Loss of functional β-cell mass is an essential feature of type 2 diabetes, and maintaining mature β-cell identity is important for preserving a functional β-cell mass. However, it is unclear how β-cells achieve and maintain their mature identity. Here we demonstrate a novel function of protein arginine methyltransferase 1 (PRMT1) in maintaining mature β-cell identity. Prmt1 knockout in fetal and adult β-cells induced diabetes, which was aggravated by high-fat diet–induced metabolic stress. Deletion of Prmt1 in adult β-cells resulted in the immediate loss of histone H4 arginine 3 asymmetric dimethylation (H4R3me2a) and the subsequent loss of β-cell identity. The expression levels of genes involved in mature β-cell function and identity were robustly downregulated as soon as Prmt1 deletion was induced in adult β-cells. Chromatin immunoprecipitation sequencing and assay for transposase-accessible chromatin sequencing analyses revealed that PRMT1-dependent H4R3me2a increases chromatin accessibility at the binding sites for CCCTC-binding factor (CTCF) and β-cell transcription factors. In addition, PRMT1-dependent open chromatin regions may show an association with the risk of diabetes in humans. Together, our results indicate that PRMT1 plays an essential role in maintaining β-cell identity by regulating chromatin accessibility.

Maintaining the functional β-cell mass is crucial for preventing diabetes, which develops when β-cells fail to meet the insulin demand (1,2). Although β-cell death is thought to be the major mechanism of β-cell failure (3), recent studies indicate that β-cell dedifferentiation can decrease the functional β-cell mass and thereby deteriorate systemic glucose homeostasis (4,5). Maintaining mature β-cell identity is also important for maintaining β-cell function (6,7). A hierarchy of transcription factor (TF) cascades directs β-cell differentiation, and β-cells require continuous activation of these TFs to maintain their function and identity (810). The genetic identity of a differentiated cell is generally controlled by the chromatin state, which is overall stable and has limited epigenomic flexibility (11,12). Likewise, epigenetic regulation plays an essential role in the postnatal maturation of β-cells and the maintenance of mature β-cell identity (1316).

Histone arginine methylation, which is regulated by protein arginine methyltransferase (PRMT), can affect chromatin structures to facilitate the recruitment of protein complexes that regulate gene transcription (17,18). PRMT4-dependent histone H3 arginine 17 asymmetric dimethylation (H3R17me2a) in β-cells has been reported to regulate glucose-stimulated insulin secretion (GSIS) (19). However, the role of PRMT-induced histone arginine methylation in regulating β-cell identity has not yet been elucidated. Among the nine members of the PRMT family, PRMT1 predominates in mammalian cells (20). It appears to be associated with diabetes, as its catalytic activity is decreased in the liver and pancreas of diabetic Goto-Kakizaki rats (21). PRMT1 has also been shown to specifically induce the active histone code, histone H4 arginine 3 asymmetric dimethylation (H4R3me2a), which potentiates subsequent histone acetylation and contributes to establishing euchromatin structure (22,23). Based on these previous findings, we herein explored the role of PRMT1-dependent H4R3me2a in mature β-cells.

Animals

Prmt1-floxed (Prmt1fl/fl) (Mouse Genome Informatics [MGI]: 4432476) mice were crossed with Rip2-Cre (MGI: 2387567) and Pdx1-CreERT2 (MGI: 2684321) mice to generate Prmt1 β-cell knockout (βKO) and inducible β-cell–specific Prmt1 KO (Prmt1 βiKO) mice, respectively. R26-eYFP (MGI: 2449038) mice were crossed for lineage-tracing experiments and β-cell sorting. All mice were backcrossed and maintained on a C57BL/6J background. Cre recombination for CreERT2 was induced by a total of five intraperitoneal injections of corn oil–dissolved tamoxifen (75 mg/kg) over 2 weeks. Mice were housed in climate-controlled, specific pathogen-free barrier facilities under a 12-h light/dark cycle, and chow and water were provided ad libitum. Mice were fed either a standard chow diet or high-fat diet (HFD) (60% kcal fat). The animal experiment protocols for this study were approved by the Institutional Animal Care and Use Committee at the Korea Advanced Institute of Science and Technology. All experiments were performed in accordance with the relevant guidelines and regulations.

Metabolic Assays

Body weight and random blood glucose levels were measured in the afternoon hours each day. The glucose tolerance test and insulin tolerance test were performed as previously described (24).

Histological Analyses

For histological analyses, formalin-fixed paraffin-embedded pancreatic slides were prepared, stained, and analyzed as described in Supplementary Data.

Pancreatic Insulin Content

Pancreatic tissues were dissected, placed in acid-ethanol (1.5% HCl in 70% ethanol), homogenized, and incubated at 4°C for 16 h. The aqueous phase of pancreatic insulin extract was neutralized with an equal amount of 1 mol/L Tris-Cl buffer (pH 7.5). The pancreatic insulin content was calculated by dividing the total pancreatic insulin by the weight of the pancreas.

GSIS

For the in vivo GSIS assay, mice were fasted for 16 h and then given an intraperitoneal injection of d-glucose in PBS (2 g/kg). For the ex vivo islet GSIS assay, pancreatic islets were isolated from mice as described previously (25), and the assay was performed as described in Supplementary Data.

Oxygen Consumption Rate

Pancreatic islets were isolated from mice as described previously (25), and the oxygen consumption rate (OCR) assay was performed as described in Supplementary Data.

Quantitative RT-PCR

Total RNA was extracted from mouse tissues, and quantitative RT-PCR (qRT-PCR) was performed as described in Supplementary Data. The sequences of the primers used are listed in Supplementary Table 1.

Chromatin Immunoprecipitation Sequencing, RNA Sequencing, and Assay for Transposase-Accessible Chromatin Sequencing Analyses

Chromatin immunoprecipitation (ChIP) experiments were performed in MIN6 cells as previously described (26) with modifications. RNA sequencing (RNA-seq) experiments were performed using wild-type (WT) and Prmt1-null islets. Assay for transposase-accessible chromatin (ATAC) experiments were performed as previously described (27), using MIN6 cells and FACS WT and Prmt1-null β-cells. ChIP sequencing (ChIP-seq), RNA-seq, and ATAC sequencing (ATAC-seq) analyses were performed as described in Supplementary Data.

Chromosome Conformation Capture PCR

Chromosome conformation capture (3C) experiments were performed in MIN6 cells as previously described (28) with modifications. The data were normalized with respect to those obtained using internal primers that recognized sequences within the Gapdh gene. At least three independent biological replicates were included for each 3C-PCR assay. The sequences of the primers used are listed in Supplementary Table 1.

Statistics

All values are expressed as mean ± SEM. The two-tailed Student t test or one-way ANOVA followed by post hoc Tukey test was used to compare groups. P values <0.05 were considered statistically significant.

Data and Resource Availability

The ChIP-seq, RNA-seq, and ATAC-seq data that support the findings of this study have been deposited in the National Center for Biotechnology Information Gene Expression Omnibus (GEO) under accession code GSE117100. All data that support the findings of this study are available from the authors on reasonable request. No applicable resources were generated or analyzed during the current study.

Prmt1 βKO Mice Develop Progressive Glucose Intolerance

We first checked the gene expression of the Prmt family genes in pancreatic islets and confirmed that Prmt1 exhibited the highest expression level among them (Supplementary Fig. 1A and B). The expression level of Prmt1 was higher in pancreatic islets than in liver or brain, and PRMT1 and H4R3me2a were enriched in pancreatic islets of both mice and humans (Supplementary Fig. 1C and D). To test the possible role of H4R3me2a in β-cells, we performed ChIP-seq for H4R3me2a and ATAC-seq in MIN6 cells. Motif analysis of the H4R3me2a ChIP-seq data suggested that H4R3me2a was significantly associated with the β-cell TFs: MAFA, NEUROD1, and FOXA1 (Supplementary Fig. 1E). Intriguingly, the most significantly associated TF was CCCTC-binding factor (CTCF), which is known to play a crucial role in regulating the chromatin architecture (29,30). In order to assess the association of H4R3me2a with CTCF and β-cell TF, we performed ChIP-seq for CTCF in MIN6 cells. Analyses of ChIP-seq data obtained for H4R3me2a and CTCF, combined with publicly available ChIP-seq data for MAFA, indicated that CTCF and MAFA bind near H4R3me2a-occupied chromatin regions (Supplementary Fig. 1F). The association of H4R3me2a with CTCF and MAFA, together with the enrichment of PRMT1 and H4R3me2a in adult β-cells, suggests that PRMT1 and H4R3me2a may play roles in β-cells.

