversity, plasticity and therapeutic opportunities. Cell Mol. Immunol. 2021, 18, 45–56. [Google Scholar] [CrossRef] [PubMed]
  111. Guilliams, M.; Bonnardel, J.; Haest, B.; Vanderborght, B.; Wagner, C.; Remmerie, A.; Bujko, A.; Martens, L.; Thone, T.; Browaeys, R.; et al. Spatial proteogenomics reveals distinct and evolutionarily conserved hepatic macrophage niches. Cell 2022, 185, 379–396.e38. [Google Scholar] [CrossRef] [PubMed]
  112. Mosser, D.M.; Edwards, J.P. Exploring the full spectrum of macrophage activation. Nat. Rev. Immunol. 2008, 8, 958–969. [Google Scholar] [CrossRef] [PubMed]
  113. Jurgens, H.; Haass, W.; Castaneda, T.R.; Schurmann, 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] [PubMed]
  114. Pal, A.; Sun, S.; Armstrong, M.; Manke, J.; Reisdorph, N.; Adams, V.R.; Kennedy, A.; Zu, Y.; Moustaid-Moussa, N.; Carroll, I.; et al. Beneficial effects of eicosapentaenoic acid on the metabolic profile of obese female mice entails upregulation of HEPEs and increased abundance of enteric Akkermansia muciniphila. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2022, 1867, 159059. [Google Scholar] [CrossRef] [PubMed]
  115. Legeza, B.; Marcolongo, P.; Gamberucci, A.; Varga, V.; Banhegyi, G.; Benedetti, A.; Odermatt, A. Fructose, Glucocorticoids and Adipose Tissue: Implications for the Metabolic Syndrome. Nutrients 2017, 9, 426. [Google Scholar] [CrossRef] [PubMed]
  116. Du, L.; Heaney, A.P. Regulation of adipose differentiation by fructose and GluT5. Mol. Endocrinol. 2012, 26, 1773–1782. [Google Scholar] [CrossRef] [PubMed]
  117. Magliano, D.C.; Penna-de-Carvalho, A.; Vazquez-Carrera, M.; Mandarim-de-Lacerda, C.A.; Aguila, M.B. Short-term administration of GW501516 improves inflammatory state in white adipose tissue and liver damage in high-fructose-fed mice through modulation of the renin-angiotensin system. Endocrine 2015, 50, 355–367. [Google Scholar] [CrossRef] [PubMed]
  118. Pektas, M.B.; Koca, H.B.; Sadi, G.; Akar, F. Dietary Fructose Activates Insulin Signaling and Inflammation in Adipose Tissue: Modulatory Role of Resveratrol. Biomed. Res. Int. 2016, 2016, 8014252. [Google Scholar] [CrossRef] [PubMed]
  119. Song, L.; Dong, G.; Guo, L.; Graves, D.T. The function of dendritic cells in modulating the host response. Mol. Oral. Microbiol. 2018, 33, 13–21. [Google Scholar] [CrossRef] [PubMed]
  120. Jaiswal, N.; Agrawal, S.; Agrawal, A. High fructose-induced metabolic changes enhance inflammation in human dendritic cells. Clin. Exp. Immunol. 2019, 197, 237–249. [Google Scholar] [CrossRef] [PubMed]
  121. Kotsiliti, E.; Leone, V.; Schuehle, S.; Govaere, O.; Li, H.; Wolf, M.J.; Horvatic, H.; Bierwirth, S.; Hundertmark, J.; Inverso, D.; et al. Intestinal B cells license metabolic T-cell activation in NASH microbiota/antigen-independently and contribute to fibrosis by IgA-FcR signalling. J. Hepatol. 2023, 79, 296–313. [Google Scholar] [CrossRef] [PubMed]
  122. Karl, M.; Hasselwander, S.; Zhou, Y.; Reifenberg, G.; Kim, Y.O.; Park, K.S.; Ridder, D.A.; Wang, X.; Seidel, E.; Hovelmeyer, N.; et al. Dual roles of B lymphocytes in mouse models of diet-induced nonalcoholic fatty liver disease. Hepatology 2022, 76, 1135–1149. [Google Scholar] [CrossRef] [PubMed]
  123. Ramachandran, P.; Dobie, R.; Wilson-Kanamori, J.R.; Dora, E.F.; Henderson, B.E.P.; Luu, N.T.; Portman, J.R.; Matchett, K.P.; Brice, M.; Marwick, J.A.; et al. Resolving the fibrotic niche of human liver cirrhosis at single-cell level. Nature 2019, 575, 512–518. [Google Scholar] [CrossRef] [PubMed]
  124. Bruzzi, 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] [PubMed]
  125. Ijare, O.B.; Baskin, D.S.; Sharpe, M.A.; Pichumani, K. Metabolism of fructose in B-cells: A (13)C NMR spectroscopy based stable isotope tracer study. Anal. Biochem. 2018, 552, 110–117. [Google Scholar] [CrossRef] [PubMed]
  126. Tan, J.; Ni, D.; Wali, J.A.; Cox, D.A.; Pinget, G.V.; Taitz, J.; Daien, C.I.; Senior, A.; Read, M.N.; Simpson, S.J.; et al. Dietary carbohydrate, particularly glucose, drives B cell lymphopoiesis and function. iScience 2021, 24, 102835. [Google Scholar] [CrossRef] [PubMed]
