Review Open Access
Copyright ©The Author(s) 2016. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Gastroenterol. Feb 28, 2016; 22(8): 2483-2493
Published online Feb 28, 2016. doi: 10.3748/wjg.v22.i8.2483
Glycosyltransferases and non-alcoholic fatty liver disease
Yu-Tao Zhan, Department of Gastroenterology, Beijing Tongren Hospital, Capital Medical University, Beijing 100730, China
Hai-Ying Su, Department of Obstetrics and Gynecology, Shandong Jining First People’s Hospital, Jining 272011, Shandong Province, China
Wei An, Department of Cell Biology, Municipal Laboratory for Liver Protection and Regulation of Regeneration, Capital Medical University, Beijing 100069, China
Author contributions: An W conceived the review and revised the paper; Zhan YT and Su HY wrote the paper; Zhan YT and Su HY contributed equally to this work; Zhan YT and An w are equally Corresponding author.
Supported by The National Natural Science Foundation of China No. 81041017; Beijing Municipal Natural Science Foundation, No. 7112032; Beijing Municipal Laboratory for Liver Protection and Regulation of Regeneration, and Scientific Research Common Program of Beijing Municipal Commission of Education No. KM201510025017.
Conflict-of-interest statement: None of the authors has any conflicts-of interest to disclose.
Open-Access: This article is an open-access article which was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/
Correspondence to: Yu-Tao Zhan, MD, Department of Gastroenterology, Beijing Tongren Hospital, Capital Medical University, No. 1 Dongjiaominxiang, Dongcheng District, Beijing 100730, China. yutaozhan@263.net
Telephone: +86-10-58268463 Fax: +86-10-58268473
Received: May 11, 2015
Peer-review started: May 12, 2015
First decision: September 11, 2015
Revised: October 22, 2015
Accepted: November 19, 2015
Article in press: November 19, 2015
Published online: February 28, 2016
Processing time: 290 Days and 3 Hours

Abstract

Non-alcoholic fatty liver disease (NAFLD) is the most common form of chronic liver disease and its incidence is increasing worldwide. However, the underlying mechanisms leading to the development of NAFLD are still not fully understood. Glycosyltransferases (GTs) are a diverse class of enzymes involved in catalyzing the transfer of one or multiple sugar residues to a wide range of acceptor molecules. GTs mediate a wide range of functions from structure and storage to signaling, and play a key role in many fundamental biological processes. Therefore, it is anticipated that GTs have a role in the pathogenesis of NAFLD. In this article, we present an overview of the basic information on NAFLD, particularly GTs and glycosylation modification of certain molecules and their association with NAFLD pathogenesis. In addition, the effects and mechanisms of some GTs in the development of NAFLD are summarized.

Key Words: Non-alcoholic fatty liver disease; Pathogenesis; Glycosyltransferases; Glycosylation

Core tip: Nonalcoholic fatty liver disease (NAFLD) is characterized by a very complicated process which is regulated by a number of protein molecules. Glycosylation, one of the most common post-translational modifications of proteins in eukaryotic cells, has been suggested to play an important role in the pathogenesis of NAFLD. As glycosylation is mainly mediated through glycosyltransferases (GTs), it seems reasonable to speculate that the GTs play an important role in the pathogenesis of NAFLD.



OVERVIEW OF NON-ALCOHOLIC FATTY LIVER DISEASE

Fatty liver is characterized by the excess accumulation of lipids including triglycerides (TGs) and cholesterol. In general, accumulation of lipids up to 5% of the liver weight results in the diagnosis of fatty liver disease or hepatic steatosis. In addition, if hepatic steatosis occurs in patients who do not consume alcohol on a daily basis, it is referred to as non-alcoholic fatty liver disease (NAFLD)[1]. Usually, NAFLD is classified as “primary” and “secondary”, depending on the underlying etiology. “Primary” NAFLD is most common, and is often associated with insulin-resistance and metabolic syndrome. Obesity, diabetes and dyslipidemia are the most common risk factors for NAFLD. The term “secondary” NAFLD is currently discouraged and the preferred nomenclature is based on the known causative factors and the resultant pathologies e.g., viral infections, autoimmune diseases, endocrine-metabolic disorders, total parenteral nutrition and drug-induced fatty liver. Therefore, the term “NAFLD” generally refers to “primary” NAFLD. The obesity and type 2 diabetes pandemic, and the improved management of chronic viral hepatitis have resulted in NAFLD being a leading cause of chronic liver disease[2-6]. It is estimated that 20% to 30% of adults in the United States and Western Europe have excess fat accumulation in the liver[7]. The prevalence of NAFLD in the general population across Asia varies from 5% to 40%[8]. A recent meta-analysis showed that the prevalence of NAFLD in China is approximately 20%[9]. With the increase in obesity and diabetes, the incidence of NAFLD is expected to rise worldwide. The prevalence of NAFLD in the United States is expected to increase by 50% in 2030[10]. Based on current information, NAFLD encompasses a spectrum of diseases ranging from simple steatosis, to inflammatory steatohepatitis (NASH) with increasing levels of fibrosis and ultimately cirrhosis[11,12]. NAFLD was initially believed to be a benign illness as its progression is quite slow and rarely results in a poor outcome. However, results from clinical studies have confirmed that NAFLD, if not properly controlled, may cause liver-related morbidity and mortality[13]. Cirrhosis is a severe disease leading to death[14], and patients with NAFLD not only progress to cirrhosis[15], but are also susceptible to cardiovascular disease/death, type 2 diabetes mellitus and diabetic nephropathy[16,17], which are dependent on the severity of liver injury[18,19]. Moreover, NAFLD has also been shown to increase the risk of colorectal cancer[20] and thus, may result in an increased overall mortality[21] (Figure 1). Although weight loss is believed to be effective in NAFLD treatment, adherence to lifestyle interventions is a limitation. Various studies have shown that of the patients scheduled for NAFLD treatment, only 15% achieved weight loss, but regained weight with time[22]. Although there is no approved drug therapy for NAFLD, many approaches appear to be beneficial, such as the use of insulin sensitizers, antioxidants and anti-inflammatory agents, and these seem to have promising effects in some patients[23,24].

Figure 1
Figure 1 Manifestations of non-alcoholic fatty liver disease. NAFLD: Non-alcoholic fatty liver disease.
NAFLD PATHOGENESIS

The pathogenesis of NAFLD has not been completely elucidated[25,26]. Based on available information, various researchers have proposed different hypotheses over time. The major hypotheses are as follows: (1) In 1998, Day et al[27] first proposed the “two-hit” hypothesis for the pathogenesis of NAFLD. The first hit represents the accumulation of lipids in hepatocytes and the induction of insulin resistance which is the key pathogenic factor for the development of hepatic steatosis. The second hit leads to hepatocyte injury, inflammation and fibrosis. Factors initiating the second hit are oxidative stress and subsequent lipid peroxidation, proinflammatory cytokines, adipokines and mitochondrial dysfunction; (2) In 2008, Jou et al[28] suggested the “three-hit” hypothesis. The first hit also involves the accumulation of lipids by the mechanisms described above. The second hit involves the initiation of an inflammatory response and cell death, while the third hit results in defective repair and the induction of a regenerative response by the proliferation and differentiation of hepatocyte progenitors; (3) In 2009, Polyzos et al[29] provided the “multi-hit process” hypothesis. The initial hit leads to the development of simple steatosis which subsequently renders hepatocytes susceptible to a variety of additional hits, eventually leading to NASH. These additional hits appear to be genetic or environmental perturbations leading to liver cell inflammation and necrosis with activation of the fibrogenic cascade. This results in the development of fibrosis or even cirrhosis in a minority of NAFLD patients. Insulin resistance (IR) and subsequent hyperinsulinemia are key pathogenetic factors in both simple steatosis and its subsequent progression to NASH; and (4) In 2010, Tilg et al[30] proposed the “multiple parallel hits” hypothesis for NAFLD and it has attracted wide attention from the research community. This hypothesis reflects more precisely the current knowledge of NASH[20,31]. According to this hypothesis, many parallel hits are derived from the gut and/or the adipose tissue that promote liver inflammation. Endoplasmic reticulum (ER) stress and its related signaling networks, adipocytokines/cytokines, and innate immunity are emerging as central pathways that regulate key features of NASH[32]. Although genetic factors play a minor role in the current obesity epidemic, they may offer explanations for a more progressive disease course in NAFLD[33]. Adipose tissue-derived factors include adipocytokines such as adiponectin and leptin, certain proinflammatory cytokines such as tumor necrosis factor α (TNF-α) or interleukin 6 (IL-6), and others such as the death receptor Fas, while gut-derived factors include endotoxin, microbiota, and various nutrients such as trans-fatty acids, fructose, and arylhydrocarbon receptor ligands. In addition, other proposed hypotheses are similar to the “multi-hit” hypothesis and four-step model[34,35].

