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.

  • Letter
  • Published:

Common genetic variants influence human subcortical brain structures

Abstract

The highly complex structure of the human brain is strongly shaped by genetic influences1. Subcortical brain regions form circuits with cortical areas to coordinate movement2, learning, memory3 and motivation4, and altered circuits can lead to abnormal behaviour and disease2. To investigate how common genetic variants affect the structure of these brain regions, here we conduct genome-wide association studies of the volumes of seven subcortical regions and the intracranial volume derived from magnetic resonance images of 30,717 individuals from 50 cohorts. We identify five novel genetic variants influencing the volumes of the putamen and caudate nucleus. We also find stronger evidence for three loci with previously established influences on hippocampal volume5 and intracranial volume6. These variants show specific volumetric effects on brain structures rather than global effects across structures. The strongest effects were found for the putamen, where a novel intergenic locus with replicable influence on volume (rs945270; P = 1.08 × 10−33; 0.52% variance explained) showed evidence of altering the expression of the KTN1 gene in both brain and blood tissue. Variants influencing putamen volume clustered near developmental genes that regulate apoptosis, axon guidance and vesicle transport. Identification of these genetic variants provides insight into the causes of variability in human brain development, and may help to determine mechanisms of neuropsychiatric dysfunction.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Common genetic variants associated with subcortical volumes and the ICV.
Figure 2: Effect of rs945270 on KTN1 expression and putamen shape.

Similar content being viewed by others

References

  1. Blokland, G. A., de Zubicaray, G. I., McMahon, K. L. & Wright, M. J. Genetic and environmental influences on neuroimaging phenotypes: a meta-analytical perspective on twin imaging studies. Twin Res. Hum. Genet. 15, 351–371 (2012)

    Article  Google Scholar 

  2. Kravitz, A. V. et al. Regulation of parkinsonian motor behaviours by optogenetic control of basal ganglia circuitry. Nature 466, 622–626 (2010)

    Article  CAS  ADS  Google Scholar 

  3. Poldrack, R. A. et al. Interactive memory systems in the human brain. Nature 414, 546–550 (2001)

    Article  CAS  Google Scholar 

  4. Pessiglione, M., Seymour, B., Flandin, G., Dolan, R. J. & Frith, C. D. Dopamine-dependent prediction errors underpin reward-seeking behaviour in humans. Nature 442, 1042–1045 (2006)

    Article  CAS  ADS  Google Scholar 

  5. Stein, J. L. et al. Identification of common variants associated with human hippocampal and intracranial volumes. Nature Genet. 44, 552–561 (2012)

    Article  CAS  ADS  Google Scholar 

  6. Ikram, M. A. et al. Common variants at 6q22 and 17q21 are associated with intracranial volume. Nature Genet. 44, 539–544 (2012)

    Article  CAS  Google Scholar 

  7. Lango Allen, H. et al. Hundreds of variants clustered in genomic loci and biological pathways affect human height. Nature 467, 832–838 (2010)

    Article  CAS  ADS  Google Scholar 

  8. Speliotes, E. K. et al. Association analyses of 249,796 individuals reveal 18 new loci associated with body mass index. Nature Genet. 42, 937–948 (2010)

    Article  CAS  Google Scholar 

  9. van der Sluis, S., Posthuma, D. & Dolan, C. V. TATES: efficient multivariate genotype-phenotype analysis for genome-wide association studies. PLoS Genet. 9, e1003235 (2013)

    Article  CAS  Google Scholar 

  10. Schizophrenia Working Group of the Psychiatric Genomics Consortium. Biological insights from 108 schizophrenia-associated genetic loci. Nature 511, 421–427 (2014)

  11. Kumar, J., Yu, H. & Sheetz, M. P. Kinectin, an Essential Anchor for Kinesin-Driven Vesicle Motility. Science 267, 1834–1837 (1995)