  127. Simons, N.; Veeraiah, P.; Simons, P.; Schaper, N.C.; Kooi, M.E.; Schrauwen-Hinderling, V.B.; Feskens, E.J.M.; van der Ploeg, E.; Van den Eynde, M.D.G.; Schalkwijk, C.G.; et al. Effects of fructose restriction on liver steatosis (FRUITLESS); a double-blind randomized controlled trial. Am. J. Clin. Nutr. 2021, 113, 391–400. [Google Scholar] [CrossRef] [PubMed]
  128. Li, X.; Joh, H.K.; Hur, J.; Song, M.; Zhang, X.; Cao, Y.; Wu, K.; Giovannucci, E.L. Fructose consumption from different food sources and cardiometabolic biomarkers: Cross-sectional associations in US men and women. Am. J. Clin. Nutr. 2023, 117, 490–498. [Google Scholar] [CrossRef] [PubMed]
  129. Yu, Z.; Ley, S.H.; Sun, Q.; Hu, F.B.; Malik, V.S. Cross-sectional association between sugar-sweetened beverage intake and cardiometabolic biomarkers in US women. Br. J. Nutr. 2018, 119, 570–580. [Google Scholar] [CrossRef] [PubMed]
  130. Buziau, A.M.; Eussen, S.; Kooi, M.E.; van der Kallen, C.J.H.; van Dongen, M.; Schaper, N.C.; Henry, R.M.A.; Schram, M.T.; Dagnelie, P.C.; van Greevenbroek, M.M.J.; et al. Fructose Intake From Fruit Juice and Sugar-Sweetened Beverages Is Associated with Higher Intrahepatic Lipid Content: The Maastricht Study. Diabetes Care 2022, 45, 1116–1123. [Google Scholar] [CrossRef]
  131. Parnell, J.A.; Raman, M.; Rioux, K.P.; Reimer, R.A. The potential role of prebiotic fibre for treatment and management of non-alcoholic fatty liver disease and associated obesity and insulin resistance. Liver Int. 2012, 32, 701–711. [Google Scholar] [CrossRef] [PubMed]
  132. Bozzetto, L.; Prinster, A.; Annuzzi, G.; Costagliola, L.; Mangione, A.; Vitelli, A.; Mazzarella, R.; Longobardo, M.; Mancini, M.; Vigorito, C.; et al. Liver fat is reduced by an isoenergetic MUFA diet in a controlled randomized study in type 2 diabetic patients. Diabetes Care 2012, 35, 1429–1435. [Google Scholar] [CrossRef] [PubMed]
  133. Cantero, I.; Abete, I.; Monreal, J.I.; Martinez, J.A.; Zulet, M.A. Fruit Fiber Consumption Specifically Improves Liver Health Status in Obese Subjects under Energy Restriction. Nutrients 2017, 9, 667. [Google Scholar] [CrossRef] [PubMed]
  134. Krawczyk, M.; Maciejewska, D.; Ryterska, K.; Czerwinka-Rogowska, M.; Jamiol-Milc, D.; Skonieczna-Zydecka, K.; Milkiewicz, P.; Raszeja-Wyszomirska, J.; Stachowska, E. Gut Permeability Might be Improved by Dietary Fiber in Individuals with Nonalcoholic Fatty Liver Disease (NAFLD) Undergoing Weight Reduction. Nutrients 2018, 10, 1793. [Google Scholar] [CrossRef] [PubMed]
  135. Howlett, J.F.; Betteridge, V.A.; Champ, M.; Craig, S.A.; Meheust, A.; Jones, J.M. The definition of dietary fiber—Discussions at the Ninth Vahouny Fiber Symposium: Building scientific agreement. Food Nutr. Res. 2010, 54, 5750. [Google Scholar] [CrossRef] [PubMed]
  136. Reiser, S.; Powell, A.S.; Scholfield, D.J.; Panda, P.; Fields, M.; Canary, J.J. Day-long glucose, insulin, and fructose responses of hyperinsulinemic and nonhyperinsulinemic men adapted to diets containing either fructose or high-amylose cornstarch. Am. J. Clin. Nutr. 1989, 50, 1008–1014. [Google Scholar] [CrossRef] [PubMed]
  137. de Melo, F.H.C.; Menezes, F.; de Sousa, J.M.B.; Dos Santos Lima, M.; da Silva Campelo Borges, G.; de Souza, E.L.; Magnani, M. Prebiotic activity of monofloral honeys produced by stingless bees in the semi-arid region of Brazilian Northeastern toward Lactobacillus acidophilus LA-05 and Bifidobacterium lactis BB-12. Food Res. Int. 2020, 128, 108809. [Google Scholar] [CrossRef] [PubMed]
  138. Gao, Y.; Zhang, W.; Zeng, L.Q.; Bai, H.; Li, J.; Zhou, J.; Zhou, G.Y.; Fang, C.W.; Wang, F.; Qin, X.J. Exercise and dietary intervention ameliorate high-fat diet-induced NAFLD and liver aging by inducing lipophagy. Redox Biol. 2020, 36, 101635. [Google Scholar] [CrossRef] [PubMed]
  139. Li, H.; Dun, Y.; Zhang, W.; You, B.; Liu, Y.; Fu, S.; Qiu, L.; Cheng, J.; Ripley-Gonzalez, J.W.; Liu, S. Exercise improves lipid droplet metabolism disorder through activation of AMPK-mediated lipophagy in NAFLD. Life Sci. 2021, 273, 119314. [Google Scholar] [CrossRef] [PubMed]
  140. Babu, A.F.; Csader, S.; Mannisto, V.; Tauriainen, M.M.; Pentikainen, H.; Savonen, K.; Klavus, A.; Koistinen, V.; Hanhineva, K.; Schwab, U. Effects of exercise on NAFLD using non-targeted metabolomics in adipose tissue, plasma, urine, and stool. Sci. Rep. 2022, 12, 6485. [Google Scholar] [CrossRef] [PubMed]
