Whole genome sequences discriminate hereditary hemorrhagic telangiectasia phenotypes by non-HHT deleterious DNA variation

Katie E Joyce, Ebun Onabanjo, Sheila Brownlow, Fadumo Nur, Kike Olupona, Kehinde Fakayode, Manveer Sroya, Geraldine A Thomas, Teena Ferguson, Julian Redhead, Carolyn M Millar, Nichola Cooper, D Mark Layton, Freya Boardman-Pretty, Mark J Caulfield, Genomics England Research Consortium, Claire L Shovlin, Katie E Joyce, Ebun Onabanjo, Sheila Brownlow, Fadumo Nur, Kike Olupona, Kehinde Fakayode, Manveer Sroya, Geraldine A Thomas, Teena Ferguson, Julian Redhead, Carolyn M Millar, Nichola Cooper, D Mark Layton, Freya Boardman-Pretty, Mark J Caulfield, Genomics England Research Consortium, Claire L Shovlin

Abstract

The abnormal vascular structures of hereditary hemorrhagic telangiectasia (HHT) often cause severe anemia due to recurrent hemorrhage, but HHT causal genes do not predict the severity of hematological complications. We tested for chance inheritance and clinical associations of rare deleterious variants in which loss-of-function causes bleeding or hemolytic disorders in the general population. In double-blinded analyses, all 104 patients with HHT from a single reference center recruited to the 100 000 Genomes Project were categorized on new MALO (more/as-expected/less/opposite) sub-phenotype severity scales, and whole genome sequencing data were tested for high impact variants in 75 HHT-independent genes encoding coagulation factors, or platelet, hemoglobin, erythrocyte enzyme, and erythrocyte membrane constituents. Rare variants (all gnomAD allele frequencies <0.003) were identified in 56 (75%) of these 75 HHT-unrelated genes. Deleteriousness assignments by Combined Annotation Dependent Depletion (CADD) scores >15 were supported by gene-level mutation significance cutoff scores. CADD >15 variants were identified in 38/104 (36.5%) patients with HHT, found for 1 in 10 patients within platelet genes; 1 in 8 within coagulation genes; and 1 in 4 within erythrocyte hemolytic genes. In blinded analyses, patients with greater hemorrhagic severity that had been attributed solely to HHT vessels had more CADD-deleterious variants in platelet (Spearman ρ = 0.25; P = .008) and coagulation (Spearman ρ = 0.21; P = .024) genes. However, the HHT cohort had 60% fewer deleterious variants in platelet and coagulation genes than expected (Mann-Whitney test P = .021). In conclusion, patients with HHT commonly have rare variants in genes of relevance to their phenotype, offering new therapeutic targets and opportunities for informed, personalized medicine strategies.

© 2022 by The American Society of Hematology. Licensed under Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International (CC BY-NC-ND 4.0), permitting only noncommercial, nonderivative use with attribution. All other rights reserved.

Figures

Graphical abstract
Graphical abstract
Figure 1.
Figure 1.
Framework for studies “toward personalized genomic interpretations” (TPGI). MALO: More severe than expected for disease; As expected for disease; Less severe than expected for disease, and Opposite phenotype. *Gene selection by any broadly relevant set of panels such as PanelApp established for 100 000 Genomes Project interpretations suggested to include additional categories less relevant to phenotypes (as in current study) to control for methodological biases. Where TPGI demonstrates evidence of relevance across a patient cohort, this supports attention to a subsequent “for personalized genomic interpretations” (FPGI) stage to be developed within mainstream medicine. WES, whole exome sequencing; WGS, whole genome sequencing.
Figure 2.
Figure 2.