    Article  CAS  ADS  Google Scholar 

  12. Hamasaki, T., Goto, S., Nishikawa, S. & Ushio, Y. A role of netrin-1 in the formation of the subcortical structure striatum: repulsive action on the migration of late-born striatal neurons. J. Neurosci. 21, 4272–4280 (2001)

    Article  CAS  Google Scholar 

  13. Kang, H. J. et al. Spatio-temporal transcriptome of the human brain. Nature 478, 483–489 (2011)

    Article  CAS  ADS  Google Scholar 

  14. Motoyama, N. et al. Massive cell death of immature hematopoietic cells and neurons in Bcl-x-deficient mice. Science 267, 1506–1510 (1995)

    Article  CAS  ADS  Google Scholar 

  15. Itoh, K. et al. Apoptosis in the basal ganglia of the developing human nervous system. Acta Neuropathol. 101, 92–100 (2001)

    CAS  PubMed  Google Scholar 

  16. Scannevin, R. H. & Huganir, R. L. Postsynaptic organization and regulation of excitatory synapses. Nature Rev. Neurosci. 1, 133–141 (2000)

    Article  CAS  Google Scholar 

  17. Nithianantharajah, J. et al. Synaptic scaffold evolution generated components of vertebrate cognitive complexity. Nature Neurosci. 16, 16–24 (2013)

    Article  CAS  Google Scholar 

  18. Kirov, G. et al. De novo CNV analysis implicates specific abnormalities of postsynaptic signalling complexes in the pathogenesis of schizophrenia. Mol. Psychiatry 17, 142–153 (2012)

    Article  CAS  Google Scholar 

  19. Bis, J. C. et al. Common variants at 12q14 and 12q24 are associated with hippocampal volume. Nature Genet. 44, 545–551 (2012)

    Article  CAS  Google Scholar 

  20. Deans, M. R. et al. Control of neuronal morphology by the atypical cadherin Fat3. Neuron 71, 820–832 (2011)

    Article  CAS  Google Scholar 

  21. Stefansson, H. et al. A common inversion under selection in Europeans. Nature Genet. 37, 129–137 (2005)

    Article  CAS  Google Scholar 

  22. Ernst, J. & Kellis, M. ChromHMM: automating chromatin-state discovery and characterization. Nature Methods 9, 215–216 (2012)

    Article  CAS  Google Scholar 

  23. Ziebarth, J. D., Bhattacharya, A. & Cui, Y. CTCFBSDB 2.0: a database for CTCF-binding sites and genome organization. Nucleic Acids Res. 41, D188–D194 (2013)

    Article  CAS  Google Scholar 

  24. Hernandez, D. G. et al. Integration of GWAS SNPs and tissue specific expression profiling reveal discrete eQTLs for human traits in blood and brain. Neurobiol. Dis. 47, 20–28 (2012)

    Article  CAS  Google Scholar 

  25. Ramasamy, A. et al. Genetic variability in the regulation of gene expression in ten regions of the human brain. Nature Neurosci. 17, 1418–1428 (2014)

    Article  CAS  Google Scholar 

  26. Westra, H. J. et al. Systematic identification of trans eQTLs as putative drivers of known disease associations. Nature Genet. 45, 1238–1243 (2013)

    Article  CAS  Google Scholar 

  27. Toyoshima, I. & Sheetz, M. P. Kinectin distribution in chicken nervous system. Neurosci. Lett. 211, 171–174 (1996)

    Article  CAS  Google Scholar 

  28. Zhang, X. et al. Kinectin-mediated endoplasmic reticulum dynamics supports focal adhesion growth in the cellular lamella. J. Cell Sci. 123, 3901–3912 (2010)

    Article  CAS  Google Scholar 

  29. Cohen, M. X., Schoene-Bake, J. C., Elger, C. E. & Weber, B. Connectivity-based segregation of the human striatum predicts personality characteristics. Nature Neurosci. 12, 32–34 (2009)

    Article  CAS  Google Scholar 

  30. Parent, A. & Hazrati, L. N. Functional anatomy of the basal ganglia. I. The cortico-basal ganglia-thalamo-cortical loop. Brain Res. Brain Res. Rev. 20, 91–127 (1995)

