Main

We conducted a meta-analysis of nine GWAS involving up to 26,356 non-diabetic individuals of European descent. We assessed eight BCF measures: (1) BCF insulin sensitivity glucose tolerance test for acute insulin response (BIGTT-AIR), (2) corrected insulin response (CIR), (3) first-phase Stumvoll (Stumvoll), (4) insulinogenic index (xinsdG30), (5) modified insulinogenic index (xinsG30), (6) disposition index for modified insulinogenic index and Matsuda (DI), (7) disposition index for BIGTT-AIR and BIGTT-SI (DIBIG) and the HOMA-β index (HOMA-β)19,20,21,22,23,24,25 (Fig. 1 and Supplementary Tables 1 and 2). After supplementing GWAS discovery with approximate conditional analyses, we identified 55 independent genetic signals at 44 GWAS loci, each detected in at least one BCF trait (Fig. 1b, Extended Data Fig. 1 and Supplementary Tables 3 and 4). Following Bonferroni correction for the eight BCF traits, 31 loci remained significant (Supplementary Table 3).

Fig. 1: GWAS discovery for eight BCF traits identifies 55 significant associations.
figure 1

a, Overview of the signal discovery strategy. We mapped genetic effects on BCF indices using oral glucose-stimulated and steady-state fasting measurements from ~26,000 non-diabetic individuals. In-house BCF groupings are highlighted on the right. b. Multi-trait Manhattan plot of single-variant genome-wide association meta-analyses for eight BCF estimates. From inner to outer tracks: HOMA-β, DI, DIBIG, BIGTT-AIR, CIR, Stumvoll, xinsdG30 and xinsG30. Genome-wide significant variants (P ≤ 5 × 10−8) are coloured in dark blue and previously unreported loci are in red. Loci meeting Bonferroni-correction thresholds are detailed in Supplementary Table 3. Y axis, association P-values on the –log10 scale; x axis, genomic position in hg19.

Next, we investigated the extent to which GWAS loci were specific to a single index or a recognizable group of BCF traits. Hierarchical clustering of genetic (rg) and phenotypic (rp) correlations clustered BCF indices into three main groups (Extended Data Fig. 2a,b). This is consistent with our expectations, as the traits are grouped according to whether the BCF indices are adjusted for insulin sensitivity or derived from dynamic (OGTT) or steady-state measurements. The groups include insulin secretion measurements during an OGTT (mean rg = 0.90, rp = 0.85), disposition indexes (rg = 0.93, rp = 0.78) and a steady-state BCF surrogate. We observed a significant negative phenotype correlation between DI and HOMA-β, which was unexpected. HOMA-β summarizes the feedback loop between β-cells and the liver that balances the hepatic glucose output and insulin secretion in the fasting state26. Previous work has reported that progressive β-cell failure in impaired glucose tolerance is captured by disposition indexes but not by HOMA-β27. This indicates that HOMA-β may inadequately reflect glucose-responsive β-cell activity28, limiting its utility in predicting the role of β-cell dysfunction in T2D progression29.

We observed that 48% of all GWAS loci were shared across all glucose-stimulated BCF-based indices (18 out of 44) or together with HOMA-β (3 out of 44). Nonetheless, 52% of them (23 out of 44) were confined to selective BCF groups or single traits (Extended Data Fig. 2c,d). Equivalent sample sizes across BCF indices indicate that differences in signal discovery between traits are not due to statistical power. Accordingly, heterogeneity tests for the effect sizes revealed high heterogeneity within BCF groups at selective signals (Supplementary Table 5). For example, GWAS associations were identified at the GIPR locus for xinsG30, CIR, Stumvoll, xinsdG30 and DI. Heterogeneity tests at rs2238691 found significant differences in the effect sizes between the two insulinogenic indexes (xinsG30 and xinsdG30) and CIR (CIR vs xinsG30, P = 1.36 × 10−8 and CIR vs xinsdG30, P = 4.37 × 10−9) and Stumvoll (Stumvoll vs xinsG30, P = 3.36 × 10−13 and Stumvoll vs xinsdG30 P = 1.15 × 106), respectively. Collectively, these results provide evidence that despite high genetic and phenotypic correlations, different estimates for BCF probably encapsulate distinct genetic profiles.

We compared BCF-GWAS with and without body mass index (BMI) adjustment to assess whether genetic effects influencing β-cell secretory function might be confounded by obesity. We found substantial concordance in effect sizes regardless of BMI adjustment (Extended Data Fig. 3 and Supplementary Table 6), despite a sizable number of shared genetic loci between BCF and BMI (Supplementary Table 7). This suggests that BCF associations are unlikely to be steered by the effects of obesity-related insulin resistance30,31.

To characterize the distinct biological mechanisms underlying genetic effects on glucose-responsive BCF, we examined their relationship with other traits involved in glucose homeostasis. We identified 14 unknown loci, eight of which had been previously associated with metabolic or anthropometric traits (for example, fasting glucose, T2D or BMI), albeit lacking prior evidence of association with BCF. Similarly, we observed a high genetic correlation between BCF and glycaemic traits as well as with other diabetes-related phenotypes (Fig. 2a). We then investigated the effects of 42 non-palindromic BCF loci on glucose metabolism and other complex traits (Fig. 2b). Colocalization evidence (posterior probability (PP) > 0.6 for at least one BCF trait) was observed between BCF associations and variants influencing T2D (25 out of 42) and/or fasting glucose (24 out of 42) (Fig. 2b, Extended Data Fig. 4a and Supplementary Table 8) across numerous loci. Interestingly, only three BCF loci show colocalization with BMI–GWAS associations, also suggesting a minor influence of BMI on BCF-GWAS signals. We further characterized the allelic effects of BCF associations on T2D risk (Fig. 2B). We observed a subset of GWAS signals in which the allele associated with decreased BCF also increased T2D risk but lowered BMI. These findings align with previous efforts identifying a similar genetic signature (in which increased adiposity is coupled with lower cardiometabolic risk) for some of these loci (for example, ARAP1, IGF2BP2, TCF7L2)32. Conversely, allelic effects of the insulin-sensitivity LOC646736;MIR5702 (IRS1) signal (Extended Data Fig. 4b,c) were associated with both reduced BCF and T2D risk but increased BMI. This locus was reported to uncouple adiposity from cardiometabolic risk by promoting a more favourable fat distribution32,33,34. Effects on BCF could be explained by an indirect influence of insulin sensitivity or a simultaneous effect on IRS1 signalling in β-cells. We then examined causal links between BCF on T2D risk using Mendelian randomization, providing evidence of a causal effect of increased BCF on lower T2D risk (inverse-variance weighted P = 8.08 × 1020 to 1.82 × 106). Interestingly, this was not observed for HOMA-β (Supplementary Table 9). Altogether, our analysis delineated key genetic determinants of BCF that modulate T2D risk through different molecular mechanisms.

Fig. 2: Impact of BCF loci on other complex traits.
figure 2

a, Heatmap of genetic correlations (Z-score) between each BCF trait and other complex traits, as calculated by LDSC regression implemented in LDHub. Asterisks indicate significant correlations after Bonferroni multiple-test correction (P ≤ 0.05 / (8 BCF traits × 9 complex trait categories)). Rows are grouped using hierarchical clustering with Euclidean distance. b, Effect sizes (Z-scores) for multiple pancreatic islet-relevant traits among 42 BCF loci. Loci are named as follows: gene | chromosome:position [effect allele]. WHR, waist-to-hip ratio; IFC, insulin fold change; ISI, Stumvoll insulin sensitivity; T1D, type 1 diabetes; IVGTT, intravenous glucose tolerance test; MRI, magnetic resonance imaging; Kidney function, estimated glomerular filtration rate based on serum creatinine; TG, triglycerides. Variant effect sizes were aligned to a negative BIGTT-AIR effect. The sentinel variant at each locus corresponds to the lead variant or a proxy (LD r2 > 0.7), excluding non-palindromic alleles and with available data among all BCF, BMI and T2D traits. **Genome-wide significant (P ≤ 5 × 108), *Bonferroni-corrected P-value (P ≤ 0.05 / 42 variants), | Nominal significance (P ≤ 0.05). Red, yellow and blue dots on the left indicate moderate evidence of colocalization at a given BCF locus between at least one BCF trait and T2D, fasting glucose or BMI, respectively.

Source data

We next identified the most relevant tissues and cell types in which genetic signals exert their effects on BCF. We examined the enrichment of BCF heritability among enhancer and promoter epigenomic annotations in human pancreatic islets13 as well as in other tissues and cell types35. All BCF traits except HOMA-β showed the strongest heritability enrichments in islet non-coding elements (Fig. 3a, Supplementary Fig. 4 and Supplementary Table 10). Likewise, we found significant BCF-GWAS enrichment in genes expressed in mouse pancreatic tissues36 as well as in intestine and brain tissues and in human endocrine cells37 (Fig. 3b,c). Notably, we observed an attenuated pattern of enrichments for HOMA-β across tissue and cell-type-specific annotations. These findings reinforce the notion that steady-state measurements are not sensitive enough to capture the glucose-responsive function of pancreatic islet cells. The adequacy of HOMA-β has been previously discussed38,39 and is also reflected in both the small genetic overlap with other BCF traits (Extended Data Fig. 2c) and the relatively low number of genetic findings in a previous meta-analysis of HOMA-β in ~95,000 individuals8. Interestingly, we observed heritability enrichments in peripheral tissues (for example, omental fat pad, small intestine, cerebellum) for insulin-sensitivity-corrected BCF measures. This further indicates that disposition indexes might capture effects from glucose metabolism in addition to BCF40. Collectively, these findings highlight pancreatic islets as the most relevant tissue of action, together with the influence of other tissues and cell types.

Fig. 3: Role of human islet gene expression and alternative splicing regulation into BCF and identity.
figure 3

a, Heatmap of heritability (h2) enrichments for BCF and other complex traits and diseases, calculated using S-LDSC, across cis-regulatory transcriptional annotations from multiple tissues and cell types. Self rep.: self-reported. b,c, Heatmap showing the association of BCF genetic signals with cell-type-specific expression in relevant cells from Tabula Muris36 (b) and pancreatic islet cells from human donors37 (c). Z-scores displayed on the heatmap were calculated using CELLECT-LDSC. **Bonferroni-corrected P ≤ 0.05 / (8 BCF traits × 7 complex trait categories) in b and P ≤ 0.05 / (8 BCF traits × 14 pancreatic cell types) in c. *Nominal significance. Rows are grouped by hierarchical clustering with Euclidean distance in ac. d, Scheme illustrating the prioritization of effector transcripts for BCF traits based on epigenomic annotations and enhancer-gene assignments in human pancreatic islets, and effects on human islet gene expression and alternative splicing through TWAS and colocalization approaches (adapted from a previous publication65). e, Enrichments of BCF candidate target genes within transcription factor (TF) co-expression networks from ARCHS4, calculated using enrichR. Barplot representation shows enrichR combined scores, with bars coloured according to the Benjamini–Hochberg-adjusted P value. *Significant enrichments at a false discovery rate of 5%. f, Illustrative examples of candidate BCF protein-coding genes contributing to the enrichment of annotations that are central for islet function and identity, and diabetes, as calculated using enrichR. The number of BCF genes contributing to each functional annotation is indicated in brackets. Adjusted P values using the Benjamini–Hochberg method are provided.

Source data

To identify molecular effectors influencing the β-cell secretory function, we integrated BCF-GWAS data with pancreatic islet epigenomic and chromatin interaction maps, as well as genetic effects on islet expression and splicing activity (Fig. 3d). We created credible sets of variants with 99% causal PP (Supplementary Table 11) and overlaid BCF fine-mapped variants on non-coding elements, connecting them to their target genes using human islet promoter capture Hi-C (pcHi-C) chromatin maps13 (Fig. 3d, left panel). From these data, we identified 72 candidate effector genes (Supplementary Table 12). For example, pcHi-C data linked enhancers harbouring BCF-GWAS variants in the CLEC14AA/LINC00639 and SLC39A11 loci to somatostatin receptor genes SSTR1 and SSTR2, respectively (Extended Data Fig. 5, Supplementary Table 12). Somatostatin inhibits insulin and glucagon secretion by activating G-protein-coupled somatostatin receptors. Although SSTR2 has a functionally dominant role in both α-cells and β-cells, selective effects on glucagon secretion have been reported via SSTR1-agonists41.

We next harnessed transcriptome-wide association studies (TWAS), which integrate transcriptomic reference panels42 with GWAS summary statistics to identify genes whose expression or splicing usage is associated with BCF traits (Fig. 3d, right panel). TWAS predicts gene expression or splicing activity within a GWAS cohort, uncovering associations between gene expression (eTWAS) or splicing junction usage (sTWAS) and GWAS traits. By reducing multiple comparison penalties, TWAS aids in identifying genetic associations that may not have reached GWAS significance43. We found 21 genes (15 eTWAS and six sTWAS) whose islet transcriptome activity is associated with BCF traits, supported by colocalization between islet expression or splicing quantitative trait loci (eQTLs and sQTLs, respectively) and BCF-GWAS signals (Supplementary Figs. 5 and 6 and Supplementary Tables 13 and 14). Among these genes, five loci harboured significant eTWAS associations but showed suggestive significance in our BCF meta-analyses. This includes UBE2E2 (rs35352848, xinsdG30 P = 2.60 × 10−7) and FAM234A (also known as ITFG3; rs56038902, Stumvoll P = 9.54 × 107) (Supplementary Fig. 5). Allelic effects increasing ubiquitin-conjugating enzyme E2E2 (UBE2E2) expression reduce BCF and are associated with higher T2D risk (Extended Data Fig. 6a–d and Supplementary Table 14). This supports the long-standing notion that β-cell homeostasis relies on the ubiquitin–proteasome system44,45. eTWAS also revealed associations between ITFG3 expression and BCF (Supplementary Fig. 5). The non-coding rs56038902-T allele, which increases islet ITFG3 expression, is also associated with lower CIR and Stumvoll measures (Extended Data Fig. 6e–h) but higher T2D risk and glucose levels46 (Supplementary Table 15). Consistent with this, exome-sequencing analysis identified significant associations between ITFG3 loss-of-function variants and lower blood glucose levels and a 36% reduction in T2D risk, respectively47. Additionally, genetic effects on splicing variation supplied eight target genes, three of which were unique to splicing regulation (Supplementary Fig. 6, Extended Data Fig. 7 and Supplementary Table 14). Overall, we identified 92 candidate molecular effectors (82 protein-coding genes and six long non-coding RNAs), 31 of which lack functional islet evidence such as that provided by previous large-scale transcriptomic and epigenomic studies (Supplementary Table 16). Protein-coding genes were enriched in islet-specific co-expression networks (Fig. 3e and Supplementary Table 17), known regulators of BCF (for example, ADCY5, KCNJ11, TCF7L2) and genes implicated in monogenic (for example, HNF1A, PDX1, ABCC8) and common (for example, IGF2BP2, C2CD4A/4B, PAM) forms of diabetes (Fig. 3f and Supplementary Table 17). Our findings provide a comprehensive list of candidate genes relevant to insulin release regulation and diabetes pathophysiology.

We next explored the mechanisms by which candidate genes at two distinct loci might influence insulin release. Disposition indexes identified the ACSL1 locus. Integrating fine-mapped DI and DIBIG associations with epigenomic annotations localized rs4862423 (DI PP = 0.23 and DIBIG PP = 0.13, linkage disequilibrium (LD) with ACSL1-lead rs10022124r2 = 0.85) and rs72695654 (DI PP = 0.00040, LD with ACSL1-lead rs10022124r2 = 0.12) to islet enhancers connected through chromatin interactions13 to the ACSL1 gene, encoding the long-chain fatty acyl-CoA synthetase 1 (Fig. 4a and Supplementary Table 12). Although TWAS did not report a significant association between ACSL1 expression and BCF, a previous large-scale human islet eQTL mapping study found strong colocalization between ACSL1 eQTLs and T2D risk through rs58730668(ref. 48), which was also prioritized through DI fine-mapping (DI PP = 0.0004, LD with ACSL1-lead rs10022124r2 = 0.12). We observed that the DI rs58730668-C allele is associated with higher ACSL1 expression, lower T2D risk and higher BCF (Supplementary Table 15). Additionally, DI and DIBIG rs4862423-T and rs72695654-T alleles, located in islet enhancers, were also associated with decreased BCF, increased glycaemic levels49 and T2D risk50 (Fig. 4b and Supplementary Table 15). ACSL1 showed robust colocalization between disposition indexes and fasting glucose (DI PP.H4 = 0.998, DIBIG PP.H4 = 0.998) and with T2D via other BCF traits (for example, BIGTT-AIR PP.H4 = 0.85) (Supplementary Table 8). We then examined the adjacent CENPU gene as a potential molecular effector in addition to ACSL1. Despite the close proximity of the enhancer containing rs4862423 to the CENPU promoter, colocalization of molecular QTLs or TWAS data from this study or others did not support a role for CENPU in BCF or diabetes-related traits. Single-cell RNA sequencing (RNA-seq) datasets specifically highlighted ACSL1 as relevant in human pancreatic endocrine cells (Extended Data Fig. 8). Based on the convergence of islet 3D-chromatin interactions, eQTLs and single-cell RNA-seq data, ACSL1 was nominated as a strong molecular mediator of this GWAS association. Earlier work demonstrated that silencing two ACSL1 paralogs impairs insulin release by altering the fatty acid composition of insulin granules and mitochondrial function51. We thus silenced ACSL1 mRNA using small interfering RNA (siRNA) in EndoC-βH1 and INS-1 832/13 cell lines (Extended Data Fig. 9a,b), resulting in impaired insulin release in response to stimulation by glucose, pyruvate and KCl (Fig. 4c–g). Acyl-CoAs have long been promoted as important metabolic coupling factors regulating fuel-mediated insulin secretion52. Hence, the abrogated formation of acyl-CoA upon ACSL1 knockdown was expected to inhibit insulin release, as we observed. Thus, the human genetic evidence presented here supports this widely held53, though only partially proven, mechanism of insulin release regulation54. We also examined the adjacent gene Cenpu by silencing it in INS-1 832/13 cells, which resulted in a significant decrease in insulin release and total insulin content, along with abnormal cell morphology (Extended Data Fig. 9e,f). However, Cenpu gene silencing did not affect fractional insulin release when adjusted for insulin content, unlike Acsl1 (Extended Data Fig. 9c,d,g,h). Given that CENPU encodes for centromere protein U, which holds a crucial role in cell proliferation, we presume that CENPU modulates growth or differentiation in the immortalized β-cell line rather than stimulus-secretion coupling. Collectively, cellular, genetic and epigenomic evidence supports the role of ACSL1 as the most likely molecular mediator of this BCF association, offering the potential for diabetes-modifying strategies.

Fig. 4: Transcriptional regulation of ACSL1 influences BCF, glucose metabolism and diabetes.
figure 4

a, Regional signal plot at the ACSL1 locus, showing P values on a −log10 scale (y axis) in hg19 locations (x axis) from DIBIG single-variant GWAS meta-analyses. Variants are coloured by their LD correlation (r2) with the lead variant (rs10022124). Below, epigenomic datasets are shown, including ATAC-seq and ChIP-seq data in human pancreatic islets. Enhancer-gene assignments from pcHi-C data are represented as pink arcs; those inferred using pcHi-C data are in purple. Fine-mapped DI and DIBIG signals located in islet active enhancers connected to ACSL1 and CENPU, or identified as T2D-colocalizing islet ACSL1 eQTLs, are highlighted. b, Forest plot of the strongest functionally prioritized signal at the ACSL1 locus (rs4862423, DIBIG). For each of the BCF traits, glycaemic levels and T2D risk, the square represents the β estimate with 95% CI error bars. The square size reflects precision. The effect allele and variant RSID are provided at the top. Summary statistics data were generated in this study or obtained from previous publications49,50 and FinnGen (release v.8). OR, odds ratio. c,d, Insulin secretion in human EndoC-β H1 cells following ACSL1 siRNA silencing in response to high glucose (c) and KCl stimuli (d). eg, Insulin secretion in INS-1 832/13 cells upon Acsl1 gene silencing in response to high glucose (e), KCl (f) and pyruvate (g). Bar plots show the means of three (in c and d) and four (in eg) independent replicates; error bars, s.e.m. Statistical significance was assessed using one-way ANOVA followed by Tukey’s post hoc test of each condition against the negative control group (scrambled siRNA).

Source data

At the BIGTT-AIR TENT5C-DT locus (also known as LOC100996263), TWAS identified associations between FAM46C expression (encoding a member of the nucleotidyltransferase gene superfamily) and several BCF indices (Supplementary Fig. 5). Fine-mapping of BIGTT-AIR signals prioritized rs1975283 (PP = 0.028, LD with GWAS-lead rs320370r2 = 0.83), which also leads eTWAS associations between BCF traits and FAM46C expression. This variant resides within an islet enhancer connected to FAM46C through chromatin-based enhancer-gene assignments (Extended Data Fig. 10a). Strong colocalization between BCF traits and T2D was detected at the FAM46C locus (Supplementary Table 8). The rs1975283-A allele was associated with increased FAM46C islet expression, negative effects on BCF, higher T2D risk (odds ratio, 1.03, P = 3.51 × 10−8) and elevated blood glucose levels (β = 0.01, P = 2.29 × 10−18)55,56 (Extended Data Fig. 10b–d and Supplementary Table 13). This epigenomic and transcriptomic evidence supports FAM46C as a strong candidate effector for this locus. The polyadenylation activity of FAM46C has been linked to enhanced expression of endoplasmic reticulum-targeted proteins57, concordant with the role of FAM46 proteins in RNA stabilization58. To delve into the effects of FAM46C on BCF, we silenced FAM46C mRNA in EndoC-βH1 and INS-1 832/13 cells (Extended Data Fig. 10e,h). This resulted in a significant inhibition of insulin release in response to all the secretagogues tested (Extended Data Fig. 10f,g,i–n). Consistently, the International Mouse Phenotyping Consortium reported mild glucose intolerance for female heterozygous Tent5c knock-out mice (area under glucose-response curve, wild-type 11,408.5 ± 2,532.2 vs heterozygous 13,763.2 ± 2,011.7 min × mg dl−1, P = 0.04)59. FAM46C mRNA levels in 188 human islet samples showed significant positive correlations with glucagon (r = 0.65, P = 6.06 × 10−24) and IAPP (r = 0.68, P = 1.14 × 10−26) expression60, alongside the co-expression of genes relevant for insulin exocytosis (SYT13, SYT16, STXBP4), proinsulin processing (PCSK1, CPE) and β-cell differentiation (NEUROD1)83 (Supplementary Table 18). These results underscore the important role of FAM46C in BCF and glucose homeostasis. However, further research is needed to fully elucidate its precise role in the insulin secretory function.

Altogether, this study offers a more comprehensive perspective of the genetic architecture of oral glucose-stimulated insulin release. Our data show that the eight BCF indices provide both shared and specific contributions to the genetics of β-cell secretory function. We report intricate patterns of shared genetic effects on BCF, T2D risk, BMI and glucose metabolism-relevant traits. Our results expand the repertoire of effector genes potentially influencing T2D risk through insulin secretory pathways, highlighting ACSL1 and FAM46C as additional BCF regulators. These findings advance the understanding of glucose-stimulated β-cell response and offer promising avenues for drug discovery in diabetes.

Methods

Ethics statement

All contributing studies included in this analysis were approved by relevant ethical committees and participants provided informed consent. Specific ethics statements for each study are provided in Supplementary Table 2.

Statistics and reproducibility

To investigate the genetic architecture of the BCF, we conducted multiple analyses, including single-variant association studies across nine cohorts assembled in eight meta-analyses for each of the BCF traits. We examined the independence of the eight BCF traits through genetic and phenotypic correlations, heterogeneity tests of the effect sizes and pleiotropic effects using colocalization, Mendelian randomization and by assessing BCF allelic effects in other relevant complex traits. To identify molecular effectors for BCF-GWAS loci, we carried out heritability enrichment analyses across tissue-specific chromatin annotations and cell-type gene expression data. We prioritized candidate target genes using epigenomic data and genetic effects on transcriptome activity. Finally, we assessed the role of selected candidate target genes in insulin secretion using gene silencing in human and rat cell lines. Detailed descriptions of all analyses are provided below.

BCF estimates

This study examined the genetic basis of eight BCF estimates derived from a 75 g OGTT. Plasma glucose and serum insulin data at 0, 30 and 120 min during the OGTT were used to calculate the BCF indices, which include BIGTT-AIR (BIGTT-AIR0–30–120)19, CIR21, Stumvoll21, xinsdG30 (ref. 22), xinsG30 (ref. 24), DI23 and DIBIG19. We also incorporated HOMA-β25, derived from steady-state measures of fasting glucose and insulin. For mathematical formulations, see Supplementary Table 1. Even though multiple indices, such as disposition indexes, have emerged to model the complex hyperbolic relationship between β-function and insulin resistance, there is no consensus on the most appropriate estimate40,61,62,63. We selected DI, which is widely accepted in the literature, and DIBIG, developed by our lab.

We categorized these eight BCF measures into three groups: insulin secretion estimates during an OGTT, insulin secretion estimates during an OGTT adjusted for insulin sensitivity (that is, disposition indexes) and a surrogate BCF index based on steady-state fasting glucose and insulin measurements.

Study-specific analyses

BCF measures were analysed alongside genetic data from 26,356 individuals of Northern European descent, recruited across nine cohorts. Participants were genotyped using various genome-wide single-nucleotide polymorphism (SNP) arrays (Supplementary Table 2). We implemented a quality control to each cohort for genotype data that excluded samples if the call rate was <95%, heterozygosity was greater than the median plus three times the interquartile range, if there was discordance between genetically inferred and self-reported sex, duplicates and outliers identified using PCA projection of the study samples onto 1 KG reference samples. Variants were excluded if the variant call rate was <98%, there was a departure from Hardy–Weinberg equilibrium of P < 105 (or <10−4 if the data were genotyped in batches; also by inspecting Hardy–Weinberg P value distribution to minimize peaks of low P values) and monomorphic markers. Variants were imputed using the Michigan Imputation Server (https://imputationserver.sph.umich.edu)64 with the HRC reference panel (GRCh37 genome build, HRC r1.1 2016 release)65 and Eagle (v.2.3) for haplotype phasing. For the PPP-Botnia Study, genotype imputation was performed using a workflow linked to the THL Biobank’s SISu v.3 WGS imputation reference panel (https://doi.org/10.17504/protocols.io.nmndc5e).

For each study, we used linear models to test genetic associations with BCF indices using inverse normal rank-transformed values adjusted by BMI, age, age2 and sex. No sex-specific analyses were performed, as the aim of this study was to map the genetic architecture of glucose-stimulated insulin secretory function independently of sex. Covariates used in BCF calculations, such as BMI and sex in BIGTT-AIR and DIBIG, were not included in the adjustments. To account for population structure and relatedness, association testing was also adjusted by genetic principal components calculated with PLINK66 (Supplementary Table 2) and kinship using a genetic relationship matrix (Balding–Nicols model) computed with vcf2kinship (https://github.com/zhanxw/rvtests/blob/master/vcfUtils/vcf2kinship.cpp). Before calculating the principal components, data pruning was performed with PLINK (v.1.9) using the settings (–indep 200 20 2–maf 0.01–thin 0.1)66. Additional association testing was conducted with linear models after removing related individuals (PI_HAT > 0.125) and adjusting for the same covariates as the linear mixed models (Supplementary Table 2). Individuals with known diabetes or those meeting the World Health Organization’s 2006 criteria for diabetes (fasting glucose of >7 mmol l−1 and 2 h glucose of ≥11.1 mmol l−1) were excluded before association testing. The Danish-based cohorts and PPP-Botnia Study were analysed using the RVTEST (v.20171009) software67, the Botnia Family Study was analysed using EPACTS (v.3.3.0)68 and METSIM was analysed using GTCA69. Complementary GWAS analyses were conducted without BMI adjustment.

Meta-analysis

Summary statistics from nine GWAS were combined using a fixed-effects meta-analysis with METAL (v.2011-03-25)70, using standard error-based weights for effect sizes. Variants with multiallelic sites, a minor allele frequency (MAF) of <0.001 (<0.01 for the GDM offspring cohort) and imputation quality of <0.6 were excluded. Genomic control was applied to standard errors to address between-study structure. METAL used Cochran's Q-test to assess heterogeneity, excluding variants with heterogeneity of >75% and n × MAF > 43. Genomic inflation was estimated using variants with MAF × n > 50, MAF > 1% and MAF > 5% (Supplementary Figs. 13). Consistent results were observed across all MAF filters. GWAS genomic loci were defined using the IRanges package in R (v.3.6.1) as non-overlapping 500 kb windows (merged otherwise) with a GWAS-significant variant (P ≤ 5 × 10−8). For multiple-test correction across eight BCF traits, significance was set at P ≤ 6.25 × 109. BCF-GWAS loci were named based on the nearest gene.

Detection of independent BCF signals

Conditional analysis to identify independent signals at each genome-wide significant BCF was performed using genome-wide complex trait analysis (GCTA (v.1.91.2 beta)71, GCTA-COJO). The Inter99 cohort (Supplementary Table 2) served as the reference dataset to model LD structure, excluding SNPs with low imputation quality (MINIMAC R2 < 0.6) and individuals with relatedness (PI_HAT > 0.125). We examined genome-wide significant variants within a ~250 kb range centred on the lead BCF signal. Signals with P < 10−5 were considered independent. Conditional analysis was iterated until no additional independent variants were found, with reciprocal conditional analysis confirming each additional signal. The final set of independent BCF variants was determined by clumping all independent lead variants across the eight BCF estimates using the R package TwoSampleMR with an R2 cut-off of 0.05, using individuals of European descent as the reference dataset.

Assessing differences between various estimates of BCF

Phenotype correlations between the eight BCF estimates were calculated at the study-specific level using Pearson correlation with rank-based inverse normal transformed values, adjusting for sex, age and BMI. Diabetic individuals were removed from this analysis. Phenotype correlations were aggregated in a meta-analysis conducted in R (v.3.4.2) using the meta package (v.4.9-2)72, applying a fixed-effect model based on Fisher’s Z-transformation of the correlation coefficients.

LD score (LDSC) regression was used to calculate pairwise genetic correlations between the eight BCF estimates based on the proportion of shared genetic variance73,74 (https://github.com/hillfung/rg-matrix). Following the developers’ guidelines, calculations were based on GWAS data without genomic control, using association estimates derived from linear models.

For clustering BCF estimates based on genetic and phenotypic correlations, we used complete hierarchical clustering with Euclidean distance, as implemented in the R (v.3.6.1) stats package.

The heterogeneity of the effect sizes between BCF estimates was assessed using estimated standardized differences:

$$\frac{{\rm{effect}}_{{t}_1}-{\rm{effect}}_{{t}_2}}{\sqrt{{\rm{{stderror}}_{{t}_1}}^{2}+{{\rm{stderror}}_{{t}_2}}^{2}-\left(2\times \rho \times {\rm{stderror}}_{{t}_1}\times {\rm{stderror}}_{{t}_2}\right)}}$$
(1)

Here, t1 and t2 represent the two traits being compared, ‘effect’ is the effect size, ‘stderror’ is the standard error of the effect size and ρ is the genetic correlation between the two traits, which was calculated considering all variants and using Pearson correlation as implemented in the R (v.4.0.2) stats package.

Known BCF-GWAS loci

We classified a BCF locus as ‘previously known’ if it met any of the following criteria: (1) association with any of the eight BCF estimates listed in the National Human Genome Research Institute–European Bioinformatics Institute (NHGRI-EBI) GWAS Catalog75 (Supplementary Table 7), (2) association reported with any of the eight BCF traits through T2D candidate gene studies or (3) identification as a T2D genetic association linked to a β-cell dysfunction genetic cluster or a fasting proinsulin genetic signal. For criterion (3), islet functional evidence was required to confirm known BCF associations.

The NHGRI-EBI GWAS Catalog was accessed on 30 January 2023 and lifted from GRCh38 to GRCh37 using LiftOver (https://genome.ucsc.edu/cgi-bin/hgLiftOver). We investigated two sets of variants: those with a P value of <5 × 10−8 or with a P value of <5 × 106 in LD with BCF lead variants (R2 > 0.6 or R2 > 0.1, respectively). Proxies were identified using the LDlinkR R (v.4.1.0) package and European individuals as the reference dataset76. We also checked LD (R2 > 0.1) using LDLINK (https://ldlink.nih.gov) between BCF lead variants of this study and those from known GWAS of insulin secretory function or related traits that account for islet functional evidence. Details on the different datasets used are provided in Supplementary Table 19. BCF loci not previously reported in GWAS datasets for BMI, T2D or glycaemic-related traits were further investigated using the T2D Knowledge Portal (https://t2d.hugeamp.org) and the MR-Base platform (https://gwas.mrcieu.ac.uk/phewas)77,78.

BCF heritability and the relationship with other traits

Genetic correlations with relevant traits were calculated using LDHub (http://ldsc.broadinstitute.org). Metabolic-related traits were investigated together with brain-related traits owing to the neuroendocrine origin of β-cells. We also excluded traits with negative heritability, those that are duplicated and traits connected to the following categories: (1) education, (2) cancer, (3) smoking and alcohol, (4) personality, (5) reproductive system, (6) skin and nails, (7) uric acid, (8) metal, (9) medication, supplements and consumption, (10) eyes and vision, (11) certain infectious and parasitic diseases, (12) other, (13) aging, (14) bone, (15) autoimmune and inflammatory, (16) sleeping, (17) lung system and (18) physical activity.

Only GWAS summary statistics data from individuals of European descent were included. LDHub leverages LDSC regression; therefore, we used the same GWAS summary statistics used for genetic correlation calculations. Multiple testing correction was applied based on eight main categories of complex traits (P < 0.05 in seven categories: P = 0.00714).

Heterogeneity tests between BMI-adjusted and BMI-unadjusted models were assessed as described in equation 1.

For the study-specific heatmap, we selected the strongest non-palindromic association for each BCF locus that was significant for most traits and had GWAS data available for BCF, T2D and BMI (LD R2 > 0.7 with BCF variant). Effect sizes were aligned to the BCF-lowering allele. Proxies could not be identified for two loci with palindromic sequences: LOC100996263 and PROX1-AS1. The allelic effects of the final set of 42 qualifying BCF signals were examined using GWAS datasets for T2D, BMI and other relevant traits (Supplementary Table 20).

GWAS summary statistics for body fat percentage were calculated for individuals of European ancestry from UK Biobank using REGENIE (v.3.0)79. Body fat percentage was rank-based inverse normal transformed per sex. The following covariates were used for association testing: age, age2, PC1-10, sex, chip and centre.

GWAS summary statistics for body incretins were calculated as follows: blood samples for measurement of GIP and GLP-1 were drawn into EDTA tubes and immediately put on ice. Radioimmunological determinations of total plasma GIP and GLP-1 concentrations were performed as previously described80,81,82,83. The analytical detection limit was 1 pmol l−1 and intra-assay and inter-assay coefficients of variation were <6% and <15%, respectively. The incremental areas under the curves (iAUCs) of plasma GIP and GLP-1 concentrations were calculated using the trapezoid rule, from 0 to 30 min (iAUC 0–30 min) and 0 to 120 min (iAUC 0–120 min) during an OGTT. Participants were genotyped by the Illumina Infinium HumanCoreExome Beadchip platform. Genotypes were called using the Genotyping module (v.1.9.4) of GenomeStudio software (v.2011.1; Illumina). Closely related individuals, individuals with an extreme inbreeding coefficient, individuals with mislabelled sex, individuals with a call rate of <95%, duplicates and individuals identified as ethnic outliers as well as genetic markers with a call rate of <98%, MAF < 0.01 and a Hardy–Weinberg equilibrium P value of <1 × 10−5 were removed during quality control. Imputation was performed based on the 2,184 haplotypes available from the 1000 Genomes Project Phase 1, version 3 (ref. 84) using IMPUTE85. Associations between genetic variants and plasma incretin hormone levels were studied using a linear mixed model (EMMAX) implemented in the EPACTS software package86 using inverse-normalized residuals of the trait adjusted for age, sex and BMI in each cohort separately. Individuals with missing information were excluded from the analysis.

Colocalization analysis with T2D, fasting glucose and BMI

Colocalization was conducted using the coloc package (v.5.2.2) in R (v.4.3.1) with default settings. We analysed the fraction of genetic variants with MAF > 0.001 within a 500 kb window around each BCF lead variant. For T2D colocalization analysis, we harnessed a large-scale European-based T2D meta-analysis55. Genetic variants were aligned with those in our BCF-GWAS meta-analysis. The same approach was applied to BMI87 and BMI-adjusted fasting glucose49 colocalization analyses.

Mendelian randomization

Mendelian randomization was conducted using the TwoSampleMR package in R (v.4.3.1; 2023-06-16)78,88 with summary statistics from a recent large-scale European-based T2D meta-analysis55. T2D GWAS data were harmonized to align with BCF-GWAS summary statistics. Steiger filtering was applied to select BCF genetic variants that explained more variance (r2) in BCF than in T2D.

Fine-mapping of variants associated with BCF estimates

We fine-mapped 500 kb genomic regions centred on each BCF-independent signal, including all variants with LD r2 ≥ 0.1 with the lead SNP. LD calculations were based on individuals of European descent from the Genomes Project (http://www.1000genomes.org)84. Summary statistics from the conditional analysis were used to fine-map independent signals identified in subsequent conditional analyses. Credible sets were built using a Bayesian approach89. For each variant, the approximate Bayes’ factor was calculated based on effect size and standard error. PPs were then computed, variants were ranked in descending order of approximate Bayes’ factor and cumulative PPs were generated. Genetic signals were included in the credible set until cumulative PP exceeded 99%.

Heritability enrichments on chromatin annotations

To estimate the enrichment for BCF heritability in functional genomic regions, we used stratified LDSC regression (S-LDSC)90. For the sake of comparison, in addition to our BCF-GWAS data, we also included additional traits connected to pancreatic islet cell biology as well as a broad range of complex traits and diseases (https://alkesgroup.broadinstitute.org/LDSCORE/all_sumstats). S-LDSC was performed across epigenomic annotations comprising enhancer and promoter elements identified in human pancreatic islets but also across a wide range of tissues and cell types. For pancreatic islets, we selected active enhancer and active or bivalent promoter states from a pancreatic islet regulome published elsewhere13. For other tissues and cell types, we relied on chromHMM states generated by EpiMap (https://personal.broadinstitute.org/cboix/epimap/ChromHMM/observed_aux_18_hg19/CALLS)35. We overlaid each set of enhancer and promoter annotations identified in each of the selected tissues with 1000 Genomes Phase 3 variants, as provided by S-LDSC authors. We next generated binary annotations and custom LDSC files. For each trait, we estimated heritability enrichments relative to the annotations included in the baseline LD model (v.2.2). We computed two-sided P values for per-SNP heritability Z-scores. Enrichment P values and per-SNP heritability P values were corrected for multiple testing using the Benjamini–Hochberg procedure (α = 0.01).

Cell-type gene expression enrichments using CELLECT

Cell-type-specific gene expression enrichments were calculated using the CELL-type Expression-specific integration for Complex Traits (CELLECT)-LDSC91, based on BCF-GWAS summary statistics and single-cell RNA-seq data from human donors37 and the Tabula Muris36. For the Tabula Muris dataset, the following tissues were excluded: lung, mammary, heart, thymus, trachea, bladder, skin, spleen, tongue and marrow.

Target gene identification using epigenomic and chromatin interaction datasets

We leveraged 99% credible sets for all our BCF sentinel SNPs and overlaid them with islet regulome annotations13. Fine-mapped variants located in islet promoter elements were directly connected to their target genes through transcription start site overlap. For fine-mapped variants within distal regulatory elements, we used previously identified enhancer-gene assignments13 to establish links to their target genes.

Target gene identification using transcriptomic datasets

Candidate genes for BCF traits were also identified by (1) TWAS, (2) colocalization between eQTLs or sQTLs and BCF associations and (3) the overlap between fine-mapped eQTLs and sQTLs and fine-mapped BCF variants. For approach (3), we overlaid BCF credible set variants with fine-mapped eQTL and sQTL variants42. For approach (1), we conducted TWAS as in FUSION43. We used the FUSION_assoc_test.R script to test associations between imputed gene expression (or splicing activity) and each BCF trait, respectively. We advocated for the FUSION software because of its ability to test expression or splicing trait associations using pre-computed SNP weights for gene expression or splicing activity (trained in a panel of 399 human islet samples with genotype and transcriptome data available42) and BCF-GWAS summary statistics data. This avoids using individual-level data to impute gene expression or splicing junction activity in our BCF-GWAS cohorts. We filtered the BCF-GWAS summary statistics data by the genomic coverage of the panel used to train per-gene predictive models of genetic effects on islet gene expression and splicing junction usage (~6.5 common genetic variants). Multiple-testing correction for the TWAS P values was applied using Bonferroni correction, considering 2,851 and 5,804 features for expression and alternative splicing, respectively. For approach (2), we also relied on the FUSION software that allowed us to implement colocalization analysis while performing TWAS. This approach looks for evidence of a shared causal genetic association between each BCF trait and islet gene expression or splicing activity. FUSION uses the coloc software to compute approximate colocalization statistics based on the marginal FUSION weights. We used the median sample size across all variants tested for each BCF study (–GWASN) and set–coloc_P to 0.00001. Note that the TWAS results are unfitted to distinguish causal relationships from linkage92. Consequently, we discarded TWAS signals with low colocalization evidence and are hereby confounded by linkage (coloc PP4 < 0.6). Finally, TWAS associations were linked to BCF-GWAS loci if the TWAS lead variant was in LD (r2 ≥ 0.1, using the genotypes of individuals with European descent from Phase 3 of the 1000 Genomes Project) or within 500 kb from any of the BCF sentinel SNPs.

To visualize colocalization between islet eQTLs and BCF associations at selective GWAS loci or TWAS associations, we used LocusCompareR (v.1.0.0) in R (v.3.6.1). If the LD (r2) correlation information for the lead QTL or TWAS association was not available in pre-calculated LD datasets, we used PLINK66 (v.1.9) with the settings –ld-window-kb 1000–ld-window 99999–ld-window-r2 0, and genotypes from 399 human islet samples used in the TWAS analysis as reference dataset. eQTL P values for the ITFG3 eTWAS association were computed here using genotype and RNA-seq data from 399 human islet samples42, as this signal was missed in the original eQTL mapping analysis. We applied the same eQTL mapping guidelines as in the original study. eQTLs for ITFG3 have also been identified in other islet eQTL mapping studies, including rs56314769 (ref. 12) and rs6600191(ref. 48), which are in LD with the lead ITFG3 eTWAS association (rs56038902r2 = 0.6 and r2 = 0.8, respectively).

We observed that pcHi-C assignments supported three and five candidate genes identified by eTWAS and sTWAS, respectively. Limited overlap between BCF candidate genes identified through eTWAS and pcHi-C approaches could be partially explained by the limited discovery power of eQTL mapping on GWAS-relevant traits, as recently discussed93. Likewise, the ability of pcHi-C assignments to identify BCF target genes from credible set variants is hampered by several factors, including stringent significance thresholds for chromatin loop calling, distance to the target gene, enhancer discovery in the disease-relevant cell state and the resolution of genetic fine-mapping to prioritize causal regulatory variants. Therefore, using complementary functional approaches could surmount the intrinsic technical limitations of each individual strategy.

Visualization of regional association plots

Regional association plots for BCF susceptibility regions were created using LocusZoom94 (v.1.4) in R (v.3.6.1).

Visualization of splicing events

IGV95 was used to visualize islet splice junctions identified elsewhere42 as arcs.