MECHANISMS OF HEPATIC FAT ACCUMULATION

NAFLD is characterized by excess fat accumulation in the liver[36], which arises from an imbalance between fat acquisition and removal (Figure 2). TGs are composed of three fatty acids (FAs) coupled to a glycerol backbone via an ester bond. The fatty acids used for hepatic TGs formation are derived from three sources; (1) adipose tissue; (2) de novo lipogenesis (DNL); and (3) dietary sources[37]. Approximately 60% of liver FAs are derived from adipose tissue, 25% are from DNL, and 15% are from the diet[38]. FAs can be stored as lipid droplets within hepatocytes or secreted into the blood as very low-density lipoprotein (VLDL). However, they can also be channeled towards the β-oxidation pathway in mitochondria. Therefore, excess hepatic lipid accumulation can be caused by the following four different metabolic perturbations: (1) an increase in free fatty acid (FFA) uptake derived from the circulation due to increased lipolysis from adipose tissue and/or from the diet in the form of chylomicrons; (2) increased DNL; (3) reduced FA oxidation; and (4) reduced lipid export in the form of VLDL[39]. Rodent studies have shown that the mechanisms leading to excess accumulation of hepatic TGs are mainly associated with an increased supply of FFAs from peripheral adipose tissue to the liver and an enhanced de novo lipid synthesis via the lipogenic pathway[40]. Conversely, their disposal from the liver viaβ-oxidation and VLDL export are moderately affected[41]. Particularly in humans, obesity increases TNF-α production in adipocytes, facilitates adipocyte IR, and increases lipolysis rate. Thus, the circulating pool of FFAs is increased in obese individuals and thus accounts for the majority of the liver TGs in NAFLD. DNL refers to the synthesis of endogenous FAs in hepatocytes. During this process, glucose is converted to acetyl-CoA by glycolysis and the oxidation of pyruvate. Acetyl-CoA carboxylase then converts acetyl-CoA into malonyl-CoA and finally, FA synthase catalyzes the formation of palmitic acid from malonyl-CoA and acetyl-CoA. The rate of DNL is regulated primarily at the transcriptional level[42]. Several nuclear transcription factors are involved such as liver X receptors, sterol regulatory element-binding protein-1c (SREBP-1c), and carbohydrate-responsive element binding protein (ChREBP). SREBP-1c can regulate more than 32 genes involved in lipid biosynthesis and transport[43]. IR may promote DNL by stimulation of hyperinsulinemia to SREBP-1c[44]. Dietary fats taken up in the intestine are packaged into TG-rich chylomicrons and delivered to the systemic circulation. About 80% of the TG components in chylomicrons are unloaded in adipose and muscle tissues. The remaining 20% are transported to the liver through the hepatic artery[45]. As a result, the FAs derived from dietary fats account for the minority of circulating FFAs in NAFLD.

Figure 2
Figure 2 Mechanism of triglyceride accumulation in hepatocytes.
BASIC INFORMATION OF GLYCOSYLTRANSFERASES

Glycosyltransferases (GTs) are a diverse class of enzymes encompassing 1% to 2% of all sequenced genomes[46]. They catalyze the transfer of one or multiple sugar residues to a wide range of acceptor molecules such as lipids, proteins, hormones, secondary metabolites, and oligosaccharides[47,48], and mediate a wide range of functions from structure and storage to signaling[49]. Thus, they play a key role in many fundamental biological processes including cell signaling, cellular adhesion, carcinogenesis, and cell wall biosynthesis in human pathogens[50-52]. GTs are present in both prokaryotes and eukaryotes. In eukaryotes, the majority of GTs exist as membrane proteins of the Golgi apparatus. The newly synthesized GTs are transported from the ER to the Golgi via COPII-transport vesicles[53,54]. All the Golgi-localized enzymes share the common topology of type II membrane proteins, consisting of a short N-terminal cytoplasmic domain, a single transmembrane segment and a stem region of variable length followed by a large C-terminal catalytic domain[55,56]. The length and amino acid composition of catalytic domains are relatively well conserved and the variations in protein sizes are generally attributed to differences in the length of the stem region. In general, robust localization of Golgi enzymes relies on the contribution from each of these domains, although the transmembrane segment for a long time was considered to be the key determinant for GTs localization. The acceptor specificity may be regulated by the stem segment in vivo, although its role in enzyme activity is still unclear. The N-terminal domain is an important feature in acceptor binding. The significant variation in C-terminal β-strands and/or loops contributes to acceptor specificity and region specificity[57]. There are different classification systems for GTs: (1) GTs are primarily classified according to the type of sugar they transfer; (2) Based on sequence similarities of amino acids (CAZy database, http://www.cazy.org/), GTs are divided into 97 families. The vast majority of these sequences (more than 90%) are uncharacterized open-reading frames[58]; (3) X-ray structural studies have revealed that there are 105 GT structures in the Protein Data Bank, representing 36 of the 89 CAZy GT families[59], most of which adopt one of two predominant structural folds: GT-A and GT-B fold[60]. The GT-A fold consists of a single α/β/α-sandwich form that resembles a Rossmann fold. The central β-sheet is flanked by a smaller one, and the association of both creates the active site. A general feature of all the enzymes with GT-A fold is the presence of a common motif, such as the DXD motif[61,62]. The DXD motif anchors the pyrophosphate moiety of the sugar-nucleotide donor via a divalent cation, such that the location of the sugar donor on the fold is conserved. The GT-B fold consists of two separate Rossmann domains with a connecting linker region and a catalytic site located between the domains. There is an excellent structural conservation between protein members of the GT-B family, particularly in the C-terminal domain which corresponds to the nucleotide-binding domain. A third family has recently emerged which comprises a bacterial sialyltransferase belonging to the GT42 family[63]. This protein displays a fold similar to the GT-A, but with some differences, thus it can be considered a new fold; and (4) Based on the outcome of the reaction, GTs are classified into two types, either inverting or retaining. Inverting GTs most likely follow a single displacement mechanism, wherein the acceptor induces a nucleophilic attack at carbon C-1 of the sugar donor somewhat analogous to the mechanism of inverting glycosidases. Retaining GTs do not operate via a two-step mechanism involving the formation of a glycosyl-enzyme intermediate analogous to glycosidases. Instead, an internal return SNi-like mechanism has been proposed, in which the departure of the leaving group and nucleophilic attack occur in a concerted, but asynchronous manner on the same face of the glycoside.

GLYCOSYLATION OF MOLECULES INVOLVED IN THE PATHOGENESIS OF NAFLD

Glycosylation is one of the most common post-translational modifications of proteins in eukaryotic cells[64,65]. Recent studies have indicated that numerous protein molecules undergoing glycosylation are involved in the pathogenesis of NAFLD.

Apolipoprotein B

The major function of VLDL is to transport endogenous TGs from the hepatocytes to the extrahepatic tissue. Apolipoprotein B (ApoB)-100, a large secretory glycoprotein with 4536 amino acid residues, is an important component of VLDL[66]. It has 19 potential glycosylation sites (Asn-X-Ser/Thr), and 16 of them have been reported to be glycosylated[67,68]. Ihara et al[69] reported that N-acetylglucosaminyltransferase III (GnT-III) is linked to the glycosylation of ApoB-100 in hepatocytes. Chylomicrons transport TGs from the gut to the periphery via intestinal lymph and the systemic circulation. ApoB-48 is a truncated segment of ApoB-100 and is homologous to the initial 2151 amino acids of ApoB-100. Studies have also confirmed that ApoB-48 can be modified by glycosylation[70,71].