    Article  CAS  Google Scholar 

  31. Patenaude, B., Smith, S. M., Kennedy, D. N. & Jenkinson, M. A Bayesian model of shape and appearance for subcortical brain segmentation. Neuroimage 56, 907–922 (2011)

    Article  Google Scholar 

  32. Fischl, B. et al. Whole brain segmentation: automated labeling of neuroanatomical structures in the human brain. Neuron 33, 341–355 (2002)

    Article  CAS  Google Scholar 

  33. Morey, R. A. et al. Scan-rescan reliability of subcortical brain volumes derived from automated segmentation. Hum. Brain Mapp. 31, 1751–1762 (2010)

    PubMed  PubMed Central  Google Scholar 

  34. Li, Y., Willer, C. J., Ding, J., Scheet, P. & Abecasis, G. R. MaCH: using sequence and genotype data to estimate haplotypes and unobserved genotypes. Genet. Epidemiol. 34, 816–834 (2010)

    Article  Google Scholar 

  35. Howie, B., Fuchsberger, C., Stephens, M., Marchini, J. & Abecasis, G. R. Fast and accurate genotype imputation in genome-wide association studies through pre-phasing. Nature Genet. 44, 955–959 (2012)

    Article  CAS  Google Scholar 

  36. Abecasis, G. R., Cherny, S. S., Cookson, W. O. & Cardon, L. R. Merlin-rapid analysis of dense genetic maps using sparse gene flow trees. Nature Genet. 30, 97–101 (2002)

    Article  CAS  Google Scholar 

  37. Willer, C. J., Li, Y. & Abecasis, G. R. METAL: fast and efficient meta-analysis of genomewide association scans. Bioinformatics 26, 2190–2191 (2010)

    Article  CAS  Google Scholar 

  38. Nyholt, D. R. A simple correction for multiple testing for single-nucleotide polymorphisms in linkage disequilibrium with each other. Am. J. Hum. Genet. 74, 765–769 (2004)

    Article  CAS  Google Scholar 

  39. Boker, S. et al. OpenMx: an open source extended structural equation modeling framework. Psychometrika 76, 306–317 (2011)

    Article  MathSciNet  Google Scholar 

  40. Walters, R., Bartels, M. & Lubke, G. Estimating variance explained by all variants in meta-analysis with heterogeneity. Behav. Genet. 43, 543 (2013)

    Google Scholar 

  41. So, H. C., Li, M. & Sham, P. C. Uncovering the total heritability explained by all true susceptibility variants in a genome-wide association study. Genet. Epidemiol. 35, 447–456 (2011)

    Article  Google Scholar 

  42. Li, M. X., Gui, H. S., Kwan, J. S. & Sham, P. C. GATES: a rapid and powerful gene-based association test using extended Simes procedure. Am. J. Hum. Genet. 88, 283–293 (2011)

    Article  CAS  Google Scholar 

  43. Li, M. X., Kwan, J. S. & Sham, P. C. HYST: a hybrid set-based test for genome-wide association studies, with application to protein-protein interaction-based association analysis. Am. J. Hum. Genet. 91, 478–488 (2012)

    Article  CAS  Google Scholar 

  44. Ramasamy, A. et al. Resolving the polymorphism-in-probe problem is critical for correct interpretation of expression QTL studies. Nucleic Acids Res. 41, e88 (2013)

    Article  CAS  Google Scholar 

  45. Trabzuni, D. et al. Quality control parameters on a large dataset of regionally dissected human control brains for whole genome expression studies. J. Neurochem. 119, 275–282 (2011)

    Article  CAS  Google Scholar 

  46. Gutman, B. A. et al. Maximizing power to track Alzheimer’s disease and MCI progression by LDA-based weighting of longitudinal ventricular surface features. Neuroimage 70, 386–401 (2013)

