Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

A non-coding variant linked to metabolic obesity with normal weight affects actin remodelling in subcutaneous adipocytes

Abstract

Recent large-scale genomic association studies found evidence for a genetic link between increased risk of type 2 diabetes and decreased risk for adiposity-related traits, reminiscent of metabolically obese normal weight (MONW) association signatures. However, the target genes and cellular mechanisms driving such MONW associations remain to be identified. Here, we systematically identify the cellular programmes of one of the top-scoring MONW risk loci, the 2q24.3 risk locus, in subcutaneous adipocytes. We identify a causal genetic variant, rs6712203, an intronic single-nucleotide polymorphism in the COBLL1 gene, which changes the conserved transcription factor motif of POU domain, class 2, transcription factor 2, and leads to differential COBLL1 gene expression by altering the enhancer activity at the locus in subcutaneous adipocytes. We then establish the cellular programme under the genetic control of the 2q24.3 MONW risk locus and the effector gene COBLL1, which is characterized by impaired actin cytoskeleton remodelling in differentiating subcutaneous adipocytes and subsequent failure of these cells to accumulate lipids and develop into metabolically active and insulin-sensitive adipocytes. Finally, we show that perturbations of the effector gene Cobll1 in a mouse model result in organismal phenotypes matching the MONW association signature, including decreased subcutaneous body fat mass and body weight along with impaired glucose tolerance. Taken together, our results provide a mechanistic link between the genetic risk for insulin resistance and low adiposity, providing a potential therapeutic hypothesis and a framework for future identification of causal relationships between genome associations and cellular programmes in other disorders.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: The pleiotropic 2q24.3 MONW locus is associated with increased risk for T2D and decreased adiposity-related traits, and maps to sparse enhancer signatures in adipocytes.
Fig. 2: rs6712203 is a functional variant at the 2q24.3 MONW locus.
Fig. 3: The 2q24.3 effector gene COBLL1 affects actin remodelling processes in differentiating adipocytes.
Fig. 4: The rs6712203 MONW risk haplotype affects actin remodelling in adipocytes and adipocyte lipid storage capacity.
Fig. 5: Cobll1-deficient mice are leaner and display metabolically dysfunctional phenotypes.

Similar content being viewed by others

Data availability

The ATAC–seq data for the immortalized human adipocytes are deposited in the database of Genotypes and Phenotypes under accession no. PRJNA539992. The high-content imaging data are available at the Cell Painting Gallery on the Registry of Open Data on Amazon Web Services (https://registry.opendata.aws/cellpainting-gallery/) under accession no. cpg0011. Source data are provided with this paper.

Code availability

The code used is publicly available on GitHub (https://github.com/ClaussnitzerLab/). The GitHub LP-2q24.3-metabolic-risk-locus repository containing the code and files used to generate LipocyteProfiles can be accessed at https://github.com/sophiestrobel/LP-2q24.3-metabolic-risk-locus.git.

References

  1. Cho, N. H. et al. IDF Diabetes Atlas: global estimates of diabetes prevalence for 2017 and projections for 2045. Diabetes Res. Clin. Pract. 138, 271–281 (2018).

    Article  CAS  PubMed  Google Scholar 

  2. Lu, Y. et al. New loci for body fat percentage reveal link between adiposity and cardiometabolic disease risk. Nat. Commun. 7, 10495 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Hosseinpanah, F., Barzin, M., Sheikholeslami, F. & Azizi, F. Effect of different obesity phenotypes on cardiovascular events in Tehran Lipid and Glucose Study (TLGS). Am. J. Cardiol. 107, 412–416 (2011).

    Article  PubMed  Google Scholar 

  4. Caleyachetty, R. et al. Metabolically healthy obese and incident cardiovascular disease events among 3.5 million men and women. J. Am. Coll. Cardiol. 70, 1429–1437 (2017).

    Article  PubMed  Google Scholar 

  5. Arnlöv, J., Ingelsson, E., Sundström, J. & Lind, L. Impact of body mass index and the metabolic syndrome on the risk of cardiovascular disease and death in middle-aged men. Circulation 121, 230–236 (2010).

    Article  PubMed  Google Scholar 

  6. Aung, K., Lorenzo, C., Hinojosa, M. A. & Haffner, S. M. Risk of developing diabetes and cardiovascular disease in metabolically unhealthy normal-weight and metabolically healthy obese individuals. J. Clin. Endocrinol. Metab. 99, 462–468 (2014).

    Article  CAS  PubMed  Google Scholar 

  7. Calori, G. et al. Prevalence, metabolic features, and prognosis of metabolically healthy obese Italian individuals: the Cremona Study. Diabetes Care 34, 210–215 (2011).