Fatty acid translocase

Fatty acid uptake into the liver contributes to the steady balance of hepatic TGs in the liver, as well as the pathogenesis of NAFLD. The cellular capacity for fatty acid uptake depends on the numbers and activities of transporter proteins on the sinusoidal plasma membrane of the hepatocytes. Fatty acid translocase is a transporter protein, and is heavily modified post-translationally by N-linked glycosylation. The 10 putative glycosylation sites located in the large extracellular loop of the protein have been identified[72].

ChREBP

ChREBP is involved in the transcriptional activation of genes encoding the aforementioned rate-limiting enzymes in lipogenesis, and has been associated with increased DNL in NAFLD[73]. Guinez et al[74] reported that ChREBP interacts with O-GlcNAcylation transferase and is subjected to O-GlcNAcylation in liver cells which in turn stabilizes it and enhances its transcriptional activity toward its target glycolytic and lipogenic genes when combined with an active glucose flux in vivo.

Fas

Hepatocyte apoptosis is the most common and well-characterized cell death pathway. Hepatic apoptosis is also confirmed to be a pathologic hallmark of NASH[75]. Alkhouri et al[76] reported that there was an increased sensitivity to Fas-mediated hepatocyte apoptosis in a dietary model of NAFLD, when mice were fed a high-fat diet. In addition, liver tissue samples from NASH patients displayed high expression of Fas protein, suggesting that it plays a role in the development of NASH. Fas is a glycosylated protein, and undergoes glycosylation in its extracellular domain during NASH[77,78].

Adiponectin

Adiponectin is an insulin-sensitizing adipocytokine that has multiple beneficial effects in obesity-related NAFLD[79]. The collagenous region of adiponectin, produced in vitro, contains four conserved lysines that are both hydroxylated and glycosylated with a glucosylgalactosyl moiety[80]. In addition, bovine and mouse plasma adiponectin contains sialic acid, possibly on O-linked glycans[81].

EFFECT AND MECHANISM OF GTs ON THE DEVELOPMENT OF NAFLD

In recent years, a number of studies have demonstrated the role of some GTs in the development of NAFLD and their different mechanisms of action.

GnT-III

GnT-III is a key enzyme in N-glycan biosynthesis, encoded by the Mgat3 gene[82], and is a mammalian Golgi-resident GT. It catalyzes the attachment of the bisecting GlcNAc residue to β-1, 4 mannose in the core structure of N-linked oligosaccharides[83]. Bisected N-glycans are involved in physiological and pathological processes through the functional regulation of their carrier proteins[84,85]. Human GnT-II contains 531 amino acids and possesses a domain structure identical to GTs[86]. The structure includes a short N-terminal cytoplasmic tail, a transmembrane region of 16-20 amino acids (as predicted by hydropathy plots), a stem region (or neck region), and a long C-terminal catalytic domain[87]. Ihara et al[69] found that the livers of GnT-III transgenic mice contained abundant lipid droplets accompanied by ballooning degeneration. Although the levels of immunoreactive ApoB were increased, the ApoB-100 was specifically decreased to undetectable levels in the serum of these transgenic mice. These results strongly suggest that aberrant glycosylation of ApoB activated by GnT-III inhibits the ApoB assembly itself and further blocks the synthesis and secretion of VLDL, which in turn leads to an accumulation of TGs within the liver.

T-synthase

T-synthase is the key β3-galactosyltransferase essential for the biosynthesis of core 1 O-glycans (Galβ1-3GalNAcα1-Ser/Thr) in the glycoproteins of animal cells[88]. It was initially purified from rat liver and subsequently cloned into the cDNA, and the genes for T-synthase were successfully identified from Caenorhabditis elegans, mouse, rat and human. The cDNA for T-synthase in mammals encodes a 363-amino acid transmembrane protein with type II topology[89]. A decrease in the expression of T-synthase alters O-glycan elongation and results in the production of abnormal and truncated carbohydrate structures, eventually leading to exposure of the Tn antigen[90]. This has been shown to be associated with several human diseases, including cancer, Tn syndrome and IgA nephropathy[91]. A recent study showed that T-synthase knockout in endothelial and hematopoietic cells (EHC T-syn-/-) of pups, resulted in the development of fatty liver disease in mice. Fu et al[92] reported that immediately after the pups began nursing on milk, the liver of postnatal 1 wk EHC T-syn-/- mice displayed an abnormal accumulation of vacuoles containing TGs, resembling microvesicular steatosis in human steatohepatitis. At postnatal 7 wk, the livers of EHC T-syn-/- mice, had extensive steatosis, inflammatory infiltrates, and hepatocyte ballooning. EHC T-syn-/- mice that survived beyond neonatal development displayed cirrhosis. EHC T-syn-/- adult mice were not obese. The lymphatic system is essential for the transport of immune cells, interstitial fluids, and dietary lipids[93]. Dietary lipids are transported in the form of chylomicrons from the small intestine to the systemic circulation via the intestinal lymphatic vessels and thoracic duct[94]. In EHC T-syn-/- mice, due to aberrant intestinal vein and lymphatic connections, chylomicrons are directly transported to the liver via the portal vein system, which causes fatty liver disease. Endothelial O-glycans control the separation of blood and lymphatic vessels during embryonic and postnatal development by regulating podoplanin expression. The abnormal O-glycosylation of endothelial podoplanin is sufficient for the formation of hybrid vessels and blood/lymphatic vessel misconnections. Therefore, the impairment of podoplanin expression/function by specific deletion of T-synthase may contribute to the aberrant connections between intestinal blood and lymphatic vessels.

α1, 6-fucosyltransferase

α1, 6-fucosyltransferase (FUT8) catalyzes the transfer of a fucosyl residue from guanine nucleotide diphosphate-β-l-fucose to the innermost GlcNAc of an asparagine-linked oligosaccharide[95]. It plays an important role in the tumorigenesis of non-small cell lung cancer and colon carcinoma[96,97]. Human Fut8 gene is located on chromosome 14q23.3, and consists of at least nine exons spanning more than a 50 kb genomic region, and the coding sequence is divided into eight exons[98,99]. FUT8 is a typical type II membrane protein localized in the Golgi apparatus[100]. It consists of 575 amino acids, and contains a catalytic domain, an N-terminal coiled-coil domain and a C-terminal SH3 domain. The catalytic domain was structurally classified as a member of the GT-B group of GTs. Wang et al[101] reported that lipid droplets in hepatocytes were significantly increased in the liver of FUT8 transgenic mice, and these lipid droplets were apparently localized within the lysosomes. Furthermore, the study showed that liver lysosomal acid lipase (LAL) activity was significantly lower in these transgenic mice compared to wild-type mice, and the level of fucosylated LAL was greater in transgenic mice. These results suggested that aberrant fucosylation of LAL causes an accumulation of inactive LAL in the lysosomes, and results in steatosis in the lysosomes of the liver in the case of FUT8 transgenic mice.

Glycosyltransferase 8 domain containing 2

As a member of the glycosyltransferase 8 family, the human glycosyltransferase 8 domain containing 2 (Glt8D2) is a 349 amino acid single-pass type II membrane protein encoded by a gene located on chromosome 12q23.3. The first six amino acid residues extend to the cytoplasm, residues 7-24 constitute the transmembrane domain and residues 25-349 are in the luminal compartments[102]. Moylan et al[103] reported that the GLT8D2 gene is up-regulated in patients with severe NAFLD. Recently, we have cloned the GLT8D2 gene and found that GLT8D2 expression increased in fatty liver compared with normal liver in rats. Our in vitro study found that GLT8D2 expression increased in steatosis HepG2 cells compared with normal cells. In addition, further study showed that plasmid transfection of GLT8D2 increased the TG content, up-regulated ApoB-100 protein, but down-regulated microsomal triglyceride transfer protein (MTP) in HepG2 cells. MTP has both apoB 100 binding and lipid transfer domains[104], and is an essential factor for VLDL assembly and secretion. As a result, we speculate that the inhibition of MTP expression by GLT8D2 may be the major mechanism resulting in accumulation of TG in HepG2 cells.

