The genetic basis of disease

  • December 2018
  • Essays in Biochemistry 62(5):643-723
  • 62(5):643-723
  • CC BY-NC-ND 4.0
  • This person is not on ResearchGate, or hasn't claimed this research yet.

Joanna Wilson at University of Glasgow

  • University of Glasgow

Abstract and Figures

Some types of variants found in human genomes Variation involving one or a few nucleotides are shown above the chromosome icon, and structural variants below; in each case the variants are depicted in relation to the reference sequence. For depiction of structural variants A, B, C and D represent large segments of DNA; Y and Z represent segments of DNA from a different chromosome. Note that differentiation between CNVs and deletions/insertions depends upon the size of the relevant DNA segment (see text for further details). Abbreviation: CNV, copy number variant. Chromosome ideogram from NCBI Genome Decoration Page.

Discover the world's research

  • 25+ million members
  • 160+ million publication pages
  • 2.3+ billion citations
  • Aliya Medetbekova

S. Zh. Kolumbayeva

  • Saniya Dauletbayeva
  • PSYCHOPHARMACOLOGY
  • Asmita Deka Dey

Ashi Mannan

  • Oscar Aguilar
  • Cheng Chang

Elsa Bismuth

  • Manuel A. Rivas
  • Xiaoling Tian
  • ENVIRON MOL MUTAGEN

Barbara L Parsons

  • Kerry L. Dearfield
  • Paul A White
  • ADV FUNCT MATER

Xinpu Yang

  • Thuy Anh Bui

Ying Liu

  • BIOMACROMOLECULES

Curt Jarand

  • Karen Baker
  • Matthew Petroff
  • Wayne F. Reed

Naveen Pant

  • Zhenzhen Deng
  • Yuanhong Liu
  • Haiyan Zhou
  • Khalid M. Sumaily

Ahmed Mujamammi

  • Somali Chaterji
  • Eun Hyun Ahn
  • Deok-Ho Kim
  • Jan Rune Aunan
  • William C. Cho
  • Kjetil Søreide
  • PHILOS T R SOC B
  • Giancarlo Bonora
  • Christine M. Disteche

Martin Fischer

  • Inês Pinheiro

Edith Heard

  • MOL CELL ENDOCRINOL

Isabelle Stévant

  • Erin Murphy
  • Recruit researchers
  • Join for free
  • Login Email Tip: Most researchers use their institutional email address as their ResearchGate login Password Forgot password? Keep me logged in Log in or Continue with Google Welcome back! Please log in. Email · Hint Tip: Most researchers use their institutional email address as their ResearchGate login Password Forgot password? Keep me logged in Log in or Continue with Google No account? Sign up
  • DOI: 10.30970/vlubs.2024.91.04
  • Corpus ID: 270350001

Snijders Blok-Campeau syndrome: a novel neurodevelopmental genetic disorder

  • O. Yushchuk , I. Ruda , V. Fedorenko
  • Published in Visnyk of Lviv University… 7 June 2024

34 References

Snijders blok–campeau syndrome: description of 20 additional individuals with variants in chd3 and literature review, a severe neurocognitive phenotype caused by biallelic chd3 variants in two siblings, hypersociability associated with developmental delay, macrocephaly and facial dysmorphism points to chd3 mutations., a novel chd3 variant in a patient with central precocious puberty: expanded phenotype of snijders blok‐campeau syndrome, inherited variants in chd3 show variable expressivity in snijders blok-campeau syndrome., a second cohort of chd3 patients expands the molecular mechanisms known to cause snijders blok-campeau syndrome, a de novo chd3 variant in a child with intellectual disability, autism, joint laxity, and dysmorphisms, mutations in chd7, encoding a chromatin-remodeling protein, cause idiopathic hypogonadotropic hypogonadism and kallmann syndrome., chd2-related cns pathologies, chd3 helicase domain mutations cause a neurodevelopmental syndrome with macrocephaly and impaired speech and language, related papers.

Showing 1 through 3 of 0 Related Papers

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

ijms-logo

Article Menu

gene disorder research paper

  • Subscribe SciFeed
  • Recommended Articles
  • Author Biographies
  • PubMed/Medline
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Unraveling the genetic landscape of neurological disorders: insights into pathogenesis, techniques for variant identification, and therapeutic approaches.

gene disorder research paper

1. Introduction

2. genetic techniques in addressing neurodegenerative disorders, 3. genetic mutations and corresponding cellular alterations in neurodegenerative disorders, 3.1. amyotrophic lateral sclerosis (als), 3.1.1. genetic and pathological overlap between als and ftd, 3.1.2. epidemiology, 3.1.3. genetic causes and risk factor.

  • Superoxide dismutase ( SOD1 )
  • Chromosome 9 open reading frame 72 ( C9ORF72 )

3.1.4. Additional Risk Loci from Genome-Wide Association Studies

3.2. alzheimer’s disease (ad), 3.2.1. epidemiology, 3.2.2. genetic causes and risk factors, 3.2.3. additional risk loci from genome-wide association studies (gwass), 3.3. parkinson’s disease (pd), 3.3.1. epidemiology, 3.3.2. genetic causes and risk factors, 3.3.3. additional risk loci from genome-wide association studies, 3.3.4. other neurodegenerative disorders, 4. gene therapy for neurodegenerative diseases, 5. gene expression, 5.1. the exogenous introduction of genes into the cns, 5.2. dna editing, 5.3. crispr-mediated base editing and prime editing, 6. genetic therapy for ad, 6.1. targeting mapt, 6.2. targeting apoe, 6.3. targeting app, 7. gene therapy for pd, 7.1. modulating neuronal signaling, 7.2. targeting disease genes—snca, gba, and lrrk2, 8. gene therapy for als, targeting sod1, 9. conclusions and future perspectives, author contributions, acknowledgments, conflicts of interest.

