13 research outputs found

    Additional file 1: of A genome-wide analysis in cluster headache points to neprilysin and PACAP receptor gene variants

    No full text
    Supplemental Methods. Table S1. list of selected candidate genes, GO terms annotations, number of rare protein altering variants (PAV) contained in each gene, and gene-level P-values obtained from SKAT analysis. Table S2. list of rare PAV (MAF < 0.05) identified in CH cases and controls in 745 candidate genes. Table S3. P-values for single marker P-values of all tested SNPs. Table S4. logistic regression analysis. P-values for additive effect of SNPs controlling for sex as a covariate (P sex), and for sex, age and cigarettes per day (P sex, age, CPD). Table S5. Association analysis in CH cases and controls for significant migraine susceptibility SNPs emerged in meta-analysis. Figure S1. Principal component analysis (PCA) plot. Figure S2. Quantile-quantile (Q-Q) plot of Fisher’s exact test P-values for association with cluster headache (CH). Figure S3. Local association plots for chromosome 14 locus (A) and the chromosome 7 locus (B) [17, 18, 24, 41]. (ZIP 6890 kb

    Table3_Dissecting the multifaceted contribution of the mitochondrial genome to autism spectrum disorder.xlsx

    No full text
    Autism spectrum disorder (ASD) is a clinically heterogeneous class of neurodevelopmental conditions with a strong, albeit complex, genetic basis. The genetic architecture of ASD includes different genetic models, from monogenic transmission at one end, to polygenic risk given by thousands of common variants with small effects at the other end. The mitochondrial DNA (mtDNA) was also proposed as a genetic modifier for ASD, mostly focusing on maternal mtDNA, since the paternal mitogenome is not transmitted to offspring. We extensively studied the potential contribution of mtDNA in ASD pathogenesis and risk through deep next generation sequencing and quantitative PCR in a cohort of 98 families. While the maternally-inherited mtDNA did not seem to predispose to ASD, neither for haplogroups nor for the presence of pathogenic mutations, an unexpected influence of paternal mtDNA, apparently centered on haplogroup U, came from the Italian families extrapolated from the test cohort (n = 74) when compared to the control population. However, this result was not replicated in an independent Italian cohort of 127 families and it is likely due to the elevated paternal age at time of conception. In addition, ASD probands showed a reduced mtDNA content when compared to their unaffected siblings. Multivariable regression analyses indicated that variants with 15%–5% heteroplasmy in probands are associated to a greater severity of ASD based on ADOS-2 criteria, whereas paternal super-haplogroups H and JT were associated with milder phenotypes. In conclusion, our results suggest that the mtDNA impacts on ASD, significantly modifying the phenotypic expression in the Italian population. The unexpected finding of protection induced by paternal mitogenome in term of severity may derive from a role of mtDNA in influencing the accumulation of nuclear de novo mutations or epigenetic alterations in fathers’ germinal cells, affecting the neurodevelopment in the offspring. This result remains preliminary and needs further confirmation in independent cohorts of larger size. If confirmed, it potentially opens a different perspective on how paternal non-inherited mtDNA may predispose or modulate other complex diseases.</p

    Table1_Dissecting the multifaceted contribution of the mitochondrial genome to autism spectrum disorder.xlsx

    No full text
    Autism spectrum disorder (ASD) is a clinically heterogeneous class of neurodevelopmental conditions with a strong, albeit complex, genetic basis. The genetic architecture of ASD includes different genetic models, from monogenic transmission at one end, to polygenic risk given by thousands of common variants with small effects at the other end. The mitochondrial DNA (mtDNA) was also proposed as a genetic modifier for ASD, mostly focusing on maternal mtDNA, since the paternal mitogenome is not transmitted to offspring. We extensively studied the potential contribution of mtDNA in ASD pathogenesis and risk through deep next generation sequencing and quantitative PCR in a cohort of 98 families. While the maternally-inherited mtDNA did not seem to predispose to ASD, neither for haplogroups nor for the presence of pathogenic mutations, an unexpected influence of paternal mtDNA, apparently centered on haplogroup U, came from the Italian families extrapolated from the test cohort (n = 74) when compared to the control population. However, this result was not replicated in an independent Italian cohort of 127 families and it is likely due to the elevated paternal age at time of conception. In addition, ASD probands showed a reduced mtDNA content when compared to their unaffected siblings. Multivariable regression analyses indicated that variants with 15%–5% heteroplasmy in probands are associated to a greater severity of ASD based on ADOS-2 criteria, whereas paternal super-haplogroups H and JT were associated with milder phenotypes. In conclusion, our results suggest that the mtDNA impacts on ASD, significantly modifying the phenotypic expression in the Italian population. The unexpected finding of protection induced by paternal mitogenome in term of severity may derive from a role of mtDNA in influencing the accumulation of nuclear de novo mutations or epigenetic alterations in fathers’ germinal cells, affecting the neurodevelopment in the offspring. This result remains preliminary and needs further confirmation in independent cohorts of larger size. If confirmed, it potentially opens a different perspective on how paternal non-inherited mtDNA may predispose or modulate other complex diseases.</p