To further investigate the physiological role of PRMT1 and H4R3me2a in β-cells, we generated β-cell–specific Prmt1 KO (Rip2-Cre; Prmt1fl/fl, herein called Prmt1 βKO) mice. Immunofluorescence staining confirmed the deletion of PRMT1 at postnatal day 7 and the subsequent removal of H4R3me2a in the β-cells of these mice around weaning at 3 weeks of age (Supplementary Fig. 2). In WT control mice, the fluorescence signals of both PRMT1 and H4R3me2a became enriched in pancreatic islets after weaning at 3 weeks of age, when the β-cells become mature. In Prmt1 βKO mice, PRMT1 was nearly undetectable at postnatal day 7, whereas H4R3me2a remained detectable in a substantial number of β-cells until 3 weeks of age, suggesting the more important role of PRMT1-dependent H4R3me2a in mature β-cells. Indeed, the pancreatic islets of Prmt1 βKO mice developed normally and did not show any abnormality in the markers of β-cell development (Supplementary Fig. 3A and B). Prmt1 βKO mice grew normally and showed normal glucose tolerance until they developed glucose intolerance at 12 weeks of age (Fig. 1A and B and Supplementary Fig. 3C). Despite this glucose intolerance, Prmt1 βKO mice showed no defects in insulin sensitivity and insulin production (Supplementary Fig. 4A–C). Instead, GSIS was impaired in Prmt1 βKO islets (Fig. 1C). Basal insulin secretion was increased when Prmt1 βKO islets were treated with 2.8 mmol/L glucose, and those treated with 20 mmol/L glucose failed to show any further increase of insulin secretion. The impairment of GSIS was further confirmed by analyses of plasma insulin levels and mitochondrial OCRs in the isolated islets (Fig. 1D and E). To investigate whether a transcriptional change was responsible for the defect of GSIS in Prmt1 βKO mice, we performed RNA-seq analysis in the islets of Prmt1 βKO mice at 12 weeks of age (Supplementary Fig. 5A). Although there was no robust change in global gene expression, Prmt1 βKO islets exhibited downregulation of mature β-cell genes that are involved in GSIS and misexpression of genes that are disallowed to be expressed in mature β-cells (3133) (Supplementary Fig. 5B). These gene expression changes were further confirmed by qRT-PCR analysis (Supplementary Fig. 5C–F). Pathway analysis showed that most of the genes downregulated in Prmt1 βKO islets were involved in pancreas development and maturity-onset diabetes of the young (Supplementary Fig. 5G). This notion was further supported by electron microscopic analysis (Fig. 1F), which showed that Prmt1-null β-cells exhibited ultrastructural changes resembling those found in the β-cells of patients with type 2 diabetes (34,35). The volume and density of insulin granules were reduced, the endoplasmic reticulum was dilated, and the mitochondria appeared round and swollen. Condensed chromatin was observed in the nuclei of Prmt1-null β-cells (Fig. 1F), but apoptosis was not observed in the islets of Prmt1 βKO mice (Supplementary Fig. 5H). In addition, islet hormones and β-cell TFs were normally expressed in these mice (Supplementary Fig. 5I and J). These data indicate that PRMT1 is needed to maintain the function, not the survival, of β-cells.

Figure 1

Prmt1 βKO mice develop a progressive diabetes phenotype. A: Intraperitoneal glucose tolerance test (IPGTT) of 6-week-old male control and Prmt1 βKO mice after a 16-h fast; n = 4/group. BF: Male control and Prmt1 βKO mice aged 12–13 weeks were fed a standard chow diet and used for experiments. B: IPGTT after a 16-h fasting; n = 5/group. C: Ex vivo islet GSIS assay; n = 4/group. D: In vivo GSIS assay after a 16-h fast; n = 3/group. E: OCR analysis of isolated islets; n = 6/group. F: Representative β-cell images obtained by transmission electron microscopy. Arrows indicate immature insulin granules (blue), dilated endoplasmic reticulum (green), and dysmorphic mitochondria (red); n = 3/group. GO: Male control and Prmt1 βKO mice (8 weeks old) were fed HFD for 18 weeks and used for experiments. G: IPGTT after a 16-h fast; n = 5/group. H: Representative islet images obtained by immunofluorescence (IF) staining of INS (green), GCG (red), and DAPI (blue) from HFD-fed 26-week-old Prmt1 βKO mice; n = 3/group. I: Representative islet images obtained by IF staining of eYFP (green), INS (blue), and GCG, SST, or PPY (red) from HFD-fed 26-week-old R26-eYFP; Rip2-Cre (control) and R26-eYFP; Prmt1 βKO mice. White arrows indicate GCG/SST/PPY-expressing eYFP+ cells. Quantification analysis of eYFP-copositive cells expressing INS (J), GCG (K), SST (L), and PPY (M) in the islets of HFD-fed 26-week-old R26-eYFP; Rip2-Cre (WT) and R26-eYFP; Prmt1 βKO (KO) mice; n = 3/group. N: Representative islet images obtained by IF staining of INS (green) and PDX1, NKX6.1, MAFA, or SLC2A2 (red); n = 3/group. O: Representative islet images obtained by IF staining of INS (green), FOXO1 (red), and DAPI (blue); n = 3/group. Scale bars, 50 μm (H, I, N, and O); 2.5 μm (F). Prmt1fl/fl or Rip2-Cre mice were used as controls (AG, N, and O). Data are expressed as mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by Student t test (A, B, E, G, and JM) or one-way ANOVA with post hoc Tukey test (C and D). CCCP, carbonyl cyanide m-chlorophenylhydrazone; Glc, glucose; Olig., oligomycin, Rot., rotenone.

Figure 1

Prmt1 βKO mice develop a progressive diabetes phenotype. A: Intraperitoneal glucose tolerance test (IPGTT) of 6-week-old male control and Prmt1 βKO mice after a 16-h fast; n = 4/group. BF: Male control and Prmt1 βKO mice aged 12–13 weeks were fed a standard chow diet and used for experiments. B: IPGTT after a 16-h fasting; n = 5/group. C: Ex vivo islet GSIS assay; n = 4/group. D: In vivo GSIS assay after a 16-h fast; n = 3/group. E: OCR analysis of isolated islets; n = 6/group. F: Representative β-cell images obtained by transmission electron microscopy. Arrows indicate immature insulin granules (blue), dilated endoplasmic reticulum (green), and dysmorphic mitochondria (red); n = 3/group. GO: Male control and Prmt1 βKO mice (8 weeks old) were fed HFD for 18 weeks and used for experiments. G: IPGTT after a 16-h fast; n = 5/group. H: Representative islet images obtained by immunofluorescence (IF) staining of INS (green), GCG (red), and DAPI (blue) from HFD-fed 26-week-old Prmt1 βKO mice; n = 3/group. I: Representative islet images obtained by IF staining of eYFP (green), INS (blue), and GCG, SST, or PPY (red) from HFD-fed 26-week-old R26-eYFP; Rip2-Cre (control) and R26-eYFP; Prmt1 βKO mice. White arrows indicate GCG/SST/PPY-expressing eYFP+ cells. Quantification analysis of eYFP-copositive cells expressing INS (J), GCG (K), SST (L), and PPY (M) in the islets of HFD-fed 26-week-old R26-eYFP; Rip2-Cre (WT) and R26-eYFP; Prmt1 βKO (KO) mice; n = 3/group. N: Representative islet images obtained by IF staining of INS (green) and PDX1, NKX6.1, MAFA, or SLC2A2 (red); n = 3/group. O: Representative islet images obtained by IF staining of INS (green), FOXO1 (red), and DAPI (blue); n = 3/group. Scale bars, 50 μm (H, I, N, and O); 2.5 μm (F). Prmt1fl/fl or Rip2-Cre mice were used as controls (AG, N, and O). Data are expressed as mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by Student t test (A, B, E, G, and JM) or one-way ANOVA with post hoc Tukey test (C and D). CCCP, carbonyl cyanide m-chlorophenylhydrazone; Glc, glucose; Olig., oligomycin, Rot., rotenone.