UDP-glucuronosyltransferases

UDP-glucuronosyltransferases (UGTs) are glycoproteins localized in the ER which catalyze the conjugation of a wide variety of lipophilic aglycon substrates with glucuronic acid using UDP-glucuronic acid as the sugar donor[105]. The human UGTs are membrane proteins with approximately 530 amino acids, of which the first 25 residues of the signal sequence are removed after the transfer of newly synthesized polypeptides to the ER. A single transmembrane helix is predicted close to the C terminus of the protein, and the membrane topology is such that the bulk of the protein is on the luminal side of the ER membrane[106]. The mammalian UGT gene superfamily currently has 117 members. On the basis of amino acid sequence similarity, the UGT superfamily is divided into four families, UGT1, UGT2, UGT3 and UGT8[107], and further subdivided into different subfamilies, respectively. It is well-known that UGTs are highly expressed in the liver, and induced by microsomal enzyme treatment through nuclear receptor- and transcription factor-dependent mechanisms[108,109]. In recent years, it was found that the UGT expression is abnormal in the liver of NAFLD subjects. Xu et al[110] reported that the mRNA expression of some of the UGT isoforms was increased in steatotic liver of ob/ob mice, and this was accompanied by increased mRNA expression of the arylhydrocarbon receptor, constitutive androstane receptor, peroxisome proliferator-activated receptor-α, pregnane X receptor, nuclear factor-like 2, and peroxisome proliferator-activated receptor-γ coactivator-1α. Zhang et al[111] also confirmed that fatty liver in rats on a high-fat diet showed increased mRNA and protein expression of UGT, and it was further enhanced by the addition of valproic acid. The induction of UGTs was accompanied by the increased expression of constitutive androstane receptor and peroxisome proliferator-activated receptor α. However, Hardwick et al[112] found that the expression of different UGT isoforms in the liver appears to be differentially regulated in human NASH. Hence, the role of UGT expression in NAFLD remains unclear. In addition, UGT also has a role in glucuronidation, which is a major detoxification pathway for exogenous compounds and is becoming increasingly important for metabolizing approximately 40%-70% of drugs[113]. Numerous xenobiotics, including acetaminophen, morphine, propofol, chloramphenicol, and nonsteroidal anti-inflammatory drugs, as well as environmental compounds, are glucuronidated by UGT[114]. Thus, it is possible that a UGT abnormality may exacerbate the side effects of the above drugs in NAFLD patients. Therefore, UGT abnormalities may also play an important role in the pathogenesis of NAFLD.

CONCLUSION

NAFLD is the most highly prevalent chronic liver disease, and its detailed mechanism remains unclear. TGs, which play an important role in many fundamental biological processes, are also confirmed to affect the development of NAFLD and play an important role in its pathogenesis. In addition, some molecules related to the pathogenesis of NAFLD are glycosylated and are modified by some GTs. However, many questions related to protein glycosylation and its role in the development of NAFLD have yet to be clarified. The precise mechanism of hepatic steatotic injury involving protein glycosylation and consequent NAFLD require further detailed investigation.

Footnotes

P- Reviewer: Bordas JM, Meshikhes AWN, Sertoglu E S- Editor: Qi Y L- Editor: Webster JR E- Editor: Zhang DN