    DataSheet1_Dissecting the multifaceted contribution of the mitochondrial genome to autism spectrum disorder.docx

    Get PDF
    Autism spectrum disorder (ASD) is a clinically heterogeneous class of neurodevelopmental conditions with a strong, albeit complex, genetic basis. The genetic architecture of ASD includes different genetic models, from monogenic transmission at one end, to polygenic risk given by thousands of common variants with small effects at the other end. The mitochondrial DNA (mtDNA) was also proposed as a genetic modifier for ASD, mostly focusing on maternal mtDNA, since the paternal mitogenome is not transmitted to offspring. We extensively studied the potential contribution of mtDNA in ASD pathogenesis and risk through deep next generation sequencing and quantitative PCR in a cohort of 98 families. While the maternally-inherited mtDNA did not seem to predispose to ASD, neither for haplogroups nor for the presence of pathogenic mutations, an unexpected influence of paternal mtDNA, apparently centered on haplogroup U, came from the Italian families extrapolated from the test cohort (n = 74) when compared to the control population. However, this result was not replicated in an independent Italian cohort of 127 families and it is likely due to the elevated paternal age at time of conception. In addition, ASD probands showed a reduced mtDNA content when compared to their unaffected siblings. Multivariable regression analyses indicated that variants with 15%–5% heteroplasmy in probands are associated to a greater severity of ASD based on ADOS-2 criteria, whereas paternal super-haplogroups H and JT were associated with milder phenotypes. In conclusion, our results suggest that the mtDNA impacts on ASD, significantly modifying the phenotypic expression in the Italian population. The unexpected finding of protection induced by paternal mitogenome in term of severity may derive from a role of mtDNA in influencing the accumulation of nuclear de novo mutations or epigenetic alterations in fathers’ germinal cells, affecting the neurodevelopment in the offspring. This result remains preliminary and needs further confirmation in independent cohorts of larger size. If confirmed, it potentially opens a different perspective on how paternal non-inherited mtDNA may predispose or modulate other complex diseases.</p

    Common and rare variants of microRNA genes in autism spectrum disorders

    No full text
    <div><p></p><p><i>Objectives.</i> MicroRNAs (miRNAs) are post-transcriptional regulators that have been shown to be involved in disease susceptibility. Here we explore the possible contribution of common and rare variants in miRNA genes in autism spectrum disorders (ASD). <i>Methods.</i> A total of 350 tag SNPs from 163 miRNA genes were genotyped in 636 ASD cases and 673 controls. A replication study was performed in a sample of 449 ASD cases and 415 controls. Additionally, rare variants in 701 miRNA genes of 41 ASD patients were examined using whole-exome sequencing. <i>Results.</i> The most significant association in the discovery sample was obtained for the miR-133b/miR-206 cluster (rs16882131, <i>P = </i>0.00037). The replication study did not reach significance. However, the pooled analysis (1,085 cases and 1,088 controls) showed association with two miRNA clusters: miR-133b/miR-206 (rs16882131, <i>permP = </i>0.037) and miR-17/miR-18a/miR-19a/miR-20a/miR-19b-1/miR92a-1 (rs6492538, <i>permP = </i>0.019). Both miR-133b and miR-206 regulate the <i>MET</i> gene, previously associated with ASD. Rare variant analysis identified mutations in several miRNA genes, among them miR-541, a brain-specific miRNA that regulates <i>SYN1</i>, found mutated in ASD. <i>Conclusions.</i> Although our results do not establish a clear role for miRNAs in ASD, we pinpointed a few candidate genes. Further exome and GWAS studies are warranted to get more insight into their potential contribution to the disorder.</p></div

    Table2_Dissecting the multifaceted contribution of the mitochondrial genome to autism spectrum disorder.xlsx