Close modal

As the phenotypes of Prmt1 βKO mice resembled the early features of type 2 diabetes (3436), we fed Prmt1 βKO mice with an HFD from 8 weeks of age to test how these mice respond to metabolic stress (Supplementary Fig. 6A). HFD exacerbated the glucose intolerance in Prmt1 βKO mice without perturbing compensatory β-cell expansion (Fig. 1G and Supplementary Fig. 6B). Interestingly, the islets of HFD-fed Prmt1 βKO mice contained polyhormonal cells that coexpressed insulin and glucagon (GCG) (Fig. 1H). A lineage-tracing analysis revealed that these polyhormonal cells were originated from insulin-producing β-cells (GCG, ∼0.4%; somatostatin [SST], ∼2.4%; and pancreatic polypeptide [PPY], ∼0.19%), suggesting that the β-cells of HFD-fed Prmt1 βKO mice had undergone some changes in their differentiated states (Fig. 1I–M). Immunofluorescence staining of mature β-cell markers further confirmed the loss of mature β-cell identity in HFD-fed Prmt1 βKO mice; such cells showed loss of MAFA and SLC2A2, cytoplasmic localization of NKX6.1, and blockage of HFD-induced FOXO1 nuclear translocation (4,37,38) (Fig. 1N and O). These phenotypes of HFD-fed Prmt1 βKO mice suggest that PRMT1 plays an essential role in maintaining the mature β-cell identity.

PRMT1 Is Required for the Maintenance of β-Cell Identity

Although Prmt1 βKO mice presented the features of loss of β-cell identity, these phenotypes were weak. This could reflect the presence of metabolic compensation, which often comes into play in genetic KO models. To minimize the involvement of any compensatory mechanism and further confirm the role of PRMT1 in mature β-cells, we generated an inducible β-cell–specific Prmt1 KO mouse model by crossing Prmt1fl/fl mice with Pdx1 promoter-driven CreERT2 (Pdx1-CreERT2) mice (hereafter called Prmt1 βiKO). Prmt1 KO was induced in adult β-cells by intraperitoneally injecting the mice with tamoxifen (75 mg/kg) five times over 2 weeks, beginning at 6 weeks of age (Supplementary Fig. 7A). At 8 weeks of age, these mice exhibited deletion of PRMT1 in β-cells and subsequent removal of H4R3me2a, but maintained normoglycemia (Fig. 2A and Supplementary Fig. 7B and C). At 12 weeks of age, Prmt1 βiKO mice developed glucose intolerance and exhibited elevated random glucose levels due to impaired GSIS (Supplementary Fig. 7C–E). Thus, acute loss of PRMT1 in adult β-cells is sufficient to induce the loss of mature β-cell function.

Figure 2

PRMT1 is required for the maintenance of mature β-cell identity. A: Representative islet images obtained by immunofluorescence staining of INS (green) and PRMT1, H4R3me2a, PDX1, NKX6.1, MAFA, or SLC2A2 (red) in 8- and 12-week-old control and Prmt1 βiKO mice; n = 3/group. B: Representative islet images obtained by immunofluorescence staining of eYFP (green), INS (blue), and GCG, SST, PPY, and UCN3 (red) from 8- and 12-week-old R26-eYFP; Pdx1-CreERT2 (control) and R26-eYFP; Prmt1 βiKO mice; n = 3/group. White arrows indicate GCG/SST/PPY-expressing or UCN3 eYFP+ cells. Quantification analysis of eYFP-copositive cells expressing GCG (C), SST (D), PPY (E), INS (F), and UCN3 (G) in the islets of 8- and 12-week-old R26-eYFP; Pdx1-CreERT2 (WT) and R26-eYFP; Prmt1 βiKO (KO) mice; n = 3/group. Scale bars, 50 μm. Tamoxifen-injected Prmt1fl/fl or Pdx1-CreERT2 mice were used as controls. Data are expressed as mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by Student t test (CG).

Figure 2

PRMT1 is required for the maintenance of mature β-cell identity. A: Representative islet images obtained by immunofluorescence staining of INS (green) and PRMT1, H4R3me2a, PDX1, NKX6.1, MAFA, or SLC2A2 (red) in 8- and 12-week-old control and Prmt1 βiKO mice; n = 3/group. B: Representative islet images obtained by immunofluorescence staining of eYFP (green), INS (blue), and GCG, SST, PPY, and UCN3 (red) from 8- and 12-week-old R26-eYFP; Pdx1-CreERT2 (control) and R26-eYFP; Prmt1 βiKO mice; n = 3/group. White arrows indicate GCG/SST/PPY-expressing or UCN3 eYFP+ cells. Quantification analysis of eYFP-copositive cells expressing GCG (C), SST (D), PPY (E), INS (F), and UCN3 (G) in the islets of 8- and 12-week-old R26-eYFP; Pdx1-CreERT2 (WT) and R26-eYFP; Prmt1 βiKO (KO) mice; n = 3/group. Scale bars, 50 μm. Tamoxifen-injected Prmt1fl/fl or Pdx1-CreERT2 mice were used as controls. Data are expressed as mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by Student t test (CG).

Close modal

Further immunofluorescence staining revealed more severe defects in the β-cells of Prmt1 βiKO mice (Fig. 2A). At 8 weeks of age, insulin expression was robustly reduced in the β-cells of Prmt1 βiKO mice, and substantial numbers of INS/PDX1+, INS/NKX6.1+, and INS/MAFA+ cells were observed in the islets. At 12 weeks of age, the insulin signals were slightly recovered in Prmt1 βiKO mice, but their intensities were still low, and most of the β-cells had lost MAFA and SLC2A2. A lineage-tracing analysis showed that a number of β-cells lost their identity and became INS cells (∼30%), urocortin 3 (UCN3) cells (∼33%), polyhormonal cells, or other endocrine cells (GCG, ∼1.5%; SST, ∼3.5%; and PPY, ∼0.24%) after induction of Prmt1 KO (Fig. 2B–G and Supplementary Fig. 8A and B). Electron microscopic analyses showed similar ultrastructural changes in Prmt1 βiKO and Prmt1 βKO mice (Fig. 1F and Supplementary Fig. 8C).

Furthermore, HFD aggravated the metabolic phenotypes of Prmt1 βiKO mice (Supplementary Fig. 9A–D). Random blood glucose levels were continuously elevated, and glucose intolerance became more severe in HFD-fed Prmt1 βiKO mice, but the insulin sensitivity and β-cell mass were comparable to those of the WT mice (Supplementary Fig. 9C–F). Consistent with these findings, β-cells lost their identity in HFD-fed Prmt1 βiKO mice: loss of insulin (∼30%) and expression of other hormones (GCG, ∼1.7%; SST, ∼3.8%; and PPY, ∼0.37%), presence of INS/PDX1+ cells, cytoplasmic localization of NKX6.1, and loss of MAFA (Fig. 3A–F). In addition, the HFD-induced nuclear translocation of FOXO1 was blocked in Prmt1 βiKO mice (Fig. 3G). These data indicate that β-cells require PRMT1 to maintain their mature identity and that the loss of PRMT1 leads to the aberrant reprogramming of β-cells to express other hormones.

Figure 3

HFD exacerbates mature β-cell identity in Prmt1 βiKO mice. Male control and Prmt1 βiKO mice (12 weeks) were fed HFD for 14 weeks and used for experiments. A: Representative islet images obtained by immunofluorescence (IF) staining of eYFP (green), INS (blue), and GCG, SST, and PPY (red) from HFD-fed 26-week-old R26-eYFP; Pdx1-CreERT2 (control) and R26-eYFP; Prmt1 βiKO mice; n = 3/group. White arrows indicate GCG/SST/PPY-expressing eYFP+ cells. Quantification analysis of eYFP-copositive cells expressing INS (B), GCG (C), SST (D), and PPY (E) in the islets of HFD-fed 26-week-old R26-eYFP; Pdx1-CreERT2 (WT) and R26-eYFP; Prmt1 βiKO (KO) mice; n = 3/group. F: Representative islet images obtained by IF staining of INS (green) and PDX1, NKX6.1, or MAFA (red); n = 3/group. G: Representative islet images obtained by IF staining of INS (green), FOXO1 (red), and DAPI (blue); n = 3/group. Scale bars, 50 μm. Tamoxifen-injected Prmt1fl/fl or Pdx1-CreERT2 mice were used as controls. Data are expressed as mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by Student t test (BE).