    Article  PubMed  Google Scholar 

  8. Wildman, R. P. et al. The obese without cardiometabolic risk factor clustering and the normal weight with cardiometabolic risk factor clustering: prevalence and correlates of 2 phenotypes among the US population (NHANES 1999–2004). Arch. Intern. Med. 168, 1617–1624 (2008).

    Article  PubMed  Google Scholar 

  9. Primeau, V. et al. Characterizing the profile of obese patients who are metabolically healthy. Int. J. Obes. 35, 971–981 (2011).

    Article  CAS  Google Scholar 

  10. Yaghootkar, H. et al. Genetic evidence for a normal-weight ‘metabolically obese’ phenotype linking insulin resistance, hypertension, coronary artery disease, and type 2 diabetes. Diabetes 63, 4369–4377 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Vujkovic, M. et al. Discovery of 318 new risk loci for type 2 diabetes and related vascular outcomes among 1.4 million participants in a multi-ancestry meta-analysis. Nat. Genet. 52, 680–691 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Spracklen, C. N. et al. Identification of type 2 diabetes loci in 433,540 East Asian individuals. Nature 582, 240–245 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Shungin, D. et al. New genetic loci link adipose and insulin biology to body fat distribution. Nature 518, 187–196 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Fox, C. S. et al. Genome-wide association for abdominal subcutaneous and visceral adipose reveals a novel locus for visceral fat in women. PLoS Genet. 8, e1002695 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Morris, A. P. et al. Large-scale association analysis provides insights into the genetic architecture and pathophysiology of type 2 diabetes. Nat. Genet. 44, 981–990 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Chu, A. Y. et al. Multiethnic genome-wide meta-analysis of ectopic fat depots identifies loci associated with adipocyte development and differentiation. Nat. Genet. 49, 125–130 (2017).

    Article  CAS  PubMed  Google Scholar 

  17. Wang, H. et al. Genotype-by-environment interactions inferred from genetic effects on phenotypic variability in the UK Biobank. Sci. Adv. 5, eaaw3538 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  18. Gao, C. et al. Genome-wide study of subcutaneous and visceral adipose tissue reveals novel sex-specific adiposity loci in Mexican Americans. Obesity 26, 202–212 (2018).

    Article  CAS  PubMed  Google Scholar 

  19. Lagou, V. et al. Sex-dimorphic genetic effects and novel loci for fasting glucose and insulin variability. Nat. Commun. 12, 24 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Chen, J. et al. The trans-ancestral genomic architecture of glycemic traits. Nat. Genet. 53, 840–860 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Loos, R. J. F. & Kilpeläinen, T. O. Genes that make you fat, but keep you healthy. J. Intern. Med. 284, 450–463 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Kilpeläinen, T. O. et al. Genetic variation near IRS1 associates with reduced adiposity and an impaired metabolic profile. Nat. Genet. 43, 753–760 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  23. Lotta, L. A. et al. Integrative genomic analysis implicates limited peripheral adipose storage capacity in the pathogenesis of human insulin resistance. Nat. Genet. 49, 17–26 (2017).

    Article  CAS  PubMed  Google Scholar 

  24. Kooner, J. S. et al. Genome-wide association study in individuals of South Asian ancestry identifies six new type 2 diabetes susceptibility loci. Nat. Genet. 43, 984–989 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. 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 et al. Genome-wide trans-ancestry meta-analysis provides insight into the genetic architecture of type 2 diabetes susceptibility. Nat. Genet. 46, 234–244 (2014).

    Article  Google Scholar 

  26. Heid, I. M. et al. Meta-analysis identifies 13 new loci associated with waist-hip ratio and reveals sexual dimorphism in the genetic basis of fat distribution. Nat. Genet. 42, 949–960 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Udler, M. S. et al. Type 2 diabetes genetic loci informed by multi-trait associations point to disease mechanisms and subtypes: a soft clustering analysis. PLoS Med. 15, e1002654 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Hindorff, L. A. et al. Potential etiologic and functional implications of genome-wide association loci for human diseases and traits. Proc. Natl Acad. Sci. USA 106, 9362–9367 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Claussnitzer, M. et al. Leveraging cross-species transcription factor binding site patterns: from diabetes risk loci to disease mechanisms. Cell 156, 343–358 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Claussnitzer, M. et al. FTO obesity variant circuitry and adipocyte browning in humans. N. Engl. J. Med. 373, 895–907 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Kelley, D. R., Snoek, J. & Rinn, J. L. Basset: learning the regulatory code of the accessible genome with deep convolutional neural networks. Genome Res. 26, 990–999 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Dixon, J. R. et al. Chromatin architecture reorganization during stem cell differentiation. Nature 518, 331–336 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Xue, R. et al. Clonal analyses and gene profiling identify genetic biomarkers of the thermogenic potential of human brown and white preadipocytes. Nat. Med. 21, 760–768 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Gilbert, L. A. et al. CRISPR-mediated modular RNA-guided regulation of transcription in eukaryotes. Cell 154, 442–451 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Cowper-Sal lari, R. et al. Breast cancer risk-associated SNPs modulate the affinity of chromatin for FOXA1 and alter gene expression. Nat. Genet. 44, 1191–1198 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Benita, Y. et al. Gene enrichment profiles reveal T-cell development, differentiation, and lineage-specific transcription factors including ZBTB25 as a novel NF-AT repressor. Blood 115, 5376–5384 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Kawaguchi, N. et al. ADAM12 induces actin cytoskeleton and extracellular matrix reorganization during early adipocyte differentiation by regulating β1 integrin function. J. Cell Sci. 116, 3893–3904 (2003).