    Article  Google Scholar 

  47. Gutman, B. A., Wang, Y. L., Rajagopalan, P., Toga, A. W. & Thompson, P. M. Shape matching with medial curves and 1-d group-wise registration. In 2012 9th IEEE International Symposium on Biomedical Imaging (ISBI). 716–719 (2012)

  48. Schumann, G. et al. The IMAGEN study: reinforcement-related behaviour in normal brain function and psychopathology. Mol. Psychiatry 15, 1128–1139 (2010)

    Article  CAS  Google Scholar 

  49. Pollard, K. S., Hubisz, M. J., Rosenbloom, K. R. & Siepel, A. Detection of nonneutral substitution rates on mammalian phylogenies. Genome Res. 20, 110–121 (2010)

    Article  CAS  Google Scholar 

  50. Cheung, I. et al. Developmental regulation and individual differences of neuronal H3K4me3 epigenomes in the prefrontal cortex. Proc. Natl Acad. Sci. USA 107, 8824–8829 (2010)

    Article  CAS  ADS  Google Scholar 

  51. Maunakea, A. K. et al. Conserved role of intragenic DNA methylation in regulating alternative promoters. Nature 466, 253–257 (2010)

    Article  CAS  ADS  Google Scholar 

  52. Boyle, A. P. et al. High-resolution mapping and characterization of open chromatin across the genome. Cell 132, 311–322 (2008)

    Article  CAS  Google Scholar 

  53. Ernst, J. et al. Mapping and analysis of chromatin state dynamics in nine human cell types. Nature 473, 43–49 (2011)

    Article  CAS  ADS  Google Scholar 

  54. Devlin, B. & Roeder, K. Genomic control for association studies. Biometrics 55, 997–1004 (1999)

    Article  CAS  Google Scholar 

  55. Hager, R., Lu, L., Rosen, G. D. & Williams, R. W. Genetic architecture supports mosaic brain evolution and independent brain-body size regulation. Nat. Commun. 3, 1079 (2012)

    Article  ADS  Google Scholar 

  56. Schmucker, D. & Chen, B. Dscam and DSCAM: complex genes in simple animals, complex animals yet simple genes. Genes Dev. 23, 147–156 (2009)

    Article  CAS  Google Scholar 

  57. Brunet, A., Datta, S. R. & Greenberg, M. E. Transcription-dependent and -independent control of neuronal survival by the PI3K-Akt signaling pathway. Curr. Opin. Neurobiol. 11, 297–305 (2001)

    Article  CAS  Google Scholar 

Download references

Acknowledgements

Funding sources for contributing sites and acknowledgments of contributing consortia authors can be found in Supplementary Note 3.

Author information

Authors and Affiliations

Authors

Consortia

Contributions

Individual author contributions are listed in Supplementary Note 4.

Corresponding authors

Correspondence to Paul M. Thompson or Sarah E. Medland.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Additional information

Summary statistics from GWAS results are available online using the ENIGMA-Vis web tool: http://enigma.ini.usc.edu/enigma-vis/.

A list of authors and affiliations appears in the Supplementary Information.

A list of authors and affiliations appears in the Supplementary Information.

A list of authors and affiliations appears in the Supplementary Information.

A list of authors and affiliations appears in the Supplementary Information.

A list of authors and affiliations appears in the Supplementary Information.

Extended data figures and tables

Extended Data Figure 1 Outline of the genome-wide association meta-analysis.

Structural T1-weighted brain MRI and biological specimens for DNA extraction were acquired from each individual at each site. Imaging protocols were distributed to and completed by each site for standardized automated segmentation of brain structures and calculation of the ICV. Volumetric phenotypes were calculated from the segmentations. Genome-wide genotyping was completed at each site using commercially available chips. Standard imputation protocols to the 1000 Genomes reference panel (phase 1, version 3) were also distributed and completed at each site. Each site completed genome-wide association for each of the eight volumetric brain phenotypes with the listed covariates. Statistical results from GWAS files were uploaded to a central site for quality checking and fixed effects meta-analysis.