Figure 3

HFD exacerbates mature β-cell identity in Prmt1 βiKO mice. Male control and Prmt1 βiKO mice (12 weeks) were fed HFD for 14 weeks and used for experiments. A: Representative islet images obtained by immunofluorescence (IF) staining of eYFP (green), INS (blue), and GCG, SST, and PPY (red) from HFD-fed 26-week-old R26-eYFP; Pdx1-CreERT2 (control) and R26-eYFP; Prmt1 βiKO mice; n = 3/group. White arrows indicate GCG/SST/PPY-expressing eYFP+ cells. Quantification analysis of eYFP-copositive cells expressing INS (B), GCG (C), SST (D), and PPY (E) in the islets of HFD-fed 26-week-old R26-eYFP; Pdx1-CreERT2 (WT) and R26-eYFP; Prmt1 βiKO (KO) mice; n = 3/group. F: Representative islet images obtained by IF staining of INS (green) and PDX1, NKX6.1, or MAFA (red); n = 3/group. G: Representative islet images obtained by IF staining of INS (green), FOXO1 (red), and DAPI (blue); n = 3/group. Scale bars, 50 μm. Tamoxifen-injected Prmt1fl/fl or Pdx1-CreERT2 mice were used as controls. Data are expressed as mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by Student t test (BE).

Close modal

PRMT1 Regulates the Transcriptomic Program of Mature β-Cells

In Prmt1 βiKO mice, β-cells lost their identity soon after Prmt1 was ablated and before the glucose homeostasis deteriorated. To examine the molecular mechanism underlying the observed loss of mature β-cell identity, we explored the global gene expression pattern in islets of Prmt1 βiKO mice at two different stages: the early stage at 8 weeks of age and the late stage at 12 weeks of age (Fig. 4A). RNA-seq analysis revealed robust changes in gene expression at both stages (Supplementary Fig. 10A and B). In particular, genes involved in cellular energy production processes, such as oxidation reduction and the electron transport chain, were commonly downregulated at both stages (Fig. 4B). Our stage-specific gene expression analysis showed that the gene expression patterns differed between the early and late stages (Fig. 4C and Supplementary Table 2). The mature β-cell genes showed various changes at the early stage, whereas most of these genes were downregulated at the late stage. In particular, the expression levels of Ins1, Ins2, Ucn3, and Pdx1 were decreased in Prmt1 βiKO mice at both stages, whereas those of Mafa and Slc2a2 were downregulated only at the late stage. These results were further confirmed by qRT-PCR analysis (Fig. 4D–I). The notable novel features in the β-cells of Prmt1 βiKO mice included the robust downregulation of Ins1 at the early stage and its recovery at the late stage and the downregulations of Ins2 and Ucn3 at the early stage (Fig. 4D–F). These findings correlated with our immunofluorescence staining observations in Prmt1 βiKO mice (Fig. 2B). Meanwhile, RNA-seq and qRT-PCR analyses revealed that most of the genes related to oxidative phosphorylation (OXPHOS) were downregulated in Prmt1 βiKO mice at both stages, suggesting that mitochondrial dysfunction could be a feature of the loss of mature β-cell identity (Fig. 4C and Supplementary Fig. 10C). Stage-specific gene ontology and pathway analyses revealed that most of the genes downregulated at the early stage were linked with the electron transport chain and maturity-onset diabetes of the young, whereas genes related to GSIS function (e.g., those related to vesicle transport and intracellular protein trafficking) were downregulated at the late stage (Fig. 4J and K). Taken together, these data indicate that PRMT1 is required to maintain the transcriptional program of mature β-cells.

Figure 4

Deletion of Prmt1 from mature β-cells induces robust transcriptomic changes. RNA-seq analysis of islets from Prmt1 βiKO mice at the early (8 weeks) and late (12 weeks) stages of loss of mature β-cell identity. Age- and sex (male)-matched, tamoxifen (TAM)-injected littermates (Prmt1fl/fl) were used as controls at the two different stages; n = 2/group. Differentially expressed genes (DEGs) were identified using the following parameters: log2(fold change of counts per million mapped reads) ≤−1.5 or ≥1.5; false discovery rate <0.05. A: Schematic representation of time points at which RNA-seq sampling was performed for islets of control and Prmt1 βiKO mice. B: Top ranked gene ontology of commonly downregulated (n = 260) and upregulated (n = 310) DEGs. C: Expression heat maps of gene subsets relative to the functional categories identified in our islet RNA-seq analysis; n = 2/group. Gene lists for the heat maps are presented in Supplementary Table 2. DI: RNA-seq and qRT-PCR analyses of islets from Prmt1 βiKO mice at the early (8 weeks) and late (12 weeks) stages of loss of mature β-cell identity; n = 2/group for RNA-seq and n = 4/group for qRT-PCR. Line (RNA-seq) and bar (qRT-PCR) graphs showing relative expressions of the representative mature β-cell genes Ins1 (D), Ins2 (E), Ucn3 (F), Pdx1 (G), Mafa (H), and Slc2a2 (I) at the two different stages of loss of mature β-cell identity. Fold changes of counts per million are plotted, and false discovery rate values are indicated in line graphs for each stage. Expression levels of genes were normalized to Actb in each sample in qRT-PCR analysis. Heat maps of gene ontology (GO) (J) and Kyoto Encyclopedia of Genes and Genomes pathway (K) analyses of the stage-specific downregulated and upregulated DEGs. Data are expressed as mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by one-way ANOVA with post hoc Tukey test (DI). 8w, 8 weeks; 12w, 12 weeks; MAPK, mitogen-activated protein kinase; NOD, nucleotide-binding oligomerization domain; TCA, tricarboxylic acid.

Figure 4

Deletion of Prmt1 from mature β-cells induces robust transcriptomic changes. RNA-seq analysis of islets from Prmt1 βiKO mice at the early (8 weeks) and late (12 weeks) stages of loss of mature β-cell identity. Age- and sex (male)-matched, tamoxifen (TAM)-injected littermates (Prmt1fl/fl) were used as controls at the two different stages; n = 2/group. Differentially expressed genes (DEGs) were identified using the following parameters: log2(fold change of counts per million mapped reads) ≤−1.5 or ≥1.5; false discovery rate <0.05. A: Schematic representation of time points at which RNA-seq sampling was performed for islets of control and Prmt1 βiKO mice. B: Top ranked gene ontology of commonly downregulated (n = 260) and upregulated (n = 310) DEGs. C: Expression heat maps of gene subsets relative to the functional categories identified in our islet RNA-seq analysis; n = 2/group. Gene lists for the heat maps are presented in Supplementary Table 2. DI: RNA-seq and qRT-PCR analyses of islets from Prmt1 βiKO mice at the early (8 weeks) and late (12 weeks) stages of loss of mature β-cell identity; n = 2/group for RNA-seq and n = 4/group for qRT-PCR. Line (RNA-seq) and bar (qRT-PCR) graphs showing relative expressions of the representative mature β-cell genes Ins1 (D), Ins2 (E), Ucn3 (F), Pdx1 (G), Mafa (H), and Slc2a2 (I) at the two different stages of loss of mature β-cell identity. Fold changes of counts per million are plotted, and false discovery rate values are indicated in line graphs for each stage. Expression levels of genes were normalized to Actb in each sample in qRT-PCR analysis. Heat maps of gene ontology (GO) (J) and Kyoto Encyclopedia of Genes and Genomes pathway (K) analyses of the stage-specific downregulated and upregulated DEGs. Data are expressed as mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001 by one-way ANOVA with post hoc Tukey test (DI). 8w, 8 weeks; 12w, 12 weeks; MAPK, mitogen-activated protein kinase; NOD, nucleotide-binding oligomerization domain; TCA, tricarboxylic acid.