    Article  CAS  PubMed  Google Scholar 

  38. Qian, Y. et al. PI3K induced actin filament remodeling through Akt and p70S6K1: implication of essential role in cell migration. Am. J. Physiol. Cell Physiol. 286, C153–C163 (2004).

    Article  CAS  PubMed  Google Scholar 

  39. Morandi, E. M. et al. ITGAV and ITGA5 diversely regulate proliferation and adipogenic differentiation of human adipose derived stem cells. Sci. Rep. 6, 28889 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Truebestein, L., Elsner, D. J., Fuchs, E. & Leonard, T. A. A molecular ruler regulates cytoskeletal remodelling by the Rho kinases. Nat. Commun. 6, 10029 (2015).

    Article  CAS  PubMed  Google Scholar 

  41. Izadi, M. et al. Cobl-like promotes actin filament formation and dendritic branching using only a single WH2 domain. J. Cell Biol. 217, 211–230 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Takayama, K.-I., Suzuki, T., Fujimura, T., Takahashi, S. & Inoue, S. COBLL1 modulates cell morphology and facilitates androgen receptor genomic binding in advanced prostate cancer. Proc. Natl Acad. Sci. USA 115, 4975–4980 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Laber, S. et al. Discovering cellular programs of intrinsic and extrinsic drivers of metabolic traits using LipocyteProfiler. Preprint at bioRxiv https://doi.org/10.1101/2021.07.17.452050 (2021).

  44. Harris, R. B. S. Direct and indirect effects of leptin on adipocyte metabolism. Biochim. Biophys. Acta 1842, 414–423 (2014).

    Article  CAS  PubMed  Google Scholar 

  45. Kilpeläinen, T. O. et al. Genome-wide meta-analysis uncovers novel loci influencing circulating leptin levels. Nat. Commun. 7, 10494 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Folkersen, L. et al. Genomic and drug target evaluation of 90 cardiovascular proteins in 30,931 individuals. Nat. Metab. 2, 1135–1148 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Small, K. S. et al. Regulatory variants at KLF14 influence type 2 diabetes risk via a female-specific effect on adipocyte size and body composition. Nat. Genet. 50, 572–580 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Fathzadeh, M. et al. FAM13A affects body fat distribution and adipocyte function. Nat. Commun. 11, 1465 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Randall, J. C. et al. Sex-stratified genome-wide association studies including 270,000 individuals show sexual dimorphism in genetic loci for anthropometric traits. PLoS Genet. 9, e1003500 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Sung, Y. J. et al. Genome-wide association studies suggest sex-specific loci associated with abdominal and visceral fat. Int. J. Obes. 40, 662–674 (2016).

    Article  CAS  Google Scholar 

  51. Palmer, B. F. & Clegg, D. J. The sexual dimorphism of obesity. Mol. Cell. Endocrinol. 402, 113–119 (2015).

    Article  CAS  PubMed  Google Scholar 

  52. Chen, Z. et al. Functional screening of candidate causal genes for insulin resistance in human preadipocytes and adipocytes. Circ. Res. 126, 330–346 (2020).

    Article  CAS  PubMed  Google Scholar 

  53. Chen, L., Hu, H., Qiu, W., Shi, K. & Kassem, M. Actin depolymerization enhances adipogenic differentiation in human stromal stem cells. Stem Cell Res. 29, 76–83 (2018).

    Article  CAS  PubMed  Google Scholar 

  54. Kanzaki, M. & Pessin, J. E. Insulin-stimulated GLUT4 translocation in adipocytes is dependent upon cortical actin remodeling. J. Biol. Chem. 276, 42436–42444 (2001).