Extended Data Figure 2 Ancestry inference via multi-dimensional scaling plots.

Multi-dimensional scaling (MDS) plots of the discovery cohorts to HapMap III reference panels of known ancestry are displayed. Ancestry is generally homogeneous within each group. In all discovery samples any individuals with non-European ancestry were excluded before association. The axes have been flipped to the same orientation for each sample for ease of comparison. ASW, African ancestry in southwest USA; CEU, Utah residents with northern and western European ancestry from the CEPH collection; CHD, Chinese in metropolitan Denver, Colorado; GIH, Gujarati Indians in Houston, Texas; LWK, Luhya in Webuye, Kenya; MEX, Mexican ancestry in Los Angeles, California; MKK, Maasai in Kinyawa, Kenya; TSI, Tuscans in Italy; YRI, Yoruba in Ibadan, Nigeria.

Extended Data Figure 3 Genomic function is annotated near novel genome-wide significant loci.

ah, For each panel, zoomed-in Manhattan plots (±400 kb from top SNP) are shown with gene models below (GENCODE version 19). Plots below are zoomed to highlight the genomic region that probably contains the causal variant(s) (r2 > 0.8 from the top SNP). Genomic annotations from the UCSC browser and ENCODE are displayed to indicate potential functionality (see Methods for detailed track information). SNP coverage is low in f owing to a common genetic inversion in the region. Each plot was made using the LocusTrack software (http://gump.qimr.edu.au/general/gabrieC/LocusTrack/).

Extended Data Figure 4 Quantile–quantile and forest plots from meta-analysis of discovery cohorts.

a, Quantile–quantile plots show that the observed P values only deviate from the expected null distribution at the most significant values, indicating that population stratification or cryptic relatedness are not unduly inflating the results. This is quantified through the genomic control parameter (λ; which evaluates whether the median test statistic deviates from expected)54. λ values near 1 indicate that the median test statistic is similar to those derived from a null distribution. Corresponding meta-analysis Manhattan plots can be found in Fig. 1. b, Forest plots show the effect at each of the contributing sites to the meta-analysis. The size of the dot is proportional to the sample size, the effect is shown by the position on the x axis, and the standard error is shown by the line. Sites with an asterisk indicate the genotyping of a proxy SNP (in perfect linkage disequilibrium calculated from 1000 Genomes) for replication.

Extended Data Figure 5 Influence of patients with neuropsychiatric disease, age and gender on association results.

a, Scatterplot of effect sizes including and excluding patients with neuropsychiatric disorders for nominally significant SNPs. For each of the eight volumetric phenotypes, SNPs with P < 1 × 10−5 in the full discovery set meta-analysis were also evaluated excluding the patients. The beta values from regression, a measure of effect size, are plotted (blue dots) along with a line of equivalence between the two conditions (red line). The correlation between effect sizes with and without patients was very high (r > 0.99), showing that the SNPs with significant effects on brain structure are unlikely to be driven by the diseased individuals. b, Meta-regression comparison of effect size with mean age at each site. Each site has a corresponding number and coloured dot in each graph. The size of each dot is based on the standard error such that bigger sites with more definitive estimates have larger dots (and more influence on the meta-regression). The age range of participants covered most of the lifespan (9–97 years), but only one of these eight loci showed a significant relationship with the mean age of each cohort (rs608771 affecting putamen volume). c, Meta-regression comparison of effect size with the proportion of females at each site. No loci showed evidence of moderation by the proportion of females in a given sample. However, the proportion of females at each site has a very restricted range, so results should be interpreted with caution. Plotted information follows the same convention as described in b. The sites are numbered in the following order: (1) AddNeuroMed, (2) ADNI, (3) ADNI2GO, (4) BETULA, (5) BFS, (6) BIG, (7) BIG-Rep, (8) BrainSCALE, (9) BRCDECC, (10) CHARGE, (11) EPIGEN, (12) GIG, (13) GSP, (14) HUBIN, (15) IMAGEN, (16) IMpACT, (17) LBC1936, (18) Lieber, (19) MAS, (20) MCIC, (21) MooDS, (22) MPIP, (23) NCNG, (24) NESDA, (25) neuroIMAGE, (26) neuroIMAGE-Rep, (27) NIMH, (28) NTR-Adults, (29) OATS, (30) PAFIP, (31) QTIM, (32) SHIP, (33) SHIP-TREND, (34) SYS, (35) TCD-NUIG, (36) TOP, (37) UCLA-BP-NL and (38) UMCU.