References
1.  Obika M, Noguchi H. Diagnosis and evaluation of nonalcoholic fatty liver disease. Exp Diabetes Res. 2012;2012:145754.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 126]  [Cited by in F6Publishing: 154]  [Article Influence: 11.8]  [Reference Citation Analysis (0)]
2.  Milić S, Lulić D, Štimac D. Non-alcoholic fatty liver disease and obesity: biochemical, metabolic and clinical presentations. World J Gastroenterol. 2014;20:9330-9337.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 207]  [Reference Citation Analysis (1)]
3.  Machado MV, Cortez-Pinto H. Non-alcoholic fatty liver disease: what the clinician needs to know. World J Gastroenterol. 2014;20:12956-12980.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 140]  [Cited by in F6Publishing: 138]  [Article Influence: 13.8]  [Reference Citation Analysis (4)]
4.  Zhu C, Xie P, Zhao F, Zhang L, An W, Zhan Y. Mechanism of the promotion of steatotic HepG2 cell apoptosis by cholesterol. Int J Clin Exp Pathol. 2014;7:6807-6813.  [PubMed]  [DOI]  [Cited in This Article: ]
5.  Zhan YT, Weng J, Li L, Xu Q, Song X, Guo XX. Protective effect of probucol on liver injury induced by carbon tetrachloride in rats. Hepatol Int. 2011;5:899-905.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6]  [Cited by in F6Publishing: 6]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
6.  Zhan YT, An W. Roles of liver innate immune cells in nonalcoholic fatty liver disease. World J Gastroenterol. 2010;16:4652-4660.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 68]  [Cited by in F6Publishing: 73]  [Article Influence: 5.2]  [Reference Citation Analysis (0)]
7.  Athyros VG, Tziomalos K, Katsiki N, Doumas M, Karagiannis A, Mikhailidis DP. Cardiovascular risk across the histological spectrum and the clinical manifestations of non-alcoholic fatty liver disease: An update. World J Gastroenterol. 2015;21:6820-6834.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 102]  [Cited by in F6Publishing: 99]  [Article Influence: 11.0]  [Reference Citation Analysis (0)]
8.  Amarapurkar DN, Hashimoto E, Lesmana LA, Sollano JD, Chen PJ, Goh KL. How common is non-alcoholic fatty liver disease in the Asia-Pacific region and are there local differences? J Gastroenterol Hepatol. 2007;22:788-793.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 271]  [Cited by in F6Publishing: 286]  [Article Influence: 16.8]  [Reference Citation Analysis (0)]
9.  Li Z, Xue J, Chen P, Chen L, Yan S, Liu L. Prevalence of nonalcoholic fatty liver disease in mainland of China: a meta-analysis of published studies. J Gastroenterol Hepatol. 2014;29:42-51.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 153]  [Cited by in F6Publishing: 154]  [Article Influence: 15.4]  [Reference Citation Analysis (0)]
10.  Fleischman MW, Budoff M, Zeb I, Li D, Foster T. NAFLD prevalence differs among hispanic subgroups: the Multi-Ethnic Study of Atherosclerosis. World J Gastroenterol. 2014;20:4987-4993.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 79]  [Cited by in F6Publishing: 90]  [Article Influence: 9.0]  [Reference Citation Analysis (0)]
11.  Dowman JK, Tomlinson JW, Newsome PN. Systematic review: the diagnosis and staging of non-alcoholic fatty liver disease and non-alcoholic steatohepatitis. Aliment Pharmacol Ther. 2011;33:525-540.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 223]  [Cited by in F6Publishing: 216]  [Article Influence: 16.6]  [Reference Citation Analysis (0)]
12.  Panera N, Gnani D, Crudele A, Ceccarelli S, Nobili V, Alisi A. MicroRNAs as controlled systems and controllers in non-alcoholic fatty liver disease. World J Gastroenterol. 2014;20:15079-15086.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 48]  [Cited by in F6Publishing: 45]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
13.  Fargion S, Porzio M, Fracanzani AL. Nonalcoholic fatty liver disease and vascular disease: state-of-the-art. World J Gastroenterol. 2014;20:13306-13324.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 144]  [Cited by in F6Publishing: 156]  [Article Influence: 15.6]  [Reference Citation Analysis (0)]
14.  Zhan YT, Li L, Weng J, Song X, Yang SQ, An W. Serum autofluorescence, a potential serum marker for the diagnosis of liver fibrosis in rats. Int J Mol Sci. 2012;13:12130-12139.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 7]  [Cited by in F6Publishing: 7]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
15.  Paschos P, Paletas K. Non alcoholic fatty liver disease and metabolic syndrome. Hippokratia. 2009;13:9-19.  [PubMed]  [DOI]  [Cited in This Article: ]
16.  Firneisz G. Non-alcoholic fatty liver disease and type 2 diabetes mellitus: the liver disease of our age? World J Gastroenterol. 2014;20:9072-9089.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 65]  [Reference Citation Analysis (0)]
17.  Zhan Y, Zhao F, Xie P, Zhong L, Li D, Gai Q, Li L, Wei H, Zhang L, An W. Mechanism of the effect of glycosyltransferase GLT8D2 on fatty liver. Lipids Health Dis. 2015;14:43.  [PubMed]  [DOI]  [Cited in This Article: ]
18.  Armstrong MJ, Adams LA, Canbay A, Syn WK. Extrahepatic complications of nonalcoholic fatty liver disease. Hepatology. 2014;59:1174-1197.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 387]  [Cited by in F6Publishing: 404]  [Article Influence: 40.4]  [Reference Citation Analysis (0)]
19.  Machado MV, Cortez-Pinto H. Non-invasive diagnosis of non-alcoholic fatty liver disease. A critical appraisal. J Hepatol. 2013;58:1007-1019.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 278]  [Cited by in F6Publishing: 259]  [Article Influence: 23.5]  [Reference Citation Analysis (0)]
20.  Lonardo A, Sookoian S, Chonchol M, Loria P, Targher G. Cardiovascular and systemic risk in nonalcoholic fatty liver disease - atherosclerosis as a major player in the natural course of NAFLD. Curr Pharm Des. 2013;19:5177-5192.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 8]  [Cited by in F6Publishing: 8]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
21.  Musso G, Gambino R, Cassader M, Pagano G. Meta-analysis: natural history of non-alcoholic fatty liver disease (NAFLD) and diagnostic accuracy of non-invasive tests for liver disease severity. Ann Med. 2011;43:617-649.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 886]  [Cited by in F6Publishing: 868]  [Article Influence: 66.8]  [Reference Citation Analysis (0)]
22.  Sasaki A, Nitta H, Otsuka K, Umemura A, Baba S, Obuchi T, Wakabayashi G. Bariatric surgery and non-alcoholic Fatty liver disease: current and potential future treatments. Front Endocrinol (Lausanne). 2014;5:164.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 73]  [Cited by in F6Publishing: 80]  [Article Influence: 8.0]  [Reference Citation Analysis (0)]
23.  Ibrahim MA, Kelleni M, Geddawy A. Nonalcoholic fatty liver disease: current and potential therapies. Life Sci. 2013;92:114-118.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 59]  [Cited by in F6Publishing: 64]  [Article Influence: 5.3]  [Reference Citation Analysis (0)]
24.  Baran B, Akyüz F. Non-alcoholic fatty liver disease: what has changed in the treatment since the beginning? World J Gastroenterol. 2014;20:14219-14229.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 31]  [Cited by in F6Publishing: 30]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
25.  Zhang SR, Fan XM. Ghrelin-ghrelin O-acyltransferase system in the pathogenesis of nonalcoholic fatty liver disease. World J Gastroenterol. 2015;21:3214-3222.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 17]  [Cited by in F6Publishing: 20]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
26.  Tziomalos K, Athyros VG, Karagiannis A. Non-alcoholic fatty liver disease in type 2 diabetes: pathogenesis and treatment options. Curr Vasc Pharmacol. 2012;10:162-172.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 32]  [Cited by in F6Publishing: 38]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
27.  Day CP, James OF. Steatohepatitis: a tale of two “hits”? Gastroenterology. 1998;114:842-845.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2953]  [Cited by in F6Publishing: 3011]  [Article Influence: 115.8]  [Reference Citation Analysis (36)]
28.  Jou J, Choi SS, Diehl AM. Mechanisms of disease progression in nonalcoholic fatty liver disease. Semin Liver Dis. 2008;28:370-379.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 308]  [Cited by in F6Publishing: 333]  [Article Influence: 20.8]  [Reference Citation Analysis (0)]
29.  Polyzos SA, Kountouras J, Zavos C. Nonalcoholic fatty liver disease: the pathogenetic roles of insulin resistance and adipocytokines. Curr Mol Med. 2009;9:299-314.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 204]  [Cited by in F6Publishing: 229]  [Article Influence: 15.3]  [Reference Citation Analysis (0)]
30.  Dai X, Wang B. Role of gut barrier function in the pathogenesis of nonalcoholic Fatty liver disease. Gastroenterol Res Pract. 2015;2015:287348.  [PubMed]  [DOI]  [Cited in This Article: ]