  141. Ezpeleta, M.; Gabel, K.; Cienfuegos, S.; Kalam, F.; Lin, S.; Pavlou, V.; Song, Z.; Haus, J.M.; Koppe, S.; Alexandria, S.J.; et al. Effect of alternate day fasting combined with aerobic exercise on non-alcoholic fatty liver disease: A randomized controlled trial. Cell Metab. 2023, 35, 56–70.e3. [Google Scholar] [CrossRef] [PubMed]
  142. Bidwell, A.J.; Fairchild, T.J.; Wang, L.; Keslacy, S.; Kanaley, J.A. Effect of increased physical activity on fructose-induced glycemic response in healthy individuals. Eur. J. Clin. Nutr. 2014, 68, 1048–1054. [Google Scholar] [CrossRef] [PubMed]
  143. Khalaf, H.M.; Ibrahim, M.A.; Amin, E.F.; Abdel-Tawab Ibrahim, S.; Abdel-Wahab, S.; Fouad, Y.M. Allopurinol potentiates the hepatoprotective effect of metformin and vitamin E in fructose-induced fatty liver in rats. Clin. Exp. Hepatol. 2019, 5, 65–74. [Google Scholar] [CrossRef] [PubMed]
  144. Abd El-Haleim, E.A.; Bahgat, A.K.; Saleh, S. Resveratrol and fenofibrate ameliorate fructose-induced nonalcoholic steatohepatitis by modulation of genes expression. World J. Gastroenterol. 2016, 22, 2931–2948. [Google Scholar] [CrossRef] [PubMed]
  145. Rai, R.C.; Bagul, P.K.; Banerjee, S.K. NLRP3 inflammasome drives inflammation in high fructose fed diabetic rat liver: Effect of resveratrol and metformin. Life Sci. 2020, 253, 117727. [Google Scholar] [CrossRef] [PubMed]
  146. Yilmaz Demirtas, C.; Bircan, F.S.; Pasaoglu, O.T.; Turkozkan, N. The effects of resveratrol on hepatic oxidative stress in metabolic syndrome model induced by high fructose diet. Bratisl. Lek. Listy 2018, 119, 36–40. [Google Scholar] [CrossRef] [PubMed]
  147. Shi, Y.S.; Li, C.B.; Li, X.Y.; Wu, J.; Li, Y.; Fu, X.; Zhang, Y.; Hu, W.Z. Fisetin Attenuates Metabolic Dysfunction in Mice Challenged with a High-Fructose Diet. J. Agric. Food Chem. 2018, 66, 8291–8298. [Google Scholar] [CrossRef] [PubMed]
  148. Hoofnagle, J.H.; Van Natta, M.L.; Kleiner, D.E.; Clark, J.M.; Kowdley, K.V.; Loomba, R.; Neuschwander-Tetri, B.A.; Sanyal, A.J.; Tonascia, J.; Non-alcoholic Steatohepatitis Clinical Research, N. Vitamin E and changes in serum alanine aminotransferase levels in patients with non-alcoholic steatohepatitis. Aliment. Pharmacol. Ther. 2013, 38, 134–143. [Google Scholar] [CrossRef] [PubMed]
  149. Aller, R.; Izaola, O.; Gomez, S.; Tafur, C.; Gonzalez, G.; Berroa, E.; Mora, N.; Gonzalez, J.M.; de Luis, D.A. Effect of silymarin plus vitamin E in patients with non-alcoholic fatty liver disease. A randomized clinical pilot study. Eur. Rev. Med. Pharmacol. Sci. 2015, 19, 3118–3124. [Google Scholar] [PubMed]
  150. Polyzos, S.A.; Kountouras, J.; Mantzoros, C.S.; Polymerou, V.; Katsinelos, P. Effects of combined low-dose spironolactone plus vitamin E vs vitamin E monotherapy on insulin resistance, non-invasive indices of steatosis and fibrosis, and adipokine levels in non-alcoholic fatty liver disease: A randomized controlled trial. Diabetes Obes. Metab. 2017, 19, 1805–1809. [Google Scholar] [CrossRef] [PubMed]
  151. Bril, F.; Biernacki, D.M.; Kalavalapalli, S.; Lomonaco, R.; Subbarayan, S.K.; Lai, J.; Tio, F.; Suman, A.; Orsak, B.K.; Hecht, J.; et al. Role of Vitamin E for Nonalcoholic Steatohepatitis in Patients with Type 2 Diabetes: A Randomized Controlled Trial. Diabetes Care 2019, 42, 1481–1488. [Google Scholar] [CrossRef] [PubMed]
  152. Liu, A.; Cai, Y.; Yuan, Y.; Liu, M.; Zhang, Z.; Xu, Y.; Jiao, P. Efficacy and safety of carnitine supplementation on NAFLD: A systematic review and meta-analysis. Syst. Rev. 2023, 12, 74. [Google Scholar] [CrossRef] [PubMed]
  153. Fujiwara, N.; Nakagawa, H.; Enooku, K.; Kudo, Y.; Hayata, Y.; Nakatsuka, T.; Tanaka, Y.; Tateishi, R.; Hikiba, Y.; Misumi, K.; et al. CPT2 downregulation adapts HCC to lipid-rich environment and promotes carcinogenesis via acylcarnitine accumulation in obesity. Gut 2018, 67, 1493–1504. [Google Scholar] [CrossRef] [PubMed]
  154. Montesano, A.; Senesi, P.; Vacante, F.; Mollica, G.; Benedini, S.; Mariotti, M.; Luzi, L.; Terruzzi, I. L-Carnitine counteracts in vitro fructose-induced hepatic steatosis through targeting oxidative stress markers. J. Endocrinol. Investig. 2020, 43, 493–503. [Google Scholar] [CrossRef] [PubMed]