Extended Data Figure 6 Cross-structure analyses.

a, Radial plots of effect sizes from the discovery sample for all genome-wide significant SNPs identified in this study. Plots indicate the effect of each genetic variant, quantified as percentage variance explained, on the eight volumetric phenotypes studied. As expected, the SNPs identified with influence on a phenotype show the highest effect size for that phenotype: putamen volume (rs945270, rs62097986, rs608771 and rs683250), hippocampal volume (rs77956314 and rs61921502), caudate volume (rs1318862) and ICV (rs17689882). In general much smaller effects are observed on other structures. b, Correlation heat map of GWAS test statistics (t-values) and hierarchical clustering55. Independent SNPs were chosen within an linkage disequilibrium block based on the highest association in the multivariate cross-structure analysis described in Extended Data Fig. 6c. Two heat maps are shown taking only independent SNPs with either P < 1 × 10−4 (left) or P < 0.01 (right) in the multivariate cross-structure analysis. Different structures are labelled in developmentally similar regions by the colour bar on the top and side of the heat map including basal ganglia (putamen, pallidum, caudate and accumbens; blue), amygdalo–hippocampal complex (hippocampus and amygdala; red), thalamus (turquoise) and ICV (black). Hierarchical clustering showed that developmentally similar regions have mostly similar genetic influences across the entire genome. The low correlation with the ICV is owing to it being used as a covariate in the subcortical structure GWAS associations. c, A multivariate cross-structure analysis of all volumetric brain traits. A Manhattan plot (left) and corresponding quantile–quantile plot (right) of multivariate GWAS analysis of all traits (volumes of the accumbens, amygdala, caudate, hippocampus, pallidum, putamen, thalamus, and ICV) in the discovery data set using the TATES method9 is shown. Multivariate cross-structure analysis confirmed the univariate analyses (see Table 1), but did not reveal any additional loci achieving cross-structure levels of significance.

Extended Data Figure 7 Pathway analysis of GWAS results for each brain structure.

A pathway analysis was performed on each brain volume GWAS using KGG42 to conduct gene-based tests and the Reactome database for pathway definition43. Pathway-wide significance was calculated using a Bonferroni correction threshold accounting for the number of pathways and traits tested such that Pthresh = 0.05/(671 pathways × 7 independent traits) = 1.06 × 10−5 and is shown here as a red line. The number of independent traits was calculated by accounting for the non-independence of each of the eight traits examined (described in the Methods). Variants that influence the putamen were clustered near genes known to be involved in DSCAM interactions, neuronal arborization and axon guidance56. Variants that influence intracranial volume are clustered near genes involved in EGFR and phosphatidylinositol-3-OH kinase (PI(3)K)/AKT signalling pathways, known to be involved in neuronal survival57. All of these represent potential mechanisms by which genetic variants influence brain structure. It is important to note that the hybrid set-based test (HYST) method for pathway analysis used here can be strongly influenced by a few highly significant genes, as was the case for putamen hits in which DCC and BCL2L1 were driving the pathway results.

Extended Data Figure 8 Spatio-temporal maps showing expression of genes near the four significant putamen loci over time and throughout regions of the brain.