HHT genes and phenotypic associations. (A) Endothelial vasculopathy genes and variants identified in current study cohort. Note that the major HHT genes are ACVRL1 (red) and ENG (green).SMAD4 is less common, and GDF2 was more recently described as HHT causal, whereas EPHB4 and RASA1 cause separate endothelial vasculopathies (CM-AVM2 and CM-AVM1) that overlap phenotypically with HHT. (B) Schematic of the cohort of 104 patients plotted on 4 separate axes for hemorrhage (H), anemia (A), thrombosis (T), and deep-seated infection (I). Ten deep-seated infections were in association with concurrent and presumed causative pulmonary AVMs (cerebral and spinal abscesses, spinal discitis, and septic arthritis due to polymicrobial flora, particularly anaerobic and aerobic commensals of the gastrointestinal and periodontal spaces). Blue numbers indicate the number of the cohort with no events/no excess (H0, A0, T0, and I0), precipitated events (H1, A1, T1), and spontaneous events (H2, A2, T2, and I1). I1 were cerebral abscess (N = 6), spinal discitis (N = 2), spinal abscess (N = 1), septic arthritis (N = 1), recurrent bacterial endocarditis (N = 1), osteomyelitis (N = 1), and recurrent sepsis (N = 1). (C) Violin plots for ACVRL1 (left, red) and ENG (right, green) patients orientated on the 3 axes as in (B).
Figure 3.
Figure 3.
Hematological genes and variants. (A) Genome positions of the variant-containing genes in the study cohort on genome ideogram indicating positions of the 56 study genes with variants (red) and the two major HHT genes ACVRL1 and ENG. Variants present in patients with an identified variant in HHT genes are indicated for ACVRL1 (purple lines) and ENG (blue lines). To preserve anonymity, data for the single SMAD4, GDF2, and EPHB4 families are not illustrated. (B) Chromosomal distributions of variant-containing genes per 100 Mb of gDNA (blue circles/lines), number of variants per Mb (red shaded background), and variants with CADD score >15 (red circles/lines) per Mb of gDNA by chromosome. (C) Population-level burden of genetic damage in the 6 HHT panel genes (blue) and 75 study hematological genes (red), as detailed in supplemental Table 1. The study genes were subcategorized by variant presence (Variant genes; N = 56) and absence (No-variant genes; N = 19) in the study cohort, and P values were calculated by Dunn’s post Kruskal Wallis: (Ci) Gene damage index (GDI) phred scores, a genome-wide, gene-level metric of the mutational damage that has accumulated in the general population that performs well at removing exome variants in genes irrelevant to disease. (Cii) Residual variation intolerance scores (RVIS) that perform better for detection of genes in which newly identified variants are more likely cause a recognized disease.
Figure 4.
Figure 4.
Burden of deleterious variants by categories of genes and phenotypic severity profiles. (A) The CADD score distribution of variants identified in the study cohort (red) compared with gene-level MSC scores, indicating the lower limit of the 90% and 95% confidence intervals for deleterious CADD scores in individual genes (gray violin plots). Vertical dotted lines represent overall CADD scores 5 and 15 and the violin plot median and interquartile ranges. Note that CADD scores rank the deleteriousness for all 9 billion single nucleotide variants, and millions of small indels and splice site variants, based on machine learning trained on diverse genomic features derived from surrounding sequence context, gene model annotations, evolutionary constraint, epigenetic measurements, and functional predictions. (B) The 5 process-level categories, respective category contributions in the study cohort, are shown for the total number of genes, and all variants,, with each box representing the category’s percentage for the row, as indicated by the heat scale. (C) The burden of deleterious (CADD >15) variants in the 5 categories of genes for each predefined phenotypic subcohort. Filled pie-chart regions quantify the number of CADD >15 variants per patient in the category (gray if pooling across all gene categories; red if gene category specific; fully filled if 1 variant per patient). The 10 variants with CADD scores of 1 to 5 (likely benign) and 36 with CADD scores between 5 and 15 (of uncertain significance) are effectively included in the remaining white pie chart regions to enhance clarity. For each set of 6 pie charts, gene category placements are consistent. The upper row indicates total number of CADD >15 variants across all categories, and the remaining upper row pie charts focus on hemorrhage (upper middle for coagulation genes, upper right for platelet genes). The lower row presents the red cell gene categories (lower left for hemoglobin genes, lower middle for erythrocyte [red cell] membrane genes, and lower right for erythrocyte enzyme genes). The set of 6 gray-lined empty pie charts indicate there were no patients in the predefined A1 category. (D) Heat maps for number of variants per patient in each phenotypic subcohort, in which the precipitated and spontaneous H1/H2, T1/T2, and A1/A2 subcohorts have been pooled. Trends highlighted in the text are denoted by a white cross or black star. *P < .0286; **P < .01.