  155. Garcia-Arroyo, F.E.; Gonzaga-Sanchez, G.; Silva-Palacios, A.; Roldan, F.J.; Loredo-Mendoza, M.L.; Alvarez-Alvarez, Y.Q.; de Los Santos Coyotl, J.A.; Velez Orozco, K.A.; Tapia, E.; Osorio-Alonso, H.; et al. Osthole Prevents Heart Damage Induced by Diet-Induced Metabolic Syndrome: Role of Fructokinase (KHK). Antioxidants 2023, 12, 1023. [Google Scholar] [CrossRef] [PubMed]
  156. Velazquez, A.M.; Bentanachs, R.; Sala-Vila, A.; Lazaro, I.; Rodriguez-Morato, J.; Sanchez, R.M.; Laguna, J.C.; Roglans, N.; Alegret, M. KHK, PNPLA3 and PPAR as Novel Targets for the Anti-Steatotic Action of Bempedoic Acid. Biomedicines 2022, 10, 1517. [Google Scholar] [CrossRef] [PubMed]
  157. Futatsugi, K.; Smith, A.C.; Tu, M.; Raymer, B.; Ahn, K.; Coffey, S.B.; Dowling, M.S.; Fernando, D.P.; Gutierrez, J.A.; Huard, K.; et al. Discovery of PF-06835919: A Potent Inhibitor of Ketohexokinase (KHK) for the Treatment of Metabolic Disorders Driven by the Overconsumption of Fructose. J. Med. Chem. 2020, 63, 13546–13560. [Google Scholar] [CrossRef] [PubMed]
  158. Kazierad, D.J.; Chidsey, K.; Somayaji, V.R.; Bergman, A.J.; Birnbaum, M.J.; Calle, R.A. Inhibition of ketohexokinase in adults with NAFLD reduces liver fat and inflammatory markers: A randomized phase 2 trial. Med. 2021, 2, 800–813.e3. [Google Scholar] [CrossRef] [PubMed]
  159. Bu, P.; Chen, K.Y.; Xiang, K.; Johnson, C.; Crown, S.B.; Rakhilin, N.; Ai, Y.; Wang, L.; Xi, R.; Astapova, I.; et al. Aldolase B-Mediated Fructose Metabolism Drives Metabolic Reprogramming of Colon Cancer Liver Metastasis. Cell Metab. 2018, 27, 1249–1262.e4. [Google Scholar] [CrossRef] [PubMed]
  160. Simons, N.; Debray, F.G.; Schaper, N.C.; Kooi, M.E.; Feskens, E.J.M.; Hollak, C.E.M.; Lindeboom, L.; Koek, G.H.; Bons, J.A.P.; Lefeber, D.J.; et al. Patients with Aldolase B Deficiency Are Characterized by Increased Intrahepatic Triglyceride Content. J. Clin. Endocrinol. Metab. 2019, 104, 5056–5064. [Google Scholar] [CrossRef] [PubMed]
  161. Karim, S.; Liaskou, E.; Fear, J.; Garg, A.; Reynolds, G.; Claridge, L.; Adams, D.H.; Newsome, P.N.; Lalor, P.F. Dysregulated hepatic expression of glucose transporters in chronic disease: Contribution of semicarbazide-sensitive amine oxidase to hepatic glucose uptake. Am. J. Physiol. Gastrointest. Liver Physiol. 2014, 307, G1180–G1190. [Google Scholar] [CrossRef] [PubMed]
  162. Tripp, J.; Essl, C.; Iancu, C.V.; Boles, E.; Choe, J.Y.; Oreb, M. Establishing a yeast-based screening system for discovery of human GLUT5 inhibitors and activators. Sci. Rep. 2017, 7, 6197. [Google Scholar] [CrossRef] [PubMed]
  163. George Thompson, A.M.; Ursu, O.; Babkin, P.; Iancu, C.V.; Whang, A.; Oprea, T.I.; Choe, J.Y. Discovery of a specific inhibitor of human GLUT5 by virtual screening and in vitro transport evaluation. Sci. Rep. 2016, 6, 24240. [Google Scholar] [CrossRef] [PubMed]
  164. Ding, S.B.; Chu, X.L.; Jin, Y.X.; Jiang, J.J.; Zhao, X.; Yu, M. Epigallocatechin gallate alleviates high-fat diet-induced hepatic lipotoxicity by targeting mitochondrial ROS-mediated ferroptosis. Front. Pharmacol. 2023, 14, 1148814. [Google Scholar] [CrossRef] [PubMed]
  165. Satsu, H.; Awara, S.; Unno, T.; Shimizu, M. Suppressive effect of nobiletin and epicatechin gallate on fructose uptake in human intestinal epithelial Caco-2 cells. Biosci. Biotechnol. Biochem. 2018, 82, 636–646. [Google Scholar] [CrossRef] [PubMed]
  166. Du, Y.; Paglicawan, L.; Soomro, S.; Abunofal, O.; Baig, S.; Vanarsa, K.; Hicks, J.; Mohan, C. Epigallocatechin-3-Gallate Dampens Non-Alcoholic Fatty Liver by Modulating Liver Function, Lipid Profile and Macrophage Polarization. Nutrients 2021, 13, 599. [Google Scholar] [CrossRef] [PubMed]
  167. McMillan, E.A.; Ryu, M.J.; Diep, C.H.; Mendiratta, S.; Clemenceau, J.R.; Vaden, R.M.; Kim, J.H.; Motoyaji, T.; Covington, K.R.; Peyton, M.; et al. Chemistry-First Approach for Nomination of Personalized Treatment in Lung Cancer. Cell 2018, 173, 864–878.e29. [Google Scholar] [CrossRef] [PubMed]
  168. Campbell, R.K.; White, J.R., Jr.; Saulie, B.A. Metformin: A new oral biguanide. Clin. Ther. 1996, 18, 360–371; discussion 359. [Google Scholar] [CrossRef] [PubMed]