    Article  CAS  PubMed  Google Scholar 

  55. Kim, J. I. et al. During adipocyte remodeling, lipid droplet configurations regulate insulin sensitivity through F-actin and G-actin reorganization. Mol. Cell. Biol. 39, e00210–e00219 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Hansson, B. et al. Adipose cell size changes are associated with a drastic actin remodeling. Sci. Rep. 9, 12941 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  57. Li, Q. et al. Obesity and hyperinsulinemia drive adipocytes to activate a cell cycle program and senesce. Nat. Med. 27, 1941–1953 (2021).

    Article  CAS  PubMed  Google Scholar 

  58. Bäckdahl, J. et al. Spatial mapping reveals human adipocyte subpopulations with distinct sensitivities to insulin. Cell Metab. 33, 1869–1882 (2021).

    Article  PubMed  Google Scholar 

  59. Emont, M. P. et al. A single-cell atlas of human and mouse white adipose tissue. Nature 603, 926–933 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Cariou, B. et al. Increased adipose tissue expression of Grb14 in several models of insulin resistance. FASEB J. 18, 965–967 (2004).

    Article  PubMed  Google Scholar 

  61. Béréziat, V. et al. Inhibition of insulin receptor catalytic activity by the molecular adapter Grb14. J. Biol. Chem. 277, 4845–4852 (2002).

    Article  PubMed  Google Scholar 

  62. Depetris, R. S. et al. Structural basis for inhibition of the insulin receptor by the adaptor protein Grb14. Mol. Cell 20, 325–333 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Cooney, G. J. et al. Improved glucose homeostasis and enhanced insulin signalling in Grb14-deficient mice. EMBO J. 23, 582–593 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Wang, L. et al. Peripheral disruption of the Grb10 gene enhances insulin signaling and sensitivity in vivo. Mol. Cell. Biol. 27, 6497–6505 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Holt, L. J. et al. Dual ablation of Grb10 and Grb14 in mice reveals their combined role in regulation of insulin signaling and glucose homeostasis. Mol. Endocrinol. 23, 1406–1414 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Shi, J. et al. Discovery of cancer drug targets by CRISPR–Cas9 screening of protein domains. Nat. Biotechnol. 33, 661–667 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. González, F. et al. An iCRISPR platform for rapid, multiplexable, and inducible genome editing in human pluripotent stem cells. Cell Stem Cell 15, 215–226 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  68. Fuchsberger, C. et al. The genetic architecture of type 2 diabetes. Nature 536, 41–47 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Albrechtsen, A. et al. Exome sequencing-driven discovery of coding polymorphisms associated with common metabolic phenotypes. Diabetologia 56, 298–310 (2013).

    Article  CAS  PubMed  Google Scholar 

  70. Wessel, J. et al. Low-frequency and rare exome chip variants associate with fasting glucose and type 2 diabetes susceptibility. Nat. Commun. 6, 5897 (2015).

    Article  CAS  PubMed  Google Scholar 

  71. Kan, M. et al. Rare variant associations with waist-to-hip ratio in European-American and African-American women from the NHLBI-Exome Sequencing Project. Eur. J. Hum. Genet. 24, 1181–1187 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  72. Claussnitzer, M. C. et al. Leveraging transcription factor binding site patterns: from diabetes risk loci to disease mechanisms. Cell 156, 343–358 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Hauner, H., Skurk, T. & Wabitsch, M. Cultures of human adipose precursor cells. Methods Mol. Biol. 155, 239–247 (2001).

    CAS  PubMed  Google Scholar 

  74. Ott, B. et al. Effect of caloric restriction on gut permeability, inflammation markers, and fecal microbiota in obese women. Sci. Rep. 7, 11955 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  75. Glastonbury, C. A. et al. Machine learning based histology phenotyping to investigate the epidemiologic and genetic basis of adipocyte morphology and cardiometabolic traits. PLoS Comput. Biol. 16, e1008044 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Buenrostro, J. D., Wu, B., Chang, H. Y. & Greenleaf, W. J. ATAC-seq: a method for assaying chromatin accessibility genome-wide. Curr. Protoc. Mol. Biol. 109, 21.29.1–21.29.9 (2015).

    PubMed  Google Scholar 

  77. Sobreira, D. R. et al. Extensive pleiotropism and allelic heterogeneity mediate metabolic effects of IRX3 and IRX5. Science 372, 1085–1091 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Sinnott-Armstrong, N. et al. A regulatory variant at 3q21.1 confers an increased pleiotropic risk for hyperglycemia and altered bone mineral density. Cell Metab. 33, 615–628 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Eastwood, S. V. et al. Algorithms for the capture and adjudication of prevalent and incident diabetes in UK Biobank. PLoS ONE 11, e0162388 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  80. Sinnott-Armstrong, N. Genetics of 35 blood and urine biomarkers in the UK Biobank. Nat. Genet. 53, 185–194 (2021).