Close modal

PRMT1-Dependent H4R3me2a Regulates Chromatin Accessibility in Mature β-Cells

Cell type–specific chromatin state is essential for maintaining the transcriptional program and identity of a differentiated cell (39). Given the extensive gene expression changes in Prmt1 βiKO islets and the association of H4R3me2a with CTCF and β-cell TFs in MIN6 cells, we speculated that PRMT1-dependent H4R3me2a could regulate gene transcription through the actions on the chromatin status of mature β-cells. To test our hypothesis, we performed ATAC-seq with β-cells purified from Prmt1 βiKO and WT mice at 8 weeks of age (Fig. 5A). Unlike MIN6 cells that showed clear and distinctive ATAC-seq peaks, ATAC-seq with purified β-cells showed some background noise in the peaks. Therefore, we analyzed ATAC-seq data from WT and Prmt1-null β-cells along with ATAC-seq data from MIN6 cells and identified 5,044 peaks corresponding to the PRMT1-dependent open chromatin regions in mature β-cells (Fig. 5B). The average ChIP-seq peak intensities of H4R3me2a and H3K27ac (active enhancer marks) were higher in PRMT1-dependent open chromatin regions compared with PRMT1-independent open chromatin regions (Fig. 5C and D). However, the average ChIP-seq peak intensity of H3K4me1 (a poised enhancer mark) was similar in PRMT1-dependent and -independent open chromatin regions (Fig. 5E). These data indicate that H4R3me2a is responsible for the PRMT1-dependent chromatin openings in mature β-cells. The PRMT1-dependent open chromatin regions were also correlated with the binding sites of CTCF and β-cell TFs, including NKX6.1, NKX2.2, MAFA, NEUROD1, and PDX1 (Fig. 5F). Further motif analysis showed that the DNA-binding motifs of CTCF and β-cell TFs were highly associated with PRMT1-dependent open chromatin regions (Fig. 5G and Supplementary Fig. 11). Intriguingly, genes near PRMT1-dependent open chromatin regions were significantly associated with mature β-cell function and identity (Fig. 5H and I). These data indicate that PRMT1-dependent H4R3me2a is needed to maintain the unique chromatin architecture of mature β-cells and that the loss of mature β-cell identity in Prmt1 βiKO mice is associated with widespread alterations of the chromatin landscape.

Figure 5

PRMT1-dependent H4R3me2a regulates the chromatin accessibility of mature β-cells. AL, N, and P: ATAC-seq analysis performed on β-cells purified from 8-week-old R26-eYFP; Pdx1-CreERT2 (WT) and R26-eYFP; Prmt1 βiKO (KO) mice; n = 12/group were used for one ATAC-seq replicate. ATAC-seq peaks of WT and KO β-cells; n = 2 were used for replicates. A: Schematic representation of the ATAC-seq analysis. B: Volcano plot showing differential ATAC-seq peaks in WT and KO β-cells. Differentially changed ATAC-seq peaks were identified using the following parameters: log2(fold change; WT/KO) ≤−1 or ≥1. P < 0.01. Average ChIP-seq peak intensities for H4R3me2a (C), H3K27ac (D), and H3K4me1 (E) in PRMT1-independent (indep.) and -dependent (dep.) open chromatin regions. F: Heat maps of normalized ATAC-seq and ChIP-seq signals at PRMT1-dependent open chromatin regions. Each row represents a peak. Normalized ChIP-seq signals of histone marks, CTCF, and β-cell TFs are shown. G: Known TF-binding motifs returned by Hypergeometric Optimization of Motif EnRichment analysis for PRMT1-dependent open chromatin regions. Gene ontology (GO) and pathway enrichment analyses of the genes nearby promoters (H) and enhancers (I) of PRMT1-dependent open chromatin regions. Scatter plots showing correlation between gene expression and open chromatin changes in WT and KO β-cells for β-cell genes (J) and OXPHOS genes (K). The x-axis represents log2(fold change; KO/WT) of RNA-seq data from 8-week-old Prmt1 βiKO islets. The y-axis is a change of the highest ATAC-seq peak from the same gene promoters (TSS ± 2 kb). Integrative maps of ChIP-seq (histone marks and β-cell TFs), ATAC-seq (MIN6, WT, and KO β-cells), and RNA-seq (WT and KO islets) data obtained for the Ins1 (L), Ucn3 (N), and Pdx1 (P) genes. Red boxes indicate the regions in which the ATAC-seq signals of KO β-cells are decreased. Asterisks (red) indicate PRMT1-dependent open chromatin regions. 3C-PCR assays were performed in MIN6 cells for Ins1 (M) and Ucn3 (O) genes. Arrows indicate 3C-PCR primers. L, N, and P: ATAC-seq peaks of MIN6 cells; n = 3 were used for replicates. C, F, L, N, and P: H4R3me2a ChIP-seq peaks of MIN6 cells; n = 3 were used for replicates. F: CTCF ChIP-seq peaks of MIN6 cells; n = 2 were used for replicates. bp, base pair; E, enhancer; gDNA, genomic DNA; HIF, hypoxia-inducible factor; MAPK, mitogen-activated protein kinase; P, promoter; TGF-β, transforming growth factor-β.

Figure 5

PRMT1-dependent H4R3me2a regulates the chromatin accessibility of mature β-cells. AL, N, and P: ATAC-seq analysis performed on β-cells purified from 8-week-old R26-eYFP; Pdx1-CreERT2 (WT) and R26-eYFP; Prmt1 βiKO (KO) mice; n = 12/group were used for one ATAC-seq replicate. ATAC-seq peaks of WT and KO β-cells; n = 2 were used for replicates. A: Schematic representation of the ATAC-seq analysis. B: Volcano plot showing differential ATAC-seq peaks in WT and KO β-cells. Differentially changed ATAC-seq peaks were identified using the following parameters: log2(fold change; WT/KO) ≤−1 or ≥1. P < 0.01. Average ChIP-seq peak intensities for H4R3me2a (C), H3K27ac (D), and H3K4me1 (E) in PRMT1-independent (indep.) and -dependent (dep.) open chromatin regions. F: Heat maps of normalized ATAC-seq and ChIP-seq signals at PRMT1-dependent open chromatin regions. Each row represents a peak. Normalized ChIP-seq signals of histone marks, CTCF, and β-cell TFs are shown. G: Known TF-binding motifs returned by Hypergeometric Optimization of Motif EnRichment analysis for PRMT1-dependent open chromatin regions. Gene ontology (GO) and pathway enrichment analyses of the genes nearby promoters (H) and enhancers (I) of PRMT1-dependent open chromatin regions. Scatter plots showing correlation between gene expression and open chromatin changes in WT and KO β-cells for β-cell genes (J) and OXPHOS genes (K). The x-axis represents log2(fold change; KO/WT) of RNA-seq data from 8-week-old Prmt1 βiKO islets. The y-axis is a change of the highest ATAC-seq peak from the same gene promoters (TSS ± 2 kb). Integrative maps of ChIP-seq (histone marks and β-cell TFs), ATAC-seq (MIN6, WT, and KO β-cells), and RNA-seq (WT and KO islets) data obtained for the Ins1 (L), Ucn3 (N), and Pdx1 (P) genes. Red boxes indicate the regions in which the ATAC-seq signals of KO β-cells are decreased. Asterisks (red) indicate PRMT1-dependent open chromatin regions. 3C-PCR assays were performed in MIN6 cells for Ins1 (M) and Ucn3 (O) genes. Arrows indicate 3C-PCR primers. L, N, and P: ATAC-seq peaks of MIN6 cells; n = 3 were used for replicates. C, F, L, N, and P: H4R3me2a ChIP-seq peaks of MIN6 cells; n = 3 were used for replicates. F: CTCF ChIP-seq peaks of MIN6 cells; n = 2 were used for replicates. bp, base pair; E, enhancer; gDNA, genomic DNA; HIF, hypoxia-inducible factor; MAPK, mitogen-activated protein kinase; P, promoter; TGF-β, transforming growth factor-β.

Close modal

To delineate how PRMT1-dependent chromatin openings relate to the transcriptional changes observed in Prmt1-null β-cells at the early stage, we closely examined the regulatory regions of genes downregulated in the islets of Prmt1 βiKO mice at 8 weeks of age in parallel with RNA-seq and ChIP-seq data of β-cell TFs. Indeed, the PRMT1-dependent open chromatin regions included multiple promoter or enhancer regions of β-cell and OXPHOS genes that were downregulated in the islets of Prmt1 βiKO mice. A comparative analysis of data from RNA-seq and ATAC-seq showed a correlation between the gene expression changes of β-cell genes and OXPHOS genes and the chromatin accessibility in the promoters of these genes (Fig. 5J and K and Supplementary Table 3). Moreover, the average ATAC-seq peak intensities of Prmt1-null β-cells were decreased in the promoters of β-cell genes and OXPHOS genes, indicating that the decreased expression of these genes at the early stage could be attributed to the loss of PRMT1-dependent H4R3me2a (Supplementary Fig. 12A and B).