  • Nichols, E.; Steinmetz, J.D.; Vollset, S.E.; Fukutaki, K.; Chalek, J.; Abd-Allah, F.; Abdoli, A.; Abualhasan, A.; Abu-Gharbieh, E.; Akram, T.T. Estimation of the global prevalence of dementia in 2019 and forecasted prevalence in 2050: An analysis for the Global Burden of Disease Study 2019. Lancet Public Health 2022 , 7 , e105–e125. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Blackhurst, B.M.; Funk, K.E. Viral pathogens increase risk of neurodegenerative disease. Nat. Rev. Neurol. 2023 , 19 , 259–260. [ Google Scholar ] [ CrossRef ]
  • Cerejeira, J.; Lagarto, L.; Mukaetova-Ladinska, E.B. Behavioral and psychological symptoms of dementia. Front. Neurol. 2012 , 3 , 73. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Calderón-Garcidueñas, L. Common Fatal Neurodegenerative Diseases Revisited: Beyond Age, Comorbidities, and Devastating Terminal Neuropathology There Is Hope with Prevention. Front. Neurol. 2022 , 13 , 901447. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Kovacs, G.G.; Botond, G.; Budka, H. Protein coding of neurodegenerative dementias: The neuropathological basis of biomarker diagnostics. Acta Neuropathol. 2010 , 119 , 389–408. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Goutman, S.A.; Hardiman, O.; Al-Chalabi, A.; Chió, A.; Savelieff, M.G.; Kiernan, M.C.; Feldman, E.L. Recent advances in the diagnosis and prognosis of amyotrophic lateral sclerosis. Lancet Neurol. 2022 , 21 , 480–493. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Spires-Jones, T.L.; Attems, J.; Thal, D.R. Interactions of pathological proteins in neurodegenerative diseases. Acta Neuropathol. 2017 , 134 , 187–205. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Wang, Y.A.; Kammenga, J.E.; Harvey, S.C. Genetic variation in neurodegenerative diseases and its accessibility in the model organism Caenorhabditis elegans. Hum. Genom. 2017 , 11 , 12. [ Google Scholar ] [ CrossRef ]
  • Bocchetta, M.; Mega, A.; Bernardi, L.; Di Maria, E.; Benussi, L.; Binetti, G.; Borroni, B.; Colao, R.; Di Fede, G.; Fostinelli, S.; et al. Genetic counseling and testing for alzheimer’s disease and frontotemporal lobar degeneration: An Italian consensus protocol. J. Alzheimer’s Dis. 2016 , 51 , 277–291. [ Google Scholar ] [ CrossRef ]
  • Fernández, M.V.; Kim, J.H.; Budde, J.P.; Black, K.; Medvedeva, A.; Saef, B.; Deming, Y.; Del-Aguila, J.; Ibañez, L.; Dube, U.; et al. Analysis of neurodegenerative Mendelian genes in clinically diagnosed Alzheimer Disease. PLoS Genet. 2017 , 13 , e1007045. [ Google Scholar ] [ CrossRef ]
  • Woollacott, I.O.C.; Rohrer, J.D. The clinical spectrum of sporadic and familial forms of frontotemporal dementia. J. Neurochem. 2016 , 138 , 6–31. [ Google Scholar ] [ CrossRef ]
  • Faghihi, M.A.; Mottagui-Tabar, S.; Wahlestedt, C. Genetics of neurological disorders. Expert Rev. Mol. Diagn. 2004 , 4 , 317–332. [ Google Scholar ] [ CrossRef ]
  • Koretsky, M.J.; Alvarado, C.; Makarious, M.B.; Vitale, D.; Levine, K.; Bandres-Ciga, S.; Dadu, A.; Scholz, S.W.; Sargent, L.; Faghri, F.; et al. Genetic risk factor clustering within and across neurodegenerative diseases. Brain 2023 , 146 , 4486–4494. [ Google Scholar ] [ CrossRef ]
  • Jiang, T.; Tan, M.-S.; Tan, L.; Yu, J.-T. Application of next-generation sequencing technologies in Neurology. Ann. Transl. Med. 2014 , 2 , 125. [ Google Scholar ] [ CrossRef ]
  • Guo, M.H.; Hirschhorn, J.N.; Dauber, A. Insights and implications of genome-wide association studies of height. J. Clin. Endocrinol. Metab. 2018 , 103 , 3155–3168. [ Google Scholar ] [ CrossRef ]
  • Craig, J. Complex Diseases: Research and Applications. Nat. Educ. 2008 , 1 , 184. [ Google Scholar ]
  • Tan, M.-S.; Jiang, T.; Tan, L.; Yu, J.-T. Genome-wide association studies in neurology. Ann. Transl. Med. 2014 , 2 , 125. [ Google Scholar ] [ CrossRef ]
  • Manolio, T.A.; Collins, F.S.; Cox, N.J.; Goldstein, D.B.; Hindorff, L.A.; Hunter, D.J.; McCarthy, M.I.; Ramos, E.M.; Cardon, L.R.; Chakravarti, A.; et al. Finding the missing heritability of complex diseases. Nature 2009 , 461 , 747–753. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Young, A.I. Solving the missing heritability problem. PLoS Genet. 2019 , 15 , e1008222. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Tsuji, S. Genetics of neurodegenerative diseases: Insights from high-throughput resequencing. Hum. Mol. Genet. 2010 , 19 , R65–R70. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Dilliott, A.A.; Abdelhady, A.; Sunderland, K.M.; Farhan, S.M.K.; Abrahao, A.; Binns, M.A.; Black, S.E.; Borrie, M.; Casaubon, L.K.; Dowlatshahi, D.; et al. Contribution of rare variant associations to neurodegenerative disease presentation. npj Genom. Med. 2021 , 6 , 80. [ Google Scholar ] [ CrossRef ]
  • Goswami, C.; Chattopadhyay, A.; Chuang, E.Y. Rare variants: Data types and analysis strategies. Ann. Transl. Med. 2021 , 9 , 961. [ Google Scholar ] [ CrossRef ]
  • Zhao, Y.; Fang, L.T.; Shen, T.-W.; Choudhari, S.; Talsania, K.; Chen, X.; Shetty, J.; Kriga, Y.; Tran, B.; Zhu, B.; et al. Whole genome and exome sequencing reference datasets from a multi-center and cross-platform benchmark study. Sci. Data 2021 , 8 , 296. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Pervez, M.T.; Hasnain, M.J.U.; Abbas, S.H.; Moustafa, M.F.; Aslam, N.; Shah, S.S.M. A comprehensive review of performance of next-generation sequencing platforms. BioMed Res. Int. 2022 , 2022 , 3457806. [ Google Scholar ] [ CrossRef ]
  • Chen, Y.-C.; Liu, T.; Yu, C.-H.; Chiang, T.-Y.; Hwang, C.-C. Effects of GC bias in next-generation-sequencing data on de novo genome assembly. PLoS ONE 2013 , 8 , e62856. [ Google Scholar ] [ CrossRef ]
  • Eid, J.; Fehr, A.; Gray, J.; Luong, K.; Lyle, J.; Otto, G.; Peluso, P.; Rank, D.; Baybayan, P.; Bettman, B.; et al. Real-time DNA sequencing from single polymerase molecules. Science 2009 , 323 , 133–138. [ Google Scholar ] [ CrossRef ]
  • Clarke, J.; Wu, H.-C.; Jayasinghe, L.; Patel, A.; Reid, S.; Bayley, H. Continuous base identification for single-molecule nanopore DNA sequencing. Nat. Nanotechnol. 2009 , 4 , 265–270. [ Google Scholar ] [ CrossRef ]
  • Su, Y.; Fan, L.; Shi, C.; Wang, T.; Zheng, H.; Luo, H.; Zhang, S.; Hu, Z.; Fan, Y.; Dong, Y.; et al. Deciphering neurodegenerative diseases using long-read sequencing. Neurology 2021 , 97 , 423–433. [ Google Scholar ] [ CrossRef ]
  • Oehler, J.B.; Wright, H.; Stark, Z.; Mallett, A.J.; Schmitz, U. The application of long-read sequencing in clinical settings. Hum. Genom. 2023 , 17 , 73. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Hafford-Tear, N.J.; Tsai, Y.-C.; Sadan, A.N.; Sanchez-Pintado, B.; Zarouchlioti, C.; Maher, G.J.; Liskova, P.; Tuft, S.J.; Hardcastle, A.J.; Clark, T.A.; et al. CRISPR/Cas9-targeted enrichment and long-read sequencing of the Fuchs endothelial corneal dystrophy–associated TCF4 triplet repeat. Genet. Med. 2019 , 21 , 2092–2102. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Miga, K.H.; Koren, S.; Rhie, A.; Vollger, M.R.; Gershman, A.; Bzikadze, A.; Brooks, S.; Howe, E.; Porubsky, D.; Logsdon, G.A.; et al. Telomere-to-telomere assembly of a complete human X chromosome. Nature 2020 , 585 , 79–84. [ Google Scholar ] [ CrossRef ]
  • Nurk, S.; Koren, S.; Rhie, A.; Rautiainen, M.; Bzikadze, A.V.; Mikheenko, A.; Vollger, M.R.; Altemose, N.; Uralsky, L.; Gershman, A.; et al. The complete sequence of a human genome. Science 2022 , 376 , 44–53. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Chaisson, M.J.P.; Huddleston, J.; Dennis, M.Y.; Sudmant, P.H.; Malig, M.; Hormozdiari, F.; Antonacci, F.; Surti, U.; Sandstrom, R.; Boitano, M.; et al. Resolving the complexity of the human genome using single-molecule sequencing. Nature 2015 , 517 , 608–611. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ebert, P.; Audano, P.A.; Zhu, Q.; Rodriguez-Martin, B.; Porubsky, D.; Bonder, M.J.; Sulovari, A.; Ebler, J.; Zhou, W.; Mari, R.S.; et al. Haplotype-resolved diverse human genomes and integrated analysis of structural variation. Science 2021 , 372 , 48. [ Google Scholar ] [ CrossRef ]
  • Theunissen, F.; Flynn, L.L.; Anderton, R.S.; Mastaglia, F.; Pytte, J.; Jiang, L.; Hodgetts, S.; Burns, D.K.; Saunders, A.; Fletcher, S.; et al. Structural variants may be a source of missing heritability in sALS. Front. Neurosci. 2020 , 14 , 47. [ Google Scholar ] [ CrossRef ]
  • Parobkova, E.; Matej, R. Amyotrophic Lateral Sclerosis and Frontotemporal Lobar Degenerations: Similarities in Genetic Background. Diagnostics 2021 , 11 , 509. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Perrone, F.; Nguyen, H.P.; Van Mossevelde, S.; Moisse, M.; Sieben, A.; Santens, P.; De Bleecker, J.; Vandenbulcke, M.; Engelborghs, S.; Baets, J.; et al. Investigating the role of ALS genes CHCHD10 and TUBA4A in Belgian FTD-ALS spectrum patients. Neurobiol. Aging 2017 , 51 , 177.e9–177.e16. [ Google Scholar ] [ CrossRef ]
  • Ahmed, R.M.; Devenney, E.M.; Strikwerda-Brown, C.; Hodges, J.R.; Piguet, O.; Kiernan, M.C. Phenotypic variability in ALS-FTD and effect on survival. Neurology 2020 , 94 , e2005–e2013. [ Google Scholar ] [ CrossRef ]
  • Abramzon, Y.A.; Fratta, P.; Traynor, B.J.; Chia, R. The Overlapping Genetics of Amyotrophic Lateral Sclerosis and Frontotemporal Dementia. Front. Neurosci. 2020 , 14 , 42. [ Google Scholar ] [ CrossRef ]
  • Olney, N.T.; Spina, S.; Miller, B.L. Frontotemporal Dementia. Neurol. Clin. 2017 , 35 , 339–374. [ Google Scholar ] [ CrossRef ]
  • Broce, I.J.; Castruita, P.A.; Yokoyama, J.S. Moving Toward Patient-Tailored Treatment in ALS and FTD: The Potential of Genomic Assessment as a Tool for Biological Discovery and Trial Recruitment. Front. Neurosci. 2021 , 15 , 639078. [ Google Scholar ] [ CrossRef ]
  • Marin, B.; Boumédiene, F.; Logroscino, G.; Couratier, P.; Babron, M.-C.; Leutenegger, A.-L.; Copetti, M.; Preux, P.-M.; Beghi, E. Variation in worldwide incidence of amyotrophic lateral sclerosis: A meta-analysis. Int. J. Epidemiol. 2017 , 46 , 57–74. [ Google Scholar ] [ CrossRef ]
  • Feldman, E.L.; Goutman, S.A.; Petri, S.; Mazzini, L.; Savelieff, M.G.; Shaw, P.J.; Sobue, G. Amyotrophic lateral sclerosis. Lancet 2022 , 400 , 1363–1380. [ Google Scholar ] [ CrossRef ]
  • Chia, R.; Chiò, A.; Traynor, B.J. Novel genes associated with amyotrophic lateral sclerosis: Diagnostic and clinical implications. Lancet Neurol. 2018 , 17 , 94–102. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Genç, B.; Gautam, M.; Helmold, B.R.; Koçak, N.; Günay, A.; Goshu, G.M.; Silverman, R.B.; Ozdinler, P.H. NU-9 improves health of hSOD1G93A mouse upper motor neurons in vitro, especially in combination with riluzole or edaravone. Sci. Rep. 2022 , 12 , 5383. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Merjane, J.; Chung, R.; Patani, R.; Lisowski, L. Molecular mechanisms of amyotrophic lateral sclerosis as broad therapeutic targets for gene therapy applications utilizing adeno-associated viral vectors. Med. Res. Rev. 2023 , 43 , 829–854. [ Google Scholar ] [ CrossRef ]
  • Aschenbrenner, D.S. New Drug Approved For ALS. AJN Am. J. Nurs. 2023 , 123 , 22–23. [ Google Scholar ] [ CrossRef ]
  • Goutman, S.A.; Hardiman, O.; Al-Chalabi, A.; Chió, A.; Savelieff, M.G.; Kiernan, M.C.; Feldman, E.L. Emerging insights into the complex genetics and pathophysiology of amyotrophic lateral sclerosis. Lancet Neurol. 2022 , 21 , 465–479. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ryan, M.; Heverin, M.; McLaughlin, R.L.; Hardiman, O. Lifetime risk and heritability of amyotrophic lateral sclerosis. JAMA Neurol. 2019 , 76 , 1367–1374. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Nicolas, A.; Kenna, K.P.; Renton, A.E.; Ticozzi, N.; Faghri, F.; Chia, R.; Dominov, J.A.; Kenna, B.J.; Nalls, M.A.; Keagle, P.; et al. Genome-wide analyses identify KIF5A as a novel ALS gene. Neuron 2018 , 97 , 1268–1283.e6. [ Google Scholar ] [ CrossRef ]
  • Shahrizaila, N.; Sobue, G.; Kuwabara, S.; Kim, S.H.; Birks, C.; Fan, D.S.; Bae, J.S.; Hu, C.J.; Gourie-Devi, M.; Noto, Y.; et al. Amyotrophic lateral sclerosis and motor neuron syndromes in Asia. J. Neurol. Neurosurg. Psychiatry 2016 , 87 , 821–830. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Zou, Z.-Y.; Zhou, Z.-R.; Che, C.-H.; Liu, C.-Y.; He, R.-L.; Huang, H.-P. Genetic epidemiology of amyotrophic lateral sclerosis: A systematic review and meta-analysis. J. Neurol. Neurosurg. Psychiatry 2017 , 88 , 540–549. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Rosen, D.R.; Siddique, T.; Patterson, D.; Figlewicz, D.A.; Sapp, P.; Hentati, A.; Donaldson, D.; Goto, J.; O’Regan, J.P.; Deng, H.-X.; et al. Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature 1993 , 362 , 59–62. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Valentine, J.S.; Doucette, P.A.; Potter, S.Z. Copper-zinc superoxide dismutase and amyotrophic lateral sclerosis. Annu. Rev. Biochem. 2005 , 74 , 563–593. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Suzuki, N.; Nishiyama, A.; Warita, H.; Aoki, M. Genetics of amyotrophic lateral sclerosis: Seeking therapeutic targets in the era of gene therapy. J. Hum. Genet. 2023 , 68 , 131–152. [ Google Scholar ] [ CrossRef ]
  • Azadmanesh, J.; Borgstahl, G.E.O. A Review of the catalytic mechanism of human manganese superoxide dismutase. Antioxidants 2018 , 7 , 25. [ Google Scholar ] [ CrossRef ]
  • Corcia, P.; Valdmanis, P.; Millecamps, S.; Lionnet, C.; Blasco, H.; Mouzat, K.; Daoud, H.; Belzil, V.; Morales, R.; Pageot, N.; et al. Phenotype and genotype analysis in amyotrophic lateral sclerosis with TARDBP gene mutations. Neurology 2012 , 78 , 1519–1526. [ Google Scholar ] [ CrossRef ]
  • Hayashi, Y.; Homma, K.; Ichijo, H. SOD1 in neurotoxicity and its controversial roles in SOD1 mutation-negative ALS. Adv. Biol. Regul. 2016 , 60 , 95–104. [ Google Scholar ] [ CrossRef ]
  • Fujisawa, T.; Homma, K.; Yamaguchi, N.; Kadowaki, H.; Tsuburaya, N.; Naguro, I.; Matsuzawa, A.; Takeda, K.; Takahashi, Y.; Goto, J.; et al. A novel monoclonal antibody reveals a conformational alteration shared by amyotrophic lateral sclerosis-linked SOD1 mutants. Ann. Neurol. 2012 , 72 , 739–749. [ Google Scholar ] [ CrossRef ]
  • Mathis, S.; Couratier, P.; Julian, A.; Vallat, J.-M.; Corcia, P.; Le Masson, G. Management and therapeutic perspectives in amyotrophic lateral sclerosis. Expert Rev. Neurother. 2017 , 17 , 263–276. [ Google Scholar ] [ CrossRef ]
  • Li, K.; Hala, T.J.; Seetharam, S.; Poulsen, D.J.; Wright, M.C.; Lepore, A.C. GLT1 overexpression in SOD1G93A mouse cervical spinal cord does not preserve diaphragm function or extend disease. Neurobiol. Dis. 2015 , 78 , 12–23. [ Google Scholar ] [ CrossRef ]
  • Szelechowski, M.; Amoedo, N.; Obre, E.; Léger, C.; Allard, L.; Bonneu, M.; Claverol, S.; Lacombe, D.; Oliet, S.; Chevallier, S.; et al. Metabolic reprogramming in amyotrophic lateral sclerosis. Sci. Rep. 2018 , 8 , 3953. [ Google Scholar ] [ CrossRef ]
  • Le Masson, G.; Przedborski, S.; Abbott, L. A computational model of motor neuron degeneration. Neuron 2014 , 83 , 975–988. [ Google Scholar ] [ CrossRef ]
  • Lamanauskas, N.; Nistri, A. Riluzole blocks persistent Na + and Ca 2+ currents and modulates release of glutamate via presynaptic NMDA receptors on neonatal rat hypoglossal motoneurons in vitro . Eur. J. Neurosci. 2008 , 27 , 2501–2514. [ Google Scholar ] [ CrossRef ]
  • Ayers, J.I.; Fromholt, S.E.; O’neal, V.M.; Diamond, J.H.; Borchelt, D.R. Prion-like propagation of mutant SOD1 misfolding and motor neuron disease spread along neuroanatomical pathways. Acta Neuropathol. 2016 , 131 , 103–114. [ Google Scholar ] [ CrossRef ]
  • Polymenidou, M.; Cleveland, D.W. The seeds of neurodegeneration: Prion-like spreading in ALS. Cell 2011 , 147 , 498–508. [ Google Scholar ] [ CrossRef ]
  • Saeed, M.; Yang, Y.; Deng, H.-X.; Hung, W.-Y.; Siddique, N.; Dellefave, L.; Gellera, C.; Andersen, P.M.; Siddique, T. Age and founder effect of SOD1 A4V mutation causing ALS. Neurology 2009 , 72 , 1634–1639. [ Google Scholar ] [ CrossRef ]
  • Zou, Z.-Y.; Liu, M.-S.; Li, X.-G.; Cui, L.-Y. H46R SOD1 mutation is consistently associated with a relatively benign form of amyotrophic lateral sclerosis with slow progression. Amyotroph. Lateral Scler. Front. Degener. 2016 , 17 , 610–613. [ Google Scholar ] [ CrossRef ]
  • Kato, M.; Aoki, M.; Ohta, M.; Nagai, M.; Ishizaki, F.; Nakamura, S.; Itoyama, Y. Marked reduction of the Cu/Zn superoxide dismutase polypeptide in a case of familial amyotrophic lateral sclerosis with the homozygous mutation. Neurosci. Lett. 2001 , 312 , 165–168. [ Google Scholar ] [ CrossRef ]
  • Hideshima, M.; Beck, G.; Yamadera, M.; Motoyama, Y.; Ikenaka, K.; Kakuda, K.; Tsuda, H.; Nagano, S.; Fujimura, H.; Morii, E.; et al. A clinicopathological study of ALS with L126S mutation in the SOD1 gene presenting with isolated inferior olivary hypertrophy. Neuropathology 2020 , 40 , 191–195. [ Google Scholar ] [ CrossRef ]
  • Gagliardi, D.; Ahmadinejad, M.; Del Bo, R.; Meneri, M.; Comi, G.P.; Corti, S.; Ronchi, D. Homozygous SOD1 variation L144S produces a severe form of amyotrophic lateral sclerosis in an Iranian family. Neurol. Genet. 2021 , 8 , e645. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Berdyński, M.; Miszta, P.; Safranow, K.; Andersen, P.M.; Morita, M.; Filipek, S.; Żekanowski, C.; Kuźma-Kozakiewicz, M. SOD1 mutations associated with amyotrophic lateral sclerosis analysis of variant severity. Sci. Rep. 2022 , 12 , 103. [ Google Scholar ] [ CrossRef ]
  • DeJesus-Hernandez, M.; Mackenzie, I.R.; Boeve, B.F.; Boxer, A.L.; Baker, M.; Rutherford, N.J.; Nicholson, A.M.; Finch, N.A.; Flynn, H.; Adamson, J.; et al. Expanded GGGGCC hexanucleotide repeat in noncoding region of C9ORF72 causes chromosome 9p-linked FTD and ALS. Neuron 2011 , 72 , 245–256. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Bertrand, A.; Wen, J.; Rinaldi, D.; Houot, M.; Sayah, S.; Camuzat, A.; Fournier, C.; Fontanella, S.; Routier, A.; Couratier, P.; et al. Early cognitive, structural, and microstructural changes in presymptomatic C9orf72 carriers younger than 40 years. JAMA Neurol. 2018 , 75 , 236–245. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Koppers, M.; Blokhuis, A.M.; Westeneng, H.; Terpstra, M.L.; Zundel, C.A.C.; de Sá, R.V.; Schellevis, R.D.; Waite, A.J.; Blake, D.J.; Veldink, J.H.; et al. C9orf72 ablation in mice does not cause motor neuron degeneration or motor deficits. Ann. Neurol. 2015 , 78 , 426–438. [ Google Scholar ] [ CrossRef ]
  • Lee, Y.-B.; Chen, H.-J.; Peres, J.N.; Gomez-Deza, J.; Attig, J.; Štalekar, M.; Troakes, C.; Nishimura, A.L.; Scotter, E.L.; Vance, C.; et al. Hexanucleotide repeats in ALS/FTD form length-dependent RNA foci, sequester RNA binding proteins, and are neurotoxic. Cell Rep. 2013 , 5 , 1178–1186. [ Google Scholar ] [ CrossRef ]
  • Lee, Y.-B.; Baskaran, P.; Gomez-Deza, J.; Chen, H.-J.; Nishimura, A.L.; Smith, B.N.; Troakes, C.; Adachi, Y.; Stepto, A.; Petrucelli, L.; et al. C9orf72 poly GA RAN-translated protein plays a key role in amyotrophic lateral sclerosis via aggregation and toxicity. Hum. Mol. Genet. 2017 , 26 , 4765–4777. [ Google Scholar ] [ CrossRef ]
  • Selvaraj, B.T.; Livesey, M.R.; Zhao, C.; Gregory, J.M.; James, O.T.; Cleary, E.M.; Chouhan, A.K.; Gane, A.B.; Perkins, E.M.; Dando, O.; et al. C9ORF72 repeat expansion causes vulnerability of motor neurons to Ca2+-permeable AMPA receptor-mediated excitotoxicity. Nat. Commun. 2018 , 9 , 347. [ Google Scholar ] [ CrossRef ]
  • Mitra, J.; Guerrero, E.N.; Hegde, P.M.; Liachko, N.F.; Wang, H.; Vasquez, V.; Gao, J.; Pandey, A.; Taylor, J.P.; Kraemer, B.C.; et al. Motor neuron disease-associated loss of nuclear TDP-43 is linked to DNA double-strand break repair defects. Proc. Natl. Acad. Sci. USA 2019 , 116 , 4696–4705. [ Google Scholar ] [ CrossRef ]
  • Kabashi, E.; Valdmanis, P.N.; Dion, P.; Spiegelman, D.; McConkey, B.J.; Velde, C.V.; Bouchard, J.-P.; Lacomblez, L.; Pochigaeva, K.; Salachas, F.; et al. TARDBP mutations in individuals with sporadic and familial amyotrophic lateral sclerosis. Nat. Genet. 2008 , 40 , 572–574. [ Google Scholar ] [ CrossRef ]
  • Braak, H.; Brettschneider, J.; Ludolph, A.C.; Lee, V.M.; Trojanowski, J.Q.; Del Tredici, K. Amyotrophic lateral sclerosis—A model of corticofugal axonal spread. Nat. Rev. Neurol. 2013 , 9 , 708–714. [ Google Scholar ] [ CrossRef ]
  • Mitsuzawa, S.; Akiyama, T.; Nishiyama, A.; Suzuki, N.; Kato, M.; Warita, H.; Izumi, R.; Osana, S.; Koyama, S.; Kato, T.; et al. TARDBP p.G376D mutation, found in rapid progressive familial ALS, induces mislocalization of TDP-43. eNeurologicalSci 2018 , 11 , 20–22. [ Google Scholar ] [ CrossRef ]
  • Kametani, F.; Obi, T.; Shishido, T.; Akatsu, H.; Murayama, S.; Saito, Y.; Yoshida, M.; Hasegawa, M. Mass spectrometric analysis of accumulated TDP-43 in amyotrophic lateral sclerosis brains. Sci. Rep. 2016 , 6 , 23281. [ Google Scholar ] [ CrossRef ]
  • Donde, A.; Sun, M.; Ling, J.P.; Braunstein, K.E.; Pang, B.; Wen, X.; Cheng, X.; Chen, L.; Wong, P.C. Splicing repression is a major function of TDP-43 in motor neurons. Acta Neuropathol. 2019 , 138 , 813–826. [ Google Scholar ] [ CrossRef ]
  • Tank, E.M.; Figueroa-Romero, C.; Hinder, L.M.; Bedi, K.; Archbold, H.C.; Li, X.; Weskamp, K.; Safren, N.; Paez-Colasante, X.; Pacut, C.; et al. Abnormal RNA stability in amyotrophic lateral sclerosis. Nat. Commun. 2018 , 9 , 2845. [ Google Scholar ] [ CrossRef ]
  • Wood, A.; Gurfinkel, Y.; Polain, N.; Lamont, W.; Rea, S.L. Molecular mechanisms underlying TDP-43 pathology in cellular and animal models of ALS and FTLD. Int. J. Mol. Sci. 2021 , 22 , 4705. [ Google Scholar ] [ CrossRef ]
  • Kwiatkowski, T.J., Jr.; Bosco, D.A.; Leclerc, A.L.; Tamrazian, E.; Vanderburg, C.R.; Russ, C.; Davis, A.; Gilchrist, J.; Kasarskis, E.J.; Munsat, T.; et al. Mutations in the FUS/TLS gene on chromosome 16 cause familial amyotrophic lateral sclerosis. Science 2009 , 323 , 1205–1208. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Huebers, A.; Just, W.; Rosenbohm, A.; Mueller, K.; Marroquin, N.; Goebel, I.; Hoegel, J.; Thiele, H.; Altmueller, J.; Nuernberg, P.; et al. De novo FUS mutations are the most frequent genetic cause in early-onset German ALS patients. Neurobiol. Aging 2015 , 36 , 3117.e1–3117.e6. [ Google Scholar ]
  • Akiyama, T.; Warita, H.; Kato, M.; Nishiyama, A.; Izumi, R.; Ikeda, C.; Kamada, M.; Suzuki, N.; Aoki, M. Genotype–phenotype relationships in familial amyotrophic lateral sclerosis with FUS / TLS mutations in Japan. Muscle Nerve 2016 , 54 , 398–404. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Neumann, M.; Rademakers, R.; Roeber, S.; Baker, M.; Kretzschmar, H.A.; Mackenzie, I.R.A. A new subtype of frontotemporal lobar degeneration with FUS pathology. Brain 2009 , 132 , 2922–2931. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Salam, S.; Tacconelli, S.; Smith, B.N.; Mitchell, J.C.; Glennon, E.; Nikolaou, N.; Houart, C.; Vance, C. Identification of a novel interaction of FUS and syntaphilin may explain synaptic and mitochondrial abnormalities caused by ALS mutations. Sci. Rep. 2021 , 11 , 13613. [ Google Scholar ] [ CrossRef ]
  • Li, Y.R.; King, O.D.; Shorter, J.; Gitler, A.D. Stress granules as crucibles of ALS pathogenesis. J. Cell Biol. 2013 , 201 , 361–372. [ Google Scholar ] [ CrossRef ]
  • van Rheenen, W.; van der Spek, R.A.A.; Bakker, M.K.; van Vugt, J.J.F.A.; Hop, P.J.; Zwamborn, R.A.J.; de Klein, N.; Westra, H.-J.; Bakker, O.B.; Deelen, P.; et al. Common and rare variant association analyses in amyotrophic lateral sclerosis identify 15 risk loci with distinct genetic architectures and neuron-specific biology. Nat. Genet. 2021 , 53 , 1636–1648. [ Google Scholar ] [ CrossRef ]
  • Nguyen, H.P.; Van Broeckhoven, C.; van der Zee, J. ALS genes in the genomic era and their implications for FTD. Trends Genet. 2018 , 34 , 404–423. [ Google Scholar ] [ CrossRef ]
  • Lipstein, N.; Verhoeven-Duif, N.M.; Michelassi, F.E.; Calloway, N.; van Hasselt, P.M.; Pienkowska, K.; van Haaften, G.; van Haelst, M.M.; van Empelen, R.; Cuppen, I.; et al. Synaptic UNC13A protein variant causes increased neurotransmission and dyskinetic movement disorder. J. Clin. Investig. 2017 , 127 , 1005–1018. [ Google Scholar ] [ CrossRef ]
  • Borg, R.; Purkiss, A.; Cacciottolo, R.; Herrera, P.; Cauchi, R.J. Loss of amyotrophic lateral sclerosis risk factor SCFD1 causes motor dysfunction in Drosophila. Neurobiol. Aging 2023 , 126 , 67–76. [ Google Scholar ] [ CrossRef ]
  • Liampas, I.; Siokas, V.; Aloizou, A.-M.; Bakirtzis, C.; Tsouris, Z.; Nousia, A.; Nasios, G.; Papadimitriou, D.; Liakos, P.; Bogdanos, D.P.; et al. MOBP rs616147 Polymorphism and Risk of Amyotrophic Lateral Sclerosis in a Greek Population: A Case-Control Study. Medicina 2021 , 57 , 1337. [ Google Scholar ] [ CrossRef ]
  • Zhang, Z.; Sun, S. Linking ALS mutations to the disruption of the MHC-II pathway. Proc. Natl. Acad. Sci. USA 2023 , 120 , e2315240120. [ Google Scholar ] [ CrossRef ]
  • Soustelle, L.; Aimond, F.; López-Andrés, C.; Brugioti, V.; Raoul, C.; Layalle, S. ALS-Associated KIF5A Mutation Causes Locomotor Deficits Associated with Cytoplasmic Inclusions, Alterations of Neuromuscular Junctions, and Motor Neuron Loss. J. Neurosci. 2023 , 43 , 8058–8072. [ Google Scholar ] [ CrossRef ]
  • De Decker, M.; Zelina, P.; Moens, T.G.; Eggermont, K.; Moisse, M.; Veldink, J.H.; Van Den Bosch, L.; Pasterkamp, R.J.; Van Damme, P. C21orf2 mutations found in ALS disrupt primary cilia function. bioRxiv 2022 . [ Google Scholar ] [ CrossRef ]
  • Restuadi, R.; Steyn, F.J.; Kabashi, E.; Ngo, S.T.; Cheng, F.-F.; Nabais, M.F.; Thompson, M.J.; Qi, T.; Wu, Y.; Henders, A.K.; et al. Functional characterisation of the amyotrophic lateral sclerosis risk locus GPX3/TNIP1. Genome Med. 2022 , 14 , 7. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Li, C.Y.; Yang, T.M.; Ou, R.W.; Wei, Q.Q.; Shang, H.F. Genome-wide genetic links between amyotrophic lateral sclerosis and autoimmune diseases. BMC Med. 2021 , 19 , 27. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Oakes, J.A.; Davies, M.C.; Collins, M.O. TBK1: A new player in ALS linking autophagy and neuroinflammation. Mol. Brain 2017 , 10 , 5. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Karch, C.M.; Wen, N.; Fan, C.C.; Yokoyama, J.S.; Kouri, N.; Ross, O.A.; Höglinger, G.; Müller, U.; Ferrari, R.; Hardy, J.; et al. Selective Genetic Overlap Between Amyotrophic Lateral Sclerosis and Diseases of the Frontotemporal Dementia Spectrum. JAMA Neurol. 2018 , 75 , 860–875. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yao, L.; He, X.; Cui, B.; Zhao, F.; Zhou, C. NEK1 mutations and the risk of amyotrophic lateral sclerosis (ALS): A meta-analysis. Neurol. Sci. 2021 , 42 , 1277–1285. [ Google Scholar ] [ CrossRef ]