    Article  CAS  PubMed  Google Scholar 

  81. Sanson, K. R. et al. Optimized libraries for CRISPR–Cas9 genetic screens with multiple modalities. Nat. Commun. 9, 5416 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Pairault, J. & Green, H. A study of the adipose conversion of suspended 3T3 cells by using glycerophosphate dehydrogenase as differentiation marker. Proc. Natl Acad. Sci. USA 76, 5138–5142 (1979).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Claussnitzer, M., Skurk, T., Hauner, H., Daniel, H. & Rist, M. J. Effect of flavonoids on basal and insulin-stimulated 2-deoxyglucose uptake in adipocytes. Mol. Nutr. Food Res. 55, S26–S34 (2011).

    Article  CAS  PubMed  Google Scholar 

  84. Untergasser, A. et al. Primer3—new capabilities and interfaces. Nucleic Acids Res. 40, e115 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Ye, J. et al. Primer-BLAST: a tool to design target-specific primers for polymerase chain reaction. BMC Bioinformatics 13, 134 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Pfaffl, M. W. A new mathematical model for relative quantification in real-time RT–PCR. Nucleic Acids Res. 29, e45 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Elliott, L. et al. The genetic basis of human brain structure and function: 1,262 genome-wide associations found from 3,144 GWAS of multimodal brain imaging phenotypes from 9,707 UK Biobank participants. Preprint at bioRxiv https://doi.org/10.1101/178806 (2017).

  88. Kundaje, A. et al. Integrative analysis of 111 reference human epigenomes. Nature 518, 317–330 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Chen, E. Y. et al. Enrichr: interactive and collaborative HTML5 gene list enrichment analysis tool. BMC Bioinformatics 14, 128 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  90. Kuleshov, M. V. et al. Enrichr: a comprehensive gene set enrichment analysis web server 2016 update. Nucleic Acids Res. 44, W90–W97 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Work in the Claussnitzer lab is supported by the Foundation for the National Institutes of Health (NIH) (AMP-T2D RFB8b), the National Institute of Diabetes and Digestive and Kidney Diseases (NIDDK) (UM1 DK126185 and NIDDK DK102173), the Novo Nordisk Foundation (NNF21SA0072102) and by a Next Generation Award from the Broad Institute of MIT and Harvard. We acknowledge support from the Center for Advanced Light Microscopy of the Technical University of Munich School of Life Sciences, T. Kufer at the University of Hohenheim for access to microscopes, the Clinical Cooperation Group ‘Nutrigenomics and Type 2 Diabetes’ from the Helmholtz Center Munich and the German Center for Diabetes Research, the Else Kröner-Fresenius-Foundation, a Stanford Graduate Fellowship, a Center for Computational, Evolutionary and Human Genomics Graduate Fellowship, the Novo Nordisk Foundation Challenge Grant (no. NNF18OC0033754), NIH grant nos. R01HL128075 and P30DK020595, the Federal Ministry of Education and Research (award no. 0315674), the German Research Foundation (award no. 338582098) and the Federal Ministry of Education and Research (grant no. 0315674). We thank A. Flaccus, A. Kozza, D. Mvondo, M. Hubersberger, E. Hofmair, J. Röttgen, C. Deuschle, A. Saadat and S. Harken for technical support.

Author information

Authors and Affiliations

Authors

Contributions

V.G. and M.C. conceptualized the study. V.G., S.L., N.S.-A., D.R.Sobreira, S.M.S., T.M.B., P.K., B.N.M., H.E., Y.H., B.B., G.G., J.H., D.R.S., N.A., V.C.-N., T.S., S.O., K.S., B.A.C., A.E.C., S.N.D., C.M.L., H.H., M.A.N. and M.C. devised the methodology. V.G., S.L., N.S.-A., D.R.Sobreira, S.M.S., T.M.B., P.K., B.N.M., H.E., Y.H., B.B., D.R.Stirling and B.A.C. carried out the formal analysis. V.G., S.L., N.S.-A., D.R.Sobreira, S.M.S., T.M.B., P.K., B.N.M., H.E., Y.H., B.B., D.R.Stirling and B.C. carried out the investigation. H.H., C.M.L., M.A.N. and M.C. obtained the resources. A.E.C., H.H., C.M.L., M.A.N. and M.C. supervised the study. H.H., C.M.L., M.A.N. and M.C. acquired the funding. V.G., N.S.-A., D.R.Sobreira, S.M.S. and M.C. wrote the original draft. V.G., S.L., N.S.-A., D.R.Sobreira, S.M.S., T.M.B., P.K., B.N.M., H.E., Y.H., B.B., G.G., J.H., D.R.Stirling, N.A., V.C.-N., T.S., S.O., K.S., B.C., A.E.C., S.N.D., C.M.L., H.H., M.A.N. and M.C. reviewed and edited the manuscript.