We also identified PRMT1-dependent open chromatin regions at 12 kb upstream of the transcription start site (TSS) for the Ins1 gene and at 40 kb (E1) and 60 kb (E2) upstream of the TSS for the Ucn3 gene (Fig. 5L and N). These regions contained binding sites for β-cell TFs, including PDX1, MAFA, NKX2.2, and NKX6.1. Although the expression levels of Ins1 and Ucn3 were robustly and rapidly downregulated in Prmt1 βiKO mice at the early stage, the ATAC-seq peaks at the promoters of both genes were not significantly reduced at this point. Instead, 3C-PCR experiments showed the long-range interactions between the upstream enhancer elements and the promoters of the Ins1 and Ucn3 genes, indicating that the loss of chromatin accessibility for the β-cell TFs at the upstream enhancer elements had reduced the promoter activities of these genes (Fig. 5M and O). We also identified a PRMT1-dependent open chromatin region at 5 kb upstream of the TSS for the Pdx1 gene (Fig. 5P); this region, which is called area IV, was recently reported to play an essential role in β-cell maturation during the weaning period (40). PDX1 has also been shown to directly regulate the gene expression of numerous mitochondrial genes that were downregulated in Prmt1-null β-cells (4144). These data suggest that PRMT1-dependent H4R3me2a plays a critical role in maintaining the unique chromatin architecture of mature β-cells and that the alteration of this chromatin architecture can result in the loss of mature β-cell identity.

In an effort to test the possible implication of PRMT1-dependent H4R3me2a in human diabetes, we performed sequence-alignment analysis of PRMT1-dependent open chromatin regions in the human genome and searched for conserved regulatory elements in these regions. The E2 element of the mouse Ucn3 gene and area IV of the mouse Pdx1 gene were highly conserved in the human UCN3 and PDX1 genes, which are found at similar genomic locations (Fig. 6A and B). A type 2 diabetes–associated locus was found near area IV of PDX1, and mice lacking endogenous area IV showed the impairment of β-cell maturation (40,45). We also found a highly conserved PRMT1-dependent open chromatin region in the human SLC30A8 gene, which is strongly associated with type 2 diabetes (4548) (Fig. 6C). These findings prompted us to examine the association of human orthologous sequences of PRMT1-dependent open chromatin regions with diabetes genome-wide association study single nucleotide polymorphisms. Interestingly, the human diabetes-associated loci were more closely localized with PRMT1-dependent open chromatin regions than with PRMT1-independent open chromatin regions (Fig. 6D). This suggests that there may be a link between PRMT1-dependent H4R3me2a and the susceptibility to type 2 diabetes in humans.

Figure 6

PRMT1-dependent open chromatin regions are conserved in the human genome. Human genome alignment of PRMT1-dependent open chromatin regions in mouse β-cells for the Ucn3 (A), Pdx1 (B), and Slc30a8 (C) genes. Asterisks (red) indicate PRMT1-dependent open chromatin regions. Red bars and numbers indicate genomic locations and percentages of sequence identity, respectively. Type 2 diabetes susceptibility loci are indicated in green. Multiple alignments (Multiz Align) were performed for the genomes of 100 vertebrate species, which were captured from the University of California Santa Cruz Genome Browser (https://genome.ucsc.edu/). D: Cumulative plot of human conserved ATAC-seq peaks with the distances to the closest diabetes susceptibility locus. Human conserved PRMT1-dependent peaks tended to be closer to the diabetes susceptibility loci compared with the PRMT1-independent peaks; P = 0.0003739 by paired Student t test, with matched sampling. E: Schematic representation describing the physiological role of PRMT1-dependent H4R3me2a in mature β-cells. Vert. Cons, vertebrate conservation.

Figure 6

PRMT1-dependent open chromatin regions are conserved in the human genome. Human genome alignment of PRMT1-dependent open chromatin regions in mouse β-cells for the Ucn3 (A), Pdx1 (B), and Slc30a8 (C) genes. Asterisks (red) indicate PRMT1-dependent open chromatin regions. Red bars and numbers indicate genomic locations and percentages of sequence identity, respectively. Type 2 diabetes susceptibility loci are indicated in green. Multiple alignments (Multiz Align) were performed for the genomes of 100 vertebrate species, which were captured from the University of California Santa Cruz Genome Browser (https://genome.ucsc.edu/). D: Cumulative plot of human conserved ATAC-seq peaks with the distances to the closest diabetes susceptibility locus. Human conserved PRMT1-dependent peaks tended to be closer to the diabetes susceptibility loci compared with the PRMT1-independent peaks; P = 0.0003739 by paired Student t test, with matched sampling. E: Schematic representation describing the physiological role of PRMT1-dependent H4R3me2a in mature β-cells. Vert. Cons, vertebrate conservation.

Close modal

Epigenetic regulation is crucial for β-cell maturation and the maintenance of mature β-cell identity (1316). As one of the major mechanisms of epigenetic regulation, histone arginine methylation plays important roles in transcriptional regulation (17,18). However, its role in β-cells has not yet been explored. Here, we demonstrate a novel function of PRMT1-dependent H4R3me2a in maintaining mature β-cell identity. Both Prmt1 βKO and Prmt1 βiKO mice developed diabetes, which was aggravated by HFD-induced metabolic stress (Supplementary Fig. 13). Deletion of Prmt1 in adult β-cells resulted in the immediate loss of H4R3me2a, which induced robust changes in the transcriptions of genes necessary for the maintenance of mature β-cell function and identity. PRMT1-dependent H4R3me2a worked as an active histone code that increased chromatin accessibility at the binding sites for CTCF and β-cell TFs, including NKX6.1, MAFA, PDX1, and NEUROD1 (Fig. 6E). Furthermore, PRMT1-dependent open chromatin regions appear to be associated with genes that have been associated with diabetes susceptibility in humans.

Genome-wide association studies have indicated that most diabetes-susceptibility genes are related to β-cells, and thus, β-cell failure is thought to be an essential feature of diabetes (49,50). The dysfunction of β-cells occurs long before hyperglycemia develops in humans (51), suggesting that β-cell dysfunction is an early feature of diabetic β-cell failure. However, due to the lack of an appropriate animal model, researchers have been limited in their ability to study how β-cell dysfunction begins in response to metabolic stress and how β-cell failure progresses before the apoptosis or dedifferentiation of β-cells is observed. In this regard, Prmt1 βiKO mice provide the following useful features: initial events of loss of β-cell identity, which has not previously been described in an animal model; loss of INS and UCN3 prior to the loss of other mature β-cell TFs; decreased expression of β-cell genes; decreased expression of mitochondrial genes; and subsequent mitochondrial dysfunction. Despite these extensive changes of gene expression in the β-cells of Prmt1 βiKO mice, their metabolic phenotype was unexpectedly mild and became severe upon HFD feeding. This discrepancy prompted us to propose the following explanations: 1) since β-cells are highly dedicated to insulin production and secretion, even though the β-cells of Prmt1 βiKO mice are not fully functional, they can maintain glycemic control as long as mice are insulin sensitive; and 2) there may be compensatory mechanisms to maintain glycemic control in Prmt1 βiKO mice. The restoration of Ins1 expression in the late stage of Prmt1 βiKO mice (12 weeks of age) supports the existence of these compensatory mechanisms.

Since the phenotype of Prmt1 βiKO mice resembles the natural history of type 2 diabetes, this mouse model may be useful for studying how β-cells behave in response to metabolic stress during the development of type 2 diabetes. Prmt1 βiKO mice showed the following features in the progression of β-cell failure: loss of β-cell identity, aberrant expression of β-cell TFs, and cell type change of β-cells to other endocrine cells. We speculate that Prmt1 deletion resulted in the loss of H4R3me2a and that this causes β-cells to lose their cell-specific chromatin architecture and gene expression program and thereby lose their identity. These β-cells that lose their identity then undergo different changes based on their genetic heterogeneity in response to metabolic stress. However, further study will be needed to elucidate the precise mechanisms underlying the severe β-cell phenotypes of HFD-fed Prmt1 βKO and Prmt1 βiKO mice. The chromatin changes driven by HFD feeding together with the losses of H4R3me2a and arginine methylation in nonhistone substrates of PRMT1 may have affected these phenotypes. FOXO1 and HNF4α are PRMT1’s nonhistone targets that are also known to play roles in β-cell function and identity (4,5254). However, the phenotypes of β-cell–specific Foxo1 or Hnf4α KO mice differed from those of Prmt1 βKO mice (4,54). In this regard, we think that the phenotypes of Prmt1 βKO and Prmt1 βiKO mice may be largely attributed to the loss of H4R3me2a in β-cells.