Figure 5.
Figure 5.
HHT patient subcategories and variation. Quantitative phenotypic measurements categorized by presence, absence, and CADD score of variants. Data points represent all datasets captured in the 104 patients across their clinical assessments, plotted according to the presence or absence of variants in the genes for (A) hemoglobin production, (B) red bood cell membrane production, (C) red blood cell enzymes, (D) coagulation and (E) platelets as listed in supplemental Table 1. Two indices are provided for each category of genes, but there were also no differences observed between categories for other indices examined (data not shown).
Figure 6.
Figure 6.
Variant burden in HHT cohort compared with general population. (A) Number of variants in the 75 study genes by category, within the gnomAD 3.1.2 dataset (blue) and current cohort (red) per genome. (B) gnomAD 3.1.2 dataset loss of function (pLOF) variants compared with CADD >15 genes in study cohort per genome, with mean and standard error illustrated. (C) Fold-enriched total number of deleterious variants per category, in all 75 genes, for HHT CADD >15 variants compared with pLOF variants in gnomAD 3.1.2 dataset. (D) Comparison of fold-enriched deleterious variants limiting to HHT CADD >15 variants and gnomAD pLOF variants in the same genes. P value calculated by Mann-Whitney test.

References

    1. Shovlin CL. Hereditary haemorrhagic telangiectasia: pathophysiology, diagnosis and treatment. Blood Rev. 2010;24(6):203-219.
    1. Faughnan ME, Mager JJ, Hetts SW, et al. . Second international guidelines for the diagnosis and management of hereditary hemorrhagic telangiectasia. Ann Intern Med. 2020;173(12):989-1001.
    1. Shovlin CL, Buscarini E, Sabbà C, et al. . The European Rare Disease Network for HHT Frameworks for management of hereditary haemorrhagic telangiectasia in general and speciality care. Eur J Med Genet. 2022;65(1):104370.
    1. Bideau A, Plauchu H, Brunet G, Robert J. Epidemiological investigation of Rendu-Osler disease in France: its geographical distribution and prevalence. Popul. 1989;44(1):3-22.
    1. Kjeldsen AD, Vase P, Green A. Hereditary haemorrhagic telangiectasia: a population-based study of prevalence and mortality in Danish patients. J Intern Med. 1999;245(1):31-39.
    1. Shovlin CL, Simeoni I, Downes K, et al. . Mutational and phenotypic characterization of hereditary hemorrhagic telangiectasia. Blood. 2020;136(17):1907-1918.
    1. McAllister KA, Grogg KM, Johnson DW, et al. . Endoglin, a TGF-β binding protein of endothelial cells, is the gene for hereditary haemorrhagic telangiectasia type 1. Nat Genet. 1994;8(4):345-351.
    1. Johnson DW, Berg JN, Baldwin MA, et al. . Mutations in the activin receptor-like kinase 1 gene in hereditary haemorrhagic telangiectasia type 2. Nat Genet. 1996;13(2):189-195.
    1. Gallione CJ, Repetto GM, Legius E, et al. . A combined syndrome of juvenile polyposis and hereditary haemorrhagic telangiectasia associated with mutations in MADH4 (SMAD4). Lancet. 2004;363(9412):852-859.
    1. Bourdeau A, Cymerman U, Paquet ME, et al. . Endoglin expression is reduced in normal vessels but still detectable in arteriovenous malformations of patients with hereditary hemorrhagic telangiectasia type 1. Am J Pathol. 2000;156(3):911-923.
    1. The HHT Mutation Database: . Accessed 23 December 2021.
    1. Bourdeau A, Dumont DJ, Letarte M. A murine model of hereditary hemorrhagic telangiectasia. J Clin Invest. 1999;104(10):1343-1351.
    1. Park SO, Wankhede M, Lee YJ, et al. . Real-time imaging of de novo arteriovenous malformation in a mouse model of hereditary hemorrhagic telangiectasia. J Clin Invest. 2009;119(11):3487-3496.
    1. Jin Y, Muhl L, Burmakin M, et al. . Endoglin prevents vascular malformation by regulating flow-induced cell migration and specification through VEGFR2 signalling. Nat Cell Biol. 2017;19(6):639-652.
    1. Ola R, Künzel SH, Zhang F, et al. . SMAD4 prevents flow induced arteriovenous malformations by inhibiting casein kinase 2. Circulation. 2018; 138(21):2379-2394.
    1. Bernabeu C, Bayrak-Toydemir P, McDonald J, Letarte M. Potential second-hits in hereditary hemorrhagic telangiectasia. J Clin Med. 2020;9(11):3571.
    1. Snodgrass RO, Chico TJA, Arthur HM. Hereditary haemorrhagic telangiectasia, an inherited vascular disorder in need of improved evidence-based pharmaceutical interventions. Genes (Basel). 2021;12(2):174.
    1. Finnamore H, Le Couteur J, Hickson M, Busbridge M, Whelan K, Shovlin CL. Hemorrhage-adjusted iron requirements, hematinics and hepcidin define hereditary hemorrhagic telangiectasia as a model of hemorrhagic iron deficiency. PLoS One. 2013;8(10):e76516.
    1. Al-Samkari H, Kasthuri RS, Parambil JG, et al. . An international, multicenter study of intravenous bevacizumab for bleeding in hereditary hemorrhagic telangiectasia: the InHIBIT-Bleed study. Haematologica. 2021;106(8):2161-2169.
    1. Al-Samkari H. Hereditary hemorrhagic telangiectasia: systemic therapies, guidelines, and an evolving standard of care. Blood. 2021;137(7): 888-895.
    1. Buscarini E, Botella LM, Geisthoff U, et al. ; VASCERN-HHT . Safety of thalidomide and bevacizumab in patients with hereditary hemorrhagic telangiectasia. Orphanet J Rare Dis. 2019;14(1):28.
    1. Kjeldsen A, Aagaard KS, Tørring PM, Möller S, Green A. 20-year follow-up study of Danish HHT patients: survival and causes of death. Orphanet J Rare Dis. 2016;11(1):157.
    1. de Gussem EM, Edwards CP, Hosman AE, et al. . Life expectancy of parents with hereditary haemorrhagic telangiectasia. Orphanet J Rare Dis. 2016;11(1):46.
    1. Shovlin CL, Hughes JM, Tuddenham EG, et al. . A gene for hereditary haemorrhagic telangiectasia maps to chromosome 9q3. Nat Genet. 1994; 6(2):205-209.
    1. Dupuis O, Delagrange L, Dupuis-Girod S. Hereditary haemorrhagic telangiectasia and pregnancy: a review of the literature. Orphanet J Rare Dis. 2020;15(1):5.
    1. Eker OF, Boccardi E, Sure U, et al. . European Reference Network for Rare Vascular Diseases (VASCERN) position statement on cerebral screening in adults and children with hereditary haemorrhagic telangiectasia (HHT). Orphanet J Rare Dis. 2020;15(1):165.
    1. Thielemans L, Layton DM, Shovlin CL. Low serum haptoglobin and blood films suggest intravascular hemolysis contributes to severe anemia in hereditary hemorrhagic telangiectasia. Haematologica. 2019;104(4):e127-e130.
    1. Devlin HL, Hosman AE, Shovlin CL. Antiplatelet and anticoagulant agents in hereditary hemorrhagic telangiectasia. N Engl J Med. 2013;368(9):876-878.
    1. Shovlin CL, Sulaiman NL, Govani FS, Jackson JE, Begbie ME. Elevated factor VIII in hereditary haemorrhagic telangiectasia (HHT): association with venous thromboembolism. Thromb Haemost. 2007;98(5):1031-1039.
    1. Livesey JA, Manning RA, Meek JH, et al. . Low serum iron levels are associated with elevated plasma levels of coagulation factor VIII and pulmonary emboli/deep venous thromboses in replicate cohorts of patients with hereditary haemorrhagic telangiectasia. Thorax. 2012;67(4):328-333.
    1. Shovlin C, Bamford K, Sabbà C, et al. ; VASCERN HHT . Prevention of serious infections in hereditary hemorrhagic telangiectasia: roles for prophylactic antibiotics, the pulmonary capillaries-but not vaccination. Haematologica. 2019;104(2):e85-e86.
    1. Boother EJ, Brownlow S, Tighe HC, Bamford KB, Jackson JE, Shovlin CL. Cerebral abscess associated with odontogenic bacteremias, hypoxemia, and iron loading in immunocompetent patients with right-to-left shunting through pulmonary arteriovenous malformations. Clin Infect Dis. 2017; 65(4):595-603.
    1. Mathis S, Dupuis-Girod S, Plauchu H, et al. . Cerebral abscesses in hereditary haemorrhagic telangiectasia: a clinical and microbiological evaluation. Clin Neurol Neurosurg. 2012;114(3):235-240.
    1. Dupuis-Girod S, Giraud S, Decullier E, et al. . Hemorrhagic hereditary telangiectasia (Rendu-Osler disease) and infectious diseases: an underestimated association. Clin Infect Dis. 2007;44(6):841-845.
    1. Topiwala KK, Patel SD, Saver JL, Streib CD, Shovlin CL. Ischemic stroke and pulmonary arteriovenous malformations: a review. Neurology. 2022;98(5):188-198.
    1. Silva BM, Hosman AE, Devlin HL, Shovlin CL. Lifestyle and dietary influences on nosebleed severity in hereditary hemorrhagic telangiectasia. Laryngoscope. 2013;123(5):1092-1099.
    1. Finnamore H, Silva BM, Hickson BM, Whelan K, Shovlin CL. 7-day weighed food diaries suggest patients with hereditary hemorrhagic telangiectasia may spontaneously modify their diet to avoid nosebleed precipitants. Orphanet J Rare Dis. 2017;12(1):60.
    1. Chamali B, Finnamore H, Manning R, et al. . Dietary supplement use and nosebleeds in hereditary haemorrhagic telangiectasia: an observational study. Intractable Rare Dis Res. 2016;5(2):109-113.
    1. Cavalcoli F, Gandini A, Matelloni IA, et al. . Dietary iron intake and anemia: food frequency questionnaire in patients with hereditary hemorrhagic telangiectasia. Orphanet J Rare Dis. 2020;15(1):295.
    1. Shovlin CL, Gilson C, Busbridge M, et al. . Can iron treatments aggravate epistaxis in some patients with hereditary hemorrhagic telangiectasia? Laryngoscope. 2016;126(11):2468-2474.
    1. 100,000 Genomes Project Pilot Investigators, Smedley D, Smith KR, et al. . 100,000 genomes pilot on rare-disease diagnosis in health care: preliminary report. N Engl J Med. 2021;385(20):1868-1880.
    1. Genomics England Confluence. Research environment user guide: Genomics England Research Environment. . Accessed 10 August 2021.
    1. Clarke JM, Alikian M, Xiao S, et al. ; Genomics England Research Consortium . Low grade mosaicism in hereditary haemorrhagic telangiectasia identified by bidirectional whole genome sequencing reads through the 100,000 Genomes Project clinical diagnostic pipeline. J Med Genet. 2020;57(12):859-862.
    1. Balachandar S, Graves TJ, Shimonty A, et al. ; Genomics England Research Consortium . Identification and validation of a novel pathogenic variant in GDF2 (BMP9) responsible for hereditary hemorrhagic telangiectasia and pulmonary arteriovenous malformations. Am J Med Genet A. 2022; 188(3):959-964.
    1. Shovlin CL, Guttmacher AE, Buscarini E, et al. . Diagnostic criteria for hereditary hemorrhagic telangiectasia (Rendu-Osler-Weber syndrome). Am J Med Genet. 2000;91(1):66-67.
    1. Shovlin CL, Chamali B, Santhirapala V, et al. . Ischaemic strokes in patients with pulmonary arteriovenous malformations and hereditary hemorrhagic telangiectasia: associations with iron deficiency and platelets. PLoS One. 2014;9(2):e88812.
    1. Gawecki F, Strangeways T, Amin A, et al. . Exercise capacity reflects airflow limitation rather than hypoxaemia in patients with pulmonary arteriovenous malformations. QJM. 2019;112(5):335-342.
    1. Köhler S, Doelken SC, Mungall CJ, et al. . The Human Phenotype Ontology project: linking molecular biology and disease through phenotype data. Nucleic Acids Res. 2014;42(database issue):D966-D974.
    1. National Center for Biotechnology Information. Genome reference consortium human build 38/hg38. . Accessed 23 December 2021.
    1. Martin AR, Williams E, Foulger RE, et al. . PanelApp crowdsources expert knowledge to establish consensus diagnostic gene panels. Nat Genet. 2019;51(11):1560-1565.
    1. Shovlin CL, Millar CM, Droege F, et al. ; VASCERN-HHT . Safety of direct oral anticoagulants in patients with hereditary hemorrhagic telangiectasia. Orphanet J Rare Dis. 2019;14(1):210.
    1. Shovlin CL, Sodhi V, McCarthy A, Lasjaunias P, Jackson JE, Sheppard MN. Estimates of maternal risks of pregnancy for women with hereditary haemorrhagic telangiectasia (Osler-Weber-Rendu syndrome): suggested approach for obstetric services. BJOG. 2008;115(9):1108-1115.
    1. Shovlin CL, Condliffe R, Donaldson JW, Kiely DG, Wort SJ; British Thoracic Society . British Thoracic Society Clinical Statement on pulmonary arteriovenous malformations. Thorax. 2017;72(12):1154-1163.
    1. Anderson E, Sharma L, Alsafi A, Shovlin CL. Pulmonary arteriovenous malformations can be the only feature of genetically-confirmed hereditary haemorrhagic telangiectasia which commonly displays few clinical manifestations [published online ahead of print 14 February 2022]. Thorax. doi: 10.1136/thoraxjnl-2021-218332.
    1. Kim Y, Park J, Kim M. Diagnostic approaches for inherited hemolytic anemia in the genetic era. Blood Res. 2017;52(2):84-94.
    1. Downes K, Megy K, Duarte D, et al. ; NIHR BioResource . Diagnostic high-throughput sequencing of 2396 patients with bleeding, thrombotic, and platelet disorders. Blood. 2019;134(23):2082-2091.
    1. Ferraro NM, Strober BJ, Einson J, et al. ; GTEx Consortium . Transcriptomic signatures across human tissues identify functional rare genetic variation. Science. 2020;369(6509):eaaz5900.
    1. Bernabeu-Herrero ME, Patel D, Bielowka A, et al. . Heterozygous transcriptional and nonsense decay signatures in blood outgrowth endothelial cells from patients with hereditary haemorrhagic telangiectasia [published online ahead of print 6 December 2021]. Biorχiv 471269. doi: 10.1101/2021.12.05.471269
    1. Richards S, Aziz N, Bale S, et al. ; ACMG Laboratory Quality Assurance Committee . Standards and guidelines for the interpretation of sequence variants: a joint consensus recommendation of the American College of Medical Genetics and Genomics and the Association for Molecular Pathology. Genet Med. 2015;17(5):405-424.
    1. Davieson CD, Joyce KE, Sharma L, Shovlin CL. DNA variant classification-reconsidering “allele rarity” and “phenotype” criteria in ACMG/AMP guidelines. Eur J Med Genet. 2021;64(10):104312.
    1. Krzywinski M, Schein J, Birol I, et al. . Circos: an information aesthetic for comparative genomics. Genome Res. 2009;19(9):1639-1645.
    1. Jalili V, Afgan E, Gu Q, et al. . The Galaxy platform for accessible, reproducible and collaborative biomedical analyses: 2020 update. Nucleic Acids Res. 2020;48(W1):W395-W402.
    1. Itan Y, Shang L, Boisson B, et al. . The human gene damage index as a gene-level approach to prioritizing exome variants. Proc Natl Acad Sci USA. 2015;112(44):13615-13620.
    1. Petrovski S, Wang Q, Heinzen EL, Allen AS, Goldstein DB. Genic intolerance to functional variation and the interpretation of personal genomes. PLoS Genet. 2013;9(8):e1003709.
    1. Kircher M, Witten DM, Jain P, O’Roak BJ, Cooper GM, Shendure J. A general framework for estimating the relative pathogenicity of human genetic variants. Nat Genet. 2014;46(3):310-315.
    1. Rentzsch P, Witten D, Cooper GM, Shendure J, Kircher M. CADD: predicting the deleteriousness of variants throughout the human genome. Nucleic Acids Res. 2019;47(D1):D886-D894.
    1. Rentzsch P, Schubach M, Shendure J, Kircher M. CADD-Splice-improving genome-wide variant effect prediction using deep learning-derived splice scores. Genome Med. 2021;13(1):31.
    1. Itan Y, Shang L, Boisson B, et al. . The mutation significance cutoff: gene-level thresholds for variant predictions. Nat Methods. 2016;13(2): 109-110.
    1. Karczewski KJ, Francioli LC, Tiao G, et al. ; Genome Aggregation Database Consortium . The mutational constraint spectrum quantified from variation in 141,456 humans. Nature. 2020;581(7809):434-443.
    1. Landrum MJ, Chitipiralla S, Brown GR, et al. . ClinVar: improvements to accessing data. Nucleic Acids Res. 2020;48(D1):D835-D844.
    1. Kent WJ, Sugnet CW, Furey TS, et al. . The human genome browser at UCSC. Genome Res. 2002;12(6):996-1006.
    1. Lee BT, Barber GP, Benet-Pagès A, et al. . The UCSC Genome Browser database: 2022 update. Nucleic Acids Res. 2022;50(D1):D1115-D1122.
    1. Benjamini Y, Hochberg Y. Controlling the False Discovery Rate: a practical and powerful approach to multiple testing. J R Stat Soc B. 1995;57(1):289-300.
    1. Shovlin CL, Jackson JE, Bamford KB, et al. . Primary determinants of ischaemic stroke/brain abscess risks are independent of severity of pulmonary arteriovenous malformations in hereditary haemorrhagic telangiectasia. Thorax. 2008;63(3):259-266.
    1. Precision medicine needs an equity agenda. Nat Med. 2021;27(5):737.
    1. Rehm HL, Berg JS, Brooks LD, et al. ; ClinGen . ClinGen: the clinical genome resource. N Engl J Med. 2015;372(23):2235-2242.
    1. Epi25 Collaborative. Sub-genic intolerance, ClinVar, and the epilepsies: a whole-exome sequencing study of 29,165 individuals. Am J Hum Genet. 2021;108(6):965-982.
    1. Keramati AR, Chen MH, Rodriguez BAT, et al. ; NHLBI Trans-Omics for Precision (TOPMed) Consortium . Genome sequencing unveils a regulatory landscape of platelet reactivity. Nat Commun. 2021;12(1):3626.
    1. Hu Y, Stilp AM, McHugh CP, et al. ; NHLBI Trans-Omics for Precision Medicine (TOPMed) Consortium . Whole-genome sequencing association analysis of quantitative red blood cell phenotypes: the NHLBI TOPMed program. Am J Hum Genet. 2021;108(5):874-893.
    1. Zaninoni A, Fermo E, Vercellati C, Marcello AP, Barcellini W, Bianchi P. Congenital hemolytic anemias: is there a role for the immune system? Front Immunol. 2020;11:1309.
    1. Limeres Posse J, Álvarez Fernández M, Fernández Feijoo J, et al. . Intravenous amoxicillin/clavulanate for the prevention of bacteraemia following dental procedures: a randomized clinical trial. J Antimicrob Chemother. 2016;71(7):2022-2030.
    1. Kanoi BN, Nagaoka H, White MT, et al. . Global repertoire of human antibodies against Plasmodium falciparum RIFINs, SURFINs, and STEVORs in a malaria exposed population. Front Immunol. 2020;11:893.

Source: PubMed

3
Abonner