    No full text
    Autism spectrum disorder (ASD) is a clinically heterogeneous class of neurodevelopmental conditions with a strong, albeit complex, genetic basis. The genetic architecture of ASD includes different genetic models, from monogenic transmission at one end, to polygenic risk given by thousands of common variants with small effects at the other end. The mitochondrial DNA (mtDNA) was also proposed as a genetic modifier for ASD, mostly focusing on maternal mtDNA, since the paternal mitogenome is not transmitted to offspring. We extensively studied the potential contribution of mtDNA in ASD pathogenesis and risk through deep next generation sequencing and quantitative PCR in a cohort of 98 families. While the maternally-inherited mtDNA did not seem to predispose to ASD, neither for haplogroups nor for the presence of pathogenic mutations, an unexpected influence of paternal mtDNA, apparently centered on haplogroup U, came from the Italian families extrapolated from the test cohort (n = 74) when compared to the control population. However, this result was not replicated in an independent Italian cohort of 127 families and it is likely due to the elevated paternal age at time of conception. In addition, ASD probands showed a reduced mtDNA content when compared to their unaffected siblings. Multivariable regression analyses indicated that variants with 15%–5% heteroplasmy in probands are associated to a greater severity of ASD based on ADOS-2 criteria, whereas paternal super-haplogroups H and JT were associated with milder phenotypes. In conclusion, our results suggest that the mtDNA impacts on ASD, significantly modifying the phenotypic expression in the Italian population. The unexpected finding of protection induced by paternal mitogenome in term of severity may derive from a role of mtDNA in influencing the accumulation of nuclear de novo mutations or epigenetic alterations in fathers’ germinal cells, affecting the neurodevelopment in the offspring. This result remains preliminary and needs further confirmation in independent cohorts of larger size. If confirmed, it potentially opens a different perspective on how paternal non-inherited mtDNA may predispose or modulate other complex diseases.</p

    Characterization of the functional impact of <i>SHANK2</i> mutations in cultured neuronal cells.

    No full text
    <p>A. The colocalization of <i>ProSAP1A/SHANK2</i>-EGFP (postsynaptic marker) and Bassoon (presynaptic marker) indicated that the mutations did not disturb the formation of SHANK2 clusters at excitatory synapses along the dendrites. B. The quantification of synapse density was performed on 20 transfected hippocampal neurons per construct from at least three independent experiments. The majority of the <i>ProSAP1A</i> variants affecting a conserved amino acid among SHANK proteins reduced significantly the synaptic density compared with the variants that affect amino acid non conserved among SHANK proteins (Mann-Whitney U-test: n<sub>WT</sub> = 20, n<sub>mut</sub> = 20; U<sub>S557N</sub> = 82.5, p<sub>S557N</sub> = 0.001; U<sub>R569H</sub> = 124, p<sub>R569H</sub> = 0.04; U<sub>L629P</sub> = 149, p<sub>L629P</sub> = 0.17; U<sub>V717F</sub> = 114, p<sub>V717F</sub> = 0.02; U<sub>A729T</sub> = 73, p<sub>A729T</sub> = 0.000; U<sub>K780Q</sub> = 154, p<sub>K780Q</sub> = 0.221; U<sub>R818H</sub> = 108, p<sub>R818H</sub> = 0.012; U<sub>A822T</sub> = 154.5, p<sub>A822T</sub> = 0.224; U<sub>V823M</sub> = 129, p<sub>V823M</sub> = 0.056; U<sub>Y967C</sub> = 134, p<sub>Y967C</sub> = 0.076; U<sub>G1170R</sub> = 78, p<sub>G1170R</sub> = 0.001; U<sub>R1290W</sub> = 142, p<sub>R1290W</sub> = 0.121; U<sub>Q1308R</sub> = 162, p<sub>Q1308R</sub> = 0.314; U<sub>D1535N</sub> = 97, p<sub>D1535N</sub> = 0.005; U<sub>P1586L</sub> = 137, p<sub>P1586L</sub> = 0.910; U<sub>L1722P</sub> = 79, p<sub>L1722P</sub> = 0.001, *p<0.05, **p<0.01, ***p<0.001). <b>C.</b> Effect of the variants on synaptic density. The y-axis represents −log P compared to WT (P obtained with Mann-Whitney test). After Bonferroni correction for 16 tests, only P values<0.003 were considered as significant. Variants represented in red were specific to ASD, in orange were shared by ASD and controls, and in green were specific to the controls. Open circles and filled circles represent non conserved and conserved amino acids, respectively. Prim, primary; second, secondary.</p

    Genetic alterations identified in the control subject SWE_Q56_508.

    No full text
    <p>A. <i>SHANK2</i> splice mutation (IVS22+1G>T) detected in a Swedish female control, SWE_Q56_508. The mutation altered the donor splicing site of exon 22 and led to a premature stop in all <i>SHANK2</i> isoforms except for the <i>AF1411901</i> isoform, where it altered the protein sequence (G263V). B. CNVs in the same individual altering <i>LOC339822</i>, <i>SNTG2</i>, <i>PXDN</i> and <i>MYT1L</i>. The two close duplications span 264 kb and 245 kb on chromosome 2 and altered <i>LOC339822</i> and <i>SNTG2</i>, and <i>PXDN</i> and <i>MYT1L</i>, respectively. Dots show the B allele frequency (BAF; in green), Log R ratio (LRR; in red), and QuantiSNP score (in blue). Lower panel: all CNVs listed in the Database of Genomic Variants (DGV) are represented: loss (in red), gain (in blue), gain or loss (in brown). H, homer binding site; D, dynamin binding site; C, cortactin binding site.</p

    <i>SHANK2</i> mutations in patients with ASD.