  169. Fan, H.; Pan, Q.; Xu, Y.; Yang, X. Exenatide improves type 2 diabetes concomitant with non-alcoholic fatty liver disease. Arq. Bras. Endocrinol. Metabol. 2013, 57, 702–708. [Google Scholar] [CrossRef] [PubMed]
  170. Feng, W.; Gao, C.; Bi, Y.; Wu, M.; Li, P.; Shen, S.; Chen, W.; Yin, T.; Zhu, D. Randomized trial comparing the effects of gliclazide, liraglutide, and metformin on diabetes with non-alcoholic fatty liver disease. J. Diabetes 2017, 9, 800–809. [Google Scholar] [CrossRef] [PubMed]
  171. Zhang, R.; Cheng, K.; Xu, S.; Li, S.; Zhou, Y.; Zhou, S.; Kong, R.; Li, L.; Li, J.; Feng, J.; et al. Metformin and Diammonium Glycyrrhizinate Enteric-Coated Capsule versus Metformin Alone versus Diammonium Glycyrrhizinate Enteric-Coated Capsule Alone in Patients with Nonalcoholic Fatty Liver Disease and Type 2 Diabetes Mellitus. Gastroenterol. Res. Pract. 2017, 2017, 8491742. [Google Scholar] [CrossRef] [PubMed]
  172. Yabiku, K.; Mutoh, A.; Miyagi, K.; Takasu, N. Effects of Oral Antidiabetic Drugs on Changes in the Liver-to-Spleen Ratio on Computed Tomography and Inflammatory Biomarkers in Patients with Type 2 Diabetes and Nonalcoholic Fatty Liver Disease. Clin. Ther. 2017, 39, 558–566. [Google Scholar] [CrossRef] [PubMed]
  173. Tian, F.; Zheng, Z.; Zhang, D.; He, S.; Shen, J. Efficacy of liraglutide in treating type 2 diabetes mellitus complicated with non-alcoholic fatty liver disease. Biosci. Rep. 2018, 38, BSR20181304. [Google Scholar] [CrossRef] [PubMed]
  174. Zsori, G.; Illes, D.; Ivany, E.; Kosar, K.; Holzinger, G.; Tajti, M.; Palinkas, E.; Szabovik, G.; Nagy, A.; Palko, A.; et al. In New-Onset Diabetes Mellitus, Metformin Reduces Fat Accumulation in the Liver, But Not in the Pancreas or Pericardium. Metab. Syndr. Relat. Disord. 2019, 17, 289–295. [Google Scholar] [CrossRef] [PubMed]
  175. Spruss, A.; Kanuri, G.; Stahl, C.; Bischoff, S.C.; Bergheim, I. Metformin protects against the development of fructose-induced steatosis in mice: Role of the intestinal barrier function. Lab. Investig. 2012, 92, 1020–1032. [Google Scholar] [CrossRef] [PubMed]
  176. Foretz, M.; Guigas, B.; Bertrand, L.; Pollak, M.; Viollet, B. Metformin: From mechanisms of action to therapies. Cell Metab. 2014, 20, 953–966. [Google Scholar] [CrossRef] [PubMed]
  177. Zhu, H.; Jia, Z.; Li, Y.R.; Danelisen, I. Molecular mechanisms of action of metformin: Latest advances and therapeutic implications. Clin. Exp. Med. 2023, 23, 2941–2951. [Google Scholar] [CrossRef] [PubMed]
  178. Karise, I.; Ornellas, F.; Barbosa-da-Silva, S.; Matsuura, C.; Del Sol, M.; Aguila, M.B.; Mandarim-de-Lacerda, C.A. Liver and Metformin: Lessons of a fructose diet in mice. Biochim. Open 2017, 4, 19–30. [Google Scholar] [CrossRef] [PubMed]
  179. Nassif, R.M.; Chalhoub, E.; Chedid, P.; Hurtado-Nedelec, M.; Raya, E.; Dang, P.M.; Marie, J.C.; El-Benna, J. Metformin Inhibits ROS Production by Human M2 Macrophages via the Activation of AMPK. Biomedicines 2022, 10, 319. [Google Scholar] [CrossRef] [PubMed]
  180. Jing, Y.; Wu, F.; Li, D.; Yang, L.; Li, Q.; Li, R. Metformin improves obesity-associated inflammation by altering macrophages polarization. Mol. Cell Endocrinol. 2018, 461, 256–264. [Google Scholar] [CrossRef] [PubMed]
  181. Qing, L.; Fu, J.; Wu, P.; Zhou, Z.; Yu, F.; Tang, J. Metformin induces the M2 macrophage polarization to accelerate the wound healing via regulating AMPK/mTOR/NLRP3 inflammasome singling pathway. Am. J. Transl. Res. 2019, 11, 655–668. [Google Scholar] [PubMed]
  182. Meloni, A.R.; DeYoung, M.B.; Lowe, C.; Parkes, D.G. GLP-1 receptor activated insulin secretion from pancreatic beta-cells: Mechanism and glucose dependence. Diabetes Obes. Metab. 2013, 15, 15–27. [Google Scholar] [CrossRef] [PubMed]
  183. Zhao, X.; Wang, M.; Wen, Z.; Lu, Z.; Cui, L.; Fu, C.; Xue, H.; Liu, Y.; Zhang, Y. GLP-1 Receptor Agonists: Beyond Their Pancreatic Effects. Front. Endocrinol 2021, 12, 721135. [Google Scholar] [CrossRef] [PubMed]
  184. Galderisi, A.; Giannini, C.; Van Name, M.; Caprio, S. Fructose Consumption Contributes to Hyperinsulinemia in Adolescents with Obesity Through a GLP-1-Mediated Mechanism. J. Clin. Endocrinol. Metab. 2019, 104, 3481–3490. [Google Scholar] [CrossRef] [PubMed]