Corresponding author

Correspondence to Melina Claussnitzer.

Ethics declarations

Competing interests

M.C. holds equity in Waypoint Bio, consults for Pfizer, and is a member of the Nestle scientific advisory board. The authors have filed a provisional patent application (63/218,656).

Peer review

Peer review information

Nature Metabolism thanks Inês Cebola, Hannah Maude, Camilla Scheele and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editor: Ashley Castellanos-Jankiewicz, in collaboration with the Nature Metabolism team.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended data Fig. 1 The pleiotropic 2q24.3 MONW locus is associated with increased risk for type 2 diabetes and decreased adiposity related traits and maps to sparse enhancer signatures in adipocytes.

(a) Schematic overview for the 2q24.3 metabolic risk locus dissection. Aim of step (top, bold); methods/experiments used (middle); key finding/result of each step (bottom). (b) PheWAS of trait associations at the rs3923113-tagged haplotype of a meta-analysis https://t2d.hugeamp.org/. Colors represent trait classes while individual rs3923113 variant association p-values are shown on the Y axis. Direction of effect is indicated by orientation of triangles, upward: increase, downward: decrease (c) The 2q24.3 MONW locus spans 19 non-coding SNPs in high linkage disequilibrium with rs3923113 (LD r2 > 0.8). The region of association localizes to a >55 kb interval in an intergenic region between COBLL1 and GRB14. (d) Annotation panel and color key for the twenty-five state chromatin model70. Rows represent chromatin states abbreviations, columns are emission parameters, corresponding to the frequency with which each mark is expected in each state (left table) and genome coverage and median enrichments of relevant genomic annotations (right panel). TssA: Active TSS, TssAFlnk: Flanking Active TSS, TxFlnk: Transcription at gene 5’ and 3’, Tx: Strong Transcription, TxWk: Weak Transcription, EnhG: Genic enhancers: Enh: Enhancers, ZNF/Rpts: ZNF genes & repeats, Het: Heterochromatin, TssBiv: Bivalent/Poised TSS, BivFlnk: Flanking Bivalent TSS/Enhancer, EnhBiv: Bivalent Enhancer, ReprPC: Repressed Polycomb, ReprPCWk: Weak Repressed Polycomb, Quies: Quiescent/Low. (e) Stranded allele-specific chromatin accessibility measures at the haplotype using ATAC-seq data in differentiating adipocytes from a heterozygous individual. For each day of differentiation of an individual heterozygous, the number of reads overlapping with 20 non-coding SNPs in the haplotype, ordered by their start position and strand relative to the position of the variant, are shown. More reads indicate higher activity in haplotype 1 (non-risk, blue) compared to haplotype 2 (risk yellow). x-axis: offset from SNP position (bp), y-axis: stranded read count. (f) Replication of the effect at time 0 (mesenchymal stem cells) with ATAC-seq. (g) BMI-dependent variant association analysis. Bar plots represent the beta of the rs6712203 association with type 2 diabetes following BMI stratification. The cohort analysed is the UK Biobank self-identified white British individuals (total N = 327,960; N = 109198 with BMI < 25, N = 140539 with BMI between 25 and 30, and N = 78223 with BMI > = 30), and overlay of data points is not practical. Betas and 95% confidence intervals are shown, derived from a two-sided generalized linear model on outcome adjusted for demographic covariates (age, sex, genotyping array, 40 PCs).

Extended data Fig. 2 Conditional ana- lyses implicating rs6712203 in the genetic control of anthropometric traits and type 2 diabetes.

a, Conditional analyses implicating rs6712203 in the genetic control of anthropometric traits and type 2 diabetes. Each panel represents a different trait/sex/conditional analysis window, and all panels have an X axis corresponding to 100 kb on either side of the rs6712203 variant. The Y axis shows, for each variant in the window, the association strength for the given trait conditioned on the variants noted in White British participants in UK Biobank with the sex shown, and red lines indicate the significance threshold 5 × 10-8). −log10 p-values are shown, derived from a two-sided generalized linear model on outcome adjusted for demographic covariates (age, sex, genotyping array, 40 PCs).

Extended data Fig. 3 Chromatin inter- actions and CRISPRi of 2q24.3 locus identify COBLL1 as target genes.