    No full text
    <p>A. A heterozygous deletion of <i>SHANK2</i> was identified with the Illumina Human 1M-Duo SNP array in a patient with autism (AU038_3). The deletion spans 421 kb on chromosome 11q13.3, covers twelve exons of the human <i>SHANK2</i> and is not present in the parents. Each dot shows Log R Ratio (LRR; in red) and B allele frequency (BAF; in green). QuantiSNP score is represented with a blue line and indicates the deletion size. B. Location of the CNV and sequence variants (from this study and Berkel <i>et al.</i> 2010) along the SHANK2 protein: in red the variations specific to ASD, in orange the variations shared by ASD and controls and in green the variations specific to controls <a href="http://www.plosgenetics.org/article/info:doi/10.1371/journal.pgen.1002521#pgen.1002521-Berkel1" target="_blank">[26]</a>. The breakpoints of the <i>SHANK2</i> deletion in AU038_3 are represented with a dotted line on the protein. Stars indicate the variants affecting conserved amino acids. C. A total of 40 variants were identified and variants affecting conserved amino acids in other SHANK proteins are enriched in patients with ASD (n<sub>conserved</sub> = 12 and n<sub>non-conserved</sub> = 3) compared with controls (n<sub>conserved</sub> = 6 and n<sub>non-conserved</sub> = 11) (Fisher's exact test 1-sided, P = 0.013, OR = 6.83, 95% IC = 1.19–53.40). D. The percentage of carriers of <i>SHANK2</i> variants in patients with ASD and Controls. Variants affecting a conserved amino acid among the SHANK proteins are enriched in patients with ASD (n<sub>conserved</sub> = 29 and n<sub>non-conserved</sub> = 822) compared with controls (n<sub>conserved</sub> = 16 and n<sub>non-conserved</sub> = 1074) (Fisher's exact test 1-sided, P = 0.004, OR = 2.37, 95% CI = 1.23–4.70). Open squares and filled squares represent the non-conserved and conserved amino acids, respectively. ANK, Ankyrin repeat domain; SH3, Src homology 3 domain; PDZ, postsynaptic density 95/Discs large/zona occludens-1 homology domain; SAM, sterile alpha motif domain; BSR, brain specific region; H, homer binding site; D, dynamin binding site; C, cortactin binding site. The proline-rich region is represented as a horizontal gray line.</p

    Characterization of CNVs in three patients carrying a <i>de novo</i> deletion of <i>SHANK2</i>.

    No full text
    <p>Paternally or maternally inherited CNVs are indicated by squares and circles, respectively. <i>De novo</i> CNVs are indicated by stars. Deletions and duplications are indicated in red and blue, respectively. CNVs hitting exons or only introns are filled with grey and white, respectively. Squares and circles within star represent <i>de novo</i> CNV of paternal or maternal origin; circles within squares represent CNV inherited by father or mother. ABCC6, ATP-binding cassette, sub-family C, member 6 pseudogene 2; ADAM, ADAM metallopeptidase; AMY1, amylase (salivary); AMY2A, amylase (pancreatic); ARHGAP11B, Rho GTPase activating protein 11B; CAMSAP1L1, calmodulin regulated spectrin-associated protein 1-like 1; CHRNA7, cholinergic receptor, nicotinic, alpha 7; CNTN4, contactin 4; CTNNA3, catenin (cadherin-associated protein), alpha 3; CYFIP1, cytoplasmic FMR1 interacting protein 1; DUSP22, dual specificity phosphatase 22; GALM, galactose mutarotase; GCNT2, glucosaminyl (N-acetyl) transferase 2; GOLGA, golgi autoantigen, golgin subfamily a; GSTT1, glutathione S-transferase theta 1; HLA-DRB, major histocompatibility complex, class II, DR beta; LAMA4, laminin, alpha 4; NIPA, non imprinted in Prader-Willi/Angelman syndrome; NLGN1, neuroligin 1; NME7, non-metastatic cells 7; OR, olfactory receptor; PCDHA, protocadherin alpha; RFPL4B, ret finger protein-like 4B; RHD, Rh blood group, D antigen; SFMBT1, Scm-like with four mbt domains 1; SHANK2, SH3 and multiple ankyrin repeat domains 2; SMC2, structural maintenance of chromosomes 2; TNS3, tensin 3; TUBGCP5, tubulin, gamma complex associated protein 5; UGT2B17, UDP glucuronosyltransferase 2 family, polypeptide B17.</p
    corecore