  185. Yunker, A.G.; Luo, S.; Jones, S.; Dorton, H.M.; Alves, J.M.; Angelo, B.; DeFendis, A.; Pickering, T.A.; Monterosso, J.R.; Page, K.A. Appetite-Regulating Hormones Are Reduced After Oral Sucrose vs Glucose: Influence of Obesity, Insulin Resistance, and Sex. J. Clin. Endocrinol. Metab. 2021, 106, 654–664. [Google Scholar] [CrossRef] [PubMed]
  186. Arakawa, M.; Mita, T.; Azuma, K.; Ebato, C.; Goto, H.; Nomiyama, T.; Fujitani, Y.; Hirose, T.; Kawamori, R.; Watada, H. Inhibition of monocyte adhesion to endothelial cells and attenuation of atherosclerotic lesion by a glucagon-like peptide-1 receptor agonist, exendin-4. Diabetes 2010, 59, 1030–1037. [Google Scholar] [CrossRef] [PubMed]
  187. Li, Q.X.; Gao, H.; Guo, Y.X.; Wang, B.Y.; Hua, R.X.; Gao, L.; Shang, H.W.; Lu, X.; Xu, J.D. GLP-1 and Underlying Beneficial Actions in Alzheimer’s Disease, Hypertension, and NASH. Front. Endocrinol. 2021, 12, 721198. [Google Scholar] [CrossRef] [PubMed]
  188. Burmeister, M.A.; Brown, J.D.; Ayala, J.E.; Stoffers, D.A.; Sandoval, D.A.; Seeley, R.J.; Ayala, J.E. The glucagon-like peptide-1 receptor in the ventromedial hypothalamus reduces short-term food intake in male mice by regulating nutrient sensor activity. Am. J. Physiol. Endocrinol. Metab. 2017, 313, E651–E662. [Google Scholar] [CrossRef] [PubMed]
  189. van Bloemendaal, L.; RG, I.J.; Ten Kulve, J.S.; Barkhof, F.; Konrad, R.J.; Drent, M.L.; Veltman, D.J.; Diamant, M. GLP-1 receptor activation modulates appetite- and reward-related brain areas in humans. Diabetes 2014, 63, 4186–4196. [Google Scholar] [CrossRef] [PubMed]
  190. Burmeister, M.A.; Ayala, J.; Drucker, D.J.; Ayala, J.E. Central glucagon-like peptide 1 receptor-induced anorexia requires glucose metabolism-mediated suppression of AMPK and is impaired by central fructose. Am. J. Physiol. Endocrinol. Metab. 2013, 304, E677–E685. [Google Scholar] [CrossRef] [PubMed]
  191. Ghidewon, M.; Wald, H.S.; McKnight, A.D.; De Jonghe, B.C.; Breen, D.M.; Alhadeff, A.L.; Borner, T.; Grill, H.J. Growth differentiation factor 15 (GDF15) and semaglutide inhibit food intake and body weight through largely distinct, additive mechanisms. Diabetes Obes. Metab. 2022, 24, 1010–1020. [Google Scholar] [CrossRef] [PubMed]
  192. Bednarz, K.; Kowalczyk, K.; Cwynar, M.; Czapla, D.; Czarkowski, W.; Kmita, D.; Nowak, A.; Madej, P. The Role of Glp-1 Receptor Agonists in Insulin Resistance with Concomitant Obesity Treatment in Polycystic Ovary Syndrome. Int. J. Mol. Sci. 2022, 23, 4334. [Google Scholar] [CrossRef] [PubMed]
  193. Gao, Z.; Song, G.Y.; Ren, L.P.; Ma, H.J.; Ma, B.Q.; Chen, S.C. beta-catenin mediates the effect of GLP-1 receptor agonist on ameliorating hepatic steatosis induced by high fructose diet. Eur. J. Histochem. 2020, 64, 3160. [Google Scholar] [CrossRef] [PubMed]
  194. Li, S.L.; Wang, Z.M.; Xu, C.; Che, F.H.; Hu, X.F.; Cao, R.; Xie, Y.N.; Qiu, Y.; Shi, H.B.; Liu, B.; et al. Liraglutide Attenuates Hepatic Ischemia-Reperfusion Injury by Modulating Macrophage Polarization. Front. Immunol. 2022, 13, 869050. [Google Scholar] [CrossRef] [PubMed]
  195. Boland, M.L.; Laker, R.C.; Mather, K.; Nawrocki, A.; Oldham, S.; Boland, B.B.; Lewis, H.; Conway, J.; Naylor, J.; Guionaud, S.; et al. Resolution of NASH and hepatic fibrosis by the GLP-1R/GcgR dual-agonist Cotadutide via modulating mitochondrial function and lipogenesis. Nat. Metab. 2020, 2, 413–431. [Google Scholar] [CrossRef] [PubMed]
  196. Zhang, D.; Wang, S.; Ospina, E.; Shabandri, O.; Lank, D.; Akakpo, J.Y.; Zhao, Z.; Yang, M.; Wu, J.; Jaeschke, H.; et al. Fructose Protects Against Acetaminophen-Induced Hepatotoxicity Mainly by Activating the Carbohydrate-Response Element-Binding Protein alpha-Fibroblast Growth Factor 21 Axis in Mice. Hepatol. Commun. 2021, 5, 992–1008. [Google Scholar] [CrossRef] [PubMed]
  197. Kanerva, N.; Sandboge, S.; Kaartinen, N.E.; Mannisto, S.; Eriksson, J.G. Higher fructose intake is inversely associated with risk of nonalcoholic fatty liver disease in older Finnish adults. Am. J. Clin. Nutr. 2014, 100, 1133–1138. [Google Scholar] [CrossRef] [PubMed]