  • Laszlo, Z.I.; Hindley, N.; Avila, A.S.; Kline, R.A.; Eaton, S.L.; Lamont, D.J.; Smith, C.; Spires-Jones, T.L.; Wishart, T.M.; Henstridge, C.M. Synaptic proteomics reveal distinct molecular signatures of cognitive change and C9ORF72 repeat expansion in the human ALS cortex. Acta Neuropathol. Commun. 2022 , 10 , 156. [ Google Scholar ] [ CrossRef ]
  • Hendriks, W.J.A.J.; van Cruchten, R.T.P.; Pulido, R. Hereditable variants of classical protein tyrosine phosphatase genes: Will they prove innocent or guilty? Front. Cell Dev. Biol. 2023 , 10 , 1051311. [ Google Scholar ] [ CrossRef ]
  • Andrés-Benito, P.; Moreno, J.; Aso, E.; Povedano, M.; Ferrer, I. Amyotrophic lateral sclerosis, gene deregulation in the anterior horn of the spinal cord and frontal cortex area 8: Implications in frontotemporal lobar degeneration. Aging 2017 , 9 , 823–851. [ Google Scholar ] [ CrossRef ]
  • Alzheimer, A. Uber eigenartige Erkrankung der Hirnrinde. All Z Psychiatr. 1907 , 64 , 146–148. [ Google Scholar ]
  • Maurer, K.; Volk, S.; Gerbaldo, H. Auguste D and Alzheimer’s disease. Lancet 1997 , 349 , 1546–1549. [ Google Scholar ] [ CrossRef ]
  • Montine, T.J.; Phelps, C.H.; Beach, T.G.; Bigio, E.H.; Cairns, N.J.; Dickson, D.W.; Duyckaerts, C.; Frosch, M.P.; Masliah, E.; Mirra, S.S.; et al. National Institute on Aging–Alzheimer’s Association guidelines for the neuropathologic assessment of Alzheimer’s disease: A practical approach. Acta Neuropathol. 2012 , 123 , 1–11. [ Google Scholar ] [ CrossRef ]
  • Roth, M. The natural history of mental disorder in old age. J. Ment. Sci. 1955 , 101 , 281–301. [ Google Scholar ] [ CrossRef ]
  • Sirkis, D.W.; Bonham, L.W.; Johnson, T.P.; La Joie, R.; Yokoyama, J.S. Dissecting the clinical heterogeneity of early-onset Alzheimer’s disease. Mol. Psychiatry 2022 , 27 , 2674–2688. [ Google Scholar ] [ CrossRef ]
  • Reitz, C.; Rogaeva, E.; Beecham, G.W. Late-onset vs nonmendelian early-onset Alzheimer disease: A distinction without a difference? Neurol. Genet. 2020 , 6 . [ Google Scholar ] [ CrossRef ]
  • Dementia, E.O. A National Challenge, a Future Crisis ; Alzheimers Association: Washington, DC, USA, 2006. [ Google Scholar ]
  • Alzheimer’s Association. 2019 Alzheimer’s disease facts and figures. Alzheimer’s Dement. 2019 , 15 , 321–387. [ Google Scholar ] [ CrossRef ]
  • Goate, A.; Chartier-Harlin, M.C.; Mullan, M.; Brown, J.; Crawford, F.; Fidani, L.; Giuffra, L.; Haynes, A.; Irving, N.; James, L.; et al. Segregation of a missense mutation in the amyloid precursor protein gene with familial Alzheimer’s disease. Nature 1991 , 349 , 704–706. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Chartier-Harlin, M.C.; Crawford, F.; Houlden, H.; Warren, A.; Hughes, D.; Fidani, L.; Goate, A.; Rossor, M.; Roques, P.; Hardy, J.; et al. Early-onset Alzheimer’s disease caused by mutations at codon 717 of the β-amyloid precursor protein gene. Nature 1991 , 353 , 844–846. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Murrell, J.; Farlow, M.; Ghetti, B.; Benson, M.D. A mutation in the amyloid precursor protein associated with hereditary Alzheimer’s disease. Science 1991 , 254 , 97–99. [ Google Scholar ] [ CrossRef ]
  • Mullan, M.; Tsuji, S.; Miki, T.; Katsuya, T.; Naruse, S.; Kaneko, K.; Shimizu, T.; Kojima, T.; Nakano, I.; Ogihara, T.; et al. Clinical comparison of Alzheimer’s disease in pedigrees with the codon 717 Val→ Ile mutation in the amyloid precursor protein gene. Neurobiol. Aging 1993 , 14 , 407–419. [ Google Scholar ] [ CrossRef ]
  • Hardy, J. The discovery of Alzheimer-causing mutations in the APP gene and the formulation of the “amyloid cascade hypothesis. FEBS J. 2017 , 284 , 1040–1044. [ Google Scholar ] [ CrossRef ]
  • Sherrington, R.; Rogaev, E.I.; Liang, Y.A.; Rogaeva, E.A.; Levesque, G.; Ikeda, M.; Chi, H.; Lin, C.; Li, G.; Holman, K.; et al. Cloning of a gene bearing missense mutations in early-onset familial Alzheimer’s disease. Nature 1995 , 375 , 754–760. [ Google Scholar ] [ CrossRef ]
  • Campion, D.; Flaman, J.M.; Brice, A.; Hannequin, D.; Dubois, B.; Martin, C.; Moreau, V.; Charbonnier, F.; Didierjean, O.; Tardieu, S.; et al. Mutations of the presenilin I gene in families with early-onset Alzheimer’s disease. Hum. Mol. Genet. 1995 , 4 , 2373–2377. [ Google Scholar ] [ CrossRef ]
  • Cacace, R.; Sleegers, K.; Van Broeckhoven, C. Molecular genetics of early-onset Alzheimer’s disease revisited. Alzheimer’s Dement. 2016 , 12 , 733–748. [ Google Scholar ] [ CrossRef ]
  • Shim, K.H.; Kang, S.; An, S.S.A.; Kang, M.J. Identification of the Third Case of PSEN1 Tyr389His Variant in Early-Onset Alzheimer’s Disease in Korea. Int. J. Mol. Sci. 2022 , 23 , 16192. [ Google Scholar ] [ CrossRef ]
  • Rogaev, E. Familial Alzheimer’s disease in kindreds with missense mutations in a gene on chromosome 1 related to the Alzheimer’s disease type 3 gene. Nature 1995 , 376 , 775–778. [ Google Scholar ] [ CrossRef ]
  • Levy-Lahad, E.; Wasco, W.; Poorkaj, P.; Romano, D.M.; Oshima, J.; Pettingell, W.H.; Yu, C.E.; Jondro, P.D.; Schmidt, S.D.; Wang, K.; et al. Candidate gene for the chromosome 1 familial Alzheimer’s disease locus. Science 1995 , 269 , 973–977. [ Google Scholar ] [ CrossRef ]
  • Finckh, U.; Müller-Thomsen, T.; Mann, U.; Eggers, C.; Marksteiner, J.; Meins, W.; Binetti, G.; Alberici, A.; Hock, C.; Nitsch, R.M.; et al. High prevalence of pathogenic mutations in patients with early-onset dementia detected by sequence analyses of four different genes. Am. J. Hum. Genet. 2000 , 66 , 110–117. [ Google Scholar ] [ CrossRef ]
  • Van Cauwenberghe, C.; Van Broeckhoven, C.; Sleegers, K. The genetic landscape of Alzheimer disease: Clinical implications and perspectives. Genet. Med. 2016 , 18 , 421–430. [ Google Scholar ] [ CrossRef ]
  • Pericak-Vance, A.M.; Bebout, J.L.; Gaskell, P.C.; Yamaoka, L.H.; Hung, W.Y.; Alberts, M.J.; Walker, A.P.; Bartlett, R.J.; A Haynes, C.; A Welsh, K. Linkage studies in familial Alzheimer disease: Evidence for chromosome 19 linkage. Am. J. Hum. Genet. 1991 , 48 , 1034–1050. [ Google Scholar ]
  • Namba, Y.; Tomonaga, M.; Kawasaki, H.; Otomo, E.; Ikeda, K. Apolipoprotein E immunoreactivity in cerebral amyloid deposits and neurofibrillary tangles in Alzheimer’s disease and kuru plaque amyloid in Creutzfeldt-Jakob disease. Brain Res. 1991 , 541 , 163–166. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Strittmatter, W.J.; Saunders, A.M.; Schmechel, D.; Pericak-Vance, M.; Enghild, J.; Salvesen, G.S.; Roses, A.D. Apolipoprotein E: High-avidity binding to beta-amyloid and increased frequency of type 4 allele in late-onset familial Alzheimer disease. Proc. Natl. Acad. Sci. USA 1993 , 90 , 1977–1981. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Corder, E.H. Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer’s disease in late onset families. Science 1993 , 261 , 921–923. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Arboleda-Velasquez, J.F.; Lopera, F.; O’hare, M.; Delgado-Tirado, S.; Marino, C.; Chmielewska, N.; Saez-Torres, K.L.; Amarnani, D.; Schultz, A.P.; Sperling, R.A.; et al. Resistance to autosomal dominant Alzheimer’s disease in an APOE3 Christchurch homozygote: A case report. Nat. Med. 2019 , 25 , 1680–1683. [ Google Scholar ] [ CrossRef ]
  • Scherzer, C.R.; Offe, K.; Gearing, M.; Rees, H.D.; Fang, G.; Heilman, C.J.; Schaller, C.; Bujo, H.; Levey, A.I.; Lah, J.J. Loss of apolipoprotein E receptor LR11 in Alzheimer disease. Arch. Neurol. 2004 , 61 , 1200–1205. [ Google Scholar ] [ CrossRef ]
  • Rogaeva, E.; Meng, Y.; Lee, J.H.; Gu, Y.; Kawarai, T.; Zou, F.; Katayama, T.; Baldwin, C.T.; Cheng, R.; Hasegawa, H.; et al. The neuronal sortilin-related receptor SORL1 is genetically associated with Alzheimer disease. Nat. Genet. 2007 , 39 , 168–177. [ Google Scholar ] [ CrossRef ]
  • Lee, J.H.; Barral, S.; Reitz, C. The neuronal sortilin-related receptor gene SORL1 and late-onset Alzheimer’s disease. Curr. Neurol. Neurosci. Rep. 2008 , 8 , 384–391. [ Google Scholar ] [ CrossRef ]
  • Strang, K.H.; Golde, T.E.; Giasson, B.I. MAPT mutations, tauopathy, and mechanisms of neurodegeneration. Lab. Investig. 2019 , 99 , 912–928. [ Google Scholar ] [ CrossRef ]
  • Roks, G.; Dermaut, B.; Heutink, P.; Julliams, A.; Backhovens, H.; Van de Broeck, M.; Serneels, S.; Hofman, A.; Van Broeckhoven, C.; van Duijn, C.M. Mutation screening of the tau gene in patients with early-onset Alzheimer’s disease. Neurosci. Lett. 1999 , 277 , 137–139. [ Google Scholar ] [ CrossRef ]
  • Baker, M.; Litvan, I.; Houlden, H.; Adamson, J.; Dickson, D.; Perez-Tur, J.; Hardy, J.; Lynch, T.; Bigio, E.; Hutton, M. Association of an extended haplotype in the tau gene with progressive supranuclear palsy. Hum. Mol. Genet. 1999 , 8 , 711–715. [ Google Scholar ] [ CrossRef ]
  • Caffrey, T.M.; Wade-Martins, R. Functional MAPT haplotypes: Bridging the gap between genotype and neuropathology. Neurobiol. Dis. 2007 , 27 , 1–10. [ Google Scholar ] [ CrossRef ]
  • Kalinderi, K.; Fidani, L.; Bostantjopoulou, S. From 1997 to 2007: A decade journey through the H1 haplotype on 17q21 chromosome. Park. Relat. Disord. 2009 , 15 , 2–5. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Sun, W.; Jia, J. The +347 C promoter allele up-regulates MAPT expression and is associated with Alzheimer’s disease among the Chinese Han. Neurosci. Lett. 2009 , 450 , 340–343. [ Google Scholar ] [ CrossRef ]
  • Filipello, F.; Morini, R.; Corradini, I.; Zerbi, V.; Canzi, A.; Michalski, B.; Erreni, M.; Markicevic, M.; Starvaggi-Cucuzza, C.; Otero, K.; et al. The microglial innate immune receptor TREM2 is required for synapse elimination and normal brain connectivity. Immunity 2018 , 48 , 979–991.e8. [ Google Scholar ] [ CrossRef ]
  • Colonna, M. TREMs in the immune system and beyond. Nat. Rev. Immunol. 2003 , 3 , 445–453. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Chouery, E.; Delague, V.; Bergougnoux, A.; Koussa, S.; Serre, J.-L.; Mégarbané, A. Mutations in TREM2 lead to pure early-onset dementia without bone cysts. Hum. Mutat. 2008 , 29 , E194–E204. [ Google Scholar ] [ CrossRef ]
  • Guerreiro, R.; Wojtas, A.; Bras, J.; Carrasquillo, M.; Rogaeva, E.; Majounie, E.; Cruchaga, C.; Sassi, C.; Kauwe, J.S.; Younkin, S.; et al. TREM2 variants in Alzheimer’s disease. N. Engl. J. Med. 2013 , 368 , 117–127. [ Google Scholar ] [ CrossRef ]
  • Carmona, S.; Zahs, K.; Wu, E.; Dakin, K.; Bras, J.; Guerreiro, R. The role of TREM2 in Alzheimer’s disease and other neurodegenerative disorders. Lancet Neurol. 2018 , 17 , 721–730. [ Google Scholar ] [ CrossRef ]
  • Colonna, M.; Wang, Y. TREM2 variants: New keys to decipher Alzheimer disease pathogenesis. Nat. Rev. Neurosci. 2016 , 17 , 201–207. [ Google Scholar ] [ CrossRef ]
  • Sakae, N.; Liu, C.-C.; Shinohara, M.; Frisch-Daiello, J.; Ma, L.; Yamazaki, Y.; Tachibana, M.; Younkin, L.; Kurti, A.; Carrasquillo, M.M.; et al. ABCA7 Deficiency Accelerates Amyloid-β Generation and Alzheimer’s Neuronal Pathology. J. Neurosci. 2016 , 36 , 3848–3859. [ Google Scholar ] [ CrossRef ]
  • Aikawa, T.; Ren, Y.; Yamazaki, Y.; Tachibana, M.; Johnson, M.R.; Anderson, C.T.; Martens, Y.A.; Holm, M.-L.; Asmann, Y.W.; Saito, T.; et al. ABCA7 haplodeficiency disturbs microglial immune responses in the mouse brain. Proc. Natl. Acad. Sci. USA 2019 , 116 , 23790–23796. [ Google Scholar ] [ CrossRef ]
  • Lyssenko, N.N.; Praticò, D. ABCA7 and the altered lipidostasis hypothesis of Alzheimer’s disease. Alzheimer’s Dement. 2021 , 17 , 164–174. [ Google Scholar ] [ CrossRef ]
  • Stepler, K.E.; Gillyard, T.R.; Reed, C.B.; Avery, T.M.; Davis, J.S.; Robinson, R.A. ABCA7, a Genetic Risk Factor Associated with Alzheimer’s Disease Risk in African Americans. J. Alzheimer’s Dis. 2022 , 86 , 5–19. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Sinha, N.; Reagh, Z.M.; Tustison, N.J.; Berg, C.N.; Shaw, A.; Myers, C.E.; Hill, D.; Yassa, M.A.; Gluck, M.A. ABCA7 risk variant in healthy older African Americans is associated with a functionally isolated entorhinal cortex mediating deficient generalization of prior discrimination training. Hippocampus 2019 , 29 , 527–538. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Bertram, L.; McQueen, M.B.; Mullin, K.; Blacker, D.; E Tanzi, R. Systematic meta-analyses of Alzheimer disease genetic association studies: The AlzGene database. Nat. Genet. 2007 , 39 , 17–23. [ Google Scholar ] [ CrossRef ]
  • Bellenguez, C.; Küçükali, F.; Jansen, I.E.; Kleineidam, L.; Moreno-Grau, S.; Amin, N.; Naj, A.C.; Campos-Martin, R.; Grenier-Boley, B.; Andrade, V.; et al. New insights into the genetic etiology of Alzheimer’s disease and related dementias. Nat. Genet. 2022 , 54 , 412–436. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Gustafsen, C.; Glerup, S.; Pallesen, L.T.; Olsen, D.; Andersen, O.M.; Nykjær, A.; Madsen, P.; Petersen, C.M. Sortilin and SorLA Display Distinct Roles in Processing and Trafficking of Amyloid Precursor Protein. J. Neurosci. 2013 , 33 , 64–71. [ Google Scholar ] [ CrossRef ]
  • Brouwers, N.; Van Cauwenberghe, C.; Engelborghs, S.; Lambert, J.-C.; Bettens, K.; Le Bastard, N.; Pasquier, F.; Gil Montoya, A.; Peeters, K.; Mattheijssens, M.; et al. Alzheimer risk associated with a copy number variation in the complement receptor 1 increasing C3b/C4b binding sites. Mol. Psychiatry 2011 , 17 , 223. [ Google Scholar ] [ CrossRef ]
  • Hartl, D.; Aesg; May, P.; Gu, W.; Mayhaus, M.; Pichler, S.; Spaniol, C.; Glaab, E.; Bobbili, D.R.; Antony, P.; et al. A rare loss-of-function variant of ADAM17 is associated with late-onset familial Alzheimer disease. Mol. Psychiatry 2020 , 25 , 629–639. [ Google Scholar ] [ CrossRef ]
  • McQuade, A.; Blurton-Jones, M. Microglia in Alzheimer’s Disease: Exploring How Genetics and Phenotype Influence Risk. J. Mol. Biol. 2019 , 431 , 1805–1817. [ Google Scholar ] [ CrossRef ]
  • Chang, C.-J.; Chang, M.-Y.; Lee, Y.-C.; Chen, K.-Y.; Hsu, T.-I.; Wu, Y.-H.; Chuang, J.-Y.; Kao, T.-J. Nck2 is essential for limb trajectory selection by spinal motor axons. Dev. Dyn. 2018 , 247 , 1043–1056. [ Google Scholar ] [ CrossRef ]
  • Tesi, N.; van der Lee, S.; Hulsman, M.; van Schoor, N.M.; Huisman, M.; Pijnenburg, Y.; van der Flier, W.M.; Reinders, M.; Holstege, H. Cognitively Healthy Centenarians are genetically protected against Alzheimer’s disease specifically in immune and endo-lysosomal systems. medRxiv 2023 . [ Google Scholar ] [ CrossRef ]
  • Lambert, E.; Saha, O.; Landeira, B.S.; de Farias, A.R.M.; Hermant, X.; Carrier, A.; Pelletier, A.; Gadaut, J.; Davoine, L.; Dupont, C.; et al. The Alzheimer susceptibility gene BIN1 induces isoform-dependent neurotoxicity through early endosome defects. Acta Neuropathol. Commun. 2022 , 10 , 4. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Chou, V.; Pearse, R.V.; Aylward, A.J.; Ashour, N.; Taga, M.; Terzioglu, G.; Fujita, M.; Fancher, S.B.; Sigalov, A.; Benoit, C.R.; et al. INPP5D regulates inflammasome activation in human microglia. Nat. Commun. 2023 , 14 , 7552. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Toca, A.B.; Arroyo, A.M.; Sainz, A.R.; Martínez, J.A.; Zárraga, I.S.; Calle, I.A.; Portillo, J.B.; Picon, S.V.; Catalli, C.; Gómez-Beldarrain, M. Mutations in MME gene causing distal hereditary motor neuropathy and MCI-AD. J. Neurol. Sci. 2021 , 429 , 118990. [ Google Scholar ] [ CrossRef ]
  • Wang, D.; El-Amouri, S.S.; Dai, M.; Kuan, C.-Y.; Hui, D.Y.; Brady, R.O.; Pan, D. Engineering a lysosomal enzyme with a derivative of receptor-binding domain of apoE enables delivery across the blood–brain barrier. Proc. Natl. Acad. Sci. USA 2013 , 110 , 2999–3004. [ Google Scholar ] [ CrossRef ]
  • Lefort, R. Reversing Synapse Loss in Alzheimer’s Disease: Rho-Guanosine Triphosphatases and Insights from Other Brain Disorders. Neurotherapeutics 2015 , 12 , 19–28. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Jorfi, M.; Maaser-Hecker, A.; Tanzi, R.E. The neuroimmune axis of Alzheimer’s disease. Genome Med. 2023 , 15 , 6. [ Google Scholar ] [ CrossRef ]
  • Bano, D.; Ehninger, D.; Bagetta, G. Decoding metabolic signatures in Alzheimer’s disease: A mitochondrial perspective. Cell Death Discov. 2023 , 9 , 432. [ Google Scholar ] [ CrossRef ]
  • Yan, H.; Liu, M.; Gao, Y.; Yuan, Y.; Liu, X.; Wang, Y.; Li, L.; Wang, Q.; Wang, Y.; Shi, C.; et al. Assessing the impact of novel risk loci on Alzheimer’s and Parkinson’s diseases in a Chinese Han cohort. Front. Neurol. 2024 , 15 , 1326692. [ Google Scholar ] [ CrossRef ]
  • Nott, A.; Holtman, I.R. Genetic insights into immune mechanisms of Alzheimer’s and Parkinson’s disease. Front. Immunol. 2023 , 14 , 1168539. [ Google Scholar ] [ CrossRef ]
  • Ferreira, A.; Royaux, I.; Liu, J.; Wang, Z.; Su, G.; Moechars, D.; Callewaert, N.; De Muynck, L. The 3-O sulfation of heparan sulfate proteoglycans contributes to the cellular internalization of tau aggregates. BMC Cell Biol. 2022 , 23 , 61. [ Google Scholar ] [ CrossRef ]
  • Zhang, X.; Zou, M.; Wu, Y.; Jiang, D.; Wu, T.; Zhao, Y.; Wu, D.; Cui, J.; Li, G. Regulation of the Late Onset alzheimer’s Disease Associated HLA-DQA1/DRB1 Expression. Am. J. Alzheimer’s Dis. Other Dementiasr 2022 , 37 , 15333175221085066. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Li, Y.; Xu, H.; Wang, H.; Yang, K.; Luan, J.; Wang, S. TREM2: Potential therapeutic targeting of microglia for Alzheimer’s disease. Biomed. Pharmacother. 2023 , 165 , 115218. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Song, Y.-N.; Li, J.-Q.; Tan, C.-C.; Wang, H.-F.; Tan, M.-S.; Cao, X.-P.; Yu, J.-T.; Tan, L.; Initiative, T.A.D.N. TREML2 Mutation Mediate Alzheimer’s Disease Risk by Altering Neuronal Degeneration. Front. Neurosci. 2019 , 13 , 455. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Tao, Q.-Q.; Chen, Y.-C.; Wu, Z.-Y. The role of CD2AP in the Pathogenesis of Alzheimer’s Disease. Aging Dis. 2019 , 10 , 901–907. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Stefani, F.; Zhang, L.; Taylor, S.; Donovan, J.; Rollinson, S.; Doyotte, A.; Brownhill, K.; Bennion, J.; Pickering-Brown, S.; Woodman, P. UBAP1 Is a Component of an Endosome-Specific ESCRT-I Complex that Is Essential for MVB Sorting. Curr. Biol. 2011 , 21 , 1245–1250. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Talwar, P.; Silla, Y.; Grover, S.; Gupta, M.; Agarwal, R.; Kushwaha, S.; Kukreti, R. Genomic convergence and network analysis approach to identify candidate genes in Alzheimer’s disease. BMC Genom. 2014 , 15 , 199. [ Google Scholar ] [ CrossRef ]
  • Lee, J.Y.; Harney, D.J.; Teo, J.D.; Kwok, J.B.; Sutherland, G.T.; Larance, M.; Don, A.S. The major TMEM106B dementia risk allele affects TMEM106B protein levels, fibril formation, and myelin lipid homeostasis in the ageing human hippocampus. Mol. Neurodegener. 2023 , 18 , 63. [ Google Scholar ] [ CrossRef ]
  • Andrews, S.J.; Renton, A.E.; Fulton-Howard, B.; Podlesny-Drabiniok, A.; Marcora, E.; Goate, A.M. The complex genetic architecture of Alzheimer’s disease: Novel insights and future directions. EBioMedicine 2023 , 90 , 104511. [ Google Scholar ] [ CrossRef ]
  • Lang, S.; Pfeffer, S.; Lee, P.-H.; Cavalié, A.; Helms, V.; Förster, F.; Zimmermann, R. An Update on Sec61 Channel Functions, Mechanisms, and Related Diseases. Front. Physiol. 2017 , 8 , 887. [ Google Scholar ] [ CrossRef ]
  • Gastwirt, R.F.; McAndrew, C.W.; Donoghue, D.J. Speedy/RINGO Regulation of CDKs in Cell Cycle, Checkpoint Activation and Apoptosis. Cell Cycle 2007 , 6 , 1188–1193. [ Google Scholar ] [ CrossRef ]
  • Kim, Y.B.; Komor, A.C.; Levy, J.M.; Packer, M.S.; Zhao, K.T.; Liu, D.R. Increasing the genome-targeting scope and precision of base editing with engineered Cas9-cytidine deaminase fusions. Nat. Biotechnol. 2017 , 35 , 371–376. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Hook, V.; Yoon, M.; Mosier, C.; Ito, G.; Podvin, S.; Head, B.P.; Rissman, R.; O’Donoghue, A.J.; Hook, G. Cathepsin B in neurodegeneration of Alzheimer’s disease, traumatic brain injury, and related brain disorders. Biochim. Biophys. Acta (BBA)-Proteins Proteom. 2020 , 1868 , 140428. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Asanomi, Y.; Shigemizu, D.; Akiyama, S.; Miyashita, A.; Mitsumori, R.; Hara, N.; Ikeuchi, T.; Niida, S.; Ozaki, K. A functional variant of SHARPIN confers increased risk of late-onset Alzheimer’s disease. J. Hum. Genet. 2022 , 67 , 203–208. [ Google Scholar ] [ CrossRef ]
  • De Farias, A.-R.M.; Saha, O.; Lambert, J.-C.; Costa, M.R. Role of the late-onset Alzheimer’s disease risk genes bin1 and ptk2b in the hyperexcitability of hiPSC-derived neurons. Alzheimer’s Dement. 2021 , 17 , e053632. [ Google Scholar ] [ CrossRef ]
  • Nelson, A.R.; Sagare, A.P.; Zlokovic, B.V. Role of clusterin in the brain vascular clearance of amyloid-β. Proc. Natl. Acad. Sci. USA 2017 , 114 , 8681–8682. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lewandowski, C.T.; Laham, M.S.; Thatcher, G.R. Remembering your A, B, C’s: Alzheimer’s disease and ABCA1. Acta Pharm. Sin. B 2022 , 12 , 995–1018. [ Google Scholar ] [ CrossRef ]
  • Becic, A.; Leifeld, J.; Shaukat, J.; Hollmann, M. Tetraspanins as Potential Modulators of Glutamatergic Synaptic Function. Front. Mol. Neurosci. 2022 , 14 , 801882. [ Google Scholar ] [ CrossRef ]
  • Xiao, B.; Li, J.; Zhou, M.; Li, X.; Huang, X.; Hang, J.; Sun, Z.; Li, L. Structure and function of B-cell linker and its role in the development of B cell-related diseases. Nan Fang Yi Ke Da Xue Xue Bao= J. South. Med. Univ. 2019 , 39 , 253–256. [ Google Scholar ] [ CrossRef ]
  • Matveeva, N.; Kiselev, I.; Baulina, N.; Semina, E.; Kakotkin, V.; Agapov, M.; Kulakova, O.; Favorova, O. Shared genetic architecture of COVID-19 and Alzheimer’s disease. Front. Aging Neurosci. 2023 , 15 , 1287322. [ Google Scholar ] [ CrossRef ]
  • Raj, T.; Li, Y.I.; Wong, G.; Humphrey, J.; Wang, M.; Ramdhani, S.; Wang, Y.-C.; Ng, B.; Gupta, I.; Haroutunian, V.; et al. Integrative transcriptome analyses of the aging brain implicate altered splicing in Alzheimer’s disease susceptibility. Nat. Genet. 2018 , 50 , 1584–1592. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Rustenhoven, J.; Smith, A.M.; Smyth, L.C.; Jansson, D.; Scotter, E.L.; Swanson, M.E.V.; Aalderink, M.; Coppieters, N.; Narayan, P.; Handley, R.; et al. PU.1 regulates Alzheimer’s disease-associated genes in primary human microglia. Mol. Neurodegener. 2018 , 13 , 44. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Wang, Y.-Y.; Deng, Y.-S.; Dai, S.-K.; Mi, T.-W.; Li, R.-Y.; Liu, P.-P.; Liu, C.; He, B.-D.; He, X.-C.; Du, H.-Z.; et al. Loss of microglial EED impairs synapse density, learning, and memory. Mol. Psychiatry 2022 , 27 , 2999–3009. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Mishra, S.; Knupp, A.; Szabo, M.P.; Williams, C.A.; Kinoshita, C.; Hailey, D.W.; Wang, Y.; Andersen, O.M.; Young, J.E. The Alzheimer’s gene SORL1 is a regulator of endosomal traffic and recycling in human neurons. Cell. Mol. Life Sci. 2022 , 79 , 162. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Foster, W.J.; Taylor, H.B.C.; Padamsey, Z.; Jeans, A.F.; Galione, A.; Emptage, N.J. Hippocampal mGluR1-dependent long-term potentiation requires NAADP-mediated acidic store Ca 2+ signaling. Sci. Signal. 2018 , 11 , eaat9093. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Dalmasso, M.C.; de Rojas, I.; Olivar, N.; Muchnik, C.; Angel, B.; Gloger, S.; Abalos, M.S.S.; Chacón, M.V.; Aránguiz, R.; Orellana, P.; et al. The first genome-wide association study in the Argentinian and Chilean populations identifies shared genetics with Europeans in Alzheimer’s disease. Alzheimer’s Dement. Early View . [ CrossRef ]
  • Eysert, F.; Coulon, A.; Boscher, E.; Vreulx, A.-C.; Flaig, A.; Mendes, T.; Hughes, S.; Grenier-Boley, B.; Hanoulle, X.; Demiautte, F.; et al. Alzheimer’s genetic risk factor FERMT2 (Kindlin-2) controls axonal growth and synaptic plasticity in an APP-dependent manner. Mol. Psychiatry 2021 , 26 , 5592–5607. [ Google Scholar ] [ CrossRef ]
  • Tan, M.-S.; Initiative, A.D.N.; Yang, Y.-X.; Xu, W.; Wang, H.-F.; Tan, L.; Zuo, C.-T.; Dong, Q.; Tan, L.; Suckling, J.; et al. Associations of Alzheimer’s disease risk variants with gene expression, amyloidosis, tauopathy, and neurodegeneration. Alzheimer’s Res. Ther. 2021 , 13 , 15. [ Google Scholar ] [ CrossRef ]
  • Park, Y.H.; Pyun, J.-M.; Hodges, A.; Jang, J.-W.; Bice, P.J.; Kim, S.; Saykin, A.J.; Kwangsik Nho for the AddNeuroMed consortium and the Alzheimer’s Disease Neuroimaging Initiative. Dysregulated expression levels of APH1B in peripheral blood are associated with brain atrophy and amyloid-β deposition in Alzheimer’s disease. Alzheimer’s Res. Ther. 2021 , 13 , 183. [ Google Scholar ] [ CrossRef ]
  • Kim, N.-Y.; Cho, M.-H.; Won, S.-H.; Kang, H.-J.; Yoon, S.-Y.; Kim, D.-H. Sorting nexin-4 regulates β-amyloid production by modulating β-site-activating cleavage enzyme-1. Alzheimer’s Res. Ther. 2017 , 9 , 4. [ Google Scholar ] [ CrossRef ]
  • Li, Y.; Xu, M.; Xiang, B.-L.; Li, X.; Zhang, D.-F.; Zhao, H.; Bi, R.; Yao, Y.-G. Functional genomics identify causal variant underlying the protective CTSH locus for Alzheimer’s disease. Neuropsychopharmacology 2023 , 48 , 1555–1566. [ Google Scholar ] [ CrossRef ]
  • Yoo, H.-S.; Shanmugalingam, U.; Smith, P.D. Potential roles of branched-chain amino acids in neurodegeneration. Nutrition 2022 , 103-104 , 111762. [ Google Scholar ] [ CrossRef ]
  • Ho, A.; Ngala, B.; Yamada, C.; Garcia, C.; Duarte, C.; Akkaoui, J.; Ciolac, D.; Nusbaum, A.; Kochen, W.; Efremova, D.; et al. IL-34 exacerbates pathogenic features of Alzheimer’s disease and calvaria osteolysis in triple transgenic (3x-Tg) female mice. Biomed. Pharmacother. 2023 , 166 , 115435. [ Google Scholar ] [ CrossRef ]