(a) Cross-cell type conserved genome-wide higher order chromatin interactions for the 2q24.3 locus analyzed by Hi-C assays in human fibroblasts (left) and NHEK primary normal human epidermal keratinocytes (right), chr2: 163,556,000 - 167,558,000 (hg19), binned at 2 kb resolution. (b) Cas9 protein expression in dCas9 hWAT compared to the parental hWAT cell line. (c) mRNA expression of COBLL1 and GRB14 in response to increasing amounts of lentiviral sgRNA vectors (2 sgRNAs, virus volume 50 μl and 500 μl) targeting TSS regions of each gene compared to non-targeting controls (NT, 2 sgRNAs). Columns are means of individual sgRNAs indicated by different symbols. (d) COBLL1 protein expression normalized to b-actin in dCas9 hWATs transduced with sgRNAs targeting COBLL1 or GRB14 compared to controls. Top panel: Image of gel of representative sgRNA targeting NT, COBLL1 or GRB14. Bottom panel: plot of protein expression; 2 sgRNA for each target in 2 replicates. (e) Representation of 1,181 bp region flanking the COBLL1 intronic variant rs6712203 at the 2q24.3 MONW locus showing individual sgRNAs (n = 6) targeting the rs6712203 flanking regulatory region used in the CRISPRi experiments. (f, g) mRNA expression of (f) COBLL1 and (g) GRB14 in undifferentiated dCas9-hWAT preadipocytes at 6 days post lentiviral transduction with sgRNAs targeting TSS regions (red: COBLL1 TSS; blue GRB14 TSS) and the rs6712203-flanking regulatory element at position 1 to 6 as depicted in (e). Data are mean +/− SEM of 3 independent experiments. **** P < 0.0001, *** P = 0.0004, ** P = 0.006, * P = 0.013 – 0.036, two-tailed Student’s t test. (h) Predicted binding of POU2F2 between the two alleles using the Intragenomic Replicate Method (Cowper-Sal lari et al. 2012). As in Fig. 2d with different kmer counts.

Extended data Fig. 4 COBLL1 regulates actin cytoskeleton remodeling.

(a) COBLL1 expression in subcutaneous and visceral AMSCs throughout adipogenic differentiation, N = 4 biologically independent experiments, t-test two-sided, data represent median + 95% CI. (b) COBLL1 gene expression enrichment across 142 tissues (A-D) from enrichment profiler36. COBLL1 probes 203641_s_at and 203642_s_at were used for coregulation analysis (E-F). (c) Correlation with COBLL1 probe ILMN_1761260 using microarray data from lean and individuals with obesity. (d) Enrichment of pathways in the HCI (upper panel) and WikiPathways (lower panel) gene set lists from Enrichr, plotted as in Fig. 3A (KEGG), with p-value thresholds corresponding to the FDR cutoffs in those data. p-values are derived from a hypergeometric test. (e) COBLL1 expression in subcutaneous adipose tissue before and after a very low caloric diet (VLCD, upper panel, n = 18), corresponding body weight (lower panel), Wilcoxon signed-rank test.

Extended data Fig. 5 Knockdown of COBLL1 affects actin remodeling processes in differentiating adipocytes along with adipocyte differentiation, insulin sensitivity and lipolysis rate.

(a) COBLL1 expression in siCOBLL1 and siNT at day 0, 3 and 14 of differentiation, N = 3 biologically independent experiments, t-test two-sided. knock-down efficiency 80%, mean values + SEM. (bd) Morphological profiles of siCOBLL compared to siNT AMSCs at day 0 (b) day 3 (c) and day 9 (d) of differentiation, t-test two-sided, significance level < 5%FDR. (e)ActinandCOBLL1staininginsiCOBLL1comparedtosiNT subcutaneous adipocytes at day 9 using phalloidin and COBLL1 antibody staining (HPA053344, Alexa-Fluor 488), magnification x63/oil. Scale bar = 52.8 um. Representative results from N = 3 independent experiments. (f) Cells_Children_LargeBODIPY_objects_count in siCOBLL1- and siNT AMSCs at day 3, 9, 14, N = 3 biologically independent experiments, t-test two-sided, significance level < 5% FDR. (G) qPCR-based gene expression of COBLL1 and adipocyte marker genes GLUT4, FASN, LIPE, PPARG, PLIN1, FABP4, CEBPA, ADIPOQ in siCOBLL1 and siNT AMSCs at day 14 of differentiation, t-test two-sided, N = 4 biologically independent experiments, mean values +/− SEM. (h) qPCR-based leptin gene expression in shCOBLL1 compared to shEV adipocytes. Data are represented as median + 95% CI, one-way ANOVA with Tukey’s HSD test, N = 4 biologically independent experiments (i) Correlation of COBLL1 mRNA with LEP mRNA in subcutaneous adipose tissue from 24 lean individuals measured by Illumina microarrays. The pearson’s correlation coefficient r and p-value are depicted (j) Schematic of siCOBLL1 KD and AMSCs differentiation. (k) UMAP-based dimensionality reduction of LipocyteProfiler features in siCOBLL1 and siNT AMSCs. (l) Actin and COBLL1 staining in siCOBLL1 and siNT visceral adipocytes at day 14 using phalloidin and COBLL1 antibody staining (HPA053344, Alexa-Fluor 488), magnification x63/oil. Representative result from N = 2 independent experiments, scale bar = 52,8um (m) Representative Oil-Red-O lipid staining in SGBS adipocytes following lentiviral COBLL1 knock-down (shCOBLL1, knock-down efficiency 69%) and GRB14 (shGRB14, knock-down efficiency 61%) compared to empty vector control (shEV), scale bar = 15 mm. (n) GPDH metabolic activity in shCOBLL1, shGRB14 and shEV SGBS adipocytes, one-way ANOVA with Tukey’s HSD test, mean + 95% CI, N = 4 biologically independent experiments (o) Basal and insulin-stimulated 3H-2-deoxyglucose uptake in shCOBLL1, shGRB14 and shEV SGBS adipocytes, one-way ANOVA with Tukey’s HSD test, mean + 95% CI, N = 4 biologically independent experiments, 1st and 3rd quartiles (box) and median (middle line) are indicated, p = 4.3 × 10-8. (p) qPCR-based GLUT4 gene expression in shCOBLL1, shGRB14 and shEV adipocytes, one-way ANOVA with Tukey’s HSD test, mean + 95% CI, N = 4 biologically independent experiments.