31.  Marcolin E, San-Miguel B, Vallejo D, Tieppo J, Marroni N, González-Gallego J, Tuñón MJ. Quercetin treatment ameliorates inflammation and fibrosis in mice with nonalcoholic steatohepatitis. J Nutr. 2012;142:1821-1828.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 120]  [Cited by in F6Publishing: 117]  [Article Influence: 9.8]  [Reference Citation Analysis (0)]
32.  Tilg H, Moschen AR. Evolution of inflammation in nonalcoholic fatty liver disease: the multiple parallel hits hypothesis. Hepatology. 2010;52:1836-1846.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1543]  [Cited by in F6Publishing: 1685]  [Article Influence: 120.4]  [Reference Citation Analysis (0)]
33.  Romeo S, Kozlitina J, Xing C, Pertsemlidis A, Cox D, Pennacchio LA, Boerwinkle E, Cohen JC, Hobbs HH. Genetic variation in PNPLA3 confers susceptibility to nonalcoholic fatty liver disease. Nat Genet. 2008;40:1461-1465.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2233]  [Cited by in F6Publishing: 2380]  [Article Influence: 148.8]  [Reference Citation Analysis (0)]
34.  Lewis JR, Mohanty SR. Nonalcoholic fatty liver disease: a review and update. Dig Dis Sci. 2010;55:560-578.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 223]  [Cited by in F6Publishing: 227]  [Article Influence: 16.2]  [Reference Citation Analysis (0)]
35.  Wanless IR, Shiota K. The pathogenesis of nonalcoholic steatohepatitis and other fatty liver diseases: a four-step model including the role of lipid release and hepatic venular obstruction in the progression to cirrhosis. Semin Liver Dis. 2004;24:99-106.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 143]  [Cited by in F6Publishing: 150]  [Article Influence: 7.5]  [Reference Citation Analysis (0)]
36.  Gusdon AM, Song KX, Qu S. Nonalcoholic Fatty liver disease: pathogenesis and therapeutics from a mitochondria-centric perspective. Oxid Med Cell Longev. 2014;2014:637027.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 101]  [Cited by in F6Publishing: 107]  [Article Influence: 10.7]  [Reference Citation Analysis (0)]
37.  Zhao F, Xie P, Jiang J, Zhang L, An W, Zhan Y. The effect and mechanism of tamoxifen-induced hepatocyte steatosis in vitro. Int J Mol Sci. 2014;15:4019-4030.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 41]  [Cited by in F6Publishing: 44]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
38.  Donnelly KL, Smith CI, Schwarzenberg SJ, Jessurun J, Boldt MD, Parks EJ. Sources of fatty acids stored in liver and secreted via lipoproteins in patients with nonalcoholic fatty liver disease. J Clin Invest. 2005;115:1343-1351.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2112]  [Cited by in F6Publishing: 2403]  [Article Influence: 126.5]  [Reference Citation Analysis (0)]
39.  Matsuzaka T, Shimano H. Molecular mechanisms involved in hepatic steatosis and insulin resistance. J Diabetes Investig. 2011;2:170-175.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 48]  [Cited by in F6Publishing: 50]  [Article Influence: 5.0]  [Reference Citation Analysis (0)]
40.  Berlanga A, Guiu-Jurado E, Porras JA, Auguet T. Molecular pathways in non-alcoholic fatty liver disease. Clin Exp Gastroenterol. 2014;7:221-239.  [PubMed]  [DOI]  [Cited in This Article: ]
41.  Lewis GF, Carpentier A, Adeli K, Giacca A. Disordered fat storage and mobilization in the pathogenesis of insulin resistance and type 2 diabetes. Endocr Rev. 2002;23:201-229.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 687]  [Cited by in F6Publishing: 744]  [Article Influence: 33.8]  [Reference Citation Analysis (0)]
42.  Kawano Y, Cohen DE. Mechanisms of hepatic triglyceride accumulation in non-alcoholic fatty liver disease. J Gastroenterol. 2013;48:434-441.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 545]  [Cited by in F6Publishing: 633]  [Article Influence: 57.5]  [Reference Citation Analysis (0)]
43.  Xu L, Kim JK, Bai Q, Zhang X, Kakiyama G, Min HK, Sanyal AJ, Pandak WM, Ren S. 5-cholesten-3β,25-diol 3-sulfate decreases lipid accumulation in diet-induced nonalcoholic fatty liver disease mouse model. Mol Pharmacol. 2013;83:648-658.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 21]  [Cited by in F6Publishing: 26]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
44.  Li S, Brown MS, Goldstein JL. Bifurcation of insulin signaling pathway in rat liver: mTORC1 required for stimulation of lipogenesis, but not inhibition of gluconeogenesis. Proc Natl Acad Sci USA. 2010;107:3441-3446.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 512]  [Cited by in F6Publishing: 552]  [Article Influence: 39.4]  [Reference Citation Analysis (0)]
45.  Cohen JC, Horton JD, Hobbs HH. Human fatty liver disease: old questions and new insights. Science. 2011;332:1519-1523.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1601]  [Cited by in F6Publishing: 1585]  [Article Influence: 121.9]  [Reference Citation Analysis (2)]
46.  Nakahara T, Hindsgaul O, Palcic MM, Nishimura S. Computational design and experimental evaluation of glycosyltransferase mutants: engineering of a blood type B galactosyltransferase with enhanced glucosyltransferase activity. Protein Eng Des Sel. 2006;19:571-578.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 9]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
47.  Moon S, Kim SR, Zhao G, Yi J, Yoo Y, Jin P, Lee SW, Jung KH, Zhang D, An G. Rice glycosyltransferase1 encodes a glycosyltransferase essential for pollen wall formation. Plant Physiol. 2013;161:663-675.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 72]  [Cited by in F6Publishing: 66]  [Article Influence: 6.0]  [Reference Citation Analysis (0)]
48.  Unligil UM, Rini JM. Glycosyltransferase structure and mechanism. Curr Opin Struct Biol. 2000;10:510-517.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 275]  [Cited by in F6Publishing: 289]  [Article Influence: 12.0]  [Reference Citation Analysis (0)]
49.  Breton C, Snajdrová L, Jeanneau C, Koca J, Imberty A. Structures and mechanisms of glycosyltransferases. Glycobiology. 2006;16:29R-37R.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 459]  [Cited by in F6Publishing: 469]  [Article Influence: 24.7]  [Reference Citation Analysis (0)]
50.  Pesnot T, Jørgensen R, Palcic MM, Wagner GK. Structural and mechanistic basis for a new mode of glycosyltransferase inhibition. Nat Chem Biol. 2010;6:321-323.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 64]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
51.  Chang A, Singh S, Phillips GN, Thorson JS. Glycosyltransferase structural biology and its role in the design of catalysts for glycosylation. Curr Opin Biotechnol. 2011;22:800-808.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 107]  [Cited by in F6Publishing: 122]  [Article Influence: 9.4]  [Reference Citation Analysis (0)]
52.  Jørgensen R, Pesnot T, Lee HJ, Palcic MM, Wagner GK. Base-modified donor analogues reveal novel dynamic features of a glycosyltransferase. J Biol Chem. 2013;288:26201-26208.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 15]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
53.  Quintero CA, Giraudo CG, Villarreal M, Montich G, Maccioni HJ. Identification of a site in Sar1 involved in the interaction with the cytoplasmic tail of glycolipid glycosyltransferases. J Biol Chem. 2010;285:30340-30346.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 20]  [Cited by in F6Publishing: 21]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
54.  Giraudo CG, Maccioni HJ. Endoplasmic reticulum export of glycosyltransferases depends on interaction of a cytoplasmic dibasic motif with Sar1. Mol Biol Cell. 2003;14:3753-3766.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 161]  [Cited by in F6Publishing: 169]  [Article Influence: 8.0]  [Reference Citation Analysis (0)]
55.  Schmitz KR, Liu J, Li S, Setty TG, Wood CS, Burd CG, Ferguson KM. Golgi localization of glycosyltransferases requires a Vps74p oligomer. Dev Cell. 2008;14:523-534.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 151]  [Cited by in F6Publishing: 158]  [Article Influence: 9.9]  [Reference Citation Analysis (0)]
56.  Schoberer J, Vavra U, Stadlmann J, Hawes C, Mach L, Steinkellner H, Strasser R. Arginine/lysine residues in the cytoplasmic tail promote ER export of plant glycosylation enzymes. Traffic. 2009;10:101-115.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 78]  [Cited by in F6Publishing: 76]  [Article Influence: 5.1]  [Reference Citation Analysis (0)]
57.  Breton C, Mucha J, Jeanneau C. Structural and functional features of glycosyltransferases. Biochimie. 2001;83:713-718.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 96]  [Cited by in F6Publishing: 95]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
58.  Breton C, Fournel-Gigleux S, Palcic MM. Recent structures, evolution and mechanisms of glycosyltransferases. Curr Opin Struct Biol. 2012;22:540-549.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 153]  [Cited by in F6Publishing: 159]  [Article Influence: 13.3]  [Reference Citation Analysis (0)]
59.  Lairson LL, Henrissat B, Davies GJ, Withers SG. Glycosyltransferases: structures, functions, and mechanisms. Annu Rev Biochem. 2008;77:521-555.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1277]  [Cited by in F6Publishing: 1370]  [Article Influence: 85.6]  [Reference Citation Analysis (0)]
60.  Franco OL, Rigden DJ. Fold recognition analysis of glycosyltransferase families: further members of structural superfamilies. Glycobiology. 2003;13:707-712.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 27]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
61.  Breton C, Imberty A. Structure/function studies of glycosyltransferases. Curr Opin Struct Biol. 1999;9:563-571.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 146]  [Cited by in F6Publishing: 151]  [Article Influence: 6.0]  [Reference Citation Analysis (0)]
62.  Breton C, Bettler E, Joziasse DH, Geremia RA, Imberty A. Sequence-function relationships of prokaryotic and eukaryotic galactosyltransferases. J Biochem. 1998;123:1000-1009.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 115]  [Cited by in F6Publishing: 119]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
63.  Chiu CP, Watts AG, Lairson LL, Gilbert M, Lim D, Wakarchuk WW, Withers SG, Strynadka NC. Structural analysis of the sialyltransferase CstII from Campylobacter jejuni in complex with a substrate analog. Nat Struct Mol Biol. 2004;11:163-170.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 167]  [Cited by in F6Publishing: 167]  [Article Influence: 8.4]  [Reference Citation Analysis (0)]
64.  Nagae M, Yamaguchi Y. Function and 3D structure of the N-glycans on glycoproteins. Int J Mol Sci. 2012;13:8398-8429.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 83]  [Cited by in F6Publishing: 90]  [Article Influence: 7.5]  [Reference Citation Analysis (0)]
65.  Kawai F, Grass S, Kim Y, Choi KJ, St Geme JW, Yeo HJ. Structural insights into the glycosyltransferase activity of the Actinobacillus pleuropneumoniae HMW1C-like protein. J Biol Chem. 2011;286:38546-38557.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 41]  [Cited by in F6Publishing: 42]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
66.  Harazono A, Kawasaki N, Kawanishi T, Hayakawa T. Site-specific glycosylation analysis of human apolipoprotein B100 using LC/ESI MS/MS. Glycobiology. 2005;15:447-462.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 65]  [Cited by in F6Publishing: 62]  [Article Influence: 3.1]  [Reference Citation Analysis (0)]
67.  Yang CY, Gu ZW, Weng SA, Kim TW, Chen SH, Pownall HJ, Sharp PM, Liu SW, Li WH, Gotto AM. Structure of apolipoprotein B-100 of human low density lipoproteins. Arteriosclerosis. 1989;9:96-108.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 167]  [Cited by in F6Publishing: 181]  [Article Influence: 5.2]  [Reference Citation Analysis (0)]
68.  Chan L. Apolipoprotein B, the major protein component of triglyceride-rich and low density lipoproteins. J Biol Chem. 1992;267:25621-25624.  [PubMed]  [DOI]  [Cited in This Article: ]
69.  Ihara Y, Yoshimura M, Miyoshi E, Nishikawa A, Sultan AS, Toyosawa S, Ohnishi A, Suzuki M, Yamamura K, Ijuhin N. Ectopic expression of N-acetylglucosaminyltransferase III in transgenic hepatocytes disrupts apolipoprotein B secretion and induces aberrant cellular morphology with lipid storage. Proc Natl Acad Sci USA. 1998;95:2526-2530.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 38]  [Cited by in F6Publishing: 37]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
70.  Gusarova V, Caplan AJ, Brodsky JL, Fisher EA. Apoprotein B degradation is promoted by the molecular chaperones hsp90 and hsp70. J Biol Chem. 2001;276:24891-24900.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 105]  [Cited by in F6Publishing: 103]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
71.  Berriot-Varoqueaux N, Dannoura AH, Moreau A, Verthier N, Sassolas A, Cadiot G, Lachaux A, Munck A, Schmitz J, Aggerbeck LP. Apolipoprotein B48 glycosylation in abetalipoproteinemia and Anderson’s disease. Gastroenterology. 2001;121:1101-1108.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 32]  [Cited by in F6Publishing: 32]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
72.  Hoosdally SJ, Andress EJ, Wooding C, Martin CA, Linton KJ. The Human Scavenger Receptor CD36: glycosylation status and its role in trafficking and function. J Biol Chem. 2009;284:16277-16288.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 79]  [Cited by in F6Publishing: 95]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
73.  Koo SH. Nonalcoholic fatty liver disease: molecular mechanisms for the hepatic steatosis. Clin Mol Hepatol. 2013;19:210-215.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 252]  [Cited by in F6Publishing: 295]  [Article Influence: 26.8]  [Reference Citation Analysis (0)]
74.  Guinez C, Filhoulaud G, Rayah-Benhamed F, Marmier S, Dubuquoy C, Dentin R, Moldes M, Burnol AF, Yang X, Lefebvre T. O-GlcNAcylation increases ChREBP protein content and transcriptional activity in the liver. Diabetes. 2011;60:1399-1413.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 159]  [Cited by in F6Publishing: 165]  [Article Influence: 12.7]  [Reference Citation Analysis (0)]
75.  Hirsova P, Gores GJ. Death Receptor-Mediated Cell Death and Proinflammatory Signaling in Nonalcoholic Steatohepatitis. Cell Mol Gastroenterol Hepatol. 2015;1:17-27.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 127]  [Cited by in F6Publishing: 147]  [Article Influence: 14.7]  [Reference Citation Analysis (0)]
76.  Alkhouri N, Carter-Kent C, Feldstein AE. Apoptosis in nonalcoholic fatty liver disease: diagnostic and therapeutic implications. Expert Rev Gastroenterol Hepatol. 2011;5:201-212.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 185]  [Cited by in F6Publishing: 176]  [Article Influence: 13.5]  [Reference Citation Analysis (0)]
77.  Akazawa Y, Gores GJ. Death receptor-mediated liver injury. Semin Liver Dis. 2007;27:327-338.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 48]  [Article Influence: 2.8]  [Reference Citation Analysis (0)]
78.  Shatnyeva OM, Kubarenko AV, Weber CE, Pappa A, Schwartz-Albiez R, Weber AN, Krammer PH, Lavrik IN. Modulation of the CD95-induced apoptosis: the role of CD95 N-glycosylation. PLoS One. 2011;6:e19927.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 50]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
79.  Gatselis NK, Ntaios G, Makaritsis K, Dalekos GN. Adiponectin: a key playmaker adipocytokine in non-alcoholic fatty liver disease. Clin Exp Med. 2014;14:121-131.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 41]  [Cited by in F6Publishing: 46]  [Article Influence: 4.2]  [Reference Citation Analysis (0)]
80.  Peake PW, Hughes JT, Shen Y, Charlesworth JA. Glycosylation of human adiponectin affects its conformation and stability. J Mol Endocrinol. 2007;39:45-52.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 29]  [Cited by in F6Publishing: 29]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
81.  Sato C, Yasukawa Z, Honda N, Matsuda T, Kitajima K. Identification and adipocyte differentiation-dependent expression of the unique disialic acid residue in an adipose tissue-specific glycoprotein, adipo Q. J Biol Chem. 2001;276:28849-28856.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 71]  [Cited by in F6Publishing: 74]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
82.  Yang X, Bhaumik M, Bhattacharyya R, Gong S, Rogler CE, Stanley P. New evidence for an extra-hepatic role of N-acetylglucosaminyltransferase III in the progression of diethylnitrosamine-induced liver tumors in mice. Cancer Res. 2000;60:3313-3319.  [PubMed]  [DOI]  [Cited in This Article: ]
83.  Okada T, Ihara H, Ito R, Taniguchi N, Ikeda Y. Bidirectional N-acetylglucosamine transfer mediated by beta-1,4-N-acetylglucosaminyltransferase III. Glycobiology. 2009;19:368-374.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 12]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
84.  Shim JK, Lee YC, Chung TH, Kim CH. Elevated expression of bisecting N-acetylglucosaminyltransferase-III gene in a human fetal hepatocyte cell line by hepatitis B virus. J Gastroenterol Hepatol. 2004;19:1374-1387.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 14]  [Cited by in F6Publishing: 14]  [Article Influence: 0.7]  [Reference Citation Analysis (0)]
85.  Hanashima S, Korekane H, Taniguchi N, Yamaguchi Y. Synthesis of N-glycan units for assessment of substrate structural requirements of N-acetylglucosaminyltransferase III. Bioorg Med Chem Lett. 2014;24:4533-4537.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 9]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
86.  Paulson JC, Colley KJ. Glycosyltransferases. Structure, localization, and control of cell type-specific glycosylation. J Biol Chem. 1989;264:17615-17618.  [PubMed]  [DOI]  [Cited in This Article: ]
87.  Ihara Y, Nishikawa A, Tohma T, Soejima H, Niikawa N, Taniguchi N. cDNA cloning, expression, and chromosomal localization of human N-acetylglucosaminyltransferase III (GnT-III). J Biochem. 1993;113:692-698.  [PubMed]  [DOI]  [Cited in This Article: ]
88.  Aryal RP, Ju T, Cummings RD. The endoplasmic reticulum chaperone Cosmc directly promotes in vitro folding of T-synthase. J Biol Chem. 2010;285:2456-2462.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 56]  [Article Influence: 3.7]  [Reference Citation Analysis (0)]