Here, we provide novel insight into the importance of epigenetic control of PRMT1-dependent H4R3me2a in maintaining mature β-cell identity. Taken together with the associations seen among CTCF, β-cell TFs, and H4R3me2a, our work reveals previously unknown functions of PRMT1-dependent open chromatin regions that govern mature β-cell identity. Thus, our phenotypic, transcriptomic, and epigenomic analyses of stage-specific Prmt1 KO in β-cells provide a new mechanistic insight into the regulation of mature β-cell identity.

Hyu. Kim, B.-H.Y., C.-M.O., and Jo. Lee contributed equally to this work.

Acknowledgments. The authors thank Hee-Saeng Jung, Jueun Kim, and Hanna Jung (Korea Advanced Institute of Science and Technology) for technical support, Dr. Dahee Choi (Korea University) and Hyun Jung Hong (Chungnam National University) for technical advice and support, Drs. Yong-Ho Ahn and Joo-Man Park (Yonsei University) for help with electron microscopic analysis, Jeong-Hwan Kim (Korea Research Institute of Bioscience and Biotechnology) for technical assistance with next-generation sequencing library preparation, Drs. Kyong Soo Park (Seoul National University), Kun-Ho Yoon (Catholic University of Korea), Soo Heon Kwak (Seoul National University), and Kyoung-Jae Won (University of Copenhagen) for helpful discussions, and Drs. Mark O. Huising (University of California, Davis) and Paul E. Sawchenko (Salk Institute) for the gift of the UCN3 antibody.

Funding. This work was supported by grants from the National Research Foundation funded by the Ministry of Science, ICT and Future Planning, Republic of Korea (grants NRF-2013M3A9D5072550 to J.K.S., NRF-2017M3C9A5028693 to M.K., and NRF-2014M3A9D5A01073546, NRF-2018R1A2A3074646, and NRF-2015M3A9B3028218 to Hai. Kim), the Korea Research Institute of Bioscience and Biotechnology Research Initiative (to M.K.), and the Korea Advanced Institute of Science and Technology Institute for the BioCentury (grant N10180027 to Hai. Kim).

Duality of Interest. No potential conflicts of interest relevant to this article were reported.

Author Contributions. Hyu. Kim, B.-H.Y., C.-M.O., Jo. Lee, M.K., and Hai. Kim generated the hypothesis, designed the experiments, and analyzed the results. Hyu. Kim, C.-M.O., Jo. Lee, K.L., H.S., M.-Y.K., Hye. Kim, Y.K.K., and J.K.S. performed the animal experiments. Hyu. Kim, C.-M.O., E.K., E.-H.S., H.H., H.-J.K., Ju. Lee, J.M.S., S.-H.K., S.K., and M.S. performed the cell and in vitro experiments. Hyu. Kim, B.-H.Y., C.-M.O., K.Y., Y.S.J., M.K., and Hai. Kim analyzed the next-generation sequencing data. Hyu. Kim, B.-H.Y., C.-M.O., Jo. Lee, M.K., and Hai. Kim wrote the manuscript. M.K. and Hai. Kim supervised the research. M.K. and Hai. Kim are the guarantors of this work and, as such, had full access to all of the data in the study and take responsibility for the integrity of the data and the accuracy of the data analysis.

Prior Presentation. Parts of this study were presented in abstract form at the Drivers of Type 2 Diabetes: From Genes to Environment (S1) conference of the Keystone Symposia, Seoul, Republic of Korea, 7–11 October 2018, and at the Third Joint EASD Islet Study Group and Beta Cell Workshop, Oxford, U.K., 1–3 April 2019.