Assessing the impact of non-coding fine-mapped variants on transcription factor binding

We used motifbreakR96 to predict the impact of BCF fine-mapped variants located in islet regulatory elements on transcription factor binding. Human motif information was sourced from the HOCOMOCO and Jaspar databases via MotifDb. The parameters used where filterp = TRUE, threshold = 5 × 10−5, method = ‘ic’. We repeated the calculations using nucleotide background frequencies estimated in islet enhancers (A = 0.270182, C = 0.2290216, G = 0.2297711, T = 0.2710253). Results were filtered for ‘strong’ effects with a score greater than 0.8 for any of the alleles tested (pctRef > 0.8 | pctAlt > 0.8).

Enrichment of functional annotations among candidate target genes

We used the R interface of enrichR97 (last accessed on 13 June 2023) to identify relevant functional annotations enriched by the protein-coding molecular effector transcripts nominated using islet epigenomic and transcriptomic datasets.

Cell culture

The EndoC-βH1 cell line (MTA from Endocells Sarl) is derived from female human foetal pancreatic buds implanted into SCID mice, maturing into pancreatic tissue and eventually giving rise to insulinoma cells after differentiation. The sex was not considered in the choice of the cell line sources. EndoC-βH1 cells were cultured as previously described98 on Matrigel-fibronectin-coated culture vessels (Matrigel at 100 μg ml−1 and fibronectin at 2 μg ml−1, both from Sigma-Aldrich) in DMEM supplemented with 5.6 mM glucose, 2% BSA fraction V (Roche Diagnostics), 10 mM nicotinamide (Merck Millipore), 50 μM 2-mercaptoethanol, 5.5 μg ml−1 transferrin and 6.7 ng ml−1 sodium selenite (Sigma-Aldrich). The INS-1 832/13 cell line is male and derived from the rat insulinoma cell line INS-1, following stable transfection with a plasmid containing the human proinsulin gene. One of the authors (H.M.) is a co-creator of the INS-1 832/13 cell line and has unrestricted access to it99. INS-1 832/13 cells were cultured in RPMI-1640 containing 11.1 mM glucose, 10% fetal bovine serum, 10 mM HEPES, 2 mM glutamine, 1 mM sodium pyruvate and 50 μM β-mercaptoethanol as previously described99. Both cell lines were cultured at 37 °C with 5% CO2. INS-1 832/13 cells were previously authenticated through PCR analysis of Ins1 and Pdx1 mRNAs. EndoC-βH1 cells were also previously authenticated through PCR analysis of insulin and β-cell transcription factors PDX1, PAX6, FOXA2, NEUROD1, MAFB and NKX6.1.

Gene silencing

Gene silencing of ACSL1, FAM46C and CENPU expression by RNA interference was performed in EndoC-βH1 and INS-1 832/13 insulin-producing cells. Lipofectamine RNAiMAX Reagent and 10 nM siRNAs specific for human ACSL1 and FAM46C, rat Acsl1, Fam46c and Cenpu, as well as a scrambled siRNA as negative control were used. After 72 h of transfection, gene expression and β-cell secretory activity were examined by quantitative PCR (qPCR) and insulin secretion assays, respectively.

Human ACSL1: Thermo Fisher Scientific; s129379 and s129380. Rat Acsl1: Thermo Fisher Scientific; s4994 and s4995. Human FAM46C: Thermo Fisher Scientific; s29596 and s29597. Rat Fam46c: Thermo Fisher Scientific; s160621 and s221157. Rat Cenpu (Thermo Fisher Scientific; s157528). Scrambled siRNA as a negative control for rat, custom select siRNA 50- GAGACCCUAUCCGUGAUUAUU-30. Scrambled siRNA as a negative control for human, Silencer Select Negative Control No. 1 siRNA (Thermo Fisher, 4390843).

mRNA expression analysis

Total RNA was extracted using the RNeasy Mini Kit (QIAGENiagen) and cDNA was generated with the RevertAid First-Strand cDNA synthesis kit. For qPCR, 10 ng of cDNA per well (in duplicates) was used with human ACSL1 and FAM46C, or rat Acsl1, Fam46c and Cenpu Taqman gene expression assays (Thermo Fisher Scientific; code: Rn01488229_m1, Rn01534138_m1, Rn01488384_m1, Hs00960561_m1, Hs01933465_s1). qPCR assays were performed using the Applied Biosystems QuantStudio7 Flex Real-Time PCR System (Thermo Fisher). Relative gene abundance was calculated using the ΔΔCt method with β-actin as the housekeeping reference gene. Gene expression values were reported as fold changes relative to the control.

Measurement of insulin secretion in EndoC-βH1 cells

The day before the experiment, EndoC-βH1 cells were switched to an overnight starvation medium containing 2.8 mM of glucose. On the experiment day, the starvation medium was replaced with HEPES-buffered Krebs–Ringer Buffer (115 mM NaCl, 24 mM NaHCO3, 5 mM KCl, 1 mM MgCl2, 1 mM CaCl2, 10 mM HEPES, 0.2% BSA pH 7.4) containing 1 mM of glucose for 1 h. The cells were then stimulated in KRB with either 1 mM or 20 mM glucose, or 35 mM KCl for an additional 1 h. Insulin secretion was measured using a human insulin ELISA (Mercodia) according to the manufacturer’s instructions.

Measurement of insulin secretion in INS-11 832/13 cells

INS-1 832/13 cells were starved for 2 h in a secretion assay buffer (114 mM NaCl, 1.2 mM KH2PO4, 4.7 mM KCl, 1.16 mM MgSO4, 2.5 mM CaCl2, 100 mM HEPES, 25 mM NaHCO3, 0.2% BSA pH 7.2) containing 2.8 mM glucose. Following starvation, the cells were stimulated for 1 h with secretion assay buffer containing either 2.8 mM or 16.7 mM glucose, 10 mM pyruvate or 35 mM KCl at 37 °C. Insulin concentration in the supernatants was measured by a high-range rat insulin ELISA (Mercodia, cat. no. 10-1145-01).

Normalization and statistical analyses

Insulin values for both INS-1 832/13 cells and EndoC-βH1 cells were normalized to the total protein and insulin content. Protein was extracted using RIPA buffer and measured with the BCA Protein Assay Kit. Total insulin content was measured in the same cell lysates to normalize insulin secretion to insulin content, which was then expressed as a percentage of secreted insulin. Analyses were conducted using GraphPad Prism software (v.9) and data are presented as means ± s.e.m. Statistical significance was assessed by Student’s t-test for single comparison and/or one-way ANOVA followed by Tukey’s post-hoc test for multiple comparisons. P values of <0.05 were considered statistically significant.

Reporting summary

Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.