  • Li, K.; Ran, B.; Wang, Y.; Liu, L.; Li, W. PLCγ2 impacts microglia-related effectors revealing variants and pathways important in Alzheimer’s disease. Front. Cell Dev. Biol. 2022 , 10 , 999061. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Wang, Q.-W.; Qin, J.; Chen, Y.-F.; Tu, Y.; Xing, Y.-Y.; Wang, Y.; Yang, L.-Y.; Lu, S.-Y.; Geng, L.; Shi, W.; et al. 16p11.2 CNV gene Doc2α functions in neurodevelopment and social behaviors through interaction with Secretagogin. Cell Rep. 2023 , 42 , 112691. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Dugan, A.J.; Nelson, P.T.; Katsumata, Y.; Shade, L.M.; Teylan, M.A.; Boehme, K.L.; Mukherjee, S.; Kauwe, J.S.; Hohman, T.J.; Schneider, J.A.; et al. Association between WWOX/MAF variants and dementia-related neuropathologic endophenotypes. Neurobiol. Aging 2022 , 111 , 95–106. [ Google Scholar ] [ CrossRef ]
  • Oli, V.; Gupta, R.; Kumar, P. FOXO and related transcription factors binding elements in the regulation of neurodegenerative disorders. J. Chem. Neuroanat. 2021 , 116 , 102012. [ Google Scholar ] [ CrossRef ]
  • Rasmi, Y.; Shokati, A.; Hassan, A.; Aziz, S.G.-G.; Bastani, S.; Jalali, L.; Moradi, F.; Alipour, S. The role of DNA methylation in progression of neurological disorders and neurodegenerative diseases as well as the prospect of using DNA methylation inhibitors as therapeutic agents for such disorders. IBRO Neurosci. Rep. 2023 , 14 , 28–37. [ Google Scholar ] [ CrossRef ]
  • Liu, X.; Yin, L.; Li, T.; Lin, L.; Zhang, J.; Li, Y. Reduction of WDR81 impairs autophagic clearance of aggregated proteins and cell viability in neurodegenerative phenotypes. PLoS Genet. 2021 , 17 , e1009415. [ Google Scholar ] [ CrossRef ]
  • Xicota, L.; Lagarde, J.; Eysert, F.; Grenier-Boley, B.; Rivals, I.; Botté, A.; Forlani, S.; Landron, S.; Gautier, C.; Gabriel, C.; et al. Modifications of the endosomal compartment in fibroblasts from sporadic Alzheimer’s disease patients are associated with cognitive impairment. Transl. Psychiatry 2023 , 13 , 54. [ Google Scholar ] [ CrossRef ]
  • Lona-Durazo, F.; Reynolds, R.H.; Scholz, S.W.; Ryten, M.; Taliun, S.A.G. Regional genetic correlations highlight relationships between neurodegenerative disease loci and the immune system. Commun. Biol. 2023 , 6 , 729. [ Google Scholar ] [ CrossRef ]
  • Aghaizu, N.D.; Jin, H.; Whiting, P.J. Dysregulated Wnt Signalling in the Alzheimer’s Brain. Brain Sci. 2020 , 10 , 902. [ Google Scholar ] [ CrossRef ]
  • Cao, M.; Liu, J.; Zhang, X.; Yang, T.; Wang, Y.; Hou, Y.; Song, Q.; Cui, Y.; Wang, Y.; Wang, P. ABI3 Is a Novel Early Biomarker of Alzheimer’s Disease. J. Alzheimer’s Dis. 2022 , 87 , 335–344. [ Google Scholar ] [ CrossRef ]
  • Wainberg, M.; Andrews, S.J.; Tripathy, S.J. Shared genetic risk loci between Alzheimer’s disease and related dementias, Parkinson’s disease, and amyotrophic lateral sclerosis. Alzheimer’s Res. Ther. 2023 , 15 , 113. [ Google Scholar ] [ CrossRef ]
  • Miners, S.; Ashby, E.; Baig, S.; Harrison, R.; Tayler, H.; Speedy, E.; Prince, J.A.; Love, S.; Kehoe, P.G. Angiotensin-converting enzyme levels and activity in Alzheimer’s disease: Differences in brain and CSF ACE and association with ACE1 genotypes. Am. J. Transl. Res. 2009 , 1 , 163–177. [ Google Scholar ]
  • Breen, C.; Papale, L.A.; Clark, L.R.; Bergmann, P.E.; Madrid, A.; Asthana, S.; Johnson, S.C.; Keleş, S.; Alisch, R.S.; Hogan, K.J. Whole genome methylation sequencing in blood identifies extensive differential DNA methylation in late-onset dementia due to Alzheimer’s disease. Alzheimer’s Dement. Early View . [ CrossRef ]
  • Siew, J.J.; Chern, Y.; Khoo, K.-H.; Angata, T. Roles of Siglecs in neurodegenerative diseases. Mol. Asp. Med. 2023 , 90 , 101141. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Kim, T.; Vidal, G.S.; Djurisic, M.; William, C.M.; Birnbaum, M.E.; Garcia, K.C.; Hyman, B.T.; Shatz, C.J. Human LilrB2 Is a β-Amyloid Receptor and Its Murine Homolog PirB Regulates Synaptic Plasticity in an Alzheimer’s Model. Science 2013 , 341 , 1399–1404. [ Google Scholar ] [ CrossRef ]
  • Zhao, Y.; Yun, D.; Zou, X.; Jiang, T.; Li, G.; Hu, L.; Chen, J.; Xu, J.; Mao, Y.; Chen, H.; et al. Whole exome-wide association study identifies a missense variant in SLC2A4RG associated with glioblastoma risk. Am. J. Cancer Res. 2017 , 7 , 1937–1947. [ Google Scholar ]
  • Giri, M.; Zhang, M.; Lü, Y. Genes associated with Alzheimer’s disease: An overview and current status. Clin. Interv. Aging 2016 , 11 , 665–681. [ Google Scholar ] [ CrossRef ]
  • Hampel, H.; Hardy, J.; Blennow, K.; Chen, C.; Perry, G.; Kim, S.H.; Villemagne, V.L.; Aisen, P.; Vendruscolo, M.; Iwatsubo, T.; et al. The Amyloid-beta Pathway in Alzheimer’s Disease. Mol. Psychiatry 2021 , 26 , 5481–5503. [ Google Scholar ] [ CrossRef ]
  • Qiu, Y.; Sha, L.; Zhang, X.; Li, G.; Zhu, W.; Xu, Q. Induction of A Disintegrin and Metalloproteinase with Thrombospondin motifs 1 by a rare variant or cognitive activities reduces hippocampal amyloid-β and consequent Alzheimer’s disease risk. Front. Aging Neurosci. 2022 , 14 , 896522. [ Google Scholar ] [ CrossRef ]
  • Jun, G.; Naj, A.C.; Beecham, G.W.; Wang, L.S.; Buros, J.; Gallins, P.J.; Buxbaum, J.D.; Ertekin-Taner, N.; Fallin, M.D.; Friedland, R.; et al. Meta-analysis confirms CR1, CLU, and PICALM as Alzheimer disease risk loci and reveals interactions with APOE genotypes. Arch. Neurol. 2010 , 67 , 1473–1484. [ Google Scholar ] [ CrossRef ]
  • Kunkle, B.W.; Grenier-Boley, B.; Sims, R.; Bis, J.C.; Damotte, V.; Naj, A.C.; Boland, A.; Vronskaya, M.; Van Der Lee, S.J.; Amlie-Wolf, A.; et al. Genetic meta-analysis of diagnosed Alzheimer’s disease identifies new risk loci and implicates Aβ, tau, immunity and lipid processing. Nat. Genet. 2019 , 51 , 414–430. [ Google Scholar ] [ CrossRef ]
  • Marioni, R.E.; Harris, S.E.; Zhang, Q.; McRae, A.F.; Hagenaars, S.P.; Hill, W.D.; Davies, G.; Ritchie, C.W.; Gale, C.R.; Starr, J.M.; et al. GWAS on family history of Alzheimer’s disease. Transl. Psychiatry 2018 , 8 , 99. [ Google Scholar ] [ CrossRef ]
  • Parkinson, J. An essay on the shaking palsy. J. Neuropsychiatry Clin. Neurosci. 2002 , 14 , 223–236. [ Google Scholar ] [ CrossRef ]
  • Poewe, W.; Seppi, K.; Tanner, C.M.; Halliday, G.M.; Brundin, P.; Volkmann, J.; Schrag, A.E.; Lang, A.E. Parkinson disease. Nat. Rev. Dis. Primers 2017 , 3 , 17013. [ Google Scholar ] [ CrossRef ]
  • Marras, C.; Beck, J.C.; Bower, J.H.; Roberts, E.; Ritz, B.; Ross, G.W.; Abbott, R.D.; Savica, R.; Van Den Eeden, S.K.; Willis, A.W.; et al. Prevalence of Parkinson’s disease across North America. NPJ Parkinson’s Dis. 2018 , 4 , 21. [ Google Scholar ] [ CrossRef ]
  • Aguirre-Mardones, C.; Iranzo, A.; Vilas, D.; Serradell, M.; Gaig, C.; Santamaría, J.; Tolosa, E. Prevalence and timeline of nonmotor symptoms in idiopathic rapid eye movement sleep behavior disorder. J. Neurol. 2015 , 262 , 1568–1578. [ Google Scholar ] [ CrossRef ]
  • Twelves, D.; Perkins, K.S.; Counsell, C. Systematic review of incidence studies of Parkinson’s disease. Mov. Disord. 2003 , 18 , 19–31. [ Google Scholar ] [ CrossRef ]
  • Dorsey, E.R.; Elbaz, A.; Nichols, E.; Abbasi, N.; Abd-Allah, F.; Abdelalim, A.; Adsuar, J.C.; Ansha, M.G.; Brayne, C.; Choi, J.Y.J.; et al. Global, regional, and national burden of Parkinson’s disease, 1990–2016: A systematic analysis for the Global Burden of Disease Study 2016. Lancet Neurol. 2018 , 17 , 939–953. [ Google Scholar ] [ CrossRef ]
  • Simon, D.K.; Tanner, C.M.; Brundin, P. Parkinson disease epidemiology, pathology, genetics, and pathophysiology. Clin. Geriatr. Med. 2020 , 36 , 1–12. [ Google Scholar ] [ CrossRef ]
  • Burré, J.; Sharma, M.; Südhof, T.C. Cell biology and pathophysiology of α-synuclein. Cold Spring Harb. Perspect. Med. 2018 , 8 , a024091. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Beyer, K.; Munoz-Marmol, A.M.; Sanz, C.; Marginet-Flinch, R.; Ferrer, I.; Ariza, A. New brain-specific beta-synuclein isoforms show expression ratio changes in Lewy body diseases. Neurogenetics 2012 , 13 , 61–72. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Polymeropoulos, M.H.; Lavedan, C.; Leroy, E.; Ide, S.E.; Dehejia, A.; Dutra, A.; Pike, B.; Root, H.; Rubenstein, J.; Boyer, R. Mutation in the α-synuclein gene identified in families with Parkinson’s disease. Science 1997 , 276 , 2045–2047. [ Google Scholar ] [ CrossRef ]
  • Spillantini, M.G.; Crowther, R.A.; Jakes, R.; Hasegawa, M.; Goedert, M. α-Synuclein in filamentous inclusions of Lewy bodies from Parkinson’s disease and dementia with Lewy bodies. Proc. Natl. Acad. Sci. USA 1998 , 95 , 6469–6473. [ Google Scholar ] [ CrossRef ]
  • Hernandez, D.G.; Reed, X.; Singleton, A.B. Genetics in Parkinson disease: Mendelian versus non-Mendelian inheritance. J. Neurochem. 2016 , 139 , 59–74. [ Google Scholar ] [ CrossRef ]
  • Vázquez-Vélez, G.E.; Zoghbi, H.Y. Parkinson’s disease genetics and pathophysiology. Annu. Rev. Neurosci. 2021 , 44 , 87–108. [ Google Scholar ] [ CrossRef ]
  • Singleton, A.B.; Farrer, M.; Johnson, J.; Singleton, A.; Hague, S.; Kachergus, J.; Hulihan, M.; Peuralinna, T.; Dutra, A.; Nussbaum, R.; et al. α-Synuclein locus triplication causes Parkinson’s disease. Science 2003 , 302 , 841. [ Google Scholar ] [ CrossRef ]
  • Konno, T.; Ross, O.A.; Puschmann, A.; Dickson, D.W.; Wszolek, Z.K. Autosomal dominant Parkinson’s disease caused by SNCA duplications. Park. Relat. Disord. 2016 , 22 , S1–S6. [ Google Scholar ] [ CrossRef ]
  • Trinh, J.; Guella, I.; Farrer, M.J. Disease penetrance of late-onset parkinsonism: A meta-analysis. JAMA Neurol. 2014 , 71 , 1535–1539. [ Google Scholar ] [ CrossRef ]
  • Soldner, F.; Stelzer, Y.; Shivalila, C.S.; Abraham, B.J.; Latourelle, J.C.; Barrasa, M.I.; Goldmann, J.; Myers, R.H.; Young, R.A.; Jaenisch, R. Parkinson-associated risk variant in distal enhancer of α-synuclein modulates target gene expression. Nature 2016 , 533 , 95–99. [ Google Scholar ] [ CrossRef ]
  • Funayama, M.; Hasegawa, K.; Kowa, H.; Saito, M.; Tsuji, S.; Obata, F. A new locus for Parkinson’s disease (PARK8) maps to chromosome 12p11. 2–q13. 1. Ann. Neurol. Off. J. Am. Neurol. Assoc. Child Neurol. Soc. 2002 , 51 , 296–301. [ Google Scholar ]
  • Araki, M.; Ito, G.; Tomita, T. Physiological and pathological functions of LRRK2: Implications from substrate proteins. Neuronal Signal. 2018 , 2 , NS20180005. [ Google Scholar ] [ CrossRef ]
  • Thaler, A.; Ash, E.; Gan-Or, Z.; Orr-Urtreger, A.; Giladi, N. The LRRK2 G2019S mutation as the cause of Parkinson’s disease in Ashkenazi Jews. J. Neural Transm. 2009 , 116 , 1473–1482. [ Google Scholar ] [ CrossRef ]
  • Healy, D.G.; Falchi, M.; O’Sullivan, S.S.; Bonifati, V.; Durr, A.; Bressman, S.; Brice, A.; Aasly, J.; Zabetian, C.P.; Goldwurm, S.; et al. Phenotype, genotype, and worldwide genetic penetrance of LRRK2-associated Parkinson’s disease: A case-control study. Lancet Neurol. 2008 , 7 , 583–590. [ Google Scholar ] [ CrossRef ]
  • Di Maio, R.; Hoffman, E.K.; Rocha, E.M.; Keeney, M.T.; Sanders, L.H.; De Miranda, B.R.; Zharikov, A.; Van Laar, A.; Stepan, A.F.; Lanz, T.A.; et al. LRRK2 activation in idiopathic Parkinson’s disease. Sci. Transl. Med. 2018 , 10 , eaar5429. [ Google Scholar ] [ CrossRef ]
  • Herzig, M.C.; Kolly, C.; Persohn, E.; Theil, D.; Schweizer, T.; Hafner, T.; Stemmelen, C.; Troxler, T.J.; Schmid, P.; Danner, S.; et al. LRRK2 protein levels are determined by kinase function and are crucial for kidney and lung homeostasis in mice. Hum. Mol. Genet. 2011 , 20 , 4209–4223. [ Google Scholar ] [ CrossRef ]
  • Vilas, D.; Gelpi, E.; Aldecoa, I.; Grau, O.; Rodriguez-Diehl, R.; Jaumà, S.; Martí, M.J.; Tolosa, E. Lack of central and peripheral nervous system synuclein pathology in R1441G LRRK2 -associated Parkinson’s disease. J. Neurol. Neurosurg. Psychiatry 2019 , 90 , 832–833. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Williams, E.T.; Chen, X.; Moore, D.J. VPS35, the retromer complex and Parkinson’s disease. J. Park. Dis. 2017 , 7 , 219–233. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Zimprich, A.; Benet-Pagès, A.; Struhal, W.; Graf, E.; Eck, S.H.; Offman, M.N.; Haubenberger, D.; Spielberger, S.; Schulte, E.C.; Lichtner, P.; et al. A mutation in VPS35, encoding a subunit of the retromer complex, causes late-onset Parkinson disease. Am. J. Hum. Genet. 2011 , 89 , 168–175. [ Google Scholar ] [ CrossRef ]
  • Sidransky, E.; Lopez, G. The link between the GBA gene and parkinsonism. Lancet Neurol. 2012 , 11 , 986–998. [ Google Scholar ] [ CrossRef ]
  • Sidransky, E.; Samaddar, T.; Tayebi, N. Mutations in GBA are associated with familial Parkinson disease susceptibility and age at onset. Neurology 2009 , 73 , 1424–1426. [ Google Scholar ] [ CrossRef ]
  • Duran, R.; Mencacci, N.E.; Angeli, A.V.; Shoai, M.; Deas, E.; Houlden, H.; Mehta, A.; Hughes, D.; Cox, T.M.; Deegan, P.; et al. The glucocerobrosidase E326K variant predisposes to Parkinson’s disease, but does not cause Gaucher’s disease. Mov. Disord. 2013 , 28 , 232–236. [ Google Scholar ] [ CrossRef ]
  • Mallett, V.; Ross, J.P.; Alcalay, R.N.; Ambalavanan, A.; Sidransky, E.; Dion, P.A.; Rouleau, G.A.; Gan-Or, Z. GBA p.T369M substitution in Parkinson disease: Polymorphism or association? A meta-analysis. Neurol. Genet. 2016 , 2 , e104. [ Google Scholar ] [ CrossRef ]
  • Riboldi, G.M.; Di Fonzo, A.B. GBA , Gaucher disease, and Parkinson’s disease: From genetic to clinic to new therapeutic approaches. Cells 2019 , 8 , 364. [ Google Scholar ] [ CrossRef ]
  • Simon-Sanchez, J.; Schulte, C.; Bras, J.M.; Sharma, M.; Gibbs, J.R.; Berg, D.; Paisan-Ruiz, C.; Lichtner, P.; Scholz, S.W.; Hernandez, D.G.; et al. Genome-wide association study reveals genetic risk underlying Parkinson’s disease. Nat. Genet. 2009 , 41 , 1308–1312. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Nalls, M.A.; Blauwendraat, C.; Vallerga, C.L.; Heilbron, K.; Bandres-Ciga, S.; Chang, D.; Tan, M.; Kia, D.A.; Noyce, A.J.; Xue, A. Identification of novel risk loci, causal insights, and heritable risk for Parkinson’s disease: A meta-analysis of genome-wide association studies. Lancet Neurol. 2019 , 18 , 1091–1102. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Buenaventura, R.G.; Harvey, A.C.; Burns, M.P.; Main, B.S. Traumatic brain injury induces an adaptive immune response in the meningeal transcriptome that is amplified by aging. Front. Neurosci. 2023 , 17 , 1210175. [ Google Scholar ] [ CrossRef ]
  • Iwaki, H.; Blauwendraat, C.; Leonard, H.L.; Liu, G.; Maple-Grødem, J.; Corvol, J.-C.; Pihlstrøm, L.; van Nimwegen, M.; Hutten, S.J.; Nguyen, K.-D.H.; et al. Genetic risk of Parkinson disease and progression: An analysis of 13 longitudinal cohorts. Neurol. Genet. 2019 , 5 , e348. [ Google Scholar ] [ CrossRef ]
  • Straniero, L.; Rimoldi, V.; Samarani, M.; Goldwurm, S.; Di Fonzo, A.; Krüger, R.; Deleidi, M.; Aureli, M.; Soldà, G.; Duga, S.; et al. The GBAP1 pseudogene acts as a ceRNA for the glucocerebrosidase gene GBA by sponging miR-22-3p. Sci. Rep. 2017 , 7 , 12702. [ Google Scholar ] [ CrossRef ]
  • Kaiser, S.; Zhang, L.; Mollenhauer, B.; Jacob, J.; Longerich, S.; Del-Aguila, J.; Marcus, J.; Raghavan, N.; Stone, D.; Fagboyegun, O.; et al. A proteogenomic view of Parkinson’s disease causality and heterogeneity. npj Park. Dis. 2023 , 9 , 24. [ Google Scholar ] [ CrossRef ]
  • Ivanova, D.; Dobson, K.L.; Gajbhiye, A.; Davenport, E.C.; Hacker, D.; Ultanir, S.K.; Trost, M.; Cousin, M.A. Control of synaptic vesicle release probability via VAMP4 targeting to endolysosomes. Sci. Adv. 2021 , 7 , eabf3873. [ Google Scholar ] [ CrossRef ]
  • Singh, S.; Park, S.-G.J.B. Functional association between NUCKS1 gene and Parkinson disease: A potential susceptibility biomarker. Bioinformation 2019 , 15 , 548–556. [ Google Scholar ] [ CrossRef ]
  • Purlyte, E.; Dhekne, H.S.; Sarhan, A.R.; Gomez, R.; Lis, P.; Wightman, M.; Martinez, T.N.; Tonelli, F.; Pfeffer, S.R.; Alessi, D.R. Rab29 activation of the Parkinson’s disease-associated LRRK2 kinase. EMBO J. 2017 , 37 , 1–18. [ Google Scholar ] [ CrossRef ]
  • Apicco, D.J.; Shlevkov, E.; Nezich, C.L.; Tran, D.T.; Guilmette, E.; Nicholatos, J.W.; Bantle, C.M.; Chen, Y.; Glajch, K.E.; Abraham, N.A.; et al. The Parkinson’s disease-associated gene ITPKB protects against α-synuclein aggregation by regulating ER-to-mitochondria calcium release. Proc. Natl. Acad. Sci. USA 2020 , 118 , e2006476118. [ Google Scholar ] [ CrossRef ]
  • Andres-Alonso, M.; Ammar, M.R.; Butnaru, I.; Gomes, G.M.; Sanhueza, G.A.; Raman, R.; Yuanxiang, P.; Borgmeyer, M.; Lopez-Rojas, J.; Raza, S.A.; et al. SIPA1L2 controls trafficking and local signaling of TrkB-containing amphisomes at presynaptic terminals. Nat. Commun. 2019 , 10 , 5448. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Pike, A.F.; Szabò, I.; Veerhuis, R.; Bubacco, L. The potential convergence of NLRP3 inflammasome, potassium, and dopamine mechanisms in Parkinson’s disease. Npj Park. Dis. 2022 , 8 , 32. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yu, E.; Larivière, R.; Thomas, R.A.; Liu, L.; Senkevich, K.; Rahayel, S.; Trempe, J.-F.; Fon, E.A.; Gan-Or, Z. Identification of novel variants, genes and pathways potentially linked to Parkinson’s disease using machine learning. medRxiv 2023 . [ Google Scholar ] [ CrossRef ]
  • Cappelletti, C.; Henriksen, S.P.; Geut, H.; Rozemuller, A.J.M.; van de Berg, W.D.J.; Pihlstrøm, L.; Toft, M. Transcriptomic profiling of Parkinson’s disease brains reveals disease stage specific gene expression changes. Acta Neuropathol. 2023 , 146 , 227–244. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Styrpejko, D.J.; Cuajungco, M.P. Transmembrane 163 (TMEM163) Protein: A New Member of the Zinc Efflux Transporter Family. Biomedicines 2021 , 9 , 220. [ Google Scholar ] [ CrossRef ]
  • Chang, K.-H.; Wu, Y.-R.; Chen, Y.-C.; Fung, H.-C.; Lee-Chen, G.-J.; Chen, C.-M. STK39, But Not BST1, HLA-DQB1, and SPPL2B Polymorphism, Is Associated With Han-Chinese Parkinson’s Disease in Taiwan. Medicine 2015 , 94 , e1690. [ Google Scholar ] [ CrossRef ]
  • Riessland, M.; Kolisnyk, B.; Kim, T.W.; Cheng, J.; Ni, J.; Pearson, J.A.; Park, E.J.; Dam, K.; Acehan, D.; Ramos-Espiritu, L.S.; et al. Loss of SATB1 Induces p21-Dependent Cellular Senescence in Post-mitotic Dopaminergic Neurons. Cell Stem Cell 2019 , 25 , 514–530.e8. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Parikshak, N.N.; Swarup, V.; Belgard, T.G.; Irimia, M.; Ramaswami, G.; Gandal, M.J.; Hartl, C.; Leppa, V.; Ubieta, L.d.l.T.; Huang, J.; et al. Genome-wide changes in lncRNA, splicing, and regional gene expression patterns in autism. Nature 2016 , 540 , 423–427. [ Google Scholar ] [ CrossRef ]
  • Nagpal, L.; Kornberg, M.D.; Snyder, S.H. Inositol hexakisphosphate kinase-2 non-catalytically regulates mitophagy by attenuating PINK1 signaling. Proc. Natl. Acad. Sci. USA 2022 , 119 , e2121946119. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Markopoulou, K.; Chase, B.A.; Premkumar, A.P.; Schoneburg, B.; Kartha, N.; Wei, J.; Yu, H.; Epshteyn, A.; Garduno, L.; Pham, A.; et al. Variable Effects of PD-Risk Associated SNPs and Variants in Parkinsonism-Associated Genes on Disease Phenotype in a Community-Based Cohort. Front. Neurol. 2021 , 12 , 662278. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Nizon, M.; Laugel, V.; Flanigan, K.M.; Pastore, M.; Waldrop, M.A.; Rosenfeld, J.A.; Marom, R.; Xiao, R.; Gerard, A.; Pichon, O.; et al. Variants in MED12L, encoding a subunit of the mediator kinase module, are responsible for intellectual disability associated with transcriptional defect. Anesthesia Analg. 2019 , 21 , 2713–2722. [ Google Scholar ] [ CrossRef ]
  • Custodia, A.; Aramburu-Núñez, M.; Correa-Paz, C.; Posado-Fernández, A.; Gómez-Larrauri, A.; Castillo, J.; Gómez-Muñoz, A.; Sobrino, T.; Ouro, A. Ceramide Metabolism and Parkinson’s Disease—Therapeutic Targets. Biomolecules 2021 , 11 , 945. [ Google Scholar ] [ CrossRef ]
  • Zhao, A.; Li, Y.; Niu, M.; Li, G.; Luo, N.; Zhou, L.; Kang, W.; Liu, J. SNPs in SNCA , MCCC1 , DLG2 , GBF1 and MBNL2 are associated with Parkinson’s disease in southern Chinese population. J. Cell. Mol. Med. 2020 , 24 , 8744–8752. [ Google Scholar ] [ CrossRef ]
  • Zhang, J.; Zeng, H.; Zhu, L.; Deng, L.; Fang, X.; Deng, X.; Liang, H.; Tang, C.; Cao, X.; Lu, Y.; et al. The Potential Mutation of GAK Gene in the Typical Sporadic Parkinson’s Disease from the Han Population of Chinese Mainland. Mol. Neurobiol. 2016 , 53 , 7119–7136. [ Google Scholar ] [ CrossRef ]
  • Hu, M.; Li, P.; Wang, C.; Feng, X.; Geng, Q.; Chen, W.; Marthi, M.; Zhang, W.; Gao, C.; Reid, W.; et al. Parkinson’s disease-risk protein TMEM175 is a proton-activated proton channel in lysosomes. Cell 2022 , 185 , 2292–2308.e20. [ Google Scholar ] [ CrossRef ]
  • Mufti, K.; Yu, E.; Rudakou, U.; Krohn, L.; Ruskey, J.A.; Asayesh, F.; Laurent, S.B.; Spiegelman, D.; Arnulf, I.; Hu, M.T.; et al. Novel Associations of BST1 and LAMP3 With REM Sleep Behavior Disorder. Neurology 2021 , 96 , e1412. [ Google Scholar ] [ CrossRef ]
  • Alcalay, R.N.; Levy, O.A.; Wolf, P.; Oliva, P.; Zhang, X.K.; Waters, C.H.; Fahn, S.; Kang, U.J.; Liong, C.; Ford, B.; et al. SCARB2 variants and glucocerebrosidase activity in Parkinson’s disease. Npj Park. Dis. 2016 , 2 , 16004. [ Google Scholar ] [ CrossRef ]
  • Siddiqui, I.J.; Pervaiz, N.; Abbasi, A.A. The Parkinson Disease gene SNCA: Evolutionary and structural insights with pathological implication. Sci. Rep. 2016 , 6 , 24475. [ Google Scholar ] [ CrossRef ]
  • Di Maio, R.; General, I.J.; Furbee, E.; Ayoob, J.C.; Castro, S.L.; Bahar, I.; Greenamyre, J.T.; Pullara, F. Disulfide bridge formation prevents CaMKII/Calmodulin interaction in Parkinson’s disease. bioRxiv 2020 . [ Google Scholar ] [ CrossRef ]
  • Fanning, S.; Cirka, H.; Thies, J.L.; Jeong, J.; Niemi, S.M.; Yoon, J.; Ho, G.P.H.; Pacheco, J.A.; Dettmer, U.; Liu, L.; et al. Lipase regulation of cellular fatty acid homeostasis as a Parkinson’s disease therapeutic strategy. npj Park. Dis. 2022 , 8 , 74. [ Google Scholar ] [ CrossRef ]
  • Charvin, D. mGlu4 allosteric modulation for treating Parkinson’s disease. Neuropharmacology 2018 , 135 , 308–315. [ Google Scholar ] [ CrossRef ]
  • Fan, L.; Shi, C.; Hu, X.; Zhang, Z.; Zheng, H.; Luo, H.; Fan, Y.; Zhang, S.; Hu, Z.; Yang, J.; et al. Analysis of 12 GWAS-Linked Loci With Parkinson’s Disease in the Chinese Han Population. Front. Neurol. 2021 , 12 , 623913. [ Google Scholar ] [ CrossRef ]
  • Zhao, C.; Jia, M.; Song, H.; Yu, Z.; Wang, W.; Li, Q.; Zhang, L.; Zhao, W.; Cao, X. The E3 Ubiquitin Ligase TRIM40 Attenuates Antiviral Immune Responses by Targeting MDA5 and RIG-I. Cell Rep. 2017 , 21 , 1613–1623. [ Google Scholar ] [ CrossRef ]
  • Su, W.-M.; Gu, X.-J.; Hou, Y.-B.; Zhang, L.-Y.; Cao, B.; Ou, R.-W.; Wu, Y.; Chen, X.-P.; Song, W.; Zhao, B.; et al. Association Analysis of WNT3, HLA-DRB5 and IL1R2 Polymorphisms in Chinese Patients With Parkinson’s Disease and Multiple System Atrophy. Front. Genet. 2021 , 12 , 765833. [ Google Scholar ] [ CrossRef ]
  • Angelopoulou, E.; Paudel, Y.N.; Julian, T.; Shaikh, M.F.; Piperi, C. Pivotal Role of Fyn Kinase in Parkinson’s Disease and Levodopa-Induced Dyskinesia: A Novel Therapeutic Target? Mol. Neurobiol. 2021 , 58 , 1372–1391. [ Google Scholar ] [ CrossRef ]
  • Chen, D.; Zhang, R.; Shen, W.; Fu, H.; Liu, S.; Sun, K.; Sun, X. RPS12-specific shRNA inhibits the proliferation, migration of BGC823 gastric cancer cells with S100A4 as a downstream effector. Int. J. Oncol. 2013 , 42 , 1763–1769. [ Google Scholar ] [ CrossRef ]
  • Diaz-Ortiz, M.E.; Seo, Y.; Posavi, M.; Cordon, M.C.; Clark, E.; Jain, N.; Charan, R.; Gallagher, M.D.; Unger, T.L.; Amari, N.; et al. GPNMB confers risk for Parkinson’s disease through interaction with α-synuclein. Science 2022 , 377 , eabk0637. [ Google Scholar ] [ CrossRef ]
  • Jones-Tabah, J.; He, K.; Senkevich, K.; Karpilovsky, N.; Deyab, G.; Cousineau, Y.; Nikanorova, D.; Goldsmith, T.; del Cid Pellitero, E.; Chen, C.X.-Q.; et al. The Parkinson’s disease risk gene cathepsin B promotes fibrillar alpha-synuclein clearance, lysosomal function and glucocerebrosidase activity in dopaminergic neurons. bioRxiv 2023 . [ Google Scholar ] [ CrossRef ]
  • Fletcher, E.J.R.; Jamieson, A.D.; Williams, G.; Doherty, P.; Duty, S. Targeted repositioning identifies drugs that increase fibroblast growth factor 20 production and protect against 6-hydroxydopamine-induced nigral cell loss in rats. Sci. Rep. 2019 , 9 , 8336. [ Google Scholar ] [ CrossRef ]
  • Cosgrove, M.S.; Ding, Y.; Rennie, W.A.; Lane, M.J.; Hanes, S.D. The Bin3 RNA methyltransferase targets 7SK RNA to control transcription and translation. Wiley Interdiscip. Rev. RNA 2012 , 3 , 633–647. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Chattaragada, M.S.; Riganti, C.; Sassoe, M.; Principe, M.; Santamorena, M.M.; Roux, C.; Curcio, C.; Evangelista, A.; Allavena, P.; Salvia, R.; et al. FAM49B, a novel regulator of mitochondrial function and integrity that suppresses tumor metastasis. Oncogene 2018 , 37 , 697–709. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Decet, M.; Soukup, S.-F. Endophilin-A/SH3GL2 calcium switch for synaptic autophagy induction is impaired by a Parkinson’s risk variant. Autophagy 2023 , 1–3. [ Google Scholar ] [ CrossRef ]
  • Kara, E.; Crimi, A.; Wiedmer, A.; Emmenegger, M.; Manzoni, C.; Bandres-Ciga, S.; D’sa, K.; Reynolds, R.H.; Botía, J.A.; Losa, M.; et al. An integrated genomic approach to dissect the genetic landscape regulating the cell-to-cell transfer of α-synuclein. Cell Rep. 2021 , 35 , 109189. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Mendoza-Ferreira, N.; Karakaya, M.; Cengiz, N.; Beijer, D.; Brigatti, K.W.; Gonzaga-Jauregui, C.; Fuhrmann, N.; Hölker, I.; Thelen, M.P.; Zetzsche, S.; et al. De Novo and Inherited Variants in GBF1 are Associated with Axonal Neuropathy Caused by Golgi Fragmentation. Am. J. Hum. Genet. 2020 , 107 , 763–777. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cao, Y.-L.; Yang, Y.-P.; Mao, C.-J.; Zhang, X.-Q.; Wang, C.-T.; Yang, J.; Lv, D.-J.; Wang, F.; Hu, L.-F.; Liu, C.-F. A role of BAG3 in regulating SNCA/α-synuclein clearance via selective macroautophagy. Neurobiol. Aging 2017 , 60 , 104–115. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cao, M.; Park, D.; Wu, Y.; De Camilli, P. Absence of Sac2/INPP5F enhances the phenotype of a Parkinson’s disease mutation of synaptojanin 1. Proc. Natl. Acad. Sci. USA 2020 , 117 , 12428–12434. [ Google Scholar ] [ CrossRef ]
  • Clarin, J.D.; Reddy, N.; Alexandropoulos, C.; Gao, W.-J. The role of cell adhesion molecule IgSF9b at the inhibitory synapse and psychiatric disease. Neurosci. Biobehav. Rev. 2024 , 156 , 105476. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Rui, Q.; Ni, H.; Li, D.; Gao, R.; Chen, G. The Role of LRRK2 in Neurodegeneration of Parkinson Disease. Curr. Neuropharmacol. 2018 , 16 , 1348–1357. [ Google Scholar ] [ CrossRef ]
  • Li, N.-N.; Tan, E.-K.; Chang, X.-L.; Mao, X.-Y.; Zhang, J.-H.; Zhao, D.-M.; Liao, Q.; Yu, W.-J.; Peng, R. Genetic Association Study betweenSTK39 and CCDC62/HIP1R and Parkinson’s Disease. PLoS ONE 2013 , 8 , e79211. [ Google Scholar ] [ CrossRef ]
  • Wang, L.-H.; Lin, Y.-M.; Lin, C.-Y.; Chern, Y.; Wang, G.-S. Neurodegeneration induces a developmental RNA processing program by calpain-mediated MBNL2 degradation. bioRxiv 2020 . [ Google Scholar ] [ CrossRef ]
  • Guella, I.; Sherman, H.E.; Appel-Cresswell, S.; Rajput, A.; Rajput, A.H.; Farrer, M.J. Parkinsonism in GTP cyclohydrolase 1 mutation carriers. Brain 2015 , 138 , e349. [ Google Scholar ] [ CrossRef ]
  • Li, G.; Cui, S.; Du, J.; Liu, J.; Zhang, P.; Fu, Y.; He, Y.; Zhou, H.; Ma, J.; Chen, S. Association of GALC, ZNF184, IL1R2 and ELOVL7 With Parkinson’s Disease in Southern Chinese. Front. Aging Neurosci. 2018 , 10 , 402. [ Google Scholar ] [ CrossRef ]
  • Lesage, S.; Drouet, V.; Majounie, E.; Deramecourt, V.; Jacoupy, M.; Nicolas, A.; Cormier-Dequaire, F.; Hassoun, S.M.; Pujol, C.; Ciura, S.; et al. Loss of VPS13C Function in Autosomal-Recessive Parkinsonism Causes Mitochondrial Dysfunction and Increases PINK1/Parkin-Dependent Mitophagy. Am. J. Hum. Genet. 2016 , 98 , 500–513. [ Google Scholar ] [ CrossRef ]
  • Ruhl, D.A.; Bomba-Warczak, E.; Watson, E.T.; Bradberry, M.M.; Peterson, T.A.; Basu, T.; Frelka, A.; Evans, C.S.; Briguglio, J.S.; Basta, T.; et al. Synaptotagmin 17 controls neurite outgrowth and synaptic physiology via distinct cellular pathways. Nat. Commun. 2019 , 10 , 3532. [ Google Scholar ] [ CrossRef ]
  • Garfias, S.; Domínguez, B.T.; Rojas, A.T.; Arroyo, M.; Rodríguez, U.; Boll, C.; Sosa, A.; Sciutto, E.; Adalid-Peralta, L.; López, Y.M.; et al. Peripheral blood lymphocyte phenotypes in Alzheimer and Parkinson’s diseases. Neurología 2022 , 37 , 110–121. [ Google Scholar ] [ CrossRef ]
  • Wang, S.; Bleeck, A.; Kasri, N.N.; Kleefstra, T.; van Rhijn, J.-R.; Schubert, D. SETD1A Mediated H3K4 Methylation and Its Role in Neurodevelopmental and Neuropsychiatric Disorders. Front. Mol. Neurosci. 2021 , 14 , 772000. [ Google Scholar ] [ CrossRef ]
  • Cheng, L.; Chen, L.; Wei, X.; Wang, Y.; Ren, Z.; Zeng, S.; Zhang, X.; Wen, H.; Gao, C.; Liu, H. NOD2 promotes dopaminergic degeneration regulated by NADPH oxidase 2 in 6-hydroxydopamine model of Parkinson’s disease. J. Neuroinflamm. 2018 , 15 , 243. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Dharshini, S.A.P.; Jemimah, S.; Taguchi, Y.H.; Gromiha, M.M. Exploring Common Therapeutic Targets for Neurodegenerative Disorders Using Transcriptome Study. Front. Genet. 2021 , 12 , 639160. [ Google Scholar ] [ CrossRef ]
  • Di Lorenzo, G.; Iavarone, F.; Maddaluno, M.; Plata-Gómez, A.B.; Aureli, S.; Meza, C.P.Q.; Cinque, L.; Palma, A.; Reggio, A.; Cirillo, C.; et al. Phosphorylation of FAM134C by CK2 controls starvation-induced ER-phagy. Sci. Adv. 2022 , 8 , eabo1215. [ Google Scholar ] [ CrossRef ]
  • Hori, R.T.; Khan, M.M.; Xiao, J.; Hargrove, P.W.; Moss, T.; LeDoux, M.S. Behavioral and molecular effects of Ubtf knockout and knockdown in mice. Brain Res. 2022 , 1793 , 148053. [ Google Scholar ] [ CrossRef ]
  • Mani, C.; Acharya, G.; Kshirsagar, S.; Vijayan, M.; Khan, H.; Reddy, P.H.; Palle, K. A Novel Role for BRIP1/FANCJ in Neuronal Cells Health and in Resolving Oxidative Stress-Induced DNA Lesions. J. Alzheimer’s Dis. 2022 , 85 , 207–221. [ Google Scholar ] [ CrossRef ]
  • Tian, J.; Vemula, S.R.; Xiao, J.; Valente, E.M.; Defazio, G.; Petrucci, S.; Gigante, A.F.; Rudzińska-Bar, M.; Wszolek, Z.K.; Kennelly, K.D.; et al. Whole-exome sequencing for variant discovery in blepharospasm. Mol. Genet. Genom. Med. 2018 , 6 , 601–626. [ Google Scholar ] [ CrossRef ]
  • Obergasteiger, J.; Castonguay, A.-M.; Pizzi, S.; Magnabosco, S.; Frapporti, G.; Lobbestael, E.; Baekelandt, V.; Hicks, A.A.; Pramstaller, P.P.; Gravel, C.; et al. The small GTPase Rit2 modulates LRRK2 kinase activity, is required for lysosomal function and protects against alpha-synuclein neuropathology. npj Park. Dis. 2023 , 9 , 44. [ Google Scholar ] [ CrossRef ]
  • Cai, C.; Tang, Y.-D.; Zhai, J.; Zheng, C. The RING finger protein family in health and disease. Signal Transduct. Target. Ther. 2022 , 7 , 300. [ Google Scholar ] [ CrossRef ]
  • Falabella, M.; Vernon, H.J.; Hanna, M.G.; Claypool, S.M.; Pitceathly, R.D. Cardiolipin, Mitochondria, and Neurological Disease. Trends Endocrinol. Metab. 2021 , 32 , 224–237. [ Google Scholar ] [ CrossRef ]
  • Yong, Y.; Wu, Q.; Meng, X.; Lu, R.; Xia, H.; Pei, F.; Yang, X. Dyrk1a Phosphorylation of α-Synuclein Mediating Apoptosis of Dopaminergic Neurons in Parkinson’s Disease. Park. Dis. 2023 , 2023 , 8848642. [ Google Scholar ] [ CrossRef ]
  • Xu, W.; Han, S.-D.; Zhang, C.; Li, J.-Q.; Wang, Y.-J.; Tan, C.-C.; Li, H.-Q.; Dong, Q.; Mei, C.; Tan, L.; et al. The FAM171A2 gene is a key regulator of progranulin expression and modifies the risk of multiple neurodegenerative diseases. Sci. Adv. 2020 , 6 , eabb3063. [ Google Scholar ] [ CrossRef ]
  • Cheng, W.-W.; Zhu, Q.; Zhang, H.-Y. Identifying Risk Genes and Interpreting Pathogenesis for Parkinson’s Disease by a Multiomics Analysis. Genes 2020 , 11 , 1100. [ Google Scholar ] [ CrossRef ]
  • Blauwendraat, C.; Nalls, M.A.; Singleton, A.B. The genetic architecture of Parkinson’s disease. Lancet Neurol. 2020 , 19 , 170–178. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Goldman, S.M.; Marek, K.; Ottman, R.; Meng, C.; Comyns, K.; Chan, P.; Ma, J.; Marras, C.; Langston, J.W.; Ross, G.W. Concordance for Parkinson’s disease in twins: A 20-year update. Ann. Neurol. 2019 , 85 , 600–605. [ Google Scholar ] [ CrossRef ]
  • Lambert, J.C.; Ibrahim-Verbaas, C.A.; Harold, D.; Naj, A.C.; Sims, R.; Bellenguez, C.; Jun, G.; DeStefano, A.L.; Bis, J.C.; Beecham, G.W.; et al. Meta-analysis of 74,046 individuals identifies 11 new susceptibility loci for Alzheimer’s disease. Nat. Genet. 2013 , 45 , 1452–1458. [ Google Scholar ] [ CrossRef ]
  • Nalls, M.A.; Pankratz, N.; Lill, C.M.; Do, C.B.; Hernandez, D.G.; Saad, M.; DeStefano, A.L.; Kara, E.; Bras, J.; Sharma, M. Large-scale meta-analysis of genome-wide association data identifies six new risk loci for Parkinson’s disease. Nat. Genet. 2014 , 46 , 989–993. [ Google Scholar ] [ CrossRef ]
  • Ridge, P.G.; Mukherjee, S.; Crane, P.K.; Kauwe, J.S.K. Alzheimer’s disease genetics consortium Alzheimer’s Disease: Analyzing the missing heritability. PLoS ONE 2013 , 8 , e79771. [ Google Scholar ] [ CrossRef ]
  • Ciani, M.; Bonvicini, C.; Scassellati, C.; Carrara, M.; Maj, C.; Fostinelli, S.; Binetti, G.; Ghidoni, R.; Benussi, L. the missing heritability of sporadic frontotemporal dementia: New insights from rare variants in neurodegenerative candidate genes. Int. J. Mol. Sci. 2019 , 20 , 3903. [ Google Scholar ] [ CrossRef ]
  • Saudou, F.; Humbert, S. The Biology of Huntingtin. Neuron 2016 , 89 , 910–926. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lee, J.-M.; Correia, K.; Loupe, J.; Kim, K.-H.; Barker, D.; Hong, E.P.; Chao, M.J.; Long, J.D.; Lucente, D.; Vonsattel, J.P.G.; et al. CAG Repeat Not Polyglutamine Length Determines Timing of Huntington’s Disease Onset. Cell 2019 , 178 , 887–900.e14. [ Google Scholar ] [ CrossRef ]
  • Harding, R.J.; Tong, Y.-F. Proteostasis in Huntington’s disease: Disease mechanisms and therapeutic opportunities. Acta Pharmacol. Sin. 2018 , 39 , 754–769. [ Google Scholar ] [ CrossRef ]
  • Folger, A.; Wang, Y. The Cytotoxicity and Clearance of Mutant Huntingtin and Other Misfolded Proteins. Cells 2021 , 10 , 2835. [ Google Scholar ] [ CrossRef ]
  • Donnelly, K.M.; Coleman, C.M.; Fuller, M.L.; Reed, V.L.; Smerina, D.; Tomlinson, D.S.; Pearce, M.M.P. Hunting for the cause: Evidence for prion-like mechanisms in Huntington’s disease. Front. Neurosci. 2022 , 16 , 946822. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ritchie, D.L.; Barria, M.A. Prion Diseases: A Unique Transmissible Agent or a Model for Neurodegenerative Diseases? Biomolecules 2021 , 11 , 207. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Watson, N.; Brandel, J.-P.; Green, A.; Hermann, P.; Ladogana, A.; Lindsay, T.; Mackenzie, J.; Pocchiari, M.; Smith, C.; Zerr, I.; et al. The importance of ongoing international surveillance for Creutzfeldt–Jakob disease. Nat. Rev. Neurol. 2021 , 17 , 362–379. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Mey, G.M.; Mahajan, K.R.; DeSilva, T.M. Neurodegeneration in multiple sclerosis. Wiley Interdiscip. Rev. Syst. Biol. Med. 2023 , 15 , e1583. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Walton, C.; King, R.; Rechtman, L.; Kaye, W.; Leray, E.; Marrie, R.A.; Robertson, N.; La Rocca, N.; Uitdehaag, B.; Van Der Mei, I.; et al. Rising prevalence of multiple sclerosis worldwide: Insights from the Atlas of MS, third edition. Mult. Scler. J. 2020 , 26 , 1816–1821. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • International Multiple Sclerosis Genetics Consortium. Multiple sclerosis genomic map implicates peripheral immune cells and microglia in susceptibility. Science 2019 , 365 , eaav7188. [ Google Scholar ] [ CrossRef ]
  • Lublin, F.D.; Reingold, S.C.; Cohen, J.A.; Cutter, G.R.; Sørensen, P.S.; Thompson, A.J.; Wolinsky, J.S.; Balcer, L.J.; Banwell, B.; Barkhof, F.; et al. Defining the clinical course of multiple sclerosis: The 2013 revisions. Neurology 2014 , 83 , 278–286. [ Google Scholar ] [ CrossRef ]
  • Das, S.K.; Menezes, M.E.; Bhatia, S.; Wang, X.; Emdad, L.; Sarkar, D.; Fisher, P.B. Gene therapies for cancer: Strategies, challenges and successes. J. Cell. Physiol. 2015 , 230 , 259–271. [ Google Scholar ] [ CrossRef ]
  • Bulcha, J.T.; Wang, Y.; Ma, H.; Tai, P.W.L.; Gao, G. Viral vector platforms within the gene therapy landscape. Signal Transduct. Target. Ther. 2021 , 6 , 53. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • O’connor, D.M.; Boulis, N.M. Gene therapy for neurodegenerative diseases. Trends Mol. Med. 2015 , 21 , 504–512. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ramamoorth, M.; Narvekar, A. Non viral vectors in gene therapy-an overview. J. Clin. Diagn. Res. JCDR 2015 , 9 , GE01. [ Google Scholar ] [ CrossRef ]
  • Puhl, D.L.; D’Amato, A.R.; Gilbert, R.J. Challenges of gene delivery to the central nervous system and the growing use of biomaterial vectors. Brain Res. Bull. 2019 , 150 , 216–230. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Liu, D.; Zhu, M.; Zhang, Y.; Diao, Y. Crossing the blood-brain barrier with AAV vectors. Metab. Brain Dis. 2020 , 36 , 45–52. [ Google Scholar ] [ CrossRef ]
  • Kumar, S.R.; Markusic, D.M.; Biswas, M.; A High, K.; Herzog, R.W. Clinical development of gene therapy: Results and lessons from recent successes. Mol. Ther.-Methods Clin. Dev. 2016 , 3 , 16034. [ Google Scholar ] [ CrossRef ]
  • Pai, S.-Y.; Logan, B.R.; Griffith, L.M.; Buckley, R.H.; Parrott, R.E.; Dvorak, C.C.; Kapoor, N.; Hanson, I.C.; Filipovich, A.H.; Jyonouchi, S.; et al. Transplantation Outcomes for Severe Combined Immunodeficiency, 2000–2009. N. Engl. J. Med. 2014 , 371 , 434–446. [ Google Scholar ] [ CrossRef ]
  • Demirci, S.; Leonard, A.; Haro-Mora, J.J.; Uchida, N.; Tisdale, J.F. CRISPR/Cas9 for Sickle Cell Disease: Applications, Future Possibilities, and Challenges. Adv. Exp. Med. Biol. 2019 , 1144 , 37–52. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Mendell, J.R.; Al-Zaidy, S.; Shell, R.; Arnold, W.D.; Rodino-Klapac, L.R.; Prior, T.W.; Lowes, L.; Alfano, L.; Berry, K.; Church, K.; et al. Single-dose gene-replacement therapy for spinal muscular atrophy. N. Engl. J. Med. 2017 , 377 , 1713–1722. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Sun, J.; Roy, S. Gene-based therapies for neurodegenerative diseases. Nat. Neurosci. 2021 , 24 , 297–311. [ Google Scholar ] [ CrossRef ]
  • Mullard, A. First in vivo CRISPR candidate enters the clinic. Nat. Rev. Drug Discov. 2019 , 18 , 656. [ Google Scholar ] [ CrossRef ]
  • Zhang, H.-X.; Zhang, Y.; Yin, H. Genome editing with mRNA encoding ZFN, TALEN, and Cas9. Mol. Ther. 2019 , 27 , 735–746. [ Google Scholar ] [ CrossRef ]
  • Jankele, R.; Svoboda, P. TAL effectors: Tools for DNA targeting. Brief. Funct. Genom. 2014 , 13 , 409–419. [ Google Scholar ] [ CrossRef ]
  • Pickar-Oliver, A.; Gersbach, C.A. The next generation of CRISPR–Cas technologies and applications. Nat. Rev. Mol. Cell Biol. 2019 , 20 , 490–507. [ Google Scholar ] [ CrossRef ]
  • Choi, E.H.; Suh, S.; Sears, A.E.; Hołubowicz, R.; Kedhar, S.R.; Browne, A.W.; Palczewski, K. Genome editing in the treatment of ocular diseases. Exp. Mol. Med. 2023 , 55 , 1678–1690. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Willems, J.; de Jong, A.P.H.; Scheefhals, N.; Mertens, E.; Catsburg, L.A.E.; Poorthuis, R.B.; de Winter, F.; Verhaagen, J.; Meye, F.J.; MacGillavry, H.D. ORANGE: A CRISPR/Cas9-based genome editing toolbox for epitope tagging of endogenous proteins in neurons. PLoS Biol. 2020 , 18 , e3000665. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Landrum, M.J.; Lee, J.M.; Riley, G.R.; Jang, W.; Rubinstein, W.S.; Church, D.M.; Maglott, D.R. ClinVar: Public archive of relationships among sequence variation and human phenotype. Nucleic Acids Res. 2014 , 42 , D980–D985. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Komor, A.C.; Kim, Y.B.; Packer, M.S.; Zuris, J.A.; Liu, D.R. Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage. Nature 2016 , 533 , 420–424. [ Google Scholar ] [ CrossRef ]
  • Gaudelli, N.M.; Komor, A.C.; Rees, H.A.; Packer, M.S.; Badran, A.H.; Bryson, D.I.; Liu, D.R. Programmable base editing of A•T to G•C in genomic DNA without DNA cleavage. Nature 2017 , 551 , 464–471. [ Google Scholar ] [ CrossRef ]
  • Levy, J.M.; Yeh, W.-H.; Pendse, N.; Davis, J.R.; Hennessey, E.; Butcher, R.; Koblan, L.W.; Comander, J.; Liu, Q.; Liu, D.R. Cytosine and adenine base editing of the brain, liver, retina, heart and skeletal muscle of mice via adeno-associated viruses. Nat. Biomed. Eng. 2020 , 4 , 97–110. [ Google Scholar ] [ CrossRef ]
  • Grünewald, J.; Zhou, R.; Garcia, S.P.; Iyer, S.; Lareau, C.A.; Aryee, M.J.; Joung, J.K. Transcriptome-wide off-target RNA editing induced by CRISPR-guided DNA base editors. Nature 2019 , 569 , 433–437. [ Google Scholar ] [ CrossRef ]
  • Zuo, E.; Sun, Y.; Yuan, T.; He, B.; Zhou, C.; Ying, W.; Liu, J.; Wei, W.; Zeng, R.; Li, Y.; et al. A rationally engineered cytosine base editor retains high on-target activity while reducing both DNA and RNA off-target effects. Nat. Methods 2020 , 17 , 600–604. [ Google Scholar ] [ CrossRef ]
  • Anzalone, A.V.; Randolph, P.B.; Davis, J.R.; Sousa, A.A.; Koblan, L.W.; Levy, J.M.; Chen, P.J.; Wilson, C.; Newby, G.A.; Raguram, A.; et al. Search-and-replace genome editing without double-strand breaks or donor DNA. Nature 2019 , 576 , 149–157. [ Google Scholar ] [ CrossRef ]
  • Rafii, M.S.; Baumann, T.L.; Bakay, R.A.; Ostrove, J.M.; Siffert, J.; Fleisher, A.S.; Herzog, C.D.; Barba, D.; Pay, M.; Salmon, D.P.; et al. A phase1 study of stereotactic gene delivery of AAV2-NGF for Alzheimer’s disease. Alzheimer’s Dement. 2014 , 10 , 571–581. [ Google Scholar ] [ CrossRef ]
  • Rocco, M.L.; Soligo, M.; Manni, L.; Aloe, L. Nerve growth factor: Early studies and recent clinical trials. Curr. Neuropharmacol. 2018 , 16 , 1455–1465. [ Google Scholar ] [ CrossRef ]
  • Tuszynski, M.H.; Yang, J.H.; Barba, D.; Hoi-Sang, U.; Bakay, R.A.; Pay, M.M.; Masliah, E.; Conner, J.M.; Kobalka, P.; Roy, S.; et al. Nerve growth factor gene therapy: Activation of neuronal responses in Alzheimer disease. JAMA Neurol. 2015 , 72 , 1139–1147. [ Google Scholar ] [ CrossRef ]
  • Rafii, M.S.; Tuszynski, M.H.; Thomas, R.G.; Barba, D.; Brewer, J.B.; Rissman, R.A.; Siffert, J.; Aisen, P.S. Adeno-associated viral vector (serotype 2)–nerve growth factor for patients with alzheimer disease: A randomized clinical trial. JAMA Neurol. 2018 , 75 , 834–841. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Castle, M.J.; Baltanás, F.C.; Kovacs, I.; Nagahara, A.H.; Barba, D.; Tuszynski, M.H. Postmortem analysis in a clinical trial of AAV2-NGF gene therapy for Alzheimer’s disease identifies a need for improved vector delivery. Hum. Gene Ther. 2020 , 31 , 415–422. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • DeVos, S.L.; Miller, R.L.; Schoch, K.M.; Holmes, B.B.; Kebodeaux, C.S.; Wegener, A.J.; Chen, G.; Shen, T.; Tran, H.; Nichols, B.; et al. Tau reduction prevents neuronal loss and reverses pathological tau deposition and seeding in mice with tauopathy. Sci. Transl. Med. 2017 , 9 , eaag0481. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Sud, R.; Geller, E.T.; Schellenberg, G.D. Antisense-mediated exon skipping decreases tau protein expression: A potential therapy for tauopathies. Mol. Ther.-Nucleic Acids 2014 , 3 , e180. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Zhou, L.; McInnes, J.; Wierda, K.; Holt, M.; Herrmann, A.G.; Jackson, R.J.; Wang, Y.-C.; Swerts, J.; Beyens, J.; Miskiewicz, K.; et al. Tau association with synaptic vesicles causes presynaptic dysfunction. Nat. Commun. 2017 , 8 , 15295. [ Google Scholar ] [ CrossRef ]
  • Velazquez, R.; Ferreira, E.; Tran, A.; Turner, E.C.; Belfiore, R.; Branca, C.; Oddo, S. Acute tau knockdown in the hippocampus of adult mice causes learning and memory deficits. Aging Cell 2018 , 17 , e12775. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Reiman, E.M.; Arboleda-Velasquez, J.F.; Quiroz, Y.T.; Huentelman, M.J.; Beach, T.G.; Caselli, R.J.; Chen, Y.; Su, Y.; Myers, A.J.; Hardy, J.; et al. Exceptionally low likelihood of Alzheimer’s dementia in APOE2 homozygotes from a 5,000-person neuropathological study. Nat. Commun. 2020 , 11 , 667. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yamazaki, Y.; Zhao, N.; Caulfield, T.R.; Liu, C.-C.; Bu, G. Apolipoprotein E and Alzheimer disease: Pathobiology and targeting strategies. Nat. Rev. Neurol. 2019 , 15 , 501–518. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Jackson, R.J.; Keiser, M.S.; Meltzer, J.C.; Fykstra, D.P.; Dierksmeier, S.E.; Melloni, A.; Nakajima, T.; Tecedor, L.; Ranum, P.T.; Carrell, E.; et al. APOE2 gene therapy reduces amyloid deposition, and improves markers of neuroinflammation and neurodegeneration in a mouse model of Alzheimer disease. bioRxiv 2023 . [ Google Scholar ] [ CrossRef ]
  • Yu, J.T.; Tan, L.; Hardy, J. Apolipoprotein E in Alzheimer’s disease: An update. Annu. Rev. Neurosci. 2014 , 37 , 79–100. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lin, Y.T.; Seo, J.; Gao, F.; Feldman, H.M.; Wen, H.L.; Penney, J.; Cam, H.P.; Gjoneska, E.; Raja, W.K.; Cheng, J.; et al. APOE4 causes widespread molecular and cellular alterations associated with Alzheimer’s disease phenotypes in human iPSC-derived brain cell types. Neuron 2018 , 98 , 1141–1154.e7. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • György, B.; Lööv, C.; Zaborowski, M.P.; Takeda, S.; Kleinstiver, B.P.; Commins, C.; Kastanenka, K.; Mu, D.; Volak, A.; Giedraitis, V.; et al. CRISPR/Cas9 mediated disruption of the Swedish APP allele as a therapeutic approach for early-onset Alzheimer’s disease. Mol. Ther.-Nucleic Acids 2018 , 11 , 429–440. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Chang, J.L.; Hinrich, A.J.; Roman, B.; Norrbom, M.; Rigo, F.; Marr, R.A.; Norstrom, E.M.; Hastings, M.L. Targeting amyloid-β precursor protein, APP, splicing with antisense oligonucleotides reduces toxic amyloid-β production. Mol. Ther. 2018 , 26 , 1539–1551. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Sun, J.; Carlson-Stevermer, J.; Das, U.; Shen, M.; Delenclos, M.; Snead, A.M.; Koo, S.Y.; Wang, L.; Qiao, D.; Loi, J.; et al. CRISPR/Cas9 editing of APP C-terminus attenuates β-cleavage and promotes α-cleavage. Nat. Commun. 2019 , 10 , 53. [ Google Scholar ] [ CrossRef ]
  • Mittermeyer, G.; Christine, C.W.; Rosenbluth, K.H.; Baker, S.L.; Starr, P.; Larson, P.; Kaplan, P.L.; Forsayeth, J.; Aminoff, M.J.; Bankiewicz, K.S. Long-term evaluation of a phase 1 study of AADC gene therapy for Parkinson’s disease. Hum. Gene Ther. 2012 , 23 , 377–381. [ Google Scholar ] [ CrossRef ]
  • Sebastian, W.S.; Richardson, R.M.; Kells, A.P.; Lamarre, C.; Bringas, J.; Pivirotto, P.; Salegio, E.A.; DeArmond, S.J.; Forsayeth, J.; Bankiewicz, K.S. Safety and tolerability of magnetic resonance imaging-guided convection-enhanced delivery of AAV2-hAADC with a novel delivery platform in nonhuman primate striatum. Hum. Gene Ther. 2012 , 23 , 210–217. [ Google Scholar ] [ CrossRef ]
  • Palfi, S.; Gurruchaga, J.M.; Ralph, G.S.; Lepetit, H.; Lavisse, S.; Buttery, P.C.; Watts, C.; Miskin, J.; Kelleher, M.; Deeley, S.; et al. Long-term safety and tolerability of ProSavin, a lentiviral vector-based gene therapy for Parkinson’s disease: A dose escalation, open-label, phase 1/2 trial. Lancet 2014 , 383 , 1138–1146. [ Google Scholar ] [ CrossRef ]
  • Emborg, M.E.; Carbon, M.; E Holden, J.; During, M.J.; Ma, Y.; Tang, C.; Moirano, J.; Fitzsimons, H.; Roitberg, B.Z.; Tuccar, E.; et al. Subthalamic glutamic acid decarboxylase gene therapy: Changes in motor function and cortical metabolism. J. Cereb. Blood Flow Metab. 2007 , 27 , 501–509. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Niethammer, M.; Tang, C.C.; LeWitt, P.A.; Rezai, A.R.; Leehey, M.A.; Ojemann, S.G.; Flaherty, A.W.; Eskandar, E.N.; Kostyk, S.K.; Sarkar, A.; et al. Long-term follow-up of a randomized AAV2-GAD gene therapy trial for Parkinson’s disease. J. Clin. Investig. 2017 , 2 , e90133. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Alarcón-Arís, D.; Pavia-Collado, R.; Miquel-Rio, L.; Coppola-Segovia, V.L.; Ferrés-Coy, A.; Ruiz-Bronchal, E.; Galofré, M.; Paz, V.; Campa, L.; Revilla, R.; et al. Anti-α-synuclein ASO delivered to monoamine neurons prevents α-synuclein accumulation in a Parkinson’s disease-like mouse model and in monkeys. EBioMedicine 2020 , 59 , 102944. [ Google Scholar ] [ CrossRef ]
  • Izco, M.; Blesa, J.; Schleef, M.; Schmeer, M.; Porcari, R.; Al-Shawi, R.; Ellmerich, S.; de Toro, M.; Gardiner, C.; Seow, Y.; et al. Systemic exosomal delivery of shRNA minicircles prevents parkinsonian pathology. Mol. Ther. 2019 , 27 , 2111–2122. [ Google Scholar ] [ CrossRef ]
  • Inoue, S.; Nishimura, K.; Gima, S.; Nakano, M.; Takata, K. CRISPR-Cas9-Edited SNCA Knockout Human Induced Pluripotent Stem Cell-Derived Dopaminergic Neurons and Their Vulnerability to Neurotoxicity. Biol. Pharm. Bull. 2023 , 46 , 517–522. [ Google Scholar ] [ CrossRef ]
  • Benskey, M.J.; Sellnow, R.C.; Sandoval, I.M.; Sortwell, C.E.; Lipton, J.W.; Manfredsson, F.P. Silencing alpha synuclein in mature nigral neurons results in rapid neuroinflammation and subsequent toxicity. Front. Mol. Neurosci. 2018 , 11 , 36. [ Google Scholar ] [ CrossRef ]
  • Rocha, E.M.; Smith, G.A.; Park, E.; Cao, H.; Brown, E.; Hayes, M.A.; Beagan, J.; McLean, J.R.; Izen, S.C.; Perez-Torres, E.; et al. Glucocerebrosidase gene therapy prevents α-synucleinopathy of midbrain dopamine neurons. Neurobiol. Dis. 2015 , 82 , 495–503. [ Google Scholar ] [ CrossRef ]
  • Morabito, G.; Giannelli, S.G.; Ordazzo, G.; Bido, S.; Castoldi, V.; Indrigo, M.; Cabassi, T.; Cattaneo, S.; Luoni, M.; Cancellieri, C.; et al. AAV-PHP.B-mediated global-scale expression in the mouse nervous system enables GBA1 gene therapy for wide protection from synucleinopathy. Mol. Ther. 2017 , 25 , 2727–2742. [ Google Scholar ] [ CrossRef ]
  • Fuji, R.N.; Flagella, M.; Baca, M.; Baptista, M.A.S.; Brodbeck, J.; Chan, B.K.; Fiske, B.K.; Honigberg, L.; Jubb, A.M.; Katavolos, P.; et al. Effect of selective LRRK2 kinase inhibition on nonhuman primate lung. Sci. Transl. Med. 2015 , 7 , ra15–ra273. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Zhao, H.T.; John, N.; Delic, V.; Ikeda-Lee, K.; Kim, A.; Weihofen, A.; Swayze, E.E.; Kordasiewicz, H.B.; West, A.B.; Volpicelli-Daley, L.A. LRRK2 antisense oligonucleotides ameliorate α-synuclein inclusion formation in a Parkinson’s disease mouse model. Mol. Ther.-Nucleic Acids 2017 , 8 , 508–519. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • McCampbell, A.; Cole, T.; Wegener, A.J.; Tomassy, G.S.; Setnicka, A.; Farley, B.J.; Schoch, K.M.; Hoye, M.L.; Shabsovich, M.; Sun, L.; et al. Antisense oligonucleotides extend survival and reverse decrement in muscle response in ALS models. J. Clin. Investig. 2018 , 128 , 3558–3567. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Miller, T.M.; Pestronk, A.; David, W.; Rothstein, J.; Simpson, E.; Appel, S.H.; Andres, P.L.; Mahoney, K.; Allred, P.; Alexander, K.; et al. A phase I, randomised, first-in-human study of an antisense oligonucleotide directed against SOD1 delivered intrathecally in SOD1-familial ALS patients. Lancet Neurol. 2013 , 12 , 435. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Miller, T.; Cudkowicz, M.; Shaw, P.J.; Andersen, P.M.; Atassi, N.; Bucelli, R.C.; Genge, A.; Glass, J.; Ladha, S.; Ludolph, A.L.; et al. Phase 1–2 trial of antisense oligonucleotide tofersen for SOD1 ALS. N. Engl. J. Med. 2020 , 383 , 109–119. [ Google Scholar ] [ CrossRef ]
  • Chen, Y.A.; Kankel, M.W.; Hana, S.; Lau, S.K.; Zavodszky, M.I.; McKissick, O.; Mastrangelo, N.; Dion, J.; Wang, B.; Ferretti, D.; et al. In vivo genome editing using novel AAV-PHP variants rescues motor function deficits and extends survival in a SOD1-ALS mouse model. Gene Ther. 2023 , 30 , 443–454. [ Google Scholar ] [ CrossRef ]
  • Mueller, C.; Berry, J.D.; McKenna-Yasek, D.M.; Gernoux, G.; Owegi, M.A.; Pothier, L.M.; Douthwright, C.L.; Gelevski, D.; Luppino, S.D.; Blackwood, M.; et al. SOD1 suppression with adeno-associated virus and microRNA in familial ALS. N. Engl. J. Med. 2020 , 383 , 151–158. [ Google Scholar ] [ CrossRef ]
  • Jiang, J.; Zhu, Q.; Gendron, T.F.; Saberi, S.; McAlonis-Downes, M.; Seelman, A.; Stauffer, J.E.; Jafar-Nejad, P.; Drenner, K.; Schulte, D.; et al. Gain of toxicity from ALS/FTD-linked repeat expansions in C9ORF72 is alleviated by antisense oligonucleotides targeting GGGGCC-containing RNAs. Neuron 2016 , 90 , 535–550. [ Google Scholar ] [ CrossRef ]

Click here to enlarge figure

S.No. Techniques MethodsBenefitsDisadvantages
1Sanger SequencingTraditional method involving the sequencing of DNA fragments using chain-termination dideoxy nucleotides.
2GWAS
(Genome-wide association studies)
Analyze the genetic variation across the entire genome to identify the link between specific genetic variants and a particular trait or disease.
3WGS
(Whole Genome Sequencing)
Involves sequencing the entire genome to identify both coding and non-coding variants associated with disease.
4WES
(Whole Exome Sequencing)
Focus on sequencing the protein coding region of genome to identify disease-associated variants.
5NGS
(Next-Generation Sequencing)
Utilizes high throughput sequencing technologies to sequence DNA or RNA molecule in parallel.
6LRS
(Long Read Sequencing)
Employs sequencing platforms that generate reads spanning hundreds to thousands of base pairs, providing more contiguous sequence information.
S.No. Gene (ALS)FunctionDisease Mechanism
1C9orf72Regulates vesicular transport and autophagyC9ORF72 haploinsufficiency (loss of function)
Sense and antisense RNA (GGGGCC) the function of RNA binding protein (gain of function)
2UNC13AFacilitates NeurotransmissionImpaired synaptic transmission [ ]
3SOD1Antioxidant roleOxidative stress, mitochondrial dysfunction, and excitotoxicity
4SCFD1Regulates ER to Golgi anterograde vesicular transportProtein misfolding and aggregation [ ]
5MOBP-RPSANeurons myelinationDemyelination of neurons [ ]
6HLAAntigen presentation and immune responseInflammation due to suppressed immune response [ ]
7KIF5AEngaged in anterograde transport of cargos along the microtubule rails in neuronsImpaired axonal transport, synaptic transmission, and motor neuronal toxicity [ ]
8CFAP410Cytoskeletal organization and ciliary function Decreased stability and increased degradation of mutant protein causes dysfunction of primary cilium [ ]
9GPX3-TNIPAntioxidantOxidative stress, mitochondrial dysfunction, and excitotoxicity [ ]
10SLC9A8Na/H exchangerExcitotoxicity and axonal degeneration [ ]
11TBK1Requires in cargo recruitment during autophagyNeuroinflammation and autophagy [ ]
12ERGIC1Maintains ER-Golgi structureDisintegration of ER and mitophagy [ ]
13NEK1A protein kinase that regulates cell cycle, DNA damage repair, apoptosis, and ciliary functionInduces DNA damage [ ]
14COG3Regulating Golgi processes, protein trafficking, and glycosylation in neuronsProtein trafficking by Golgi fragmentation [ ]
15PTPRN2Involved in vesicle-mediated secretory process in hippocampus [ ]Not clear. Probably motor neuron dysfunction [ ]
S.No.Gene (AD)FunctionDisease Mechanism
1SORT1Directs trafficking of APP into recycling pathwaysLow level of SORT1 in AD causes increased Aβ deposition [ ]
2CR1Immune complement cascadeRegulates Aβ metabolism [ ]
3ADAM17Alpha-secretase imparts a role in APP processingCauses increased APP production [ ]
4PRKD3Cell proliferationCauses neuroinflammation [ ]
5NCK2Axon growth and synapse formation and Epinephrin-mediated axon guidanceDisturbs motor axon trajectory selection [ ]
6WDR12Ribosome biogenesis and cell proliferationPossibly causing neuroinflammation [ ]
7BIN1Endocytosis and intracellular traffickingEndosome defect [ ]
8INPP5DImmune signalingInflammasome activation in microglia [ ]
9MMECleaves and degrades beta-amyloidIncreased Aβ deposition and axonal neuropathy [ ]
10IDUALysosomal protein acts in degradation of misfolded proteinLysosomal dysfunction and increased proteinopathy [ ]
11RHOHRegulation of actin cytoskeleton, and dendrites formationSynaptic loss and spinal dysfunction [ ]
12CLNKImmunomodulatory functionDisturbed immune signaling and neuroinflammation [ ]
13ANKHRegulating inflammation NF-κB-mediated neuroinflammation [ ]
14COX7CMitochondrial bioenergeticsMitochondrial respiratory defects [ ]
15TNIP1Inhibition of the TNF-α signaling pathway and NF-κB activation/translocationMicroglial activation and inflammation [ ]
16RASGEF1CAssociated with immune functionNeuroinflammation [ ]
17HS3ST5Cellular uptake and distribution of molecules like growth factors and morphogens Promotes tau fibrillation into NFTs [ ]
18HLA-DQA1Dendritic cells, macrophages and B cells and involved in adaptive immune responsesStimulates adaptive immune signaling in AD and also activates PKC and TLR signaling [ ]
19UNC5CLInvolved in mediating axon growth, neuronal migration in neuronal development, regulation of cell apoptosis.Contributes to AD pathogenesis by activating DAPK1 which in turn causes aberrant tau, Aβ and neuronal apoptosis/autophagy
20TREM2Regulates microglia proliferation, survival, migration, and phagocytosis.Downregulation induces neuroinflammation [ ]
21TREML2Regulates microglial proliferationImmune-related neuroinflammatory and increased tau deposition [ ]
22CD2APEarly endosome formation and protein traffickingRegulates Aβ generation by a neuron-specific polarization of Aβ in dendritic early endosomes [ ]
23UMAD1Involved in endosome-ubiquitin homeostasis [ ]Possibly defects in protein degradation cascade and increased deposit of Aβ and tau
24ICA1ICA1 regulates AMPA receptor trafficking [ ]Possibly disturb synaptic signaling
25TMEM106BBrain lipid metabolism,Disturbed lipid homeostasis [ ]
26JAZF1Lipid/cholesterol metabolism and microglial efferocytosis [ ]Neuroinflammation by defective efferocytosis and defective lipid metabolism (not clear)
27SEC61GProtein trafficking, ER calcium leak channel [ ]-
28EPDR1Neurogenesis and synaptic signaling [ ]Not clear
29SPDYE3Cell cycle regulator [ ]-
30EPHA1Immune response, cholesterol metabolism, and synaptic functionSpine morphology abnormalities and synaptic dysfunction [ ]
31CTSBRegulates apoptosis, neuroinflammation, and autophagylysosomal leakage of cathepsin B to the cytosol leads to neurodegeneration and behavioral deficits [ ]
32SHARPINInflammation and immune system activation
Synaptic signaling
Attenuated inflammatory/immune response [ ]
33PTK2BCa -activated non-receptor tyrosine kinase, involved in synaptic plasticityNeuronal hyperexcitability by neuronal differentiation and electrical maturation [ ]
34CLUSecreted by glia binds to Aβ and plays a protective role by preventing Aβ aggregationAβ clearance [ ]
35ABCA1Cholesterol mobilizationDefective lipid metabolism, and neuroinflammation [ ]
36ANK3Scaffolding proteins recruit diverse membrane proteins, (ion channels and cell adhesion molecules) into subcellular membrane domainsAltered neuronal excitability and altered neuronal connectivity [ ]
37TSPAN14Regulates maturation and trafficking of the transmembrane metalloprotease ADAM10 [ ]Not clear
38BLNKParticipates in the regulation of PLC-γ activity and the activation of Ras pathway [ ]Not clear. Possibly involved in immune regulation
39PLEKHA1Adaptive immunityInflammatory responses [ ]
40USP6NLGTPase-activating protein involved in control of endocytosisDysfunction of the myeloid endolysosomal system [ ]
41SPI1Controls microglial development and function Regulating neuroinflammation [ ]
42EEDCatalyzes the methylation of histone and mediates the repressive chromatinSynaptic dysfunction due to upregulation of synapse related gene [ ]
43SORL1Regulates the recycling of the APP out of the endosomeEndosomal swelling and APP misprocessing [ ]
44TPCN1Encodes a voltage-dependent calcium channel and involved in long-term potentiation in hippocampal neurons,Altered calcium signaling and cognitive dysfunction [ ]
45IGH gene clusterImmune response [ ]Not clear
46FERMT2APP metabolism and axonal growthImpaired synaptic connectivity, and long-term potentiation in an APP-dependent manner [ ]
47SLC24A4Neural development and cholesterol metabolism Increased deposition of Ab and tau [ ]
48SPPL2AEngaged in the function of B-cells and dendritic cells.Activates TNF-α signaling [ ]
49MINDY2DeubiquitinationNot clear
50APH1Bγ-secretaseBrain atrophy and amyloid-β deposition [ ]
51SNX1Endosome traffickingPrevents BACE1 trafficking to the lysosomal degradation system, resulting in increased production of Aβ [ ]
52CTSHImmune regulationRole in neuroinflammation and amyloid β production [ ]
53BCKDKRegulation of neurotransmitter synthesis, and mTOR activity.Causes hyperexcitability, neuroinflammation, and dysregulation of neurotrophic factors [ ]
54IL34Stimulates proliferation of monocytes and macrophages Triggers neuroinflammation via colony-stimulating factor-1 receptor (CSF-1r) [ ].
55PLCG2Present on microglia and function as immune regulator Upregulated and activates inflammation related pathway [ ]
56DOC2AA calcium sensor, facilitates neurotranbsmitter release in Ca -dependent manner Abnormality in synaptic transmission [ ]
57MAFRegulates T-cell susceptibility to apoptosisProbably immune cell dysfunction and neuroinflammation [ ]
58FOXF1Cell proliferation, cell cycle, and regulatory protein Activated by PI3K/AKT and stress response and may cause inflammation [ ]
59PRDM7Methyltransferases induce trimethylation Possibly suppresses the synaptic gene [ ]
60WDR81Facilitates the recruitment of autophagic protein aggregates and promotes autophagic clearance Impaired autophagy [ ]
61MYO15AA myosin involved in actin organizationRetromer dysfunction [ ]
62GRNRegulates lysosomal biogenesis, inflammation, repair, stress response, and aging.Neuroinflammation [ ]
63SCIMPImmune regulation via major histocompatibility complex class II signaling. Neuroinflammation [ ]
64WNT3Synaptic function and
immune regulation
Causes synaptic dysfunction and inflammation via Wnt3/β-catenin/GSK3β signaling pathway [ ]
65ABI3Regulator of microglia Neuroinflammation [ ]
66TSPOAP1TSPO-associated protein 1, interacts with translocator protein (TSPO) and act indirectly to activate microglia Neuroinflammation [ ]
67ACEAn endopeptidaseACE has been shown to cleave amyloid-β (Aβ) [ ]
68KLF16Regulates dopamine receptorsModulates dopaminergic transmission in the brain [ ]
69SIGLEC11Immune regulation Proinflammation and phagocytosis [ ]
70LILRB2Aβ receptorPerturbance in synaptic signaling and cognitive impairment [ ]
71ABCA7Lipid homeostasis and phagocytosis.Disturbed lipid metabolism, ER stress. Impaired microglial response to inflammation
72RBCK1Involved in ubiquitination TNF-α-mediated activation of NF-κB pathway.
73SLC2A4RGEncodes solute carrier protein. Involved in cell cycle via CDK1 pathway [ ]Not clear. Possibly induces proliferation
74CASS4Role in inflammation, calcium signaling, and microtubule stabilization. Disturbed synaptic signaling and neuroinflammation [ ]
75APPProliferation, differentiation, and maturation of neural stem cells.Abnormal cleavage causes plaque deposition [ ]
76ADAMTS1APP hydrolysisIncreased Aβ generation through β-secretase-mediated cleavage [ ]
S.No.Gene (PD)FunctionMechanism
1KRTCAP2Dementia-related geneInflammation and neurodegeneration [ ]
2PMVKInvolved in mevalonate pathwayNot clear. Possibly same as GBA [ ]
3GBAP1Encodes for the enzyme glucocerebrosidase (GCase), a lysosomal enzymeLysosomal dysfunction [ ]
4FCGR2APhagocytosis and modulates inflammatory responsesBinds with IgG-specific immune complexes and activates signaling [ ]
5VAMP4Endosomal trafficking of synaptic proteins Impaired synaptic signaling and lysosomal degradation [ ]
6NUCKS1Cell growth and proliferation [ ]Not clear
7RAB29Lysosome-related organelle biogenesisLysosomal dysfunction
Axon termination [ ]
8ITPKBInvolved in inositol metabolism and calcium release from ERCauses α-synuclein aggregation by dysregulated calcium release from ER-to-mitochondria [ ]
9SIPA1L2Controls protein trafficking and BDNF/TrkB signaling [ ]Not clear; possibly abrupt synaptic signaling
10KCNS3Potassium channel Neuroinflammation [ ]
11KCNIP3Associated with inositol biosynthetic pathway [ ]Not clear
12MAP4K4Cell proliferation, inflammation, and stress responseCytokine activation and neuroinflammation [ ]
13TMEM163Influx or efflux transporter particularly Zn transport [ ]Not clear
14STK39Immune regulationInflammatory pathway [ ]
15SATB1Transcriptional response in dopaminergic neurons Senescence-mediated neuroinflammation [ ]
16LINC00693Involved in miRNA processing complex [ ]Might affect protein expression and accumulation
17IP6K2Mitochondrial respirationMitophagy via PINK1 signaling [ ]
18KPNA1Encodes importin α5 and is involved in lysosomal biogenesis and autophagy Disturbed protein degradation [ ]
19MED12LTranscriptional coactivation of nearly all RNA polymerase II-dependent genes, Wnt/beta-catenin pathway, and immune response [ ]Possibly transcriptional defects
20SPTSSBRegulates de novo synthesis of ceramidesNeuronal signaling, synaptic transmission, cell metabolism [ ]
21MCCC1Mitochondrial enzyme and involved in leucine catabolism [ ]Possibly associated with mitochondrial dysfunction
22GAKAssociated with lysosomal and chaperons Defected lysosomal-mediated protein degradation [ ]
23TMEM175Proton channel to maintain optimum pH in lysosomes Downregulation of TMEM175 causes lysosomal over-acidification, impaired proteolytic activity, and facilitated a-synuclein aggregation [ ]
24BST1Serves as a receptor that regulates leukocyte adhesion and migration Immune regulation and inflammation [ ]
25LCORL--
26SCARB2Encodes a receptor responsible for the transport of glucocerebrosidase (GCase) to the lysosomeAssociated with lysosomal defects [ ]
27FAM47EPresent in close proximity to SCARB2 Lysosome/autophagy dysfunction
28FAM47E-STBD1--
29SNCADopamine release and transport, fibrillization of MAPT, and suppression of both p53 expression and transactivation of proapoptotic genes leading to decreased caspase-3 activationSynuclein aggregation and induction of apoptosis [ ]
30CAMK2DCalcium/calmodulin-dependent protein kinase ii delta.
Involved in synaptic plasticity
Disturbed calcium signaling and synaptic function [ ]
31CLCN3Ion channel transporter and neurotransmitter signaling [ ].Disturbed synaptic signaling
32ELOVL7Catalyzing the elongation of very long-chain fatty acids.Possibly disturbed lipid metabolism and oxidative stress [ ]
33PAMGlutamate receptor at parasynapses, associated with anxiety and hyperexcitation.Disturbed glutamatergic and GABA signaling [ ]
34C5orf24-Function and mechanism are not clear. However, upregulated expression and DNA methylation in disease condition [ ]
35LOC100131289--
36TRIM40E3 ubiquitin-protein ligase and inhibits NF-κB activityProtein degradation and inflammation [ ]
37HLA-DRB5Immune regulationInflammation [ ]
38RIMS1Encodes a synaptic protein and involved in neurotransmitter release and synaptic transmission [ ]Possibly perturbance in synaptic signaling
39FYNIon channel function, growth factor receptor signaling, immune system regulation.Activates BDNF/TrkB, PKCδ, NF-κB, MAPK, Nrf2, and NMDAR signaling pathway and induces synuclein phosphorylation, inflammation and excitotoxicity [ ]
40RPS12A special function in cell competition that defines the competitiveness of cells.Not clear in PD; however, reported to cause inflammation [ ]
41GPNMBCell differentiation, migration, proliferationInteracts with a-synuclein and induces its phosphorylation, cellular internalization, and fibrillization [ ]
42GS1-124K5.11--
43CTSBLysosomal hydrolase cathepsin B involved in waste degradation in cells Autophagy and lysosomal dysfunction causing a-synuclein aggregation [ ]
44FGF20Maintenance of dopaminergic neuronsAffects dopaminergic neurons in paracrine manner [ ]
45BIN3Cytokinesis and RNA methyltransferase Probably target transcription and translation step [ ]
46FAM49BRegulates mitochondrial functionMitochondrial fission, oxidative stress, and inflammation [ ]
47SH3GL2Encodes Endophilin A which regulates autophagy in calcium dependent mannerAutophagy dysfunction at synapses [ ]
48UBAP2Synapse formation, maintenance, and signaling -
49ITGA8Alpha8 integrin, cell adhesion, cell signaling, and cytoskeletal organizationIncreases cell-to-cell transfer of a-Syn [ ]
50GBF1Maintenance and function of the Golgi apparatus, and mitochondria migration and positioningIncrease in Golgi fragmentation [ ]
51BAG3A chaperone and regulates autophagy Its downregulation promotes autophagy dysfunction and disease progression [ ]
52INPP5FPI4P-phosphatase
Involved in endocytic pathway
Disturbed endocytosis [ ]
53RNF141--
54DLG2DLG2-encoded protein involved in glutamate receptor phosphorylationPhosphorylation of NR2 subunit and hyperexcitability [ ]
55IGSF9BCell adhesion molecule at inhibitory synapses and plays role in neuroplasticity and synaptic transmission Any disturbance in inhibitory synapses causes dysregulation of information flow and cognitive defects [ ]
56LRRK2Associated with intracellular membranes and vesicular structuresCauses accumulation of a-synuclein, which activates MAPK signaling and microglial activation leading to inflammation [ ]
57SCAF11Encodes a caspase-
58HIP1RClathrin-mediated endocytosis, actin dynamics, intrinsic cell death pathway [ ]Not clear. Probably affects the endocytosis of a-synuclein and activate caspase response
59FBRSL1--
60CAB39LEncodes calcium binding 39-like protein-
61MBNL2Encodes for the muscleblind-like protein 2, which belongs to a conserved family of RNA-binding proteinsReduced MBNL2 expression accompanied by the reduction in a developmental RNA processing [ ]
62MIPOL1--
63GCH1Essential for dopamine productionAffects dopaminergic signaling [ ]
64RPS6KL1--
65GALCEncodes galactocerebrosidaseImpaired autophagy and disturbed protein trafficking causes a-synuclein deposition [ ]
66VPS13CLocalized to the outer membrane of mitochondriaPINK1/Parkin-dependent mitophagy [ ]
67SYT17Encodes synaptotagmin-17,
associated with vesicle trafficking and transport at synapses
Disturbed synaptic trafficking [ ]
68CD19Immune regulatory molecule presents on B lymphocyte Neuroinflammation by suppressing local immune response [ ]
69SETD1AHistone methyltransferase Might affect synaptic signaling, excitation and glutamatergic signaling [ ]
70NOD2Immune homeostasisNox-2-mediated oxidative stress and neuroinflammation followed by loss of dopaminergic neurons [ ]
71CASC16--
72CHD9Activates transcription factor CREBPP, Involve in Notch signalingAberrant survival signaling pathway [ ]
73CHRNB1Encodes subunit of the n-acetylcholine receptor, Ion channels, transporters, and neurotransmitter signaling [ ]Disturbed cholinergic signaling
74RETREG3Involved in ER autophagy Activation of ER autophagy by mTOR signaling [ ]
75UBTFTranscription factor associated with ds-DNA break and apoptosis Altered protein expression [ ]
76BRIP1Facilitates repair of SSBs and DSBsExcitotoxicity, mitochondrial damage, and cell death [ ]
77DNAH17Encodes dynein axonemal heavy chain 17 involved in cytokinesis, microtubule-based movement, mitotic spindle organization, meiotic nuclear division [ ]-
78ASXL3--
79RIT2Involved in lysosomal activityActivate LRRK2 gene and lysosomal dysfunction and leads to a-synuclein deposition [ ]
80MEX3CEncodes RNF 194, an RNA binding protein impart immunoregulatory role Neuroinflammation [ ]
81SPPL2B--
82CRLS1Encodes cardiolipin synthase 1, involved in mitochondrial membrane formation Mitophagy [ ]
83DYRK1ASynaptic and nuclear proteins, including transcription factorsCauses phosphorylation of a-synuclein and downregulates PI3K/AKT pathway to induce apoptosis [ ]
84FAM171A2Downstream of GRN, is a novel genetic regulator of progranulin production expressed on microglial surfaceDownregulates progranulin level in brain [ ]
85CRHR1Encodes corticotropin-releasing hormone receptor, involved in regulation of stress and immune responsesDownregulates CREB signaling [ ]
86WNT3Immune regulationPD-related gene expression in immune cells [ ]
The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

Firdaus, Z.; Li, X. Unraveling the Genetic Landscape of Neurological Disorders: Insights into Pathogenesis, Techniques for Variant Identification, and Therapeutic Approaches. Int. J. Mol. Sci. 2024 , 25 , 2320. https://doi.org/10.3390/ijms25042320

Firdaus Z, Li X. Unraveling the Genetic Landscape of Neurological Disorders: Insights into Pathogenesis, Techniques for Variant Identification, and Therapeutic Approaches. International Journal of Molecular Sciences . 2024; 25(4):2320. https://doi.org/10.3390/ijms25042320

Firdaus, Zeba, and Xiaogang Li. 2024. "Unraveling the Genetic Landscape of Neurological Disorders: Insights into Pathogenesis, Techniques for Variant Identification, and Therapeutic Approaches" International Journal of Molecular Sciences 25, no. 4: 2320. https://doi.org/10.3390/ijms25042320

Article Metrics

Article access statistics, further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

A general passion

May 30, 2024

A general passion

Image credit: pgen.1011291

research article

Conserved signalling functions for Mps1, Mad1 and Mad2 in the Cryptococcus neoformans spindle checkpoint

Mps1-dependent phosphorylation of C-terminal Mad1 residues is a critical step in Cryptococcus spindle checkpoint signalling. 

Image credit: pgen.1011302

Conserved signalling functions for Mps1, Mad1 and Mad2 in the Cryptococcus neoformans spindle checkpoint

Recently Published Articles

  • Escherichia coli adaptation under prolonged resource exhaustion is characterized by extreme parallelism and frequent historical contingency"> Escherichia coli adaptation under prolonged resource exhaustion is characterized by extreme parallelism and frequent historical contingency
  • An endothelial regulatory module links blood pressure regulation with elite athletic performance
  • Single-step generation of homozygous knockout/knock-in individuals in an extremotolerant parthenogenetic tardigrade using DIPA-CRISPR

Current Issue

Current Issue May 2024

Adaptations to nitrogen availability drive ecological divergence of chemosynthetic symbionts

The importance of nitrogen availability in driving the ecological diversification of chemosynthetic symbiont species and the role that bacterial symbionts may play in the adaptation of marine organisms to changing environmental conditions.

Image credit: pgen.1011295

Adaptations to nitrogen availability drive ecological divergence of chemosynthetic symbionts

Paramutation at the maize pl1 locus is associated with RdDM activity at distal tandem repeats

pl1 paramutation depends on trans-chromosomal RNA-directed DNA methylation operating at a discrete cis-linked and copy-number-dependent transcriptional regulatory element.

Image credit: pgen.1011296

Paramutation at the maize pl1 locus is associated with RdDM activity at distal tandem repeats

Research Article

Genomic analyses of Symbiomonas scintillans show no evidence for endosymbiotic bacteria but does reveal the presence of giant viruses

A multi-gene tree showed the three SsV genome types branched within highly supported clades with each of BpV2, OlVs, and MpVs, respectively.

Genomic analyses of Symbiomonas scintillans show no evidence for endosymbiotic bacteria but does reveal the presence of giant viruses

Image credit: pgen.1011218

A natural bacterial pathogen of C . elegans uses a small RNA to induce transgenerational inheritance of learned avoidance

A mechanism of learning and remembering pathogen avoidance likely happens in the wild. 

A natural bacterial pathogen of C. elegans uses a small RNA to induce transgenerational inheritance of learned avoidance

Image credit: pgen.1011178

Spoink , a LTR retrotransposon, invaded D. melanogaster populations in the 1990s

Evidence of Spoink retrotransposon's horizontal transfer into D. melanogaster populations post-1993, suggesting its origin from D.willistoni .

Spoink, a LTR retrotransposon, invaded D. melanogaster populations in the 1990s

Image credit: pgen.1011201

Comparison of clinical geneticist and computer visual attention in assessing genetic conditions

Understanding AI, specifically Deep Learning, in facial diagnostics for genetic conditions can enhance the design and utilization of AI tools.

Comparison of clinical geneticist and computer visual attention in assessing genetic conditions

Image credit: pgen.1011168

Maintenance of proteostasis by Drosophila Rer1 is essential for competitive cell survival and Myc-driven overgrowth

Loss of Rer1 induces proteotoxic stress, leading to cell competition and elimination ...

Maintenance of proteostasis by Drosophila Rer1 is essential for competitive cell survival and Myc-driven overgrowth

Image credit: pgen.1011171

Anthracyclines induce cardiotoxicity through a shared gene expression response signature

TOP2i induce thousands of shared gene expression changes in cardiomyocytes.

Anthracyclines induce cardiotoxicity through a shared gene expression response signature

Image credit: pgen.1011164

New PLOS journals accepting submissions

Five new journals unified in addressing global health and environmental challenges are now ready to receive submissions: PLOS Climate , PLOS Sustainability and Transformation , PLOS Water , PLOS Digital Health , and PLOS Global Public Health

COVID-19 Collection

The COVID-19 Collection highlights all content published across the PLOS journals relating to the COVID-19 pandemic.

Submit your Lab and Study Protocols to PLOS ONE !

PLOS ONE is now accepting submissions of Lab Protocols, a peer-reviewed article collaboration with protocols.io, and Study Protocols, an article that credits the work done prior to producing and publishing results.

PLOS Reviewer Center

A collection of free training and resources for peer reviewers of PLOS journals—and for the peer review community more broadly—drawn from research and interviews with staff editors, editorial board members, and experienced reviewers.

Ten Simple Rules

PLOS Computational Biology 's "Ten Simple Rules" articles provide quick, concentrated guides for mastering some of the professional challenges research scientists face in their careers.

Welcome New Associate Editors!

PLOS Genetics welcomes several new Associate Editors to our board: Nicolas Bierne, Julie Simpson, Yun Li, Hongbin Ji, Hongbing Zhang, Bertrand Servin, & Benjamin Schwessinger

Expanding human variation at PLOS Genetics

The former Natural Variation section at PLOS Genetics relaunches as Human Genetic Variation and Disease. Read the editors' reasoning behind this change.

PLOS Genetics welcomes new Section Editors

Quanjiang Ji (ShanghaiTech University) joined the editorial board and Xiaofeng Zhu (Case Western Reserve University) was promoted as new Section Editors for the PLOS Genetics Methods section.

PLOS Genetics editors elected to National Academy of Sciences

Congratulations to Associate Editor Michael Lichten and Consulting Editor Nicole King, who are newly elected members of the National Academy of Sciences.

Harmit Malik receives Novitski Prize

Congratulations to Associate Editor Harmit Malik, who was awarded the Edward Novitski Prize by the Genetics Society of America for his work on genetic conflict. Harmit has also been elected as a new member of the American Academy of Arts & Sciences.

Publish with PLOS

  • Submission Instructions
  • Submit Your Manuscript

Connect with Us

  • PLOS Genetics on Twitter
  • PLOS on Facebook

Get new content from PLOS Genetics in your inbox

Thank you you have successfully subscribed to the plos genetics newsletter., sorry, an error occurred while sending your subscription. please try again later..

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.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Published: 08 January 2020

A brief history of human disease genetics

  • Melina Claussnitzer 1 , 2 , 3 ,
  • Judy H. Cho 4 , 5 , 6 ,
  • Rory Collins 7 , 8 ,
  • Nancy J. Cox 9 ,
  • Emmanouil T. Dermitzakis 10 , 11 ,
  • Matthew E. Hurles 12 ,
  • Sekar Kathiresan 2 , 13 , 14 ,
  • Eimear E. Kenny 4 , 6 , 15 ,
  • Cecilia M. Lindgren 2 , 16 , 17 ,
  • Daniel G. MacArthur 2 , 13 , 18 ,
  • Kathryn N. North 19 , 20 ,
  • Sharon E. Plon 21 , 22 ,
  • Heidi L. Rehm 2 , 13 , 18 , 23 ,
  • Neil Risch 24 ,
  • Charles N. Rotimi 25 ,
  • Jay Shendure 26 , 27 , 28 ,
  • Nicole Soranzo 12 , 29 &
  • Mark I. McCarthy 17 , 30 , 31 , 32  

Nature volume  577 ,  pages 179–189 ( 2020 ) Cite this article

115k Accesses

325 Citations

653 Altmetric

Metrics details

  • Functional genomics
  • Medical genetics

A primary goal of human genetics is to identify DNA sequence variants that influence biomedical traits, particularly those related to the onset and progression of human disease. Over the past 25 years, progress in realizing this objective has been transformed by advances in technology, foundational genomic resources and analytical tools, and by access to vast amounts of genotype and phenotype data. Genetic discoveries have substantially improved our understanding of the mechanisms responsible for many rare and common diseases and driven development of novel preventative and therapeutic strategies. Medical innovation will increasingly focus on delivering care tailored to individual patterns of genetic predisposition.

You have full access to this article via your institution.

Similar content being viewed by others

gene disorder research paper

The road ahead in genetics and genomics

gene disorder research paper

Genetic association analysis of 77,539 genomes reveals rare disease etiologies

gene disorder research paper

Genetic associations of protein-coding variants in human disease

For almost all human diseases, individual susceptibility is, to some degree, influenced by genetic variation. Consequently, characterizing the relationship between sequence variation and disease predisposition provides a powerful tool for identifying processes fundamental to disease pathogenesis and highlighting novel strategies for prevention and treatment.

Over the past 25 years, advances in technology and analytical approaches, often building on major community projects—such as those that generated the human genome sequence 1 and elaborated on that reference to capture sites of genetic variation 2 , 3 , 4 , 5 , 6 —have enabled many of the genes and variants that are causal for rare diseases to be identified and enabled a systematic dissection of the genetic basis of common multifactorial traits. There is growing momentum behind the application of this knowledge to drive innovation in clinical care, most obviously through developments in precision medicine. Genomic medicine, which was previously restricted to a few specific clinical indications, is poised to go mainstream.

This Review charts recent milestones in the history of human disease genetics and provides an opportunity to reflect on lessons learned by the human genetics community. We focus first on the long-standing division between genetic discovery efforts targeting rare variants with large effects and those seeking alleles that influence predisposition to common diseases. We describe how this division, with its echoes of the century-old debate between Mendelian and biometric views of human genetics, has obscured the continuous spectrum of disease risk alleles—across the range of frequencies and effect sizes—observed in the population, and outline how genome-wide analyses in large biobanks are transforming genetic research by enabling a comprehensive perspective on genotype–phenotype relationships. We describe how the expansion in the scale and scope of strategies for enumerating the functional consequences of genetic variation is transforming the torrent of genetic discoveries from the past decade into mechanistic insights, and the ways in which this knowledge increasingly underpins advances in clinical care. Finally, we reflect on some of the challenges and opportunities that confront the field, and the principles that will, over the coming decade, drive the application of human genetics to enhance understanding of health and disease and maximize clinical benefit.

Rare diseases, rare variants

During the 1980s and 1990s, efforts to map disease genes were focused on rare, monogenic and syndromic diseases and were mostly driven by linkage analysis and fine mapping within large multiplex pedigrees. Localization of genetic signals was typically followed by Sanger sequencing of the genes found to map within the linked locus to identify disease-causing alleles. Assessments of pathogenicity, based on segregation of a putatively causal variant with disease across multiple families and evidence that the risk genotype was absent in healthy individuals, were typically followed by confirmatory functional studies in cellular and animal models. This path to gene identification was laborious; nevertheless, by 2000, around 1,000 of the estimated 7,000 single-gene inherited diseases had been characterized, including many with substantial biomedical impact, such as Huntington’s disease and cystic fibrosis 7 , 8 , 9 .

Completion of the draft human genome sequence 1 reduced many of the obstacles to disease-gene mapping and propelled a fourfold increase in the genes implicated as causal for rare, single-gene disorders (Fig. 1 ). Microarray-based detection of structural variation 10 and exome- and genome-wide sequencing 11 , 12 have been pivotal, bolstered by in silico analysis and prioritization of the discovered genetic variants. Increasing availability of reference datasets cataloguing population genetic variation across diverse ethnic backgrounds has supported robust causal inference 2 , 3 , 5 , 6 . More recently, the adoption of high-throughput sequencing technologies has enabled the full range of causal genetic variation, from single mutations to large structural rearrangements, to be identified in a single assay. These technologies have extended from research into clinical usage, driving earlier and faster diagnosis for genetic disorders.

figure 1

The cumulative numbers of genes harbouring variants causal for rare, monogenic diseases and traits and of significant GWAS associations implicated in common, complex diseases and traits are shown. Left, the advent of high-throughput sequencing technologies and availability of reference genomes from diverse populations has supported a fourfold increase in the discovery of rare disease-causing genes between 1999 and 2019. Right, international efforts such as the Human Genome Project and the HapMap Project, combined with GWAS and sequencing studies, have supported identification of more than 60,000 genetic associations across thousands of human diseases and traits. Centre, more recent developments have brought a synthesis of the rare- and common-variant approaches based around the combination of sequence-informed analyses in large cohorts. Key events contributing to these themes are depicted in the timeline. GA4GH, Global Alliance for Genomics and Health 160 ; ExAC, Exome Aggregation Consortium 5 .

Reduced reliance on multiplex pedigrees in favour of collections of affected cases, often with parents 13 , has proven decisive in identifying new dominant disorders, many of which were previously considered recessive 14 . Increasingly, discovery of rare disease genes has transitioned from genetic characterization of small numbers of individuals with similar clinical presentations to genome-wide sequencing of larger cohorts of phenotypically diverse patients. This genotype-driven approach has revealed new disorders associated with more variable clinical presentation 15 , 16 .

A more systematic approach to data sharing has been critical, both for the characterization of new disorders and diagnostic interpretation of potential causal alleles. The value of sharing genetic and phenotypic data from those thought to harbour rare undiagnosed genetic diseases has fostered global collaborative networks (for example, Matchmaker Exchange, DECIPHER and GeneMatcher) designed to match patients with similar genetic variants and/or phenotypic manifestations, even across continents 17 , 18 , 19 . Interactions between researchers and families with rare disease have enabled natural history studies to be driven by family support groups positioned to initiate data collection from patient cohorts once a causal gene is discovered 20 .

Clinical translation of these technologies has benefited from a series of information resources, including open databases of genes associated with rare disorders (for example, OMIM and ORPHANET) 21 , clinically interpreted variants (for example, ClinVar and ClinGen) 22 , 23 and patient records (for example, DECIPHER and MyGene2 ( https://mygene2.org/MyGene2 )) 17 . Access to resources that catalogue genetic variation across populations (such as ExAC and its successor gnomAD) 5 , 6 has enabled the confident exclusion of genetic variants too common in population-level data to be plausible causes of rare, penetrant early-onset genetic diseases 24 . These analyses have reduced the contamination of databases with variants erroneously interpreted as causal for disease, and are addressing the overestimation of disease penetrance arising from the historical focus on multiplex pedigrees 25 . Improved recognition of the variable penetrance of many ‘monogenic’ disease alleles has invigorated efforts to identify the genetic and environmental modifiers responsible 26 , 27 .

Although huge strides have been made in associating specific genes with particular disorders, establishing the causal role of individual variants within those genes remains problematic, and many patients with suspected rare genetic diseases are left without a definitive diagnosis 28 . Even for variants with established causality, the penetrance is often unclear. Resolving these uncertainties represents the central challenge for the field. Aggregation of sequencing data from large numbers of affected cases and population reference samples will provide the evidence base required for robust interpretation of variants. Highly parallelized in vitro cellular assays that allow assessment of the functional effects of all variants in a disease-associated gene can transform interpretation of novel variants 29 , although developing well-calibrated functional assays predictive of pathogenicity for all disease genes represents a daunting prospect. Direct functional genomic exploration of accessible and disease-relevant tissues from patients using RNA sequencing and DNA methylation assays 30 , 31 can identify previously cryptic causal genetic variants, particularly in under-explored regions outside protein-coding genes 32 , 33 . Developments in each of these areas will extend the range of variants and genes for which diagnostic and prognostic clinical information can be provided to patients and their families.

Common diseases, common variants

Efforts to apply the approach—linkage analysis in multiplex pedigrees—that had been so successful for the high-penetrance variants responsible for Mendelian disease were, with notable exceptions 34 , 35 , 36 , largely unsuccessful for common, later-onset traits with more complex multifactorial aetiologies, such as asthma, diabetes and depression. Recognition that association-based methods, focused on detecting phenotype-related differences in variant allele frequencies might have greater traction for identifying less penetrant common alleles redirected attention to analysis of case–control samples 37 . However, initial efforts targeting variants within ‘candidate’ genes were plagued by inadequate power, unduly liberal thresholds for declaring significance and scant attention to sources of bias and confounding, resulting in overblown claims and failed replication.

Systematic efforts to characterize genome-wide patterns of genomic variation, initially through the HapMap Consortium 2 , proved catalytic, demonstrating that the allelic structure of the genome was segmented into haplotype blocks, each containing sets of correlated variants. Recognition that this configuration could support genome-wide surveys of association energized the technological innovation—in the form of massively parallel genotyping arrays—to make such studies possible (Fig. 1 ). Early wins in acute macular degeneration 38 and inflammatory bowel disease 39 were encouraging, and progress on several fronts—expansion of study size, denser genotyping arrays, novel strategies for imputation, attention to biases and appropriate significance thresholds—delivered robust associations across a range of diseases 40 . Most variants uncovered by these early genome-wide association studies (GWAS) were common, with more subtle effects than many had anticipated. A host of trait-specific consortia formed, covering diverse dichotomous and quantitative phenotypes, to accelerate genetic discovery through the aggregation and meta-analysis of data from multiple GWAS studies 41 , 42 , 43 . Many tens of thousands of robust associations were identified 44 . Recently, increased access to exome and whole-genome sequence data has, through both direct association analysis 45 , 46 and imputation 3 , 4 , extended discovery to low-frequency and rare alleles previously inaccessible to GWAS.

In the decade since the first GWAS, understanding of the genetic basis of common human disease has been transformed. The disparity between the observed effects of the variants first identified by GWAS and estimates of overall trait heritability (the ‘missing heritability’ conundrum) is now largely resolved 47 . Common diseases are not simply aggregations of related Mendelian conditions: for most complex traits, genetic predisposition is shared across thousands of mostly common variants with individually modest effects on population risk 41 , 43 .

Although the collective contribution of low-frequency and rare risk alleles to overall trait variability appears modest compared with that attributable to common variants 45 , 48 , the rare risk alleles detected in current sample sizes necessarily have large phenotypic effects and are proportionately more likely to be coding, enhancing their value for biological inference. Founder populations (such as those from Finland and Iceland) have provided multiple examples of otherwise rare risk alleles driven to higher frequency locally through drift and/or selection 49 , 50 , 51 , 52 . In addition, studies in populations with high rates of consanguinity make it possible to identify individuals homozygous for otherwise rare loss-of-function alleles, the basis for a ‘human knockout’ project to systematically investigate the phenotypic consequences of gene disruption in humans 53 , 54 .

For most diseases, large-scale GWAS-aggregation efforts have been disproportionately powered by information from individuals of European descent 55 . Whereas patterns of genetic predisposition appear broadly similar across major population groups and many common risk alleles discovered in one population group are detectable in others, allele frequencies can vary substantially; extending GWAS and sequencing studies to diverse populations will surely generate a rich harvest of novel risk alleles.

The relative contributions of common and rare variants indicate that, for many traits, particularly those with post-reproductive onset, purifying selection has had only limited effect 45 , 56 . For a few risk alleles, hallmarks of balancing selection reflect increased carrier survival, usually through protection from infectious diseases. This includes well-known examples of alleles maintained at high frequency in populations of African descent 57 , 58 .

While the extensive linkage disequilibrium within human populations has been essential to discovery in GWAS, high correlation between adjacent variants frustrates mapping of the specific variants responsible for these associations. Increasing sample size, improved access to trans-ethnic data, and more representative imputation reference panels 3 provide a path to improved resolution of the causal variants 59 and clues to the molecular mechanisms through which they operate. Functional interpretation is easiest for causal variants within coding sequences; however, most common disease-risk variants map to noncoding sequences, and are presumed to influence predisposition through effects on transcriptional regulation. In these cases, mechanistic inference depends on connecting association signals to their downstream targets (see below). For many traits, there is clear convergence between common-variant association signals and genes implicated in monogenic forms of the same disease, as well as enrichment of GWAS signals in regulatory elements specifically active in cell types consistent with known disease biology 60 , 61 . This provides reassurance that, even as the number of association signals for a given disease proliferates, the genetic associations uncovered will coalesce around molecular and cellular processes with a core role in pathogenesis 62 , 63 .

Importantly, the signals discovered by GWAS have revealed many unexpected insights into the biological basis of complex disease. Examples include the role of complement in the pathogenesis of acute macular degeneration 38 , synaptic pruning in schizophrenia 64 and autophagy in inflammatory bowel disease 65 . In addition, as inherited sequence variation is a prominent cause of phenotypic variation (but the reverse is not true), risk variants identified by GWAS have value as genetic instruments, mapping causal relationships between traits and inferring contributions made by circulating biomarkers and environmental exposures to disease development 66 .

As described below, findings from GWAS have increasing translational impact through identification of novel therapeutic targets 67 , prioritization (and deprioritization) of existing ones 68 and development of polygenic scores that quantify individual genetic risk 69 .

Comprehensive genotype–phenotype maps

The historical division of disease-gene discovery into monogenic and polygenic strands arose from development and implementation of analytical approaches—family-based linkage and case–control association 37 —that are best-suited for detecting particular subsets of causal alleles. This obscured the true state of nature, with disease-risk alleles being distributed across a continuous spectrum of frequencies and effect sizes. In addition, the trait- and disease-specific perspective of early GWAS discovery (mostly reliant on case–control studies) was poorly equipped to investigate the contribution of genetic variants to phenotypic effects that are nested within or spread across classical disease definitions. Recent developments have enabled a more holistic perspective on genotype–phenotype relationships (Fig. 1 ).

One major advance has been the increasing availability of large prospective population-based cohorts. These biobank efforts, pioneered in studies such as the Framingham Cohort 70 and the efforts of DeCODE in Iceland 71 , 72 , now encompass a growing inventory of national cohorts in North America, Europe, Asia and beyond 73 , 74 , 75 , 76 . The UK Biobank study, including 500,000 largely healthy, middle-aged participants has been particularly influential, transforming human genetic research in part through permissive data-sharing policies that have allowed multiple research groups to analyse the data 74 . Efforts to make clinical data embedded in electronic health records and registries available for research 77 , 78 mean that biobanks increasingly provide access to a wide range of demographic, clinical and lifestyle data, captured in harmonized, systematic fashion from large, often multi-ethnic collections of individuals. For millions of biobank participants, this rich phenotypic information has been combined with genome-wide genetic data. There are nascent efforts to capture transcriptomic, proteomic and metabolomic phenotypes, although these are not yet at equivalent scale to the genetic data 79 , 80 . Biobank analyses have provided more generalizable estimates of the relevance of genetic risk factors in the context of the separate and joint effects of non-genetic factors 81 . Increasingly, integration with healthcare data brings a longitudinal dimension to phenotypic characterization, which facilitates analyses of disease progression and lifelong disease risk 82 .

The rich phenotypic scope of these cohorts has enabled variants of interest to be interrogated for associations across the gamut of available phenotypes. These phenome-wide association studies (PheWAS) have revealed the extent to which many variants have pleiotropic effects across multiple traits 83 . Some of these relationships are expected, such as the impact of obesity variants on risk of hepatic steatosis and type 2 diabetes 84 or variants that influence multiple autoimmune conditions 85 . Others connect diseases and traits in surprising ways, highlighting shared polygenic, pleiotropic effects and cell-type specificity, and delivering insights into shared biology and overlapping mechanisms 86 , 87 . These findings inform the prioritization of therapeutic targets, providing clues to potential on-target side effects and opportunities for drug repurposing 87 , 88 , 89 .

The second enabler of inclusive, systematic analysis of genotype–phenotype relationships has been access to whole-genome sequence data. The scale of genetic analysis based on sequence data still lags behind that of genome-wide genotyping data (the largest sequence-based datasets are one tenth the size of the largest GWAS 90 , 91 , 92 ), although reductions in sequencing costs are decreasing the differential. Most direct analysis of high-throughput sequence data has focused on the coding regions. Strategies for assigning variant function and jointly analysing sets of variants of similar functional effect have enabled aggregate, gene-level tests of rare functional-variant association that are often better powered than single-variant tests 91 , 92 . However, the principal benefit to date of whole-genome sequence data to genetic discovery has been to bolster array-based access to lower-frequency alleles, either directly, through their inclusion on genotyping platforms, or indirectly, through imputation from sequence-based reference samples 3 , 4 .

These developments have enabled researchers to bridge the gap between the monogenic and polygenic realms, identifying common variant modifiers of monogenic phenotypes contributing to the variable expression of rare, large-effect alleles 26 , 93 , and low-frequency and rare variants that influence common multifactorial traits 94 , 95 . This enables more rigorous evaluation of the contribution of rare and common variants to trait susceptibility 48 and supports the enumeration of ‘allelic series’ (sets of alleles of varying frequency, effect size and direction that disrupt the same gene) critical for studies of disease mechanism and therapeutic target optimization 89 , 96 . These developments are rapidly converging towards the ultimate destination: a comprehensive matrix of the effect of all observable genetic variants across the widest possible range of cross-sectional and longitudinal biomedical phenotypes. Success in this endeavour depends on ever greater harmonization between, and integration of results from, individual studies through sustained investments in data sharing.

Adding function

From the first linkage maps to whole-genome sequencing of large cohorts, human genetics has deployed increasingly sophisticated and inherently systematic approaches for mapping the genetic factors that underlie traits and diseases. However, progress in determining how these variants influence disease, through systematic interrogation of their functional effects on molecular, cellular and physiological processes, has been far slower.

For monogenic diseases, for which the alleles responsible are typically rare, penetrant and coding, genetic approaches have generally been both necessary and sufficient to implicate a gene as causal 28 . However, as efforts to elucidate the genetic basis of Mendelian disorders progress towards completion 97 , functional studies remain important to understand the mechanisms by which disruptive variation within a causal gene leads to disease phenotypes. Unlike common diseases, the clarity of causation for Mendelian disorders usually simplifies the task of generating models (including human cells and organoids or rodents) to connect genotype to organismal phenotype; these have led to many critical insights into the biology of health and disease in humans 98 , 99 . In addition, for genes harbouring variants with medically actionable consequences (as with the BRCA1 and BRCA2 mutations that are causal for early-onset breast and ovarian cancer), functional studies can support the translational interpretation of novel alleles identified by medical sequencing 29 .