1.
Halban
PA
,
Polonsky
KS
,
Bowden
DW
, et al
.
β-Cell failure in type 2 diabetes: postulated mechanisms and prospects for prevention and treatment
.
Diabetes Care
2014
;
37
:
1751
1758
2.
Prentki
M
,
Nolan
CJ
.
Islet β cell failure in type 2 diabetes
.
J Clin Invest
2006
;
116
:
1802
1812
3.
Donath
MY
,
Ehses
JA
,
Maedler
K
, et al
.
Mechanisms of β-cell death in type 2 diabetes
.
Diabetes
2005
;
54
(
Suppl. 2
):
S108
S113
4.
Talchai
C
,
Xuan
S
,
Lin
HV
,
Sussel
L
,
Accili
D
.
Pancreatic β cell dedifferentiation as a mechanism of diabetic β cell failure
.
Cell
2012
;
150
:
1223
1234
5.
Cigliola
V
,
Thorel
F
,
Chera
S
,
Herrera
PL
.
Stress-induced adaptive islet cell identity changes
.
Diabetes Obes Metab
2016
;
18
(
Suppl. 1
):
87
96
6.
Nishimura
W
,
Takahashi
S
,
Yasuda
K
.
MafA is critical for maintenance of the mature beta cell phenotype in mice
.
Diabetologia
2015
;
58
:
566
574
7.
Gu
C
,
Stein
GH
,
Pan
N
, et al
.
Pancreatic β cells require NeuroD to achieve and maintain functional maturity
.
Cell Metab
2010
;
11
:
298
310
8.
Swisa
A
,
Avrahami
D
,
Eden
N
, et al
.
PAX6 maintains β cell identity by repressing genes of alternative islet cell types
.
J Clin Invest
2017
;
127
:
230
243
9.
Ediger
BN
,
Lim
H-W
,
Juliana
C
, et al
.
LIM domain-binding 1 maintains the terminally differentiated state of pancreatic β cells
.
J Clin Invest
2017
;
127
:
215
229
10.
Gutiérrez
GD
,
Bender
AS
,
Cirulli
V
, et al
.
Pancreatic β cell identity requires continual repression of non-β cell programs
.
J Clin Invest
2017
;
127
:
244
259
11.
Dixon
JR
,
Jung
I
,
Selvaraj
S
, et al
.
Chromatin architecture reorganization during stem cell differentiation
.
Nature
2015
;
518
:
331
336
12.
Ernst
J
,
Kheradpour
P
,
Mikkelsen
TS
, et al
.
Mapping and analysis of chromatin state dynamics in nine human cell types
.
Nature
2011
;
473
:
43
49
13.
Campbell
SA
,
Hoffman
BG
.
Chromatin regulators in pancreas development and diabetes
.
Trends Endocrinol Metab
2016
;
27
:
142
152
14.
Avrahami
D
,
Li
C
,
Zhang
J
, et al
.
Aging-dependent demethylation of regulatory elements correlates with chromatin state and improved β cell function
.
Cell Metab
2015
;
22
:
619
632
15.
Arda
HE
,
Li
L
,
Tsai
J
, et al
.
Age-dependent pancreatic gene regulation reveals mechanisms governing human β cell function
.
Cell Metab
2016
;
23
:
909
920
16.
Lu
TT-H
,
Heyne
S
,
Dror
E
, et al
.
The polycomb-dependent epigenome controls β cell dysfunction, dedifferentiation, and diabetes
.
Cell Metab
2018
;
27
:
1294
1308.e7
17.
Di Lorenzo
A
,
Bedford
MT
.
Histone arginine methylation
.
FEBS Lett
2011
;
585
:
2024
2031
18.
Blanc
RS
,
Richard
S
.
Arginine methylation: the coming of age
.
Mol Cell
2017
;
65
:
8
24
19.
Kim
JK
,
Lim
Y
,
Lee
JO
, et al
.
PRMT4 is involved in insulin secretion via methylation of histone H3 in pancreatic β cells
.
J Mol Endocrinol
2015
;
54
:
315
324
20.
Tang
J
,
Frankel
A
,
Cook
RJ
, et al
.
PRMT1 is the predominant type I protein arginine methyltransferase in mammalian cells
.
J Biol Chem
2000
;
275
:
7723
7730
21.
Iwasaki
H
.
Impaired PRMT1 activity in the liver and pancreas of type 2 diabetic Goto-Kakizaki rats
.
Life Sci
2009
;
85
:
161
166
22.
Wang
H
,
Huang
ZQ
,
Xia
L
, et al
.
Methylation of histone H4 at arginine 3 facilitating transcriptional activation by nuclear hormone receptor
.
Science
2001
;
293
:
853
857
23.
Huang
S
,
Litt
M
,
Felsenfeld
G
.
Methylation of histone H4 by arginine methyltransferase PRMT1 is essential in vivo for many subsequent histone modifications
.
Genes Dev
2005
;
19
:
1885
1893
24.
Kim
K
,
Oh
CM
,
Ohara-Imaizumi
M
, et al
.
Functional role of serotonin in insulin secretion in a diet-induced insulin-resistant state
.
Endocrinology
2015
;
156
:
444
452
25.
Szot
GL
,
Koudria
P
,
Bluestone
JA
.
Murine pancreatic islet isolation
.
J Vis Exp
2007
;
7
:
255
26.
Baek
S-J
,
Kim
M
,
Bae
D-H
, et al
.
Integrated epigenomic analyses of enhancer as well as promoter regions in gastric cancer
.
Oncotarget
2016
;
7
:
25620
25631
27.
Buenrostro
JD
,
Wu
B
,
Chang
HY
,
Greenleaf
WJ
.
ATAC-seq: a method for assaying chromatin accessibility genome-wide
.
Curr Protoc Mol Biol
2015
;
109
:
21.29.1
21.29.9
28.
Hagège
H
,
Klous
P
,
Braem
C
, et al
.
Quantitative analysis of chromosome conformation capture assays (3C-qPCR)
.
Nat Protoc
2007
;
2
:
1722
1733
29.
Tang
Z
,
Luo
OJ
,
Li
X
, et al
.
CTCF-mediated human 3D genome architecture reveals chromatin topology for transcription
.
Cell
2015
;
163
:
1611
1627
30.
Ong
C-T
,
Corces
VG
.
CTCF: an architectural protein bridging genome topology and function
.
Nat Rev Genet
2014
;
15
:
234
246
31.
Dhawan
S
,
Tschen
S-I
,
Zeng
C
, et al
.
DNA methylation directs functional maturation of pancreatic β cells
.
J Clin Invest
2015
;
125
:
2851
2860
32.
Yoshihara
E
,
Wei
Z
,
Lin
CS
, et al
.
ERRγ is required for the metabolic maturation of therapeutically functional glucose-responsive β cells
.
Cell Metab
2016
;
23
:
622
634
33.
Pullen
TJ
,
Rutter
GA
.
When less is more: the forbidden fruits of gene repression in the adult β-cell
.
Diabetes Obes Metab
2013
;
15
:
503
512
34.
Marchetti
P
,
Bugliani
M
,
Lupi
R
, et al
.
The endoplasmic reticulum in pancreatic beta cells of type 2 diabetes patients
.
Diabetologia
2007
;
50
:
2486
2494
35.
Masini
M
,
Martino
L
,
Marselli
L
, et al
.
Ultrastructural alterations of pancreatic beta cells in human diabetes mellitus
.
Diabetes Metab Res Rev
2017
;
33
:
e2894
36.
Weir
GC
,
Bonner-Weir
S
.
Five stages of evolving beta-cell dysfunction during progression to diabetes
.
Diabetes
2004
;
53
(
Suppl. 3
):
S16
S21
37.
Guo
S
,
Dai
C
,
Guo
M
, et al
.
Inactivation of specific β cell transcription factors in type 2 diabetes
.
J Clin Invest
2013
;
123
:
3305
3316
38.
Cinti
F
,
Bouchi
R
,
Kim-Muller
JY
, et al
.
Evidence of β-cell dedifferentiation in human type 2 diabetes
.
J Clin Endocrinol Metab
2016
;
101
:
1044
1054
39.
Dowen
JM
,
Fan
ZP
,
Hnisz
D
, et al
.
Control of cell identity genes occurs in insulated neighborhoods in mammalian chromosomes
.
Cell
2014
;
159
:
374
387
40.
Spaeth
JM
,
Gupte
M
,
Perelis
M
, et al
.
Defining a novel role for the Pdx1 transcription factor in islet β-cell maturation and proliferation during weaning
.
Diabetes
2017
;
66
:
2830
2839
41.
Brissova
M
,
Shiota
M
,
Nicholson
WE
, et al
.
Reduction in pancreatic transcription factor PDX-1 impairs glucose-stimulated insulin secretion
.
J Biol Chem
2002
;
277
:
11225
11232
42.
Gauthier
BR
,
Wiederkehr
A
,
Baquié
M
, et al
.
PDX1 deficiency causes mitochondrial dysfunction and defective insulin secretion through TFAM suppression
.
Cell Metab
2009
;
10
:
110
118
43.
Fujimoto
K
,
Chen
Y
,
Polonsky
KS
,
Dorn
GW
.
Targeting cyclophilin D and the mitochondrial permeability transition enhances β-cell survival and prevents diabetes in Pdx1 deficiency
.
Proc Natl Acad Sci U S A
2010
;
107
:
10214
10219
44.
Soleimanpour
SA
,
Ferrari
AM
,
Raum
JC
, et al
.
Diabetes susceptibility genes Pdx1 and Clec16a function in a pathway regulating mitophagy in β-cells
.
Diabetes
2015
;
64
:
3475
3484
45.
Wheeler
E
,
Leong
A
,
Liu
C-T
, et al.;
EPIC-CVD Consortium
;
EPIC-InterAct Consortium
;
Lifelines Cohort Study
.
Impact of common genetic determinants of Hemoglobin A1c on type 2 diabetes risk and diagnosis in ancestrally diverse populations: a transethnic genome-wide meta-analysis
.
PLoS Med
2017
;
14
:
e1002383
46.
Rutter
GA
,
Chimienti
F
.
SLC30A8 mutations in type 2 diabetes
.
Diabetologia
2015
;
58
:
31
36
47.
Zhao
W
,
Rasheed
A
,
Tikkanen
E
, et al.;
CHD Exome+ Consortium
;
EPIC-CVD Consortium
;
EPIC-Interact Consortium
;
Michigan Biobank
.
Identification of new susceptibility loci for type 2 diabetes and shared etiological pathways with coronary heart disease
.
Nat Genet
2017
;
49
:
1450
1457
48.
Mahajan
A
,
Go
MJ
,
Zhang
W
, et al.;
DIAbetes Genetics Replication And Meta-analysis (DIAGRAM) Consortium
;
Asian Genetic Epidemiology Network Type 2 Diabetes (AGEN-T2D) Consortium
;
South Asian Type 2 Diabetes (SAT2D) Consortium
;
Mexican American Type 2 Diabetes (MAT2D) Consortium
;
Type 2 Diabetes Genetic Exploration by Next-generation sequencing in multi-Ethnic Samples (T2D-GENES) Consortium
.
Genome-wide trans-ancestry meta-analysis provides insight into the genetic architecture of type 2 diabetes susceptibility
.
Nat Genet
2014
;
46
:
234
244
49.
Scott
LJ
,
Mohlke
KL
,
Bonnycastle
LL
, et al
.
A genome-wide association study of type 2 diabetes in finns detects multiple susceptibility variants
.
Science
2007
;
316
:
1341
1345
50.
Tsai
F-J
,
Yang
C-F
,
Chen
C-C
, et al
.
A genome-wide association study identifies susceptibility variants for type 2 diabetes in Han Chinese
.
PLoS Genet
2010
;
6
:
e1000847
51.
Brunzell
JD
,
Robertson
RP
,
Lerner
RL
, et al
.
Relationships between fasting plasma glucose levels and insulin secretion during intravenous glucose tolerance tests
.
J Clin Endocrinol Metab
1976
;
42
:
222
229
52.
Yamagata
K
,
Daitoku
H
,
Takahashi
Y
, et al
.
Arginine methylation of FOXO transcription factors inhibits their phosphorylation by Akt
.
Mol Cell
2008
;
32
:
221
231
53.
Barrero
MJ
,
Malik
S
.
Two functional modes of a nuclear receptor-recruited arginine methyltransferase in transcriptional activation
.
Mol Cell
2006
;
24
:
233
243
54.
Miura
A
,
Yamagata
K
,
Kakei
M
, et al
.
Hepatocyte nuclear factor-4α is essential for glucose-stimulated insulin secretion by pancreatic β-cells
.
J Biol Chem
2006
;
281
:
5246
5257
Readers may use this article as long as the work is properly cited, the use is educational and not for profit, and the work is not altered. More information is available at https://www.diabetesjournals.org/content/license.

Supplementary data