  198. Azevedo, V.Z.; Dall’Alba, V. Fructose intake is not associated to the risk of hepatic fibrosis in patients with Non-Alcoholic Fatty Liver Disease (NAFLD). Clin. Nutr. 2021, 40, 4275–4283. [Google Scholar] [CrossRef] [PubMed]
  199. Sharma, N.; Li, L.; Ecelbarger, C.M. Sex differences in renal and metabolic responses to a high-fructose diet in mice. Am. J. Physiol. Renal Physiol. 2015, 308, F400–F410. [Google Scholar] [CrossRef] [PubMed]
  200. Tran, C.; Jacot-Descombes, D.; Lecoultre, V.; Fielding, B.A.; Carrel, G.; Le, K.A.; Schneiter, P.; Bortolotti, M.; Frayn, K.N.; Tappy, L. Sex differences in lipid and glucose kinetics after ingestion of an acute oral fructose load. Br. J. Nutr. 2010, 104, 1139–1147. [Google Scholar] [CrossRef] [PubMed]
  201. Pektas, M.B.; Sadi, G.; Akar, F. Long-Term Dietary Fructose Causes Gender-Different Metabolic and Vascular Dysfunction in Rats: Modulatory Effects of Resveratrol. Cell Physiol. Biochem. 2015, 37, 1407–1420. [Google Scholar] [CrossRef] [PubMed]
  202. Hyer, M.M.; Dyer, S.K.; Kloster, A.; Adrees, A.; Taetzsch, T.; Feaster, J.; Valdez, G.; Neigh, G.N. Sex modifies the consequences of extended fructose consumption on liver health, motor function, and physiological damage in rats. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2019, 317, R903–R911. [Google Scholar] [CrossRef] [PubMed]
  203. Cui, P.; Hu, W.; Ma, T.; Hu, M.; Tong, X.; Zhang, F.; Shi, J.; Xu, X.; Li, X.; Shao, L.R.; et al. Long-term androgen excess induces insulin resistance and non-alcoholic fatty liver disease in PCOS-like rats. J. Steroid Biochem. Mol. Biol. 2021, 208, 105829. [Google Scholar] [CrossRef] [PubMed]
  204. Ohashi, T.; Kato, M.; Yamasaki, A.; Kuwano, A.; Suzuki, H.; Kohjima, M.; Ogawa, Y. Effects of high fructose intake on liver injury progression in high fat diet induced fatty liver disease in ovariectomized female mice. Food Chem. Toxicol. 2018, 118, 190–197. [Google Scholar] [CrossRef] [PubMed]
  205. Hsia, S.M.; Chiang, Y.F.; Chen, H.Y.; Ali, M.; Wang, P.S.; Wang, K.L. Effect of High-Fructose Diet-Induced Metabolic Syndrome on the Pituitary-Gonadal Axis in Male Rats. Biomedicines 2022, 10, 3009. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Direct and indirect effects of fructose in metabolic tissues. Fructose exposure can lead to recruitment of immune cells to metabolic tissues such as intestine, liver, and adipose. This recruitment leads to increased secretion of pro-inflammatory cytokines, aiding in cell-to-cell signaling. Fructose activates adipogenesis via upregulation of proteins involved in differentiation, monocyte and macrophage polarization, and cytokine secretion.
Figure 1. Direct and indirect effects of fructose in metabolic tissues. Fructose exposure can lead to recruitment of immune cells to metabolic tissues such as intestine, liver, and adipose. This recruitment leads to increased secretion of pro-inflammatory cytokines, aiding in cell-to-cell signaling. Fructose activates adipogenesis via upregulation of proteins involved in differentiation, monocyte and macrophage polarization, and cytokine secretion.
Biomolecules 14 00845 g001
Figure 2. Fructose metabolism and immune cells. Bone marrow-derived immune cells, such as macrophages, B cells, and dendritic cells, display a pro-inflammatory phenotype when exposed to fructose; however, the mechanism for this response is unique for each cell type. Macrophages rely on oxidative phosphorylation and TCA cycle passes for energy and cytokine production. Interestingly, B cells and dendritic cells utilize intermediates of glycolysis conversion to keep up with the energy demands of the cell. Secretion of the cytokines IL-6, TNFα, IL-18, and IL-1β aid in cell-to-cell communication as well as priming an immune response. This signaling can lead to recruitment of immune cells within metabolic tissues such as liver, intestine, and adipose, thus increasing overall inflammation. Cytokine secretion can directly activate T cells through TNFα inducing secretion of IFNγ. Kupffer cells depend on glycolysis and the pentose phosphate pathway for regulation of genes associated with wound healing and anti-inflammatory response such as MMP12, GPNMB, RSAD2, TNFα, and IL1RN.