89.  Aryal RP, Ju T, Cummings RD. Identification of a novel protein binding motif within the T-synthase for the molecular chaperone Cosmc. J Biol Chem. 2014;289:11630-11641.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 17]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
90.  Perrine C, Ju T, Cummings RD, Gerken TA. Systematic determination of the peptide acceptor preferences for the human UDP-Gal: glycoprotein-alpha-GalNAc beta 3 galactosyltransferase (T-synthase). Glycobiology. 2009;19:321-328.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 24]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
91.  Ju T, Xia B, Aryal RP, Wang W, Wang Y, Ding X, Mi R, He M, Cummings RD. A novel fluorescent assay for T-synthase activity. Glycobiology. 2011;21:352-362.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 29]  [Cited by in F6Publishing: 38]  [Article Influence: 2.7]  [Reference Citation Analysis (0)]
92.  Fu J, Gerhardt H, McDaniel JM, Xia B, Liu X, Ivanciu L, Ny A, Hermans K, Silasi-Mansat R, McGee S. Endothelial cell O-glycan deficiency causes blood/lymphatic misconnections and consequent fatty liver disease in mice. J Clin Invest. 2008;118:3725-3737.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 193]  [Cited by in F6Publishing: 213]  [Article Influence: 13.3]  [Reference Citation Analysis (0)]
93.  Carmeliet P. Angiogenesis in life, disease and medicine. Nature. 2005;438:932-936.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2451]  [Cited by in F6Publishing: 2456]  [Article Influence: 136.4]  [Reference Citation Analysis (0)]
94.  Black DD. Development and physiological regulation of intestinal lipid absorption. I. Development of intestinal lipid absorption: cellular events in chylomicron assembly and secretion. Am J Physiol Gastrointest Liver Physiol. 2007;293:G519-G524.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 85]  [Cited by in F6Publishing: 81]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
95.  Ihara H, Hanashima S, Okada T, Ito R, Yamaguchi Y, Taniguchi N, Ikeda Y. Fucosylation of chitooligosaccharides by human alpha1,6-fucosyltransferase requires a nonreducing terminal chitotriose unit as a minimal structure. Glycobiology. 2010;20:1021-1033.  [PubMed]  [DOI]  [Cited in This Article: ]
96.  Li W, Nakagawa T, Koyama N, Wang X, Jin J, Mizuno-Horikawa Y, Gu J, Miyoshi E, Kato I, Honke K. Down-regulation of trypsinogen expression is associated with growth retardation in alpha1,6-fucosyltransferase-deficient mice: attenuation of proteinase-activated receptor 2 activity. Glycobiology. 2006;16:1007-1019.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 35]  [Cited by in F6Publishing: 38]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
97.  Kötzler MP, Blank S, Bantleon FI, Wienke M, Spillner E, Meyer B. Donor assists acceptor binding and catalysis of human α1,6-fucosyltransferase. ACS Chem Biol. 2013;8:1830-1840.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 17]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
98.  Miyoshi E, Noda K, Yamaguchi Y, Inoue S, Ikeda Y, Wang W, Ko JH, Uozumi N, Li W, Taniguchi N. The alpha1-6-fucosyltransferase gene and its biological significance. Biochim Biophys Acta. 1999;1473:9-20.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 171]  [Cited by in F6Publishing: 166]  [Article Influence: 6.6]  [Reference Citation Analysis (0)]
99.  Yamaguchi Y, Ikeda Y, Takahashi T, Ihara H, Tanaka T, Sasho C, Uozumi N, Yanagidani S, Inoue S, Fujii J. Genomic structure and promoter analysis of the human alpha1, 6-fucosyltransferase gene (FUT8). Glycobiology. 2000;10:637-643.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 21]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
100.  Ihara H, Ikeda Y, Taniguchi N. Reaction mechanism and substrate specificity for nucleotide sugar of mammalian alpha1,6-fucosyltransferase--a large-scale preparation and characterization of recombinant human FUT8. Glycobiology. 2006;16:333-342.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 59]  [Cited by in F6Publishing: 55]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
101.  Wang W, Li W, Ikeda Y, Miyagawa JI, Taniguchi M, Miyoshi E, Sheng Y, Ekuni A, Ko JH, Yamamoto Y. Ectopic expression of alpha1,6 fucosyltransferase in mice causes steatosis in the liver and kidney accompanied by a modification of lysosomal acid lipase. Glycobiology. 2001;11:165-174.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 29]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
102.  Wei HS, Wei HL, Zhao F, Zhong LP, Zhan YT. Glycosyltransferase GLT8D2 positively regulates ApoB100 protein expression in hepatocytes. Int J Mol Sci. 2013;14:21435-21446.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 19]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
103.  Moylan CA, Pang H, Dellinger A, Suzuki A, Garrett ME, Guy CD, Murphy SK, Ashley-Koch AE, Choi SS, Michelotti GA. Hepatic gene expression profiles differentiate presymptomatic patients with mild versus severe nonalcoholic fatty liver disease. Hepatology. 2014;59:471-482.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 215]  [Cited by in F6Publishing: 227]  [Article Influence: 22.7]  [Reference Citation Analysis (0)]
104.  Hussain MM, Shi J, Dreizen P. Microsomal triglyceride transfer protein and its role in apoB-lipoprotein assembly. J Lipid Res. 2003;44:22-32.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 425]  [Cited by in F6Publishing: 415]  [Article Influence: 19.8]  [Reference Citation Analysis (0)]
105.  Radominska-Pandya A, Czernik PJ, Little JM, Battaglia E, Mackenzie PI. Structural and functional studies of UDP-glucuronosyltransferases. Drug Metab Rev. 1999;31:817-899.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 394]  [Cited by in F6Publishing: 360]  [Article Influence: 14.4]  [Reference Citation Analysis (0)]
106.  Itäaho K, Laakkonen L, Finel M. How many and which amino acids are responsible for the large activity differences between the highly homologous UDP-glucuronosyltransferases (UGT) 1A9 and UGT1A10? Drug Metab Dispos. 2010;38:687-696.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 15]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
107.  Mackenzie PI, Bock KW, Burchell B, Guillemette C, Ikushiro S, Iyanagi T, Miners JO, Owens IS, Nebert DW. Nomenclature update for the mammalian UDP glycosyltransferase (UGT) gene superfamily. Pharmacogenet Genomics. 2005;15:677-685.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 640]  [Cited by in F6Publishing: 613]  [Article Influence: 32.3]  [Reference Citation Analysis (0)]
108.  Ziegler K, Tumova S, Kerimi A, Williamson G. Cellular asymmetric catalysis by UDP-glucuronosyltransferase 1A8 shows functional localization to the basolateral plasma membrane. J Biol Chem. 2015;290:7622-7633.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 8]  [Cited by in F6Publishing: 8]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
109.  Buckley DB, Klaassen CD. Induction of mouse UDP-glucuronosyltransferase mRNA expression in liver and intestine by activators of aryl-hydrocarbon receptor, constitutive androstane receptor, pregnane X receptor, peroxisome proliferator-activated receptor alpha, and nuclear factor erythroid 2-related factor 2. Drug Metab Dispos. 2009;37:847-856.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 115]  [Cited by in F6Publishing: 117]  [Article Influence: 7.8]  [Reference Citation Analysis (0)]
110.  Xu J, Kulkarni SR, Li L, Slitt AL. UDP-glucuronosyltransferase expression in mouse liver is increased in obesity- and fasting-induced steatosis. Drug Metab Dispos. 2012;40:259-266.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 53]  [Cited by in F6Publishing: 63]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
111.  Zhang L, Chu X, Wang H, Xie H, Guo C, Cao L, Zhou X, Wang G, Hao H. Dysregulations of UDP-glucuronosyltransferases in rats with valproic acid and high fat diet induced fatty liver. Eur J Pharmacol. 2013;721:277-285.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 17]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
112.  Hardwick RN, Ferreira DW, More VR, Lake AD, Lu Z, Manautou JE, Slitt AL, Cherrington NJ. Altered UDP-glucuronosyltransferase and sulfotransferase expression and function during progressive stages of human nonalcoholic fatty liver disease. Drug Metab Dispos. 2013;41:554-561.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 97]  [Cited by in F6Publishing: 89]  [Article Influence: 8.1]  [Reference Citation Analysis (0)]
113.  Jancova P, Anzenbacher P, Anzenbacherova E. Phase II drug metabolizing enzymes. Biomed Pap Med Fac Univ Palacky Olomouc Czech Repub. 2010;154:103-116.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 334]  [Cited by in F6Publishing: 352]  [Article Influence: 25.1]  [Reference Citation Analysis (0)]
114.  Hanioka N, Naito T, Narimatsu S. Human UDP-glucuronosyltransferase isoforms involved in bisphenol A glucuronidation. Chemosphere. 2008;74:33-36.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 99]  [Cited by in F6Publishing: 105]  [Article Influence: 6.6]  [Reference Citation Analysis (0)]