For common diseases, functional studies have a more fundamental role. Although tens of thousands of associations have been discovered across thousands of common human diseases and traits 44 , multiple factors have frustrated efforts to convert these genetic signals to knowledge about causal variants, genes and mechanisms. For the common variants that underlie the bulk of complex-disease risk, the resolution of association mapping is often limited by the haplotype structure of the human genome 2 , 3 , 4 . Furthermore, most GWAS associations map to the noncoding genome and thus lack a direct address to the gene that mediates their effects. Growing appreciation of the pervasive role of pleiotropy complicates matters: many variants identified by GWAS are associated with multiple traits and exert diverse effects across multiple cell types 100 .

To date, relatively few studies have achieved the goal of connecting variants causal for complex traits to the molecular and cellular functions that mediate that predisposition. One early success described how regulatory variants that modulate SORT1 expression influence low-density lipoprotein cholesterol and myocardial infarction risk 101 . More recent examples have focused on the relationship between obesity-associated variants intronic to FTO , altered expression of IRX3 and IRX5 , and adipocyte 102 and hypothalamic 103 function. Similar functional descriptions have been reported for individual loci implicated in schizophrenia 64 , cardiovascular disease 104 , type 2 diabetes 105 and Alzheimer’s disease 106 , among others.

Over the past decade, the challenge for the functional genomics community has been to convert this ‘one-locus-at-a-time’ workflow to a systematic, multidimensional, integrative approach able to deliver genome-scale functional analyses to match genome-wide variant discovery (Fig. 2 ). At the molecular level, one cornerstone has been generation of genome-wide catalogues of functional activity. For example, the ENCODE and Roadmap Epigenomics projects have generated maps of histone modifications, transcription-factor binding, chromatin accessibility, three-dimensional genome structure and other regulatory annotations across hundreds of cell types and tissues 107 , 108 . The patterns of genomic overlap between these data and GWAS results enable the functional inference of risk variants, deliver clues to the specific cell types driving disease pathogenesis 60 , 109 and accelerate locus-specific mechanistic insights.

figure 2

Top, the growth in the number of genetic loci associated by GWAS with human traits and diseases (bars) and of variant-to-function studies (area under line, not to scale). Bottom, foundational technological and computational advances over the last decade that enabled (1) development of systematic, genome-wide catalogues of functional elements across multiple cell types and tissues (blue); (2) mapping of QTLs in the context of gene expression, metabolites, proteins and regulatory elements (red); (3) engineering of genes, genetic elements and genetic variation at increasing scale (orange); and (4) systematic tissue-specific surveys of regulatory elements and transcription (grey). scRNA-seq, single-cell RNA-sequencing analysis; ChIA-PET, chromatin interaction analysis by paired-end tag sequencing; ChIP–seq, chromatin immunoprecipitation followed by sequencing; FAIRE-seq, formaldehyde-assisted isolation of regulatory elements with sequencing; DHS-seq, DNase I-hypersensitive sites sequencing; ATAC-seq, assay for transposase-accessible chromatin using sequencing; MPRA, massively parallel reporter assay; STARR-seq, self-transcribing active regulatory region sequencing; CNN: convolutional neural networks. For further details and primary literature on many of these assays, see ref. 173 .

In parallel, there has been a scaling of efforts to connect trait-associated regulatory variants to the genes and processes that they regulate in cell types relevant to the disease of interest 110 , 111 . For example, the GTEx (Genotype-Tissue Expression) consortium has mapped thousands of expression quantitative trait loci (QTLs) across hundreds of individuals and dozens of tissues 112 . Further clues to the relationships between regulatory variants and their effector genes can be gathered from DNA proximity assays (such as Hi-C) and single-cell data 113 (Fig. 2 ). Programs such as HubMAP 114 and the Human Cell Atlas 115 are set to deliver comprehensive, high-resolution reference maps of individual human cell types across diverse developmental stages, providing new opportunities to understand how regulatory genetic variation results in cellular and organismal phenotypes.

Efforts to probe the clinical consequences of coding alleles with large phenotypic effects (particularly null alleles) in humans 53 , 54 and across diverse animal models 116 represent powerful strategies for extending functional analyses to the whole-body level. Connections between genetic variation and circulating proteomic and metabolomic data provide additional mechanistic links between cellular events and whole-body physiology 79 , 80 . These efforts are paralleled by PheWAS approaches 83 , which, by mapping variant effects across the range of traits available in biobanks and EMRs, can inform priors for cell types and pathways at individual loci. Importantly, whereas early studies typically linked GWAS risk alleles to data from a single functional assay, the focus is increasingly on maximizing biological insight through the multi-dimensional integration of multiple genome-wide data types using approaches such as heritability partitioning 117 , functional enrichment analyses 60 , 109 , integration of the three-dimensional genome structure 118 and deep convolutional neural networks 119 , 120 .

Although QTL analyses can implicate a haplotype in a molecular, cellular or organismal phenotype, they are, in isolation, insufficient to define the specific causal variants responsible. To address this, there has been rapid maturation of technologies, such as massively parallel reporter assays 121 , 122 , 123 and CRISPR genome editing, to support functional characterization of targeted sequence perturbations at scale. Variations on these methods enable the functional evaluation of genes (via knockout screens 124 ), regulatory elements (using CRISPR interference and CRISPR activation screens 125 , 126 ), and genetic variants (base editors 127 ) at increasing scale and resolution 29 . Combined with complex readouts—including high-content imaging 128 and single-cell transcriptomics and epigenomics 129 , 130 —these methods can generate empirical ‘truth’ data, supporting the development of in silico models to predict causal variants, effector transcripts 126 and cellular effects. In due course, such models should reduce the need for exhaustive experimental characterization of function for all variants across all cell types.

The goal of such efforts is to enumerate the cascade of molecular events that underlie observed genotype–phenotype associations using physiologically relevant cellular systems (from primary cells to organoids and ‘organ-on-chip’ designs) and whole-body assays appropriate to the disease of interest. Collectively, strategies that offer large-scale functional evaluation of variants and genes of interest will reduce (but probably not eliminate) the intensive effort required for ‘final mile’ validation of disease mechanisms in dedicated systems, thereby accelerating downstream translational application.

Clinical implementation

Medical genetics, as applied to rare diseases, has been characterized by the rapid application in the clinic of the transformative genomic technologies that drove initial research discoveries. There are now targeted genetic tests for nearly all clinical presentations attributable to large-impact alleles, alongside more extensive genome-sequencing assays that, when necessary, enable interrogation of a longer list of relevant genes. Genetic testing for symptomatic individuals and at-risk relatives occurs routinely in many medical specialties. In parallel, the use of somatic cancer testing has increased as therapies targeted to specific mutational events have entered clinical practice (these developments are reviewed elsewhere 131 , 132 ).

For patients with symptoms that indicate a probable monogenic aetiology (such as retinal degeneration, hearing loss or cardiomyopathy), targeted panels are typically the platform of choice 133 , although they are increasingly performed on a more extensive sequence backbone. For more complex phenotypes—those without a clear match to a specific syndrome, such as neurodevelopmental disorders and multiple congenital anomalies—testing has gravitated towards early deployment of exome and genome-sequencing platforms that offer speedy resolution of what has historically often been a traumatic diagnostic odyssey 15 , 134 . The power of genomic diagnosis is especially clear for those presenting with monogenic neurodevelopmental disorders and critically ill infants 135 , 136 . Sequencing of the parent–offspring trio can detect de novo variation in dominant disorders and phase biallelic rare variants in recessive disease 13 .

The transition from targeted gene tests to genomic sequencing enables recursive reanalysis, including reinterpretation of individual sequences on the basis of subsequent discoveries regarding causal disease alleles and their phenotypic consequences 137 . However, improved molecular diagnostics are required to ensure reliable detection of a subset of genetic disorders, including those arising from triplet repeats and complex rearrangements 138 . Deep sequencing of affected tissues for mosaic variants and the use of RNA sequencing to detect noncoding variants that drive early-onset disease (for example, through effects on splicing) represent new fronts for clinical diagnostics 30 .

Other examples of the rapid adoption of new genomic technologies include noninvasive prenatal testing (more than ten million tests by 2018 across multiple countries 139 , 140 , 141 ) and the use of recessive carrier panels for couples planning pregnancies. Newborn screening is now universal in many countries, although it is limited to disorders combining high-throughput low-cost detection with effective early interventions (such as diet restrictions or enzyme replacement) 142 . Genetic diagnostics are also increasingly applied to newborn screening as a reflex test following an abnormal (for example, metabolic) screening test 143 . Over the next decade, the repertoire of disorders captured by neonatal screening and prenatal testing is likely to expand markedly. Whereas prenatal testing may be more effective at avoiding disease, the associated ethical issues are more complex 144 .

Although genetic testing for rare disease and cancer has exploded, there has been more limited uptake of genetic information in other aspects of healthcare. For example, despite multiple examples of clinically important genetic markers related to drug efficacy and side-effect profile 145 , the roll-out of pharmacogenetics has been hampered by a range of factors, including lack of clinical decision support in electronic medical systems to guide the drug choice or dosing by the physician. This has been compounded by challenges in diagnostic testing: complex haplotype structures and structural variants at some key drug metabolism loci necessitate genome sequencing or specific targeted panels to detect all clinically relevant variants.

For common diseases, translational attention is currently focused on the clinical potential of polygenic risk scores. The development of robust polygenic scores for several common diseases has been catalysed by more precise per-variant effect estimates from larger GWAS datasets, improved algorithms for combining information across millions of single-nucleotide polymorphisms, and large-scale biobanks that support score validation 69 , 146 , 147 . For example, a genome-wide polygenic score for heart attack, incorporating 6.6 million variants, indicates that 5% of European-descent individuals have a risk of future cardiac events equivalent to that seen in those with less frequent monogenic forms of hypercholesterolaemia 69 . Increasingly, the shift from array-based genotyping to sequence-based analysis is facilitating risk prediction, which integrates information from rare, large-effect alleles with that from polygenic scores 93 . By improving the capture of genetic risk, particularly in non-European populations, and integrating environmental and biomarker data to quantify aspects of non-genetic risk, it should be possible to achieve increasingly accurate prediction of individual disease risk, and to use this information to tailor screening, prevention and treatment. Success will depend on developing models of risk that robustly integrate these diverse data types and on optimizing the strategies deployed to ensure effective implementation.

The absence of evidence-based guidelines to support healthcare recommendations continues to hinder the clinical applications of genetic data. In some countries, this is compounded by confusion over reimbursement and disparities in testing across society 148 . Many healthcare professionals lack experience in genomic medicine and need education and guidance to practice in the rapidly evolving space of genetic and genomic testing 149 . One consequence of these difficulties has been an expanding direct-to-consumer testing market, variably controlled by country-specific regulations 150 , which is moving beyond a focus on ancestry and personal traits, towards models in which individuals have direct access to ordering physicians and genetic counselors 151 . The risk of commercial influence in this model remains high. There are concerns about the consequences of unfettered release of genetic data of dubious or inflated clinical relevance, and limited infrastructure to pull these results into mainstream medical systems.

These advances have fostered debate about the value of genetics for population screening, for both monogenic and complex disorders. Population screening for monogenic disorders is most likely to be initiated for conditions for which risk estimates are well-understood and there are actionable interventions (for example, Lynch syndrome and familial hypercholesterolaemia). Expansion to other disorders requires better understanding of the penetrance of pathogenic alleles in unselected populations 152 and caution before extending screening to longer lists of genes that are less securely implicated in disease causation 153 . As certain countries consider universal capture of genome-wide genetic data at birth or later in life, key questions concern the strategies for releasing this information to citizens and their medical teams to support individual healthcare.

Ultimately, barriers to genomic medicine are most directly overcome by demonstrating clinical utility in disease management and therapeutic decision-making, with evidence for improved patient outcomes. Hereditary cancers provide multiple examples, such as the use of BRCA1/BRCA2 testing to inform PARP inhibitor treatment in patients with cancer 154 . There is a growing list of diseases for which a molecular diagnosis results in specific interventions designed to improve patient outcomes ( https://www.ncbi.nlm.nih.gov/books/NBK1116/ ) (some examples are listed in Table 1 ), and there are currently more than 50 FDA-approved drugs for genetic disorders 155 . Although gene therapy has been slow to evolve since its early introduction, recent advances in gene editing are reinvigorating approaches to treat disorders by manipulation of the underlying genetic defects 156 .

Looking forward

Over the coming decade, the challenge will be to optimize and to implement at scale, strategies that use human genetics to further the understanding of health and disease, and to maximize the clinical benefit of those discoveries. Realizing these goals will require the concerted effort of researchers in academia and industry to bring about transformational change across a range of highly interconnected domains, for example, through the auspices of the recently established International Common Disease Alliance ( https://www.icda.bio ). Such efforts will be directed towards establishing: (a) comprehensive inventories of genotype–phenotype relationships across populations and environments; (b) systematic assays of variant- and gene-level function across cell types, states and exposures; (c) improved scalable strategies for turning this basic knowledge into fully developed molecular, cellular and physiological models of disease pathogenesis; and (d) application of those biological insights to drive novel preventative and therapeutic options.

The first of these will involve documenting the full spectrum of natural genetic variation across all human populations, including capture of structural variants, and somatic mutations that accumulate with aging 157 , 158 , and associating these variations with the ever-richer disease-related intermediate and clinical traits available through biobanks and electronic health records. It will be particularly important to include populations historically under-represented in genomic research, following the pioneering work of the H3Africa consortium 159 . As over time, clinically sequenced genomes will outnumber those collected in academia, research and healthcare communities will need to develop a harmonized approach to genomics to transcend historical boundaries. Progress will be critically dependent on platforms and governance that lower barriers to the integration of genetic and phenotypic data across studies and countries, along with technical standards that are reliable, secure and compatible with the international regulatory landscape 160 .

Mechanistic interpretation of genetic associations, particularly those in regulatory regions, will be driven by the systematic annotation of sequence variants and genes for functional impact across disease-relevant cell types, enabling mapping of processes contributing to disease development with respect to place (tissue and cell type), time (developmental stage) and context (external influences) 161 . Accelerating efforts to characterize the cellular composition of tissues through single-cell assays 115 will increase the granularity of these observations. Large-scale perturbation studies across diverse cellular and animal models will, together with analyses of coding variants in humans 53 , 54 , provide confidence in causal inference. Large-scale proteomic and metabolomic analyses (in tissues and biological fluids) will provide a bridge to downstream pathways 79 , 80 . Research access to such functional data, generated at scale, should lower the barriers to mechanistic inference, provide system-wide context and enable researchers to focus wet-laboratory validation on the most critical experiments. Collectively, these efforts will support compilation of a systematic catalogue of key networks and processes that influence normal physiology and disease development and inform a revised molecular taxonomy of disease.

This knowledge will reinforce the essential contribution of human genetics to the identification and prioritization of targets for therapeutic development 89 , 162 . Insights into the efficacy of target perturbation and potential for adverse events, allied to characterization of translatable biomarkers, provide ways to boost the efficiency of drug-development pipelines 162 . Given the clinical importance of slowing disease progression 163 , target-discovery efforts will increasingly need to embrace the genetics of disease progression and treatment response, as these may involve processes distinct from those captured by studies of disease onset.

In parallel, the clinical use of human genetics will benefit from progress towards universal determination of individual genome sequences built through a combination of biobank expansion and direct access within healthcare systems. This will power clinical applications that extend beyond the current focus on neonatal sequencing, Mendelian diagnostics and somatic tumour sequencing 164 . In particular, improvements in polygenic score derivation will boost risk prediction for multifactorial traits, provide a molecular basis for disease classification, support biomarker discovery and therapeutic optimization and contribute to understanding of the variable penetrance of monogenic conditions 69 . Implementing genomic medicine as a routine component of clinical care across diverse healthcare environments will inevitably require investment in the training of healthcare professionals and attention to optimal strategies for returning genetic findings to patients.

The limited heritability of many multifactorial traits constrains the clinical precision available from genetic data alone. This will drive efforts to integrate information on personal environment, lifestyle and behaviour, and to combine prognostic, predictive information on disease risk with longitudinal measures of molecular and clinical state that track an individual’s journey from health to disease. Human genetics will also, given its unique potential for causal inference, support identification of the non-genetic risk factors (often modifiable) that directly contribute to disease predisposition and development 165 . As polygenic score performance improves, analysis of individuals who show marked divergence between genetic predisposition and real-world clinical outcomes should define exposures (such as lifestyle choices or gut microbiome) the contribution of which to disease causation remains unclear 166 .

Collectively, these developments can be expected to accelerate personalization of healthcare delivery. Provided costs are sustainable, a more preventative perspective on health could emerge, managed through proactive genomic, clinical and lifestyle surveillance using risk scores, complex biomarkers, liquid biopsies and wearables. Improved understanding of aetiological heterogeneity, patterns of sharing of genetic risk across diseases, variation in therapeutic response and risk of adverse events will enhance targeting of preventative and therapeutic interventions 167 . At the population level, intervention strategies will seek to combine population-wide and targeted strategies to best effect 168 . It will be critical to ensure that these benefits are available to as many as possible, so that genomics reduces, rather than exacerbates, national and global health disparities 55 , 169 (Box 1 ).

The developments described above, represent variations on the theme of ‘reading’ the genome. The emerging capacity to block this reading (for example, through siRNA therapies 170 ) or even to ‘write’ the genome (through CRISPR editing) promises to be equally transformative, providing new opportunities to correct, and even cure, Mendelian disease. Spectacular advances in developing novel therapeutic strategies are likely for many diseases, based, for example, on ex vivo cellular manipulation 171 or in vivo somatic cell editing 172 .

Importantly, developments in genomic medicine need to proceed in a bioethical framework for research and clinical use that recognizes the personal relevance of human genetics and the critical importance of autonomous consent and the protection of privacy, while minimizing the adverse consequences of genetic exceptionalism. Governance needs to reaffirm the rights of citizens to make individual contributions to scientific progress through research participation and encourage the responsible exchange of data for clinical and research purposes.

Box 1 Global genomics

Present and future advances in genetics and genomics have the potential to provide benefits to individuals and societies across the world, but equitable and fair access to those benefits will require proactive measures to address entrenched disparities in scientific capacity and clinical opportunities. This includes:

Global characterization of genetic variation. Systematic catalogues of human genome variation from a broad range of global populations will maximize the genetic diversity available for genetic discovery and clinical implementation. This will support more accurate imputation, enable more effective use of polygenic scores, and improve detection and interpretation of rare large-effect alleles.

Genetic discovery for diseases with restricted geographical coverage. Genetic studies conducted in diverse populations will support genetic discovery as well as functional characterization of hitherto-neglected diseases with disproportionate regional impact and may lead to novel preventative and therapeutic strategies.

An ethical framework for global research. Genetic research needs to proceed on the expectation that the benefits of research will be available to those who have participated, and that relationships between researchers and patients are based on robust expressions of accountability and governance. Participation should be based on appropriate informed individual consent in the context of community involvement, with mechanisms to support equitable data sharing with researchers in participating countries.

Support for local research capacity. Collection of biological specimens and data are dependent on adequate support for data collection, generation and analysis in participating countries, with infrastructure, training and career structures in place to support local researchers and equitable credit for shared research outputs.

Equitable translation. Participants and their communities should have fair and equitable access to the biomedical information arising from the studies in which they are involved to support population-relevant clinical diagnostics and implementation (for example, of polygenic scores). There should be commitment to the development of technologies and strategies that enable the clinical benefits of genomic medicine to be accessible to those in less-developed healthcare environments.

Future prospects

Over the past two decades, understanding of the genetic basis of human disease has been transformed by a combination of spectacular technological and analytical advances, collaborative commitment to the development of foundational resources and the collection and analysis of vast amounts of genetic, molecular and clinical data. The biological insights derived from these data are, increasingly, drivers of translational innovation, and widening personal access to large-scale genetic and molecular data promises to reshape medical care.

However, for the full potential of genomic medicine to be realized, there will need to be sustained collaborative endeavour on several fronts to ensure that the capacity to generate ever more detailed maps of the relationships between sequence variation and biomedical phenotypes delivers a comprehensive understanding of disease mechanisms that can be translated into the medicines of tomorrow.

International Human Genome Sequencing Consortium. Initial sequencing and analysis of the human genome. Nature 409 , 860–921 (2001). This paper describes the first analyses from the draft human genome sequence assembled over the previous decade: it launched modern human genetics and represents a tribute to the power of collaborative science.

Article   ADS   Google Scholar  

International HapMap Consortium. The International HapMap Project. Nature 426 , 789–796 (2003). The HapMap Consortium developed the first genome-wide maps of common sequence variation, using this information to lay out the haplotypic structure of this variation across three major ancestral groupings (from Europe, East Asia and Africa).

Article   CAS   Google Scholar  

1000 Genomes Project Consortium. A global reference for human genetic variation. Nature 526 , 68–74 (2015).

Article   ADS   CAS   Google Scholar  

The Haplotype Reference Consortium. A reference panel of 64,976 haplotypes for genotype imputation. Nat. Genet . 48 , 1279–1283 (2016).

Lek, M. et al. Analysis of protein-coding genetic variation in 60,706 humans. Nature 536 , 285–291 (2016).

Article   PubMed   PubMed Central   CAS   Google Scholar  

Karczewski, K. J. et al. Variation across 141,456 human exomes and genomes reveals the spectrum of loss-of-function intolerance across human protein-coding genes. Preprint at bioRxiv https://doi.org/10.1101/531210 (2019). The most recent report from the genome aggregation database (gnomAD) project combining human sequence data on an unprecedented scale to characterize rare, high-impact variants and their relationship to health and disease.

Kremer, B. et al. A worldwide study of the Huntington’s disease mutation. The sensitivity and specificity of measuring CAG repeats. N. Engl. J. Med . 330 , 1401–1406 (1994).

Article   PubMed   CAS   Google Scholar  

Collins, F. S. Identifying human disease genes by positional cloning. Harvey Lect . 86 , 149–164 (1990–1991).

PubMed   Google Scholar  

Gusella, J. F. & MacDonald, M. E. Huntington’s disease and repeating trinucleotides. N. Engl. J. Med . 330 , 1450–1451 (1994).

Vissers, L. E. et al. Array-based comparative genomic hybridization for the genomewide detection of submicroscopic chromosomal abnormalities. Am. J. Hum. Genet . 73 , 1261–1270 (2003).

Ng, S. B. et al. Targeted capture and massively parallel sequencing of 12 human exomes. Nature 461 , 272–276 (2009).

Article   ADS   PubMed   PubMed Central   CAS   Google Scholar  

Wheeler, D. A. et al. The complete genome of an individual by massively parallel DNA sequencing. Nature 452 , 872–876 (2008).

Article   ADS   PubMed   CAS   Google Scholar  

Vissers, L. E. et al. A de novo paradigm for mental retardation. Nat. Genet . 42 , 1109–1112 (2010).

Deciphering Developmental Disorders Study. Prevalence and architecture of de novo mutations in developmental disorders. Nature 542 , 433–438 (2017).

Splinter, K. et al. Effect of genetic diagnosis on patients with previously undiagnosed disease. N. Engl. J. Med . 379 , 2131–2139 (2018).

Deciphering Developmental Disorders Study. Large-scale discovery of novel genetic causes of developmental disorders. Nature 519 , 223–228 (2015).

Firth, H. V. et al. DECIPHER: database of chromosomal imbalance and phenotype in humans using ensembl resources. Am. J. Hum. Genet . 84 , 524–533 (2009).

Philippakis, A.A. et al. The Matchmaker Exchange: a platform for rare disease gene discovery. Hum. Mutat. 36 , 915–921 (2015).

Google Scholar  

Boycott, K. M. et al. International cooperation to enable the diagnosis of all rare genetic diseases. Am. J. Hum. Genet . 100 , 695–705 (2017).

Kennedy, J. et al. KAT6A syndrome: genotype–phenotype correlation in 76 patients with pathogenic KAT6A variants. Genet. Med . 21 , 850–860 (2019).

Amberger, J. S., Bocchini, C. A., Schiettecatte, F., Scott, A. F. & Hamosh, A. OMIM.org: Online Mendelian Inheritance in Man (OMIM), an online catalog of human genes and genetic disorders. Nucleic Acids Res . 43 , D789–D798 (2015).

Landrum, M. J. et al. ClinVar: public archive of relationships among sequence variation and human phenotype. Nucleic Acids Res . 42 , D980–D985 (2014).

Article   CAS   PubMed   Google Scholar  

Rehm, H. L. et al. ClinGen—the Clinical Genome Resource. N. Engl. J. Med . 372 , 2235–2242 (2015).

Minikel, E. V. et al. Quantifying prion disease penetrance using large population control cohorts. Sci. Transl. Med . 8 , 322ra9 (2016).

Flannick, J. et al. Assessing the phenotypic effects in the general population of rare variants in genes for a dominant Mendelian form of diabetes. Nat. Genet . 45 , 1380–1385 (2013).

Bečanović, K. et al. A SNP in the HTT promoter alters NF-κB binding and is a bidirectional genetic modifier of Huntington disease. Nat. Neurosci . 18 , 807–816 (2015).

Castel, S. E. et al. Modified penetrance of coding variants by cis -regulatory variation contributes to disease risk. Nat. Genet . 50 , 1327–1334 (2018).

MacArthur, D. G. et al. Guidelines for investigating causality of sequence variants in human disease. Nature 508 , 469–476 (2014).

Findlay, G. M. et al. Accurate classification of BRCA1 variants with saturation genome editing. Nature 562 , 217–222 (2018). This paper describes an effort to obtain massively parallel functional assessments of all potential sequence variants in a gene causal for familial breast cancer, and thereby provide more confident predictions of clinical significance for those found to carry a previously unseen mutation.

Cummings, B. B. et al. Improving genetic diagnosis in Mendelian disease with transcriptome sequencing. Sci. Transl. Med . 9 , eaal5209 (2017).

Article   PubMed   PubMed Central   Google Scholar  

Butcher, D. T. et al. CHARGE and Kabuki syndromes: gene-specific DNA methylation signatures identify epigenetic mechanisms linking these clinically overlapping conditions. Am. J. Hum. Genet . 100 , 773–788 (2017).

Short, P. J. et al. De novo mutations in regulatory elements in neurodevelopmental disorders. Nature 555 , 611–616 (2018).

Jaganathan, K. et al. Predicting splicing from primary sequence with deep learning. Cell 176 , 535–548 (2019).

Hugot, J. P. et al. Association of NOD2 leucine-rich repeat variants with susceptibility to Crohn’s disease. Nature 411 , 599–603 (2001).

Ogura, Y. et al. A frameshift mutation in NOD2 associated with susceptibility to Crohn’s disease. Nature 411 , 603–606 (2001).

Barbosa, J., Chern, M. M., Noreen, H. & Anderson, V. E. Analysis of linkage between the major histocompatibility system and juvenile, insulin-dependent diabetes in multiplex families. Reanalysis of data. J. Clin. Invest . 62 , 492–495 (1978).

Risch, N. & Merikangas, K. The future of genetic studies of complex human diseases. Science 273 , 1516–1517 (1996). A highly influential commentary that switched much of the complex trait genetics field from linkage to association and imagined GWAS several years before the technology made them a reality.

Klein, R. J. et al. Complement factor H polymorphism in age-related macular degeneration. Science 308 , 385–389 (2005). Arguably the first full GWAS, this study demonstrated the potential of agnostic genomic surveys to highlight entirely novel biology—in