Spatio-temporal gene expression13 was plotted as normalized log2 expression. Different areas of the neocortex (A1C, primary auditory cortex; DFC, dorsolateral prefrontal cortex; IPC, posterior inferior parietal cortex; ITC, inferior temporal cortex; MFC, medial prefrontal cortex; M1C, primary motor cortex; OFC, orbital prefrontal cortex; STC, superior temporal cortex; S1C, primary somatosensory cortex; VFC, ventrolateral prefrontal cortex; V1C, primary visual cortex) as well as subcortical areas (AMY, amygdala; CBC, cerebellar cortex; HIP, hippocampus; MD, mediodorsal nucleus of the thalamus; STR, striatum) are plotted from 10 post-conception weeks (PCW) to more than 60 years old. Genes that probably influence putamen volume are expressed in the striatum at some point during the lifespan. After late fetal development, KTN1 is expressed in the human thalamus, striatum and hippocampus and is more highly expressed in the striatum than the cortex. Most genes seem to have strong gradients of expression across time, with DCC most highly expressed during early prenatal life, and DLG2 most highly expressed at mid-fetal periods and throughout adulthood. BCL2L1, which inhibits programmed cell death, has decreased striatal expression at the end of neurogenesis (24–38 PCW), a period marked by increased apoptosis in the putamen15.

Extended Data Figure 9 CTCF-binding sites in the vicinity of the putamen locus marked by rs945270.

CTCF-binding sites from the ENCODE project are displayed from the database CTCFBSDB 2.0 (ref. 23) from two different cell types: embryonic stem cells (track ENCODE_Broad_H1-hESC_99540) and a neuroblastoma cell line differentiated with retinoic acid (ENCODE_UW_SK-N-SH_RA_97826). A proxy SNP to the top hit within the locus, rs8017172 (r2 = 1.0 to rs945270), lies within a CTCF-binding site called based on ChIP-seq data in the embryonic stem cells and near the binding site in neural SK-N-SH cells. As this is the lone chromatin mark in the intergenic region (see Extended Data Fig. 3), it suggests that the variant may disrupt a CTCF-binding site and thereby influence transcription of surrounding genes.

Extended Data Figure 10 Shape analysis in 1,541 young healthy subjects shows consistent deformations of the putamen regardless of segmentation protocol.

a, b, The distance from a medial core to surfaces derived from FSL FIRST (a; identical to Fig. 2c) or FreeSurfer (b) segmentations was derived in the same 1,541 subjects. Each copy of the rs945270-C allele was significantly associated with an increased width in coloured areas (false discovery rate corrected at q = 0.05) and the degree of deformation is labelled by colour. The orientation is indicated by arrows. A, anterior; I, inferior; P, posterior; S, superior. Shape analysis in both software suites gives statistically significant associations in the same direction. Although the effects are more widespread in the FSL segmentations, FreeSurfer segmentations also show overlapping regions of effect, which appears strongest in anterior and superior sections.

Extended Data Figure 11 The phenotypic variance explained by all common variants in this study.

a, Twin-based heritability (with 95% confidence intervals), measuring additive genetic influences from both common and rare variation, is shown for comparison with common variant based heritability (see Methods). b, The median estimated percentage of phenotypic variance explained by all SNPs (and 95% confidence interval) is given for each brain structure studied41. The full genome-wide association results from common variants explain approximately 7–15% of variance depending on the phenotype. c, The median estimated variance explained by each chromosome is shown for each phenotype. d, Some chromosomes explain more variance than would be expected by their length, for example chromosome 18 in the case of the putamen, which contains the DCC gene.

Supplementary information

Supplementary Information

This file contains authorship information regarding all the authors, contributing consortia authors, acknowledgements of funding sources and a list of author contributions. This file was replaced on 8 April 2015 to reflect changes to the IMAGEN and ADNI consortia lists. (PDF 224 kb)

Supplementary Tables

This file contains Supplementary Tables 1-9, detailing site-specific demographics and methods, consistency of MRI segmentation across software, independent genome-wide significant SNPs and genes, generalization sample association results, and look-ups of previous candidate gene associations. (XLSX 114 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hibar, D., Stein, J., Renteria, M. et al. Common genetic variants influence human subcortical brain structures. Nature 520, 224–229 (2015). https://doi.org/10.1038/nature14101

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature14101

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing