Next Article in Journal
Oncogenic Pathways in Neurodegenerative Diseases
Next Article in Special Issue
Protein Kinase C (Pkc)-δ Mediates Arginine-Induced Glucagon Secretion in Pancreatic α-Cells
Previous Article in Journal
Characterization of a Novel Aspect of Tissue Scarring Following Experimental Spinal Cord Injury and the Implantation of Bioengineered Type-I Collagen Scaffolds in the Adult Rat: Involvement of Perineurial-like Cells?
Previous Article in Special Issue
Stem Cell-Derived β Cells: A Versatile Research Platform to Interrogate the Genetic Basis of β Cell Dysfunction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

HNF1A Mutations and Beta Cell Dysfunction in Diabetes

Department of Medicine and Bioregulatory Science, Graduate School of Medical Sciences, Kyushu University, Fukuoka 812-8582, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(6), 3222; https://doi.org/10.3390/ijms23063222
Submission received: 28 February 2022 / Revised: 14 March 2022 / Accepted: 16 March 2022 / Published: 16 March 2022
(This article belongs to the Special Issue Fate of Pancreatic Islets in Type 2 Diabetes)

Abstract

:
Understanding the genetic factors of diabetes is essential for addressing the global increase in type 2 diabetes. HNF1A mutations cause a monogenic form of diabetes called maturity-onset diabetes of the young (MODY), and HNF1A single-nucleotide polymorphisms are associated with the development of type 2 diabetes. Numerous studies have been conducted, mainly using genetically modified mice, to explore the molecular basis for the development of diabetes caused by HNF1A mutations, and to reveal the roles of HNF1A in multiple organs, including insulin secretion from pancreatic beta cells, lipid metabolism and protein synthesis in the liver, and urinary glucose reabsorption in the kidneys. Recent studies using human stem cells that mimic MODY have provided new insights into beta cell dysfunction. In this article, we discuss the involvement of HNF1A in beta cell dysfunction by reviewing previous studies using genetically modified mice and recent findings in human stem cell-derived beta cells.

1. Introduction

The number of patients with type 2 diabetes (T2DM) has increased four-fold worldwide over the past 30 years, and is expected to increase by another 1.5 times in the next 20 years [1]. In dealing with the T2DM epidemic, it is crucial to consider the etiology and pathology of the disease.
T2DM is a multifactorial disease caused by both genetic and environmental factors, including diet, physical activity, and age. It is well recognized that obesity and aging contribute to diabetes development, however, there are two populations with the same weight and age: one with diabetes and the other with normal glucose levels. This difference in glucose tolerance is derived from genetic factors.
Genetic factors in diabetes have been investigated globally in multiple regions and races for (1) gene mutations in both nuclear and mitochondrial DNA, and (2) single nucleotide polymorphisms (SNPs) associated with diabetes. The most common form of monogenic diabetes is MODY, which accounts for 1–2.5% of all diabetes cases [2,3]. MODY is characterized by autosomal dominant inheritance, early onset of diabetes under 25 years of age, and impaired insulin secretion due to mutations in transcription factors involved in beta cell differentiation and function. To date, 14 MODY subtypes have been reported, including MODY1 (HNF4A), MODY2 (GK), MODY3 (HNF1A), MODY4 (PDX1), MODY5 (HNF1B), and MODY6 (NeuroD1) [4]. Among these, HNF1A gene mutations are the most common cause of MODY. In addition, SNPs in and near HNF1A are associated with an increased risk of T2DM.
Although an association between T2DM and HNF1A SNPs has been reported, it is unclear whether this relationship is causal, similar to that observed between MODY3 and HNF1A mutations. To clarify the pathophysiological role of HNF1A in diabetes, multiple studies have been conducted using Hnf1a-null mice; however, due to the inconsistency of phenotypes between Hnf1a-null mice and people with diabetes, the molecular basis has not been fully elucidated.
Undoubtedly, genetically modified animals will continue to be a powerful tool for elucidating human pathology. Nevertheless, recent human stem cell-derived beta cell research has provided new insights into the pathogenesis of diabetes caused by genetic variants in humans [5]. Here, we review the current understanding of HNF1A in multiple organs, including the pancreas, and discuss how HNF1A mutations contribute to the pathogenesis of diabetes, primarily focusing on beta cell dysfunction.

2. Extrapancreatic Organs and HNF1A

HNF1A is a transcription factor that belongs to the HNF1 homeobox family. The endogenous ligands of HNF1A are unknown, whereas aspirin and resveratrol have been identified as candidates for exogenous ligands by in silico analysis [6]. Aspirin and resveratrol are not specific ligands for HNF1A because they affect various molecules, including cyclooxygenases and sirtuins, respectively; however, these substances can potentially modulate the activity of HNF1A. Furthermore, protein–protein interactions have been identified between HNF1A and sirtuins [7,8]. For example, sirtuin 1 physically interacts with HNF1A in the liver to regulate HNF1A target genes [8].
HNF1A is mainly expressed in the liver, gut, pancreas, and kidneys. The physiological roles of HNF1A in each organ are summarized in Figure 1. During mouse embryo development, HNF1A mRNA was detected in the yolk sac at embryonic day (E) 8.5, and in the liver, intestine, and kidneys at E14.5 [9]. HNF1A has three isoforms: A (encoded by exons 1–10), B (exons 1–7), and C (exons 1–6). Isoform A is predominant in the liver, kidneys, and fetal pancreas, whereas isoforms B and C are abundant in the adult pancreas and islets [10].

2.1. Liver and HNF1A

HNF families, including HNF1A, are involved in liver development, function, and tumor growth [11]. For example, hepatocellular adenomas (HCAs), benign liver tumors, are classified into eight subgroups based on genotypes; one subgroup is defined by inactivating HNF1A mutations with a histological feature of tumor steatosis [12]. The majority of HCAs in this subgroup exhibit biallelic somatic inactivation of HNF1A. Certain HCAs found in MODY3 families contain a somatic mutation in one allele within the tumor and a heterozygous germline mutation in the other allele [13,14,15], showing the tumor-suppressive role of HNF1A in the liver.
Lipid metabolism plays a critical role in liver function. Homozygous Hnf1a-deficient mice exhibited growth retardation and hepatomegaly [16]. Hnf1a-null mice also had hyperbileacidemia and hypercholesterolemia, along with altered gene expression related to the synthesis and uptake of bile acids and de novo biosynthesis of cholesterol [17]. In another study, Hnf1a-null mice had elevated blood triglyceride levels and fatty liver accompanied by increased fatty acid synthesis [18]. Furthermore, lipidomic analysis of HCAs caused by HNF1A inactivation showed that de novo lipogenesis is enhanced in tumors, revealing the underlying mechanism of lipid accumulation in tumors [19]. However, the lipid profile of MODY3 and non-diabetic participants was similar, characterized by higher high-density lipoprotein cholesterol levels and lower triglyceride levels compared to T2DM participants [20,21]. Thus, heterozygous HNF1A mutations are unlikely to be sufficient to cause abnormal lipid metabolism.
Protein synthesis plays another important role in liver function. Albumin and C-reactive protein (CRP) are typical proteins produced by the liver. High CRP levels have been associated with various diseases, including cardiovascular disease [22], T2DM [23], obesity [24], psychological distress and depression [25], and cancer [26]. Plasma CRP levels are also associated with SNPs in and near the HNF1A gene [27]. Moreover, high-sensitivity CRP levels are lower in MODY3 groups than in a non-diabetic group and other diabetic groups, including type 1 and type 2 diabetes, MODY1, MODY2, and MODY5 [28,29]. There are two HNF1A-binding sites within the human CRP promoter and HNF1A is capable of activating CRP transcription [30]. Therefore, it is conceivable that the transcriptional activity of HNF1A is reduced in the liver of patients with MODY3.

2.2. Kidney and HNF1A

The expression of HNF1A and HNF1B was detectable in the kidneys during early development [9]. HNF1B mutations cause developmental kidney diseases, including renal cysts, single kidneys, renal hypoplasia, and electrolyte abnormalities [31], whereas the majority of HNF1A mutations are not associated with abnormal kidney morphology [32,33]. It has been observed that the reabsorption of urinary glucose is lower in patients with MODY3 than in those with type 1 and type 2 diabetes [34,35]. Intriguingly, the binding of phlorizin, a competitive inhibitor of sodium–glucose cotransporter (SGLT) 1 and 2, to the cell surface of proximal tubules was diminished in Hnf1a-null mice [16]. Furthermore, the SGLT2 expression was decreased in the kidneys of Hnf1a-null mice [35], and HNF1A directly promoted SGLT2 transcription in rodents and humans [35,36,37].
SGLT2 inhibitors have mainly been used for T2DM treatment. In one study, patients with MODY3 received a single dose of dapagliflozin, an SGLT2 inhibitor [38]. Surprisingly, changes in urinary glucose excretion following drug administration were more substantial in patients with MODY3 than in those with T2DM, contrary to the expectation that the drug effect would be attenuated in patients with MODY3, owing to the decreased SGLT2 expression. The precise reason for this result is unclear; however, it is possible that urinary glucose reabsorption via SGLT1, other than SGLT2, is enhanced in patients with T2DM [39,40,41].
Hnf1a-null mice also exhibited phenylketonuria and Fanconi syndrome [16]. However, these phenotypes were not reproduced in another study [42]. In addition, there have been few reports of renal dysfunction in patients with MODY3, whereas HNF4A mutations (MODY1) can cause Fanconi syndrome [43,44,45].

2.3. Gut and HNF1A

HNF1A and HNF1B are highly expressed in the crypts of the small intestine [46], and cooperatively regulate intestinal cell differentiation [46,47]. Mice with Hnf1a or Hnf1b deficiency in the gut had no severe phenotypes, however, mice with Hnf1a and Hnf1b double deficiencies in the gut died of dehydration soon after being born because of their inability to absorb water in the intestine [47].
Lactase-phlorizin hydrolase (LPH) is expressed in the small intestine, and plays an important role in the digestion of lactose. The physical interaction between the DNA-binding domain of HNF1A and the C-terminal zinc finger of GATA5 synergistically activates the promoter of human LPH [48].
The role of HNF1A in the gut in diabetes has not been well investigated; however, ghrelin, an appetite-stimulating hormone produced primarily in the stomach and intestine, is a potential target of HNF1A. Hnf1a-null mice possessed a higher number of ghrelin-positive cells in the intestine, accompanied by increased ghrelin gene expression, with elevated total and active ghrelin levels in the blood [49]. Likewise, fasting total ghrelin levels were elevated in patients with MODY3 compared to those with type 1 and type 2 diabetes [50]. In contrast, acylated ghrelin, or active ghrelin, was comparable in patients with MODY3 to healthy individuals and those with T2DM [51]. These inconsistent results may partly reflect differences in the measurement of ghrelin (i.e., total or acylated ghrelin).

3. Pancreas and HNF1A

3.1. Physiological and Pathophysiological Roles of HNF1A in Endocrine and Exocrine Cells of the Pancreas

HNF1A is expressed in endocrine cells, including alpha, beta, delta, and pancreatic polypeptide (PP) cells, and exocrine cells, including acinar and duct cells, in mouse pancreatic tissue [52]. HNF1A was detected in most pancreatic epithelial cells at E10.5, and hormone-positive cells and amylase-positive cells at E15.5, whereas the HFN1A expression was weak in duct cells [52]. During liver development, a transcriptional hierarchy exists, in which HNF4A positively regulates HNF1A expression [53,54], whereas HNF1A controls HNF4A by directly binding to the P2 promoter of HNF4A in differentiated pancreatic cells [55,56]. In addition, HNF1A regulates the PDX1 expression, which is essential for pancreatic development and the maintenance of beta cell function, in a cooperative manner with PDX1 itself in rodent beta cells [57]. The interaction between HNF1A and HNF4A is important in the differentiation of human induced pluripotent stem (iPS) cells into beta cells, supporting the relevance of the HNF1A–HNF4A axis during pancreatic development [58].
Pancreatic endocrine cells consist of alpha, beta, delta, PP, and other cells (e.g., ghrelin-producing cells). HNF1A was involved in glucagon secretion during hypoglycemia by directly regulating SGLT1 expression in mouse alpha cells [59]. Hnf1a-null mice showed impaired insulin secretion and subsequent hyperglycemia, suggesting that HNF1A was involved in insulin secretion in mouse beta cells [42,60]. However, heterozygous Hnf1a-deficient mice have normal blood glucose levels [60]. In addition, Hnf1a-null mice on a C3H or CBA background were non-diabetic [61], demonstrating phenotypic differences between patients with MODY3 and Hnf1a-null mice. The role of HNF1A in delta, PP, and other cells remains unknown.
Hnf1a-null mice showed a high acinar cell proliferation rate, and both endocrine and exocrine granules were observed in acinar cells, suggesting that HNF1A is required for normal exocrine differentiation [62]. Although it is undetermined whether HNF1A mutations lead to disorders in human exocrine cells, genome-wide association studies (GWAS) have revealed an association between HNF1A SNPs and pancreatic cancer [63,64]. Pancreatic cancer generally develops from the exocrine gland, and pancreatic ductal adenocarcinoma (PDAC) is the most common pancreatic cancer. A recent study has shown that HNF1A is highly expressed in pancreatic cancer stem cells (PCSCs), suggesting that HFN1A is a potential central regulator of PCSC function [65]. In that study, the knockdown of HNF1A in multiple PDAC cell lines resulted in tumor growth inhibition and apoptosis, accompanied by a reduction in stem cell markers. In contrast, HNF1A overexpression in PDAC cell lines promoted tumor growth, accompanied by the elevation of stem cell markers. These data indicate that the activation of HNF1A in pancreatic cancer promotes tumor growth, while HNF1A has also been shown to have a tumor-suppressive role in pancreatic cancer [66,67,68]. For example, HNF1A exerts its tumor-suppressive effect by forming a complex with lysine-specific demethylase 6A, encoded by the KDM6A gene, to inhibit the oncogenic pathway in a mouse acinar cell line [68]. These discrepancies may be due to the use of different cell lines in each experiment.

3.2. Type 2 Diabetes and the HNF1A Gene

HNF1A consists of 631 amino acids, including amino-terminal dimerization (residues 1–32), DNA-binding (203–276), and transactivation domains (281–631) [69]. The most frequent HNF1A mutation in MODY3 is P291fsinsC (p.G292fs), which has a dominant negative effect on wild-type HNF1A by lacking most of the transactivation domain [70,71,72]. Missense mutations are abundant in the amino-terminal dimerization and DNA-binding domains, and deletion mutations are frequently found in the transactivation domain [73,74], potentially leading to clinical heterogeneity in patients with MODY3 [75,76].
SNPs are the most common DNA sequence variants in the genome, with an estimated 10 million SNPs in the human genome [77]. Statistical analysis of the association between diseases and SNPs allows us to infer disease risk, the molecular basis of the disease, and drug targets for patients with certain SNPs. GWAS have identified over 100 common variants that cause T2DM in humans (minor allele frequency (MAF) > 1%) [78]. According to the simulated models of T2DM, these common T2DM variants can potentially account for approximately 75% of the heritability of the disease [79]. The individual common variants are weakly associated with T2DM (odds ratio < 1.3), however, people with more disease-associated variants are expected to have an increased risk of developing the disease [80]. In contrast, rare coding variants (MAF < 0.5%) have a negligible effect on the heritability of T2DM [81]. Considering the small contribution of rare variants, the common variants appear to play a significant role in the heritability of T2DM.
HNF1A SNPs are associated with the risk of T2DM, and this finding was confirmed across different ethnic groups [82,83,84]. In detail, rs1169288 (I27L) [85,86], rs1800574 (A98V) [85,86], rs140730081 [85,86], G319S [87], rs2464196 (S487N) [88], M490T [89], E508K [89], rs7957197 [82], and rs12427353 [90] were associated with T2DM. Among these, I27L, A98V, and S487N are common coding variants of HNF1A, and may have a small impact on the risk of T2DM [91,92].
An oral glucose tolerance test (OGTT) was performed on 17 participants with rare coding variants of HNF1A (MAF < 1%), and only one showed an aberrant increase in 2-h glucose levels after 75 g OGTT (defined as a difference of >90 mg/dl between 2-h glucose and fasting glucose levels) [93]. Although insulin concentrations were not measured in these rare variant carriers, it is possible that the reduction in insulin secretion was partly compensated by an increase in insulin sensitivity and a decrease in the threshold for urinary glucose excretion, as observed in patients with MODY3 [88,94]. Furthermore, the results suggest that the functional validation of HNF1A variants is necessary [91], and also demonstrate that the prediction of T2DM in the general population based on rare variants obtained from GWAS can overestimate the contribution of rare variants, leading to an increase in the incidence of false positives [93,95]. Rare variants of a disease are generally assumed to have a significant effect on the phenotype, however, potential ascertainment biases in Mendelian genetic studies occur, even for rare variants of HNF1A that have been shown to cause MODY3 owing to the selection of patients with overt diabetes or a family history of MODY [76,93,95].
Despite the frequency of MODY3 and the association of HNF1A SNPs with T2DM, the molecular basis of HNF1A mutations leading to diabetes remains elusive. In the following sections, we will discuss how HNF1A controls insulin secretion and regulates downstream genes in beta cells.

3.3. Roles of HNF1A in Beta Cells

The microarray analysis of pancreatic islets from Hnf1a-null mice showed that HNF1A regulated the genes involved in glucose and amino acid metabolism, including glycolysis, tricarboxylic acid cycle, and oxidative phosphorylation [96]. MODY3 is characterized by reduced insulin secretion before the onset of diabetes and displays a distinct phenotype from type 1 and type 2 diabetes [97,98]. An impaired insulin secretion response to high glucose or arginine was revealed in ex vivo islets from Hnf1a-null mice [60]. The reduction in arginine-induced insulin secretion was also observed in MIN6 cells, a mouse beta cell line, expressing a dominant negative mutant of HNF1A (P291fsinsC) [99]. Human iPS cell-derived beta cells with a MODY3 mutation (HNF1A+/H126D) showed impaired glucose-stimulated insulin secretion (GSIS) after six months of maturation by transplantation into mice [100]. In addition, impaired GSIS due to HNF1A dysfunction was confirmed in islets from a diabetic patient harboring HNF1A+/T260M [101].
T2DM affects insulin secretion through a combination of genetic factors, including the accumulation of disease-associated common variants, and environmental factors, including metabolic overload and aging. HNF1A mutations (MODY3) reduce insulin secretion by altering glucose metabolism in beta cells [96,102,103]. HNF1A mutations result in decreased glucose transporter 2 (GLUT2) and glucose uptake [100,102]. HNF1A regulates liver-type pyruvate kinase, a rate-limiting enzyme in glycolysis, and GLUT2 in rodent beta cells [104,105,106,107]. Furthermore, the most common HNF1A mutation in MODY3 (P291fsinsC) reduces mitochondrial ATP production in mouse beta cells [107]. Therefore, HNF1A is likely to be involved in ATP production, predominantly in the mitochondria, and subsequent insulin secretion, by maintaining glucose flux (Figure 2).
It is controversial whether HNF1A directly regulates mitochondrial function-related genes; these genes might be regulated through a network of transcription factors, such as PDX1 and HNF4A [104]. Loss of HNF1A in human embryonic stem (ES) cell-derived beta cells impairs mitochondrial respiration accompanied by a decrease in LINKA, a human-specific long non-coding RNA, suggesting the involvement of HNF1A in mitochondrial function [108]. Further studies are required to elucidate the role of HNF1A in regulating mitochondrial metabolism in beta cells.
In addition to the central metabolic pathway, HNF1A directly regulates TMEM27 [109,110] and hepatocyte growth factor activator [111], which promote proliferation in mouse beta cells. Collectrin (TMEM27) also controls insulin exocytosis through the formation of the SNARE complex [110]. In addition, HNF1A regulates genes whose functions in beta cells are poorly understood, including AKR1C19 [96,112].

3.4. Epigenetics and HNF1A in Beta Cells

GWAS have provided us with a better understanding of the heritability of T2DM over the past 15 years; however, improving the accuracy of personalized risk prediction for T2DM remains a challenge [113,114]. The difficulty in accurately predicting the onset of T2DM is likely due to a complex combination of genetic and environmental factors, including individual lifestyle [115]. Moreover, much remains to be explored in terms of genetic factors. For example, epigenetic changes, including histone modification, DNA methylation, and non-coding RNA, can alter transcription factor binding to the genome and gene expression, independently of the DNA sequence. Although there is accumulating evidence that epigenetic changes are associated with the development of T2DM [116,117,118], the underlying molecular mechanisms are not fully understood.
Histone modification plays an important role in beta cell development and function [119]. The acetylation of histones loosens the chromatin structure and facilitates the binding of transcription factors to DNA, and is thus considered an active epigenetic marker. CREB-binding protein (CBP) and P300 are the major histone acetyltransferases; they have similar structures and are involved in the acetylation of H3K27. HNF1A is required for the maintenance of histone acetylation in the promoter regions of its target genes, including GLUT2, in mouse islets [105]. It was also found that CBP and P300/CBP associated factor interacted with the N- and C-terminal domains of HNF1A, respectively [120]. In addition, HNF1A recruits P300 to the promoter of human GLUT2 [121]. Interestingly, the genes that decreased in CBP- or P300-null islets overlapped with those in Hnf1a-null islets [122]. Taken together, these results indicate that HNF1A and CBP/P300 cooperate to activate gene expression, possibly through histone modification, including the acetylation of H3K27.

3.5. Beta Cell Mass and HNF1A

Beta cell dysfunction is a pathological hallmark of T2DM, and the preservation or restoration of beta cell mass is important for T2DM treatment. The decrease in beta cell mass in T2DM was considered to be caused by beta cell death [123]. However, within five years of diagnosing T2DM, the difference in beta cell mass between healthy individuals and patients with T2DM was small, with a large overlap, suggesting that the decrease in beta cell mass was a result of disease progression rather than a contributor to the development of the disease [124,125].
Although the data on beta cell mass in patients with MODY3 are inconclusive, beta cells from a patient with HNF1A+/T260M showed impaired GSIS but no obvious decrease in the cell mass compared to the average beta cell mass of healthy individuals [101]. The beta cell mass in Hnf1a-null mice was smaller than that in heterozygous Hnf1a-deficient mice, and this difference disappeared after adjusting for body weight [60]. Alternatively, the common HNF1a mutation (P291fsinsC) decreased mouse beta cell proliferation [126]. Likewise, a dominant negative mutant of HNF1A increased sensitivity to endoplasmic reticulum (ER) stress in mouse islets and promoted drug-induced apoptosis in mouse beta cells [127,128]. In summary, HNF1A mutations may not have a significant impact on beta cell mass per se, however, it is possible that certain triggers, including metabolic stress, reduce beta cell proliferation and cause beta cell death in patients with HNF1A mutations.
Beta cell proliferation in adults is considerably low under normal conditions [129]; however, obesity expands beta cell mass in response to increased insulin demand [130]. Various explanations have been proposed for changes in beta cell mass, including beta cell replication [131,132,133], neogenesis [134,135], and transdifferentiation [134,136], which have been linked to the therapeutic promise of beta cell regeneration [137].
Cellular plasticity has been shown in the islets of patients with diabetes [125,138]. In Korea, a decrease in beta cell mass and an increase in the relative volume of alpha cells in islets were observed in non-obese patients with T2DM [139]. The mechanism underlying the increase in alpha cells in the islets of T2DM is unknown, however, the concept of beta cell dedifferentiation has recently been proposed [140]. Beta cell dedifferentiation is an adaptive process in which beta cells return to a progenitor-like stage after prolonged metabolic stress, and a portion of the immature beta cells convert to other endocrine cells, such as alpha and delta cells [141]. The concept of beta cell dedifferentiation explains the mechanisms of beta cell decline and alpha cell increase in T2DM [139,142,143,144] and supports the observation that beta cell loss progresses after the onset of T2DM [124]. In contrast, dedifferentiation plays a minor role in beta cell decline in obese T2DM patients with an average BMI > 40 [145]. Moreover, it remains unclear whether beta cell differentiation and ER stress-induced apoptosis can coexist in the same cells, or whether they are mutually exclusive.
The molecular basis of dedifferentiation remains to be fully elucidated [146], however, it has been shown that FoxO1 migrates to the nucleus in beta cells prior to dedifferentiation, maintaining the MODY network, including HNF1A and HNF4A [147]. The molecular basis of MODY-related genes causing diabetes and their involvement in dedifferentiation have rarely been investigated because of the difficulties in obtaining pancreatic tissue or beta cells from patients with MODY. Recent studies have attempted to overcome these obstacles by generating iPS cell-derived beta cells from MODY [100,148,149,150]. Studies using stem cell-derived beta cells have indicated a possible involvement of HNF1A in beta cell dedifferentiation [108,151]. HNF1A deficiency in human ES cell-derived beta cells results in increased alpha cell markers, including glucagon, while decreasing the expression of PAX4, which is involved in beta cell development [108]. Decreased PAX4 levels were also observed in iPS cell-derived beta cells lacking HNF1A [151]. Despite this evidence, the most compelling approach to the involvement of HNF1A in dedifferentiation is to determine whether beta cell dedifferentiation occurs in the pancreatic tissue of patients with MODY3. Interestingly, the alpha cell mass was higher in a diabetic patient harboring HNF1A+/T260M than in healthy individuals [101]. It is currently unknown whether common variants associated with T2DM are involved in dedifferentiation. The roles of HNF1A variants in MODY3 and T2DM are summarized in Table 1.

4. Pharmacologic Treatment for Diabetic Patients with HNF1A Mutations

The glucose-lowering effect of metformin was comparable in patients with MODY3 and those with T2DM, while the sensitivity to sulfonylureas (SU) was higher in patients with MODY3 than in those with T2DM [152,153]. The high sensitivity of patients with MODY3 to SU may be due to their increased insulin sensitivity [94]; therefore, patients with MODY3 are likely to be more susceptible to hypoglycemia with SU [154].
Monotherapy with glucagon-like peptide-1 receptor agonists (GLP-1 RA) [153,155] or dipeptidyl peptidase-4 (DPP4) inhibitors [156,157] may prevent hypoglycemia and achieve good glycemic control in patients with MODY3. Moreover, combination therapy with GLP-1 RA and SU may increase insulin secretion in patients with MODY3 [158], and combination therapy with DPP4 inhibitors and SU may improve glycemic control without significantly increasing hypoglycemia in patients with MODY3 [154].
Given the response to SU treatment for patients with MODY3, it is possible that HNF1A mutations do not substantially impair the insulin secretory pathway after membrane depolarization in beta cells. However, HNF1A mutations have been shown to alter the formation of mature insulin secretory granules [110,151,159], and HNF1A may also regulate insulin exocytosis. In addition, the impact of HNF1A mutations on other endocrine cells, including alpha cells, requires further investigation in humans.

5. Concluding Remarks

Recent findings indicate that HNF1A mutations are associated with an increase in alpha cells in human pancreatic islets and differentiation of stem cell-derived beta cells toward alpha cells, showing the relevance of studies using human islets and beta cells. The molecular basis of HNF1A abnormality in insulin secretion in human beta cells and the pathophysiological role of HNF1A in the liver, kidneys, and gut in diabetes requires further investigation. In this review, we discuss the similarities and differences in the pathophysiological role of HNF1A in patients with MODY and T2DM, however, few reports were found directly comparing the effects of HNF1A mutations in MODY3 and T2DM-related HNF1A common variants. Therefore, it is unclear whether the findings of HNF1A mutations in MODY3 can be extrapolated to the pathogenesis of T2DM. It is expected that in addition to conventional animal experiments, human stem cells and new technologies will reveal new molecular bases and therapeutic targets for diabetes (Table 2).

Author Contributions

Conceptualization, Y.M.; writing, Y.M., T.M. and Y.O.; supervision, Y.O. All authors have read and agreed to the published version of the manuscript.

Funding

Y.M. was supported by JSPS KAKENHI grant number 21K20934.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chatterjee, S.; Khunti, K.; Davies, M.J. Type 2 diabetes. Lancet 2017, 389, 2239–2251. [Google Scholar] [CrossRef]
  2. Pihoker, C.; Gilliam, L.K.; Ellard, S.; Dabelea, D.; Davis, C.; Dolan, L.M.; Greenbaum, C.J.; Imperatore, G.; Lawrence, J.M.; Marcovina, S.M.; et al. Prevalence, characteristics and clinical diagnosis of maturity onset diabetes of the young due to mutations in HNF1A, HNF4A, and glucokinase: Results from the SEARCH for Diabetes in Youth. J. Clin. Endocrinol. Metab. 2013, 98, 4055–4062. [Google Scholar] [CrossRef] [Green Version]
  3. Shepherd, M.; Shields, B.; Hammersley, S.; Hudson, M.; McDonald, T.J.; Colclough, K.; Oram, R.A.; Knight, B.; Hyde, C.; Cox, J.; et al. Systematic Population Screening, Using Biomarkers and Genetic Testing, Identifies 2.5% of the U.K. Pediatric Diabetes Population With Monogenic Diabetes. Diabetes Care 2016, 39, 1879–1888. [Google Scholar] [CrossRef] [Green Version]
  4. Urakami, T. Maturity-onset diabetes of the young (MODY): Current perspectives on diagnosis and treatment. Diabetes Metab. Syndr. Obes. 2019, 12, 1047–1056. [Google Scholar] [CrossRef] [Green Version]
  5. Bartolome, A. Stem Cell-Derived beta Cells: A Versatile Research Platform to Interrogate the Genetic Basis of beta Cell Dysfunction. Int. J. Mol. Sci. 2022, 23, 501. [Google Scholar] [CrossRef]
  6. Sridhar, G.R.; Nageswara Rao, P.V.; Kaladhar, D.S.; Devi, T.U.; Kumar, S.V. In Silico Docking of HNF-1a Receptor Ligands. Adv. Bioinform. 2012, 2012, 705435. [Google Scholar] [CrossRef] [Green Version]
  7. Tao, R.; Xiong, X.; DePinho, R.A.; Deng, C.X.; Dong, X.C. FoxO3 transcription factor and Sirt6 deacetylase regulate low density lipoprotein (LDL)-cholesterol homeostasis via control of the proprotein convertase subtilisin/kexin type 9 (Pcsk9) gene expression. J. Biol. Chem. 2013, 288, 29252–29259. [Google Scholar] [CrossRef] [Green Version]
  8. Grimm, A.A.; Brace, C.S.; Wang, T.; Stormo, G.D.; Imai, S. A nutrient-sensitive interaction between Sirt1 and HNF-1α regulates Crp expression. Aging Cell 2011, 10, 305–317. [Google Scholar] [CrossRef] [Green Version]
  9. Cereghini, S.; Ott, M.; Power, S.; Maury, M. Expression patterns of vHNF1 and HNF1 homeoproteins in early postimplantation embryos suggest distinct and sequential developmental roles. Development 1992, 116, 783–797. [Google Scholar] [CrossRef]
  10. Harries, L.W.; Ellard, S.; Stride, A.; Morgan, N.G.; Hattersley, A.T. Isomers of the TCF1 gene encoding hepatocyte nuclear factor-1 alpha show differential expression in the pancreas and define the relationship between mutation position and clinical phenotype in monogenic diabetes. Hum. Mol. Genet 2006, 15, 2216–2224. [Google Scholar] [CrossRef] [Green Version]
  11. Lau, H.H.; Ng, N.H.J.; Loo, L.S.W.; Jasmen, J.B.; Teo, A.K.K. The molecular functions of hepatocyte nuclear factors—In and beyond the liver. J. Hepatol. 2018, 68, 1033–1048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Nault, J.C.; Couchy, G.; Balabaud, C.; Morcrette, G.; Caruso, S.; Blanc, J.F.; Bacq, Y.; Calderaro, J.; Paradis, V.; Ramos, J.; et al. Molecular Classification of Hepatocellular Adenoma Associates With Risk Factors, Bleeding, and Malignant Transformation. Gastroenterology 2017, 152, 880–894.e6. [Google Scholar] [CrossRef] [Green Version]
  13. Bluteau, O.; Jeannot, E.; Bioulac-Sage, P.; Marques, J.M.; Blanc, J.F.; Bui, H.; Beaudoin, J.C.; Franco, D.; Balabaud, C.; Laurent-Puig, P.; et al. Bi-allelic inactivation of TCF1 in hepatic adenomas. Nat. Genet. 2002, 32, 312–315. [Google Scholar] [CrossRef] [PubMed]
  14. Fu, J.; Wang, T.; Zhai, X.; Xiao, X. Primary hepatocellular adenoma due to biallelic HNF1A mutations and its co-occurrence with MODY 3: Case-report and review of the literature. Endocrine 2020, 67, 544–551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Reznik, Y.; Dao, T.; Coutant, R.; Chiche, L.; Jeannot, E.; Clauin, S.; Rousselot, P.; Fabre, M.; Oberti, F.; Fatome, A.; et al. Hepatocyte nuclear factor-1 alpha gene inactivation: Cosegregation between liver adenomatosis and diabetes phenotypes in two maturity-onset diabetes of the young (MODY)3 families. J. Clin. Endocrinol. Metab. 2004, 89, 1476–1480. [Google Scholar] [CrossRef] [PubMed]
  16. Pontoglio, M.; Barra, J.; Hadchouel, M.; Doyen, A.; Kress, C.; Bach, J.P.; Babinet, C.; Yaniv, M. Hepatocyte nuclear factor 1 inactivation results in hepatic dysfunction, phenylketonuria, and renal Fanconi syndrome. Cell 1996, 84, 575–585. [Google Scholar] [CrossRef] [Green Version]
  17. Shih, D.Q.; Bussen, M.; Sehayek, E.; Ananthanarayanan, M.; Shneider, B.L.; Suchy, F.J.; Shefer, S.; Bollileni, J.S.; Gonzalez, F.J.; Breslow, J.L.; et al. Hepatocyte nuclear factor-1alpha is an essential regulator of bile acid and plasma cholesterol metabolism. Nat. Genet. 2001, 27, 375–382. [Google Scholar] [CrossRef]
  18. Akiyama, T.E.; Ward, J.M.; Gonzalez, F.J. Regulation of the Liver Fatty Acid-binding Protein Gene by Hepatocyte Nuclear Factor 1α (HNF1α). J. Biol. Chem. 2000, 275, 27117–27122. [Google Scholar] [CrossRef]
  19. Rebouissou, S.; Imbeaud, S.; Balabaud, C.; Boulanger, V.; Bertrand-Michel, J.; Terce, F.; Auffray, C.; Bioulac-Sage, P.; Zucman-Rossi, J. HNF1alpha inactivation promotes lipogenesis in human hepatocellular adenoma independently of SREBP-1 and carbohydrate-response element-binding protein (ChREBP) activation. J. Biol. Chem. 2007, 282, 14437–14446. [Google Scholar] [CrossRef] [Green Version]
  20. Pearson, E.R.; Badman, M.K.; Lockwood, C.R.; Clark, P.M.; Ellard, S.; Bingham, C.; Hattersley, A.T. Contrasting diabetes phenotypes associated with hepatocyte nuclear factor-1alpha and -1beta mutations. Diabetes Care 2004, 27, 1102–1107. [Google Scholar] [CrossRef] [Green Version]
  21. McDonald, T.J.; McEneny, J.; Pearson, E.R.; Thanabalasingham, G.; Szopa, M.; Shields, B.M.; Ellard, S.; Owen, K.R.; Malecki, M.T.; Hattersley, A.T.; et al. Lipoprotein composition in HNF1A-MODY: Differentiating between HNF1A-MODY and type 2 diabetes. Clin. Chim. Acta 2012, 413, 927–932. [Google Scholar] [CrossRef]
  22. Lagrand, W.K.; Visser, C.A.; Hermens, W.T.; Niessen, H.W.; Verheugt, F.W.; Wolbink, G.J.; Hack, C.E. C-reactive protein as a cardiovascular risk factor: More than an epiphenomenon? Circulation 1999, 100, 96–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Hu, F.B.; Meigs, J.B.; Li, T.Y.; Rifai, N.; Manson, J.E. Inflammatory markers and risk of developing type 2 diabetes in women. Diabetes 2004, 53, 693–700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Visser, M.; Bouter, L.M.; McQuillan, G.M.; Wener, M.H.; Harris, T.B. Elevated C-reactive protein levels in overweight and obese adults. JAMA 1999, 282, 2131–2135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Wium-Andersen, M.K.; Orsted, D.D.; Nielsen, S.F.; Nordestgaard, B.G. Elevated C-reactive protein levels, psychological distress, and depression in 73, 131 individuals. JAMA Psychiatry 2013, 70, 176–184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Erlinger, T.P.; Platz, E.A.; Rifai, N.; Helzlsouer, K.J. C-reactive protein and the risk of incident colorectal cancer. JAMA 2004, 291, 585–590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Ridker, P.M.; Pare, G.; Parker, A.; Zee, R.Y.; Danik, J.S.; Buring, J.E.; Kwiatkowski, D.; Cook, N.R.; Miletich, J.P.; Chasman, D.I. Loci related to metabolic-syndrome pathways including LEPR, HNF1A, IL6R, and GCKR associate with plasma C-reactive protein: The Women’s Genome Health Study. Am. J. Hum. Genet 2008, 82, 1185–1192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Owen, K.R.; Thanabalasingham, G.; James, T.J.; Karpe, F.; Farmer, A.J.; McCarthy, M.I.; Gloyn, A.L. Assessment of high-sensitivity C-reactive protein levels as diagnostic discriminator of maturity-onset diabetes of the young due to HNF1A mutations. Diabetes Care 2010, 33, 1919–1924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. McDonald, T.J.; Shields, B.M.; Lawry, J.; Owen, K.R.; Gloyn, A.L.; Ellard, S.; Hattersley, A.T. High-sensitivity CRP discriminates HNF1A-MODY from other subtypes of diabetes. Diabetes Care 2011, 34, 1860–1862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Toniatti, C.; Demartis, A.; Monaci, P.; Nicosia, A.; Ciliberto, G. Synergistic trans-activation of the human C-reactive protein promoter by transcription factor HNF-1 binding at two distinct sites. EMBO J. 1990, 9, 4467–4475. [Google Scholar] [CrossRef] [PubMed]
  31. Clissold, R.L.; Hamilton, A.J.; Hattersley, A.T.; Ellard, S.; Bingham, C. HNF1B-associated renal and extra-renal disease-an expanding clinical spectrum. Nat. Rev. Nephrol. 2015, 11, 102–112. [Google Scholar] [CrossRef]
  32. Simms, R.J.; Sayer, J.A.; Quinton, R.; Walker, M.; Ellard, S.; Goodship, T.H. Monogenic diabetes, renal dysplasia and hypopituitarism: A patient with a HNF1A mutation. QJM 2011, 104, 881–883. [Google Scholar] [CrossRef] [Green Version]
  33. Malecki, M.T.; Skupien, J.; Gorczynska-Kosiorz, S.; Klupa, T.; Nazim, J.; Moczulski, D.K.; Sieradzki, J. Renal malformations may be linked to mutations in the hepatocyte nuclear factor-1alpha (MODY3) gene. Diabetes Care 2005, 28, 2774–2776. [Google Scholar] [CrossRef] [Green Version]
  34. Menzel, R.; Kaisaki, P.J.; Rjasanowski, I.; Heinke, P.; Kerner, W.; Menzel, S. A low renal threshold for glucose in diabetic patients with a mutation in the hepatocyte nuclear factor-1alpha (HNF-1alpha) gene. Diabet. Med. J. Br. Diabet. Assoc. 1998, 15, 816–820. [Google Scholar] [CrossRef]
  35. Pontoglio, M.; Prié, D.; Cheret, C.; Doyen, A.; Leroy, C.; Froguel, P.; Velho, G.; Yaniv, M.; Friedlander, G. HNF1α controls renal glucose reabsorption in mouse and man. EMBO Rep. 2000, 1, 359–365. [Google Scholar] [CrossRef] [Green Version]
  36. Freitas, H.S.; Anhe, G.F.; Melo, K.F.; Okamoto, M.M.; Oliveira-Souza, M.; Bordin, S.; Machado, U.F. Na(+) -glucose transporter-2 messenger ribonucleic acid expression in kidney of diabetic rats correlates with glycemic levels: Involvement of hepatocyte nuclear factor-1alpha expression and activity. Endocrinology 2008, 149, 717–724. [Google Scholar] [CrossRef] [Green Version]
  37. Takesue, H.; Hirota, T.; Tachimura, M.; Tokashiki, A.; Ieiri, I. Nucleosome Positioning and Gene Regulation of the SGLT2 Gene in the Renal Proximal Tubular Epithelial Cells. Mol. Pharm. 2018, 94, 953–962. [Google Scholar] [CrossRef] [Green Version]
  38. Hohendorff, J.; Szopa, M.; Skupien, J.; Kapusta, M.; Zapala, B.; Platek, T.; Mrozinska, S.; Parpan, T.; Glodzik, W.; Ludwig-Galezowska, A.; et al. A single dose of dapagliflozin, an SGLT-2 inhibitor, induces higher glycosuria in GCK- and HNF1A-MODY than in type 2 diabetes mellitus. Endocrine 2017, 57, 272–279. [Google Scholar] [CrossRef] [Green Version]
  39. Chichger, H.; Cleasby, M.E.; Srai, S.K.; Unwin, R.J.; Debnam, E.S.; Marks, J. Experimental type II diabetes and related models of impaired glucose metabolism differentially regulate glucose transporters at the proximal tubule brush border membrane. Exp. Physiol. 2016, 101, 731–742. [Google Scholar] [CrossRef]
  40. Norton, L.; Shannon, C.E.; Fourcaudot, M.; Hu, C.; Wang, N.; Ren, W.; Song, J.; Abdul-Ghani, M.; DeFronzo, R.A.; Ren, J.; et al. Sodium-glucose co-transporter (SGLT) and glucose transporter (GLUT) expression in the kidney of type 2 diabetic subjects. Diabetes Obes. Metab. 2017, 19, 1322–1326. [Google Scholar] [CrossRef]
  41. Yakovleva, T.; Sokolov, V.; Chu, L.; Tang, W.; Greasley, P.J.; Peilot Sjogren, H.; Johansson, S.; Peskov, K.; Helmlinger, G.; Boulton, D.W.; et al. Comparison of the urinary glucose excretion contributions of SGLT2 and SGLT1: A quantitative systems pharmacology analysis in healthy individuals and patients with type 2 diabetes treated with SGLT2 inhibitors. Diabetes Obes. Metab. 2019, 21, 2684–2693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Lee, Y.H.; Sauer, B.; Gonzalez, F.J. Laron dwarfism and non-insulin-dependent diabetes mellitus in the Hnf-1alpha knockout mouse. Mol. Cell. Biol. 1998, 18, 3059–3068. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Hamilton, A.J.; Bingham, C.; McDonald, T.J.; Cook, P.R.; Caswell, R.C.; Weedon, M.N.; Oram, R.A.; Shields, B.M.; Shepherd, M.; Inward, C.D.; et al. The HNF4A R76W mutation causes atypical dominant Fanconi syndrome in addition to a beta cell phenotype. J. Med. Genet. 2014, 51, 165–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Liu, J.; Shen, Q.; Li, G.; Xu, H. HNF4A-related Fanconi syndrome in a Chinese patient: A case report and review of the literature. J. Med. Case Rep. 2018, 12, 203. [Google Scholar] [CrossRef]
  45. Marchesin, V.; Perez-Marti, A.; Le Meur, G.; Pichler, R.; Grand, K.; Klootwijk, E.D.; Kesselheim, A.; Kleta, R.; Lienkamp, S.; Simons, M. Molecular Basis for Autosomal-Dominant Renal Fanconi Syndrome Caused by HNF4A. Cell Rep. 2019, 29, 4407–4421. [Google Scholar] [CrossRef] [Green Version]
  46. Serfas, M.S.; Tyner, A.L. HNF-1 alpha and HNF-1 beta expression in mouse intestinal crypts. Am. J. Physiol. 1993, 265, G506–G513. [Google Scholar] [CrossRef] [PubMed]
  47. D’Angelo, A.; Bluteau, O.; Garcia-Gonzalez, M.A.; Gresh, L.; Doyen, A.; Garbay, S.; Robine, S.; Pontoglio, M. Hepatocyte nuclear factor 1alpha and beta control terminal differentiation and cell fate commitment in the gut epithelium. Development 2010, 137, 1573–1582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. van Wering, H.M.; Huibregtse, I.L.; van der Zwan, S.M.; de Bie, M.S.; Dowling, L.N.; Boudreau, F.; Rings, E.H.; Grand, R.J.; Krasinski, S.D. Physical interaction between GATA-5 and hepatocyte nuclear factor-1alpha results in synergistic activation of the human lactase-phlorizin hydrolase promoter. J. Biol. Chem. 2002, 277, 27659–27667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Brial, F.; Lussier, C.R.; Belleville, K.; Sarret, P.; Boudreau, F. Ghrelin Inhibition Restores Glucose Homeostasis in Hepatocyte Nuclear Factor-1alpha (MODY3)-Deficient Mice. Diabetes 2015, 64, 3314–3320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Nowak, N.; Hohendorff, J.; Solecka, I.; Szopa, M.; Skupien, J.; Kiec-Wilk, B.; Mlynarski, W.; Malecki, M.T. Circulating ghrelin level is higher in HNF1A-MODY and GCK-MODY than in polygenic forms of diabetes mellitus. Endocrine 2015, 50, 643–649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. St-Jean, M.; Boudreau, F.; Carpentier, A.C.; Hivert, M.F. HNF1alpha defect influences post-prandial lipid regulation. PLoS ONE 2017, 12, e0177110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Nammo, T.; Yamagata, K.; Hamaoka, R.; Zhu, Q.; Akiyama, T.E.; Gonzalez, F.J.; Miyagawa, J.; Matsuzawa, Y. Expression profile of MODY3/HNF-1alpha protein in the developing mouse pancreas. Diabetologia 2002, 45, 1142–1153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Li, J.; Ning, G.; Duncan, S.A. Mammalian hepatocyte differentiation requires the transcription factor HNF-4alpha. Genes Dev. 2000, 14, 464–474. [Google Scholar] [CrossRef] [PubMed]
  54. Kuo, C.J.; Conley, P.B.; Chen, L.; Sladek, F.M.; Darnell, J.E., Jr.; Crabtree, G.R. A transcriptional hierarchy involved in mammalian cell-type specification. Nature 1992, 355, 457–461. [Google Scholar] [CrossRef] [PubMed]
  55. Boj, S.F.; Parrizas, M.; Maestro, M.A.; Ferrer, J. A transcription factor regulatory circuit in differentiated pancreatic cells. Proc. Natl. Acad. Sci. USA 2001, 98, 14481–14486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Hansen, S.K.; Parrizas, M.; Jensen, M.L.; Pruhova, S.; Ek, J.; Boj, S.F.; Johansen, A.; Maestro, M.A.; Rivera, F.; Eiberg, H.; et al. Genetic evidence that HNF-1alpha-dependent transcriptional control of HNF-4alpha is essential for human pancreatic beta cell function. J. Clin. Investig. 2002, 110, 827–833. [Google Scholar] [CrossRef] [PubMed]
  57. Gerrish, K.; Cissell, M.A.; Stein, R. The role of hepatic nuclear factor 1 alpha and PDX-1 in transcriptional regulation of the pdx-1 gene. J. Biol. Chem. 2001, 276, 47775–47784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Legoy, T.A.; Mathisen, A.F.; Salim, Z.; Vethe, H.; Bjorlykke, Y.; Abadpour, S.; Paulo, J.A.; Scholz, H.; Raeder, H.; Ghila, L.; et al. In vivo Environment Swiftly Restricts Human Pancreatic Progenitors Toward Mono-Hormonal Identity via a HNF1A/HNF4A Mechanism. Front. Cell Dev. Biol. 2020, 8, 109. [Google Scholar] [CrossRef] [PubMed]
  59. Sato, Y.; Rahman, M.M.; Haneda, M.; Tsuyama, T.; Mizumoto, T.; Yoshizawa, T.; Kitamura, T.; Gonzalez, F.J.; Yamamura, K.I.; Yamagata, K. HNF1alpha controls glucagon secretion in pancreatic alpha-cells through modulation of SGLT1. Biochim. Biophys. Acta Mol. Basis Dis. 2020, 1866, 165898. [Google Scholar] [CrossRef] [PubMed]
  60. Pontoglio, M.; Sreenan, S.; Roe, M.; Pugh, W.; Ostrega, D.; Doyen, A.; Pick, A.J.; Baldwin, A.; Velho, G.; Froguel, P.; et al. Defective insulin secretion in hepatocyte nuclear factor 1alpha-deficient mice. J. Clin. Investig. 1998, 101, 2215–2222. [Google Scholar] [CrossRef]
  61. Garcia-Gonzalez, M.A.; Carette, C.; Bagattin, A.; Chiral, M.; Makinistoglu, M.P.; Garbay, S.; Prevost, G.; Madaras, C.; Herault, Y.; Leibovici, M.; et al. A suppressor locus for MODY3-diabetes. Sci. Rep. 2016, 6, 33087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Molero, X.; Vaquero, E.C.; Flandez, M.; Gonzalez, A.M.; Ortiz, M.A.; Cibrian-Uhalte, E.; Servitja, J.M.; Merlos, A.; Juanpere, N.; Massumi, M.; et al. Gene expression dynamics after murine pancreatitis unveils novel roles for Hnf1alpha in acinar cell homeostasis. Gut 2012, 61, 1187–1196. [Google Scholar] [CrossRef] [PubMed]
  63. Pierce, B.L.; Ahsan, H. Genome-wide “pleiotropy scan” identifies HNF1A region as a novel pancreatic cancer susceptibility locus. Cancer Res. 2011, 71, 4352–4358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Childs, E.J.; Mocci, E.; Campa, D.; Bracci, P.M.; Gallinger, S.; Goggins, M.; Li, D.; Neale, R.E.; Olson, S.H.; Scelo, G.; et al. Common variation at 2p13.3, 3q29, 7p13 and 17q25.1 associated with susceptibility to pancreatic cancer. Nat. Genet. 2015, 47, 911–916. [Google Scholar] [CrossRef] [PubMed]
  65. Abel, E.V.; Goto, M.; Magnuson, B.; Abraham, S.; Ramanathan, N.; Hotaling, E.; Alaniz, A.A.; Kumar-Sinha, C.; Dziubinski, M.L.; Urs, S.; et al. HNF1A is a novel oncogene that regulates human pancreatic cancer stem cell properties. eLife 2018, 7, e33947. [Google Scholar] [CrossRef]
  66. Hoskins, J.W.; Jia, J.; Flandez, M.; Parikh, H.; Xiao, W.; Collins, I.; Emmanuel, M.A.; Ibrahim, A.; Powell, J.; Zhang, L.; et al. Transcriptome analysis of pancreatic cancer reveals a tumor suppressor function for HNF1A. Carcinogenesis 2014, 35, 2670–2678. [Google Scholar] [CrossRef] [Green Version]
  67. Luo, Z.; Li, Y.; Wang, H.; Fleming, J.; Li, M.; Kang, Y.; Zhang, R.; Li, D. Hepatocyte nuclear factor 1A (HNF1A) as a possible tumor suppressor in pancreatic cancer. PLoS ONE 2015, 10, e0121082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Kalisz, M.; Bernardo, E.; Beucher, A.; Maestro, M.A.; Del Pozo, N.; Millan, I.; Haeberle, L.; Schlensog, M.; Safi, S.A.; Knoefel, W.T.; et al. HNF1A recruits KDM6A to activate differentiated acinar cell programs that suppress pancreatic cancer. EMBO J. 2020, 39, e102808. [Google Scholar] [CrossRef] [PubMed]
  69. Sneha, P.; Kumar, D.T.; Doss, C.G.P.; Siva, R.; Zayed, H. Determining the role of missense mutations in the POU domain of HNF1A that reduce the DNA-binding affinity: A computational approach. PLoS ONE 2017, 12, e0174953. [Google Scholar] [CrossRef] [Green Version]
  70. Yamagata, K.; Yang, Q.; Yamamoto, K.; Iwahashi, H.; Miyagawa, J.; Okita, K.; Yoshiuchi, I.; Miyazaki, J.; Noguchi, T.; Nakajima, H.; et al. Mutation P291fsinsC in the transcription factor hepatocyte nuclear factor-1alpha is dominant negative. Diabetes 1998, 47, 1231–1235. [Google Scholar] [CrossRef] [PubMed]
  71. Vaxillaire, M.; Abderrahmani, A.; Boutin, P.; Bailleul, B.; Froguel, P.; Yaniv, M.; Pontoglio, M. Anatomy of a homeoprotein revealed by the analysis of human MODY3 mutations. J. Biol. Chem. 1999, 274, 35639–35646. [Google Scholar] [CrossRef] [Green Version]
  72. Colclough, K.; Bellanne-Chantelot, C.; Saint-Martin, C.; Flanagan, S.E.; Ellard, S. Mutations in the genes encoding the transcription factors hepatocyte nuclear factor 1 alpha and 4 alpha in maturity-onset diabetes of the young and hyperinsulinemic hypoglycemia. Hum. Mutat. 2013, 34, 669–685. [Google Scholar] [CrossRef] [PubMed]
  73. Chi, Y.I.; Frantz, J.D.; Oh, B.C.; Hansen, L.; Dhe-Paganon, S.; Shoelson, S.E. Diabetes mutations delineate an atypical POU domain in HNF-1alpha. Mol. Cell 2002, 10, 1129–1137. [Google Scholar] [CrossRef]
  74. Ellard, S.; Colclough, K. Mutations in the genes encoding the transcription factors hepatocyte nuclear factor 1 alpha (HNF1A) and 4 alpha (HNF4A) in maturity-onset diabetes of the young. Hum. Mutat. 2006, 27, 854–869. [Google Scholar] [CrossRef] [PubMed]
  75. Bellanne-Chantelot, C.; Carette, C.; Riveline, J.P.; Valero, R.; Gautier, J.F.; Larger, E.; Reznik, Y.; Ducluzeau, P.H.; Sola, A.; Hartemann-Heurtier, A.; et al. The type and the position of HNF1A mutation modulate age at diagnosis of diabetes in patients with maturity-onset diabetes of the young (MODY)-3. Diabetes 2008, 57, 503–508. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Awa, W.L.; Thon, A.; Raile, K.; Grulich-Henn, J.; Meissner, T.; Schober, E.; Holl, R.W.; Group, D.P.-W.S. Genetic and clinical characteristics of patients with HNF1A gene variations from the German-Austrian DPV database. Eur. J. Endocrinol. 2011, 164, 513–520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Kruglyak, L.; Nickerson, D.A. Variation is the spice of life. Nat. Genet. 2001, 27, 234–236. [Google Scholar] [CrossRef] [PubMed]
  78. Visscher, P.M.; Wray, N.R.; Zhang, Q.; Sklar, P.; McCarthy, M.I.; Brown, M.A.; Yang, J. 10 Years of GWAS Discovery: Biology, Function, and Translation. Am. J. Hum. Genet. 2017, 101, 5–22. [Google Scholar] [CrossRef] [Green Version]
  79. Fuchsberger, C.; Flannick, J.; Teslovich, T.M.; Mahajan, A.; Agarwala, V.; Gaulton, K.J.; Ma, C.; Fontanillas, P.; Moutsianas, L.; McCarthy, D.J.; et al. The genetic architecture of type 2 diabetes. Nature 2016, 536, 41–47. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Lango, H.; Consortium, U.K.T.D.G.; Palmer, C.N.; Morris, A.D.; Zeggini, E.; Hattersley, A.T.; McCarthy, M.I.; Frayling, T.M.; Weedon, M.N. Assessing the combined impact of 18 common genetic variants of modest effect sizes on type 2 diabetes risk. Diabetes 2008, 57, 3129–3135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Flannick, J.; Mercader, J.M.; Fuchsberger, C.; Udler, M.S.; Mahajan, A.; Wessel, J.; Teslovich, T.M.; Caulkins, L.; Koesterer, R.; Barajas-Olmos, F.; et al. Exome sequencing of 20,791 cases of type 2 diabetes and 24,440 controls. Nature 2019, 570, 71–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Voight, B.F.; Scott, L.J.; Steinthorsdottir, V.; Morris, A.P.; Dina, C.; Welch, R.P.; Zeggini, E.; Huth, C.; Aulchenko, Y.S.; Thorleifsson, G.; et al. Twelve type 2 diabetes susceptibility loci identified through large-scale association analysis. Nat. Genet. 2010, 42, 579–589. [Google Scholar] [CrossRef] [PubMed]
  83. Mahajan, A.; Go, M.J.; Zhang, W.; Below, J.E.; Gaulton, K.J.; Ferreira, T.; Horikoshi, M.; Johnson, A.D.; Ng, M.C.; Prokopenko, I.; et al. Genome-wide trans-ancestry meta-analysis provides insight into the genetic architecture of type 2 diabetes susceptibility. Nat. Genet. 2014, 46, 234–244. [Google Scholar] [CrossRef]
  84. Imamura, M.; Takahashi, A.; Yamauchi, T.; Hara, K.; Yasuda, K.; Grarup, N.; Zhao, W.; Wang, X.; Huerta-Chagoya, A.; Hu, C.; et al. Genome-wide association studies in the Japanese population identify seven novel loci for type 2 diabetes. Nat. Commun. 2016, 7, 10531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Gaulton, K.J.; Ferreira, T.; Lee, Y.; Raimondo, A.; Magi, R.; Reschen, M.E.; Mahajan, A.; Locke, A.; Rayner, N.W.; Robertson, N.; et al. Genetic fine mapping and genomic annotation defines causal mechanisms at type 2 diabetes susceptibility loci. Nat. Genet. 2015, 47, 1415–1425. [Google Scholar] [CrossRef] [PubMed]
  86. Locke, J.M.; Saint-Martin, C.; Laver, T.W.; Patel, K.A.; Wood, A.R.; Sharp, S.A.; Ellard, S.; Bellanne-Chantelot, C.; Hattersley, A.T.; Harries, L.W.; et al. The Common HNF1A Variant I27L Is a Modifier of Age at Diabetes Diagnosis in Individuals with HNF1A-MODY. Diabetes 2018, 67, 1903–1907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Hegele, R.A.; Cao, H.; Harris, S.B.; Hanley, A.J.; Zinman, B. The hepatic nuclear factor-1alpha G319S variant is associated with early-onset type 2 diabetes in Canadian Oji-Cree. J. Clin. Endocrinol. Metab. 1999, 84, 1077–1082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Holmkvist, J.; Cervin, C.; Lyssenko, V.; Winckler, W.; Anevski, D.; Cilio, C.; Almgren, P.; Berglund, G.; Nilsson, P.; Tuomi, T.; et al. Common variants in HNF-1 alpha and risk of type 2 diabetes. Diabetologia 2006, 49, 2882–2891. [Google Scholar] [CrossRef] [Green Version]
  89. Consortium, S.T.D.; Estrada, K.; Aukrust, I.; Bjorkhaug, L.; Burtt, N.P.; Mercader, J.M.; Garcia-Ortiz, H.; Huerta-Chagoya, A.; Moreno-Macias, H.; Walford, G.; et al. Association of a low-frequency variant in HNF1A with type 2 diabetes in a Latino population. JAMA 2014, 311, 2305–2314. [Google Scholar] [CrossRef] [Green Version]
  90. Morris, A.P.; Voight, B.F.; Teslovich, T.M.; Ferreira, T.; Segre, A.V.; Steinthorsdottir, V.; Strawbridge, R.J.; Khan, H.; Grallert, H.; Mahajan, A.; et al. Large-scale association analysis provides insights into the genetic architecture and pathophysiology of type 2 diabetes. Nat. Genet. 2012, 44, 981–990. [Google Scholar] [CrossRef] [PubMed]
  91. Najmi, L.A.; Aukrust, I.; Flannick, J.; Molnes, J.; Burtt, N.; Molven, A.; Groop, L.; Altshuler, D.; Johansson, S.; Bjørkhaug, L.; et al. Functional Investigations of HNF1A Identify Rare Variants as Risk Factors for Type 2 Diabetes in the General Population. Diabetes 2017, 66, 335–346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Winckler, W.; Burtt, N.P.; Holmkvist, J.; Cervin, C.; de Bakker, P.I.; Sun, M.; Almgren, P.; Tuomi, T.; Gaudet, D.; Hudson, T.J.; et al. Association of common variation in the HNF1alpha gene region with risk of type 2 diabetes. Diabetes 2005, 54, 2336–2342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Flannick, J.; Beer, N.L.; Bick, A.G.; Agarwala, V.; Molnes, J.; Gupta, N.; Burtt, N.P.; Florez, J.C.; Meigs, J.B.; Taylor, H.; et al. Assessing the phenotypic effects in the general population of rare variants in genes for a dominant Mendelian form of diabetes. Nat. Genet. 2013, 45, 1380–1385. [Google Scholar] [CrossRef] [PubMed]
  94. Stride, A.; Ellard, S.; Clark, P.; Shakespeare, L.; Salzmann, M.; Shepherd, M.; Hattersley, A.T. Beta-cell dysfunction, insulin sensitivity, and glycosuria precede diabetes in hepatocyte nuclear factor-1alpha mutation carriers. Diabetes Care 2005, 28, 1751–1756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Mahajan, A.; Wessel, J.; Willems, S.M.; Zhao, W.; Robertson, N.R.; Chu, A.Y.; Gan, W.; Kitajima, H.; Taliun, D.; Rayner, N.W.; et al. Refining the accuracy of validated target identification through coding variant fine-mapping in type 2 diabetes. Nat. Genet. 2018, 50, 559–571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Servitja, J.M.; Pignatelli, M.; Maestro, M.A.; Cardalda, C.; Boj, S.F.; Lozano, J.; Blanco, E.; Lafuente, A.; McCarthy, M.I.; Sumoy, L.; et al. Hnf1alpha (MODY3) controls tissue-specific transcriptional programs and exerts opposed effects on cell growth in pancreatic islets and liver. Mol. Cell. Biol. 2009, 29, 2945–2959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Lehto, M.; Tuomi, T.; Mahtani, M.M.; Widen, E.; Forsblom, C.; Sarelin, L.; Gullstrom, M.; Isomaa, B.; Lehtovirta, M.; Hyrkko, A.; et al. Characterization of the MODY3 phenotype. Early-onset diabetes caused by an insulin secretion defect. J. Clin. Investig. 1997, 99, 582–591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Byrne, M.M.; Sturis, J.; Menzel, S.; Yamagata, K.; Fajans, S.S.; Dronsfield, M.J.; Bain, S.C.; Hattersley, A.T.; Velho, G.; Froguel, P.; et al. Altered insulin secretory responses to glucose in diabetic and nondiabetic subjects with mutations in the diabetes susceptibility gene MODY3 on chromosome 12. Diabetes 1996, 45, 1503–1510. [Google Scholar] [CrossRef] [PubMed]
  99. Tanizawa, Y.; Ohta, Y.; Nomiyama, J.; Matsuda, K.; Tanabe, K.; Inoue, H.; Matsutani, A.; Okuya, S.; Oka, Y. Overexpression of dominant negative mutant hepatocyte nuclear factor (HNF)-1alpha inhibits arginine-induced insulin secretion in MIN6 cells. Diabetologia 1999, 42, 887–891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Low, B.S.J.; Lim, C.S.; Ding, S.S.L.; Tan, Y.S.; Ng, N.H.J.; Krishnan, V.G.; Ang, S.F.; Neo, C.W.Y.; Verma, C.S.; Hoon, S.; et al. Decreased GLUT2 and glucose uptake contribute to insulin secretion defects in MODY3/HNF1A hiPSC-derived mutant beta cells. Nat. Commun. 2021, 12, 3133. [Google Scholar] [CrossRef] [PubMed]
  101. Haliyur, R.; Tong, X.; Sanyoura, M.; Shrestha, S.; Lindner, J.; Saunders, D.C.; Aramandla, R.; Poffenberger, G.; Redick, S.D.; Bottino, R.; et al. Human islets expressing HNF1A variant have defective beta cell transcriptional regulatory networks. J. Clin. Investig. 2019, 129, 246–251. [Google Scholar] [CrossRef] [Green Version]
  102. Dukes, I.D.; Sreenan, S.; Roe, M.W.; Levisetti, M.; Zhou, Y.P.; Ostrega, D.; Bell, G.I.; Pontoglio, M.; Yaniv, M.; Philipson, L.; et al. Defective pancreatic beta-cell glycolytic signaling in hepatocyte nuclear factor-1alpha-deficient mice. J. Biol. Chem. 1998, 273, 24457–24464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Yamagata, K. Roles of HNF1α and HNF4α in pancreatic β-cells: Lessons from a monogenic form of diabetes (MODY). Vitam. Horm. 2014, 95, 407–423. [Google Scholar] [CrossRef] [PubMed]
  104. Shih, D.Q.; Screenan, S.; Munoz, K.N.; Philipson, L.; Pontoglio, M.; Yaniv, M.; Polonsky, K.S.; Stoffel, M. Loss of HNF-1alpha function in mice leads to abnormal expression of genes involved in pancreatic islet development and metabolism. Diabetes 2001, 50, 2472–2480. [Google Scholar] [CrossRef] [Green Version]
  105. Parrizas, M.; Maestro, M.A.; Boj, S.F.; Paniagua, A.; Casamitjana, R.; Gomis, R.; Rivera, F.; Ferrer, J. Hepatic nuclear factor 1-alpha directs nucleosomal hyperacetylation to its tissue-specific transcriptional targets. Mol. Cell. Biol. 2001, 21, 3234–3243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Wang, H.; Maechler, P.; Hagenfeldt, K.A.; Wollheim, C.B. Dominant-negative suppression of HNF-1alpha function results in defective insulin gene transcription and impaired metabolism-secretion coupling in a pancreatic beta-cell line. EMBO J. 1998, 17, 6701–6713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Wang, H.; Antinozzi, P.A.; Hagenfeldt, K.A.; Maechler, P.; Wollheim, C.B. Molecular targets of a human HNF1 alpha mutation responsible for pancreatic beta-cell dysfunction. EMBO J. 2000, 19, 4257–4264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Cardenas-Diaz, F.L.; Osorio-Quintero, C.; Diaz-Miranda, M.A.; Kishore, S.; Leavens, K.; Jobaliya, C.; Stanescu, D.; Ortiz-Gonzalez, X.; Yoon, C.; Chen, C.S.; et al. Modeling Monogenic Diabetes using Human ESCs Reveals Developmental and Metabolic Deficiencies Caused by Mutations in HNF1A. Cell Stem Cell 2019, 25, 273–289.e5. [Google Scholar] [CrossRef] [PubMed]
  109. Akpinar, P.; Kuwajima, S.; Krutzfeldt, J.; Stoffel, M. Tmem27: A cleaved and shed plasma membrane protein that stimulates pancreatic beta cell proliferation. Cell Metab. 2005, 2, 385–397. [Google Scholar] [CrossRef] [Green Version]
  110. Fukui, K.; Yang, Q.; Cao, Y.; Takahashi, N.; Hatakeyama, H.; Wang, H.; Wada, J.; Zhang, Y.; Marselli, L.; Nammo, T.; et al. The HNF-1 target collectrin controls insulin exocytosis by SNARE complex formation. Cell Metab. 2005, 2, 373–384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Ohki, T.; Sato, Y.; Yoshizawa, T.; Yamamura, K.; Yamada, K.; Yamagata, K. Identification of hepatocyte growth factor activator (Hgfac) gene as a target of HNF1alpha in mouse beta-cells. Biochem. Biophys. Res. Commun. 2012, 425, 619–624. [Google Scholar] [CrossRef] [PubMed]
  112. Miyachi, Y.; Kuo, T.; Son, J.; Accili, D. Aldo-ketoreductase 1c19 ablation does not affect insulin secretion in murine islets. PLoS ONE 2021, 16, e0260526. [Google Scholar] [CrossRef] [PubMed]
  113. Parast, L.; Mathews, M.; Friedberg, M.W. Dynamic risk prediction for diabetes using biomarker change measurements. BMC Med. Res. Methodol. 2019, 19, 175. [Google Scholar] [CrossRef] [Green Version]
  114. Wang, Y.; Zhang, L.; Niu, M.; Li, R.; Tu, R.; Liu, X.; Hou, J.; Mao, Z.; Wang, Z.; Wang, C. Genetic Risk Score Increased Discriminant Efficiency of Predictive Models for Type 2 Diabetes Mellitus Using Machine Learning: Cohort Study. Front. Public Health 2021, 9, 606711. [Google Scholar] [CrossRef] [PubMed]
  115. Lall, K.; Magi, R.; Morris, A.; Metspalu, A.; Fischer, K. Personalized risk prediction for type 2 diabetes: The potential of genetic risk scores. Genet. Med. 2017, 19, 322–329. [Google Scholar] [CrossRef] [Green Version]
  116. Kwak, S.H.; Park, K.S. Recent progress in genetic and epigenetic research on type 2 diabetes. Exp. Mol. Med. 2016, 48, e220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Xu, F.; Liu, J.; Na, L.; Chen, L. Roles of Epigenetic Modifications in the Differentiation and Function of Pancreatic beta-Cells. Front. Cell Dev. Biol. 2020, 8, 748. [Google Scholar] [CrossRef]
  118. Ling, C.; Groop, L. Epigenetics: A molecular link between environmental factors and type 2 diabetes. Diabetes 2009, 58, 2718–2725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Lenoir, O.; Flosseau, K.; Ma, F.X.; Blondeau, B.; Mai, A.; Bassel-Duby, R.; Ravassard, P.; Olson, E.N.; Haumaitre, C.; Scharfmann, R. Specific control of pancreatic endocrine β- and δ-cell mass by class IIa histone deacetylases HDAC4, HDAC5, and HDAC9. Diabetes 2011, 60, 2861–2871. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Soutoglou, E.; Papafotiou, G.; Katrakili, N.; Talianidis, I. Transcriptional activation by hepatocyte nuclear factor-1 requires synergism between multiple coactivator proteins. J. Biol. Chem. 2000, 275, 12515–12520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Ban, N.; Yamada, Y.; Someya, Y.; Miyawaki, K.; Ihara, Y.; Hosokawa, M.; Toyokuni, S.; Tsuda, K.; Seino, Y. Hepatocyte nuclear factor-1alpha recruits the transcriptional co-activator p300 on the GLUT2 gene promoter. Diabetes 2002, 51, 1409–1418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Wong, C.K.; Wade-Vallance, A.K.; Luciani, D.S.; Brindle, P.K.; Lynn, F.C.; Gibson, W.T. The p300 and CBP Transcriptional Coactivators Are Required for beta-Cell and alpha-Cell Proliferation. Diabetes 2018, 67, 412–422. [Google Scholar] [CrossRef] [Green Version]
  123. Butler, A.E.; Janson, J.; Bonner-Weir, S.; Ritzel, R.; Rizza, R.A.; Butler, P.C. Beta-cell deficit and increased beta-cell apoptosis in humans with type 2 diabetes. Diabetes 2003, 52, 102–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Rahier, J.; Guiot, Y.; Goebbels, R.M.; Sempoux, C.; Henquin, J.C. Pancreatic beta-cell mass in European subjects with type 2 diabetes. Diabetes Obes. Metab. 2008, 10 (Suppl. S4), 32–42. [Google Scholar] [CrossRef] [PubMed]
  125. Hanley, S.C.; Austin, E.; Assouline-Thomas, B.; Kapeluto, J.; Blaichman, J.; Moosavi, M.; Petropavlovskaia, M.; Rosenberg, L. {beta}-Cell mass dynamics and islet cell plasticity in human type 2 diabetes. Endocrinology 2010, 151, 1462–1472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Yamagata, K.; Nammo, T.; Moriwaki, M.; Ihara, A.; Iizuka, K.; Yang, Q.; Satoh, T.; Li, M.; Uenaka, R.; Okita, K.; et al. Overexpression of dominant-negative mutant hepatocyte nuclear fctor-1 alpha in pancreatic beta-cells causes abnormal islet architecture with decreased expression of E-cadherin, reduced beta-cell proliferation, and diabetes. Diabetes 2002, 51, 114–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Kirkpatrick, C.L.; Wiederkehr, A.; Baquie, M.; Akhmedov, D.; Wang, H.; Gauthier, B.R.; Akerman, I.; Ishihara, H.; Ferrer, J.; Wollheim, C.B. Hepatic nuclear factor 1alpha (HNF1alpha) dysfunction down-regulates X-box-binding protein 1 (XBP1) and sensitizes beta-cells to endoplasmic reticulum stress. J. Biol. Chem. 2011, 286, 32300–32312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Wobser, H.; Dussmann, H.; Kogel, D.; Wang, H.; Reimertz, C.; Wollheim, C.B.; Byrne, M.M.; Prehn, J.H. Dominant-negative suppression of HNF-1 alpha results in mitochondrial dysfunction, INS-1 cell apoptosis, and increased sensitivity to ceramide-, but not to high glucose-induced cell death. J. Biol. Chem. 2002, 277, 6413–6421. [Google Scholar] [CrossRef] [Green Version]
  129. Perl, S.; Kushner, J.A.; Buchholz, B.A.; Meeker, A.K.; Stein, G.M.; Hsieh, M.; Kirby, M.; Pechhold, S.; Liu, E.H.; Harlan, D.M.; et al. Significant human beta-cell turnover is limited to the first three decades of life as determined by in vivo thymidine analog incorporation and radiocarbon dating. J. Clin. Endocrinol. Metab. 2010, 95, E234–E239. [Google Scholar] [CrossRef] [Green Version]
  130. Linnemann, A.K.; Baan, M.; Davis, D.B. Pancreatic beta-cell proliferation in obesity. Adv. Nutr. 2014, 5, 278–288. [Google Scholar] [CrossRef] [Green Version]
  131. Georgia, S.; Bhushan, A. Beta cell replication is the primary mechanism for maintaining postnatal beta cell mass. J. Clin. Investig. 2004, 114, 963–968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Dor, Y.; Brown, J.; Martinez, O.I.; Melton, D.A. Adult pancreatic beta-cells are formed by self-duplication rather than stem-cell differentiation. Nature 2004, 429, 41–46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Teta, M.; Rankin, M.M.; Long, S.Y.; Stein, G.M.; Kushner, J.A. Growth and regeneration of adult beta cells does not involve specialized progenitors. Dev. Cell 2007, 12, 817–826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Aguayo-Mazzucato, C.; Bonner-Weir, S. Pancreatic beta Cell Regeneration as a Possible Therapy for Diabetes. Cell Metab. 2018, 27, 57–67. [Google Scholar] [CrossRef] [Green Version]
  135. Yoneda, S.; Uno, S.; Iwahashi, H.; Fujita, Y.; Yoshikawa, A.; Kozawa, J.; Okita, K.; Takiuchi, D.; Eguchi, H.; Nagano, H.; et al. Predominance of beta-cell neogenesis rather than replication in humans with an impaired glucose tolerance and newly diagnosed diabetes. J. Clin. Endocrinol. Metab. 2013, 98, 2053–2061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Houbracken, I.; de Waele, E.; Lardon, J.; Ling, Z.; Heimberg, H.; Rooman, I.; Bouwens, L. Lineage tracing evidence for transdifferentiation of acinar to duct cells and plasticity of human pancreas. Gastroenterology 2011, 141, 731–741. [Google Scholar] [CrossRef] [PubMed]
  137. Meier, J.J. Beta cell mass in diabetes: A realistic therapeutic target? Diabetologia 2008, 51, 703–713. [Google Scholar] [CrossRef] [PubMed]
  138. Fukuda, T.; Bouchi, R.; Takeuchi, T.; Amo-Shiinoki, K.; Kudo, A.; Tanaka, S.; Tanabe, M.; Akashi, T.; Hirayama, K.; Odamaki, T.; et al. Importance of Intestinal Environment and Cellular Plasticity of Islets in the Development of Postpancreatectomy Diabetes. Diabetes Care 2021, 44, 1002–1011. [Google Scholar] [CrossRef] [PubMed]
  139. Yoon, K.H.; Ko, S.H.; Cho, J.H.; Lee, J.M.; Ahn, Y.B.; Song, K.H.; Yoo, S.J.; Kang, M.I.; Cha, B.Y.; Lee, K.W.; et al. Selective beta-cell loss and alpha-cell expansion in patients with type 2 diabetes mellitus in Korea. J. Clin. Endocrinol. Metab. 2003, 88, 2300–2308. [Google Scholar] [CrossRef] [PubMed]
  140. Accili, D.; Talchai, S.C.; Kim-Muller, J.Y.; Cinti, F.; Ishida, E.; Ordelheide, A.M.; Kuo, T.; Fan, J.; Son, J. When beta-cells fail: Lessons from dedifferentiation. Diabetes Obes. Metab. 2016, 18 (Suppl. S1), 117–122. [Google Scholar] [CrossRef] [PubMed]
  141. Talchai, C.; Xuan, S.; Lin, H.V.; Sussel, L.; Accili, D. Pancreatic beta cell dedifferentiation as a mechanism of diabetic beta cell failure. Cell 2012, 150, 1223–1234. [Google Scholar] [CrossRef] [Green Version]
  142. Clark, A.; Wells, C.A.; Buley, I.D.; Cruickshank, J.K.; Vanhegan, R.I.; Matthews, D.R.; Cooper, G.J.; Holman, R.R.; Turner, R.C. Islet amyloid, increased A-cells, reduced B-cells and exocrine fibrosis: Quantitative changes in the pancreas in type 2 diabetes. Diabetes Res. 1988, 9, 151–159. [Google Scholar] [PubMed]
  143. Amo-Shiinoki, K.; Tanabe, K.; Hoshii, Y.; Matsui, H.; Harano, R.; Fukuda, T.; Takeuchi, T.; Bouchi, R.; Takagi, T.; Hatanaka, M.; et al. Islet cell dedifferentiation is a pathologic mechanism of long-standing progression of type 2 diabetes. JCI Insight 2021, 6, e143791. [Google Scholar] [CrossRef]
  144. Deng, S.; Vatamaniuk, M.; Huang, X.; Doliba, N.; Lian, M.M.; Frank, A.; Velidedeoglu, E.; Desai, N.M.; Koeberlein, B.; Wolf, B.; et al. Structural and functional abnormalities in the islets isolated from type 2 diabetic subjects. Diabetes 2004, 53, 624–632. [Google Scholar] [CrossRef] [Green Version]
  145. Butler, A.E.; Dhawan, S.; Hoang, J.; Cory, M.; Zeng, K.; Fritsch, H.; Meier, J.J.; Rizza, R.A.; Butler, P.C. beta-Cell Deficit in Obese Type 2 Diabetes, a Minor Role of beta-Cell Dedifferentiation and Degranulation. J. Clin. Endocrinol. Metab. 2016, 101, 523–532. [Google Scholar] [CrossRef] [Green Version]
  146. Son, J.; Ding, H.; Farb, T.B.; Efanov, A.M.; Sun, J.; Gore, J.L.; Syed, S.K.; Lei, Z.; Wang, Q.; Accili, D.; et al. BACH2 inhibition reverses beta cell failure in type 2 diabetes models. J. Clin. Investig. 2021, 131, e153876. [Google Scholar] [CrossRef]
  147. Kim-Muller, J.Y.; Zhao, S.; Srivastava, S.; Mugabo, Y.; Noh, H.L.; Kim, Y.R.; Madiraju, S.R.; Ferrante, A.W.; Skolnik, E.Y.; Prentki, M.; et al. Metabolic inflexibility impairs insulin secretion and results in MODY-like diabetes in triple FoxO-deficient mice. Cell Metab. 2014, 20, 593–602. [Google Scholar] [CrossRef] [Green Version]
  148. Yabe, S.G.; Nishida, J.; Fukuda, S.; Takeda, F.; Nasiro, K.; Yasuda, K.; Iwasaki, N.; Okochi, H. Expression of mutant mRNA and protein in pancreatic cells derived from MODY3-iPS cells. PLoS ONE 2019, 14, e0217110. [Google Scholar] [CrossRef] [Green Version]
  149. Teo, A.K.; Windmueller, R.; Johansson, B.B.; Dirice, E.; Njolstad, P.R.; Tjora, E.; Raeder, H.; Kulkarni, R.N. Derivation of human induced pluripotent stem cells from patients with maturity onset diabetes of the young. J. Biol. Chem. 2013, 288, 5353–5356. [Google Scholar] [CrossRef] [Green Version]
  150. Stepniewski, J.; Kachamakova-Trojanowska, N.; Ogrocki, D.; Szopa, M.; Matlok, M.; Beilharz, M.; Dyduch, G.; Malecki, M.T.; Jozkowicz, A.; Dulak, J. Induced pluripotent stem cells as a model for diabetes investigation. Sci. Rep. 2015, 5, 8597. [Google Scholar] [CrossRef] [Green Version]
  151. González, B.J.; Zhao, H.; Niu, J.; Williams, D.J.; Lee, J.; Goulbourne, C.N.; Xing, Y.; Wang, Y.; Oberholzer, J.; Chen, X.; et al. Human stem cell model of HNF1A deficiency shows uncoupled insulin to C-peptide secretion with accumulation of abnormal insulin granules. bioRxiv 2021. [Google Scholar] [CrossRef]
  152. Pearson, E.R.; Starkey, B.J.; Powell, R.J.; Gribble, F.M.; Clark, P.M.; Hattersley, A.T. Genetic cause of hyperglycaemia and response to treatment in diabetes. Lancet 2003, 362, 1275–1281. [Google Scholar] [CrossRef]
  153. Ostoft, S.H.; Bagger, J.I.; Hansen, T.; Pedersen, O.; Faber, J.; Holst, J.J.; Knop, F.K.; Vilsboll, T. Glucose-lowering effects and low risk of hypoglycemia in patients with maturity-onset diabetes of the young when treated with a GLP-1 receptor agonist: A double-blind, randomized, crossover trial. Diabetes Care 2014, 37, 1797–1805. [Google Scholar] [CrossRef] [Green Version]
  154. Christensen, A.S.; Haedersdal, S.; Stoy, J.; Storgaard, H.; Kampmann, U.; Forman, J.L.; Seghieri, M.; Holst, J.J.; Hansen, T.; Knop, F.K.; et al. Efficacy and Safety of Glimepiride With or Without Linagliptin Treatment in Patients with HNF1A Diabetes (Maturity-Onset Diabetes of the Young Type 3): A Randomized, Double-Blinded, Placebo-Controlled, Crossover Trial (GLIMLINA). Diabetes Care 2020, 43, 2025–2033. [Google Scholar] [CrossRef] [PubMed]
  155. Fantasia, K.L.; Steenkamp, D.W. Optimal Glycemic Control in a Patient With HNF1A MODY With GLP-1 RA Monotherapy: Implications for Future Therapy. J. Endocr. Soc. 2019, 3, 2286–2289. [Google Scholar] [CrossRef] [Green Version]
  156. Lumb, A.N.; Gallen, I.W. Treatment of HNF1-alpha MODY with the DPP-4 inhibitor Sitagliptin(1). Diabet. Med. 2009, 26, 189–190. [Google Scholar] [CrossRef]
  157. Katra, B.; Klupa, T.; Skupien, J.; Szopa, M.; Nowak, N.; Borowiec, M.; Kozek, E.; Malecki, M.T. Dipeptidyl peptidase-IV inhibitors are efficient adjunct therapy in HNF1A maturity-onset diabetes of the young patients—Report of two cases. Diabetes Technol. Ther. 2010, 12, 313–316. [Google Scholar] [CrossRef]
  158. Christensen, A.S.; Haedersdal, S.; Storgaard, H.; Rose, K.; Hansen, N.L.; Holst, J.J.; Hansen, T.; Knop, F.K.; Vilsboll, T. GIP and GLP-1 Potentiate Sulfonylurea-Induced Insulin Secretion in Hepatocyte Nuclear Factor 1alpha Mutation Carriers. Diabetes 2020, 69, 1989–2002. [Google Scholar] [CrossRef] [PubMed]
  159. Hagenfeldt-Johansson, K.A.; Herrera, P.L.; Wang, H.; Gjinovci, A.; Ishihara, H.; Wollheim, C.B. Beta-cell-targeted expression of a dominant-negative hepatocyte nuclear factor-1 alpha induces a maturity-onset diabetes of the young (MODY)3-like phenotype in transgenic mice. Endocrinology 2001, 142, 5311–5320. [Google Scholar] [CrossRef]
Figure 1. Physiological role of HNF1A in the liver, pancreas, kidneys, and intestine.
Figure 1. Physiological role of HNF1A in the liver, pancreas, kidneys, and intestine.
Ijms 23 03222 g001
Figure 2. Schematic representation of beta cell dysfunction caused by HNF1A mutations.
Figure 2. Schematic representation of beta cell dysfunction caused by HNF1A mutations.
Ijms 23 03222 g002
Table 1. Comparison of the roles of HNF1A variants in MODY3 and T2DM. hs-CRP, high-sensitivity CRP; SU, sulfonylureas.
Table 1. Comparison of the roles of HNF1A variants in MODY3 and T2DM. hs-CRP, high-sensitivity CRP; SU, sulfonylureas.
HNF1A VariantsTypes of VariantsInsulin Sensitivity Beta Cell Fate Sensitivity to SU
MODY3Pathogenic roleMissense [73,74] and deletion [70,71,72,73,74] mutationsLow hs-CRP levels [28,29]
Low renal threshold for glucose [34,35]
Reduced insulin secretion [97,98]
Dedifferentiation [108,151]
High [152,153]
T2DMDisease riskSNPs [82,83,84,85,86,87,88,89,90]Not reportedCertain SNPs (e.g., coding variants) may affect insulin secretion [91,92]Not reported
Table 2. Future research directions.
Table 2. Future research directions.
AimsObjectives
Functional analysis of HNF1A variantsMolecular mechanism
Therapeutic targets
Mainly use human stem cell-derived beta cells with HNF1A variants.
Insulin secretionAssess insulin secretion, insulin granules, and mitochondrial metabolism in beta cells.
EpigeneticsInvestigate changes in histone modification, DNA methylation, and non-coding RNA in beta cells.
Beta cell dedifferentiationConfirm a decrease in beta cell mass and an increase in alpha cell mass in the pancreatic tissue harboring HNF1A variants.
Extrapancreatic organsMeasure hs-CRP levels and urinary glucose reabsorption in individuals with HNF1A SNPs.
Treatment for patients with HNF1A variantsPersonalized medicineAssess the sensitivity to SU using iPS cell-derived beta cells from individuals with HNF1A SNPs.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Miyachi, Y.; Miyazawa, T.; Ogawa, Y. HNF1A Mutations and Beta Cell Dysfunction in Diabetes. Int. J. Mol. Sci. 2022, 23, 3222. https://doi.org/10.3390/ijms23063222

AMA Style

Miyachi Y, Miyazawa T, Ogawa Y. HNF1A Mutations and Beta Cell Dysfunction in Diabetes. International Journal of Molecular Sciences. 2022; 23(6):3222. https://doi.org/10.3390/ijms23063222

Chicago/Turabian Style

Miyachi, Yasutaka, Takashi Miyazawa, and Yoshihiro Ogawa. 2022. "HNF1A Mutations and Beta Cell Dysfunction in Diabetes" International Journal of Molecular Sciences 23, no. 6: 3222. https://doi.org/10.3390/ijms23063222

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