Figure 2. Fructose metabolism and immune cells. Bone marrow-derived immune cells, such as macrophages, B cells, and dendritic cells, display a pro-inflammatory phenotype when exposed to fructose; however, the mechanism for this response is unique for each cell type. Macrophages rely on oxidative phosphorylation and TCA cycle passes for energy and cytokine production. Interestingly, B cells and dendritic cells utilize intermediates of glycolysis conversion to keep up with the energy demands of the cell. Secretion of the cytokines IL-6, TNFα, IL-18, and IL-1β aid in cell-to-cell communication as well as priming an immune response. This signaling can lead to recruitment of immune cells within metabolic tissues such as liver, intestine, and adipose, thus increasing overall inflammation. Cytokine secretion can directly activate T cells through TNFα inducing secretion of IFNγ. Kupffer cells depend on glycolysis and the pentose phosphate pathway for regulation of genes associated with wound healing and anti-inflammatory response such as MMP12, GPNMB, RSAD2, TNFα, and IL1RN.
Biomolecules 14 00845 g002
Figure 3. The hepatic effects of nutritional supplements and pharmacological interventions in fructose-induced NAFLD. (A) The hepatic effects of nutritional supplements in fructose-induced NAFLD. Antioxidant supplements, such as vitamin E and resveratrol, are shown to alleviate fructose-induced hepatic steatosis and liver damage through the attenuation of lipid peroxidation and inflammation. L-carnitine also facilitates antioxidant properties through increased expression of superoxide dismutase 2 (SOD2), which catalyzes the dismutation of ROS, reducing oxidative stress. L-carnitine also is suggested to increase levels of liver enzymes and insulin sensitivity in hepatocytes and alleviate hepatic lipid accumulation. Additionally, the flavonoid fisetin is suggested to alleviate fructose-induced liver damage and inflammation through suppressing NFkB. Fisetin also activates Nrf2, a transcription factor that binds the antioxidant response element (ARE), a DNA sequence found in the promoter regions of several genes encoding antioxidant enzymes, increasing the expression in the liver. (B) The hepatic effects of pharmacological interventions in fructose-induced NAFLD. Several pharmacological interventions target fructose-induced hepatic DNL. GLUT5 inhibitors alleviate fructose-induced DNL through decreasing fructose uptake and metabolism, resulting in increased insulin sensitivity. Likewise, ketohexokinase (KHK) and hexokinase (HK) inhibitors decrease fructose-induced DNL by reducing hepatic fructose metabolism, improving insulin sensitivity. Similarly, GLP-1 receptor agonists (GLP-1-RAs) function to decrease hepatic DNL by increasing gene expression of the transcription factor β-catenin, which translocates to the nucleus and mediates the downregulation of acetyl-CoA carboxylase (ACC), fatty acid synthase (FAS), and stearoyl-CoA desaturase-1 (SCD-1). Apart from its effects in increasing insulin sensitivity, metformin plays a protective role against NAFLD by reducing hepatic expression of TNFα and inhibiting the overactivation of hepatic Kupffer cells, reducing inflammation.
Figure 3. The hepatic effects of nutritional supplements and pharmacological interventions in fructose-induced NAFLD. (A) The hepatic effects of nutritional supplements in fructose-induced NAFLD. Antioxidant supplements, such as vitamin E and resveratrol, are shown to alleviate fructose-induced hepatic steatosis and liver damage through the attenuation of lipid peroxidation and inflammation. L-carnitine also facilitates antioxidant properties through increased expression of superoxide dismutase 2 (SOD2), which catalyzes the dismutation of ROS, reducing oxidative stress. L-carnitine also is suggested to increase levels of liver enzymes and insulin sensitivity in hepatocytes and alleviate hepatic lipid accumulation. Additionally, the flavonoid fisetin is suggested to alleviate fructose-induced liver damage and inflammation through suppressing NFkB. Fisetin also activates Nrf2, a transcription factor that binds the antioxidant response element (ARE), a DNA sequence found in the promoter regions of several genes encoding antioxidant enzymes, increasing the expression in the liver. (B) The hepatic effects of pharmacological interventions in fructose-induced NAFLD. Several pharmacological interventions target fructose-induced hepatic DNL. GLUT5 inhibitors alleviate fructose-induced DNL through decreasing fructose uptake and metabolism, resulting in increased insulin sensitivity. Likewise, ketohexokinase (KHK) and hexokinase (HK) inhibitors decrease fructose-induced DNL by reducing hepatic fructose metabolism, improving insulin sensitivity. Similarly, GLP-1 receptor agonists (GLP-1-RAs) function to decrease hepatic DNL by increasing gene expression of the transcription factor β-catenin, which translocates to the nucleus and mediates the downregulation of acetyl-CoA carboxylase (ACC), fatty acid synthase (FAS), and stearoyl-CoA desaturase-1 (SCD-1). Apart from its effects in increasing insulin sensitivity, metformin plays a protective role against NAFLD by reducing hepatic expression of TNFα and inhibiting the overactivation of hepatic Kupffer cells, reducing inflammation.
Biomolecules 14 00845 g003
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

Lodge, M.; Dykes, R.; Kennedy, A. Regulation of Fructose Metabolism in Nonalcoholic Fatty Liver Disease. Biomolecules 2024, 14, 845. https://doi.org/10.3390/biom14070845

AMA Style

Lodge M, Dykes R, Kennedy A. Regulation of Fructose Metabolism in Nonalcoholic Fatty Liver Disease. Biomolecules. 2024; 14(7):845. https://doi.org/10.3390/biom14070845

Chicago/Turabian Style

Lodge, Mareca, Rachel Dykes, and Arion Kennedy. 2024. "Regulation of Fructose Metabolism in Nonalcoholic Fatty Liver Disease" Biomolecules 14, no. 7: 845. https://doi.org/10.3390/biom14070845

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