Source data

Extended data Fig. 6 Lipocyte profiles of risk versus non-risk haplotype carriers.

(ac) Differences in morphological profiles between TT (n = 7) and CC (n = 6) allele carriers at day 0 (a), day 3 (b) and day 8 (c) in subcutaneous AMSCs (multi-way ANOVA, significance level < 5% FDR). (df) Differences in morphological profiles between TT (n = 7) and CC (n = 6) allele carriers at (d) day 0, (e) day 3 and (f) day 8 in visceral AMSCs (multi-way ANOVA, significance level < 5% FDR).

Extended data Fig. 7 Generation of COBLL1 mutant mice using CRISPR/Cas9 editing.

(a) Overview of the CRISPR/- Cas9 strategy to delete ~20 kb of the Cobll1 gene. The gRNA-targeting sequences (gRNAs) are underlined, and the PAM sequences are indicated in bold. Exons are represented as thick black boxes, introns are indicated as black lines with arrows, and the yellow boxes indicate the DNA-targeting region. Red hexagon indicates a stop codon gen- erating a Cobll1 truncated protein. Agarose gel showing the PCR products generated from DNA containing success- fully targeted Cobll1 from F0 mouse tail genomic DNA. The 308 bp band corresponds to the genomic deletion. (b) A real-time quantitative PCR of levels of Cobll1 mRNA in white adipose tissue (WAT), liver and kidney of Cobll1 WT, Cobll1 heterozygous (+/−) and null knockout Cobll1 (−/−) animals to confirm the Cobll1 ablation in knockout animals. Each group was analyzed using 5 different mice and the values were expressed as the mean ± s.e.m and P values by Student’s t-test. the experiment was repeated independently two times with similar results. (c) Pie chart illustrating non-redundant differential features per channel and class of measurement at day 8 of subcutaneous adipocyte differentiation in rs6712003 homozygous risk compared to non-risk carriers. (d, e) Differences in morphological profiles between AMSCs from Cobll1−/− mice (n = 3) and WT (n = 4) at (h) day 0 (i) day 2 (t-test two-sided, significance level < 5%FDR).

Supplementary information

Source data

Source Data Fig. 3

The source data for Fig1a is related to Fig. 3j: Images of Oil Red O lipid staining in SGBS adipocytes after lentiviral COBLL1 and GRB14 knockdown; the source data in Fig1a is related to Fig. 3n: western blots for lipolysis-relevant proteins assayed in basal or isoproterenol/IBMX-stimulated differentiated shCOBLL1 compared to shEV SGBS adipocytes.

Source Data Extended Data Fig. 5

The source data for Fig2a is related to Extended Data Fig. 5m: Images of Oil Red O lipid staining in SGBS adipocytes after lentiviral COBLL1 and GRB14 knockdown; the source data for Fig2b–h is related to Extended Data Fig. 5i: microscopy images of COBLL1 antibody (green), actin (phalloidin, red) and nuclei (Hoechst, blue) staining in siCOBLL1 and siNT visceral adipocytes.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Glunk, V., Laber, S., Sinnott-Armstrong, N. et al. A non-coding variant linked to metabolic obesity with normal weight affects actin remodelling in subcutaneous adipocytes. Nat Metab 5, 861–879 (2023). https://doi.org/10.1038/s42255-023-00807-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s42255-023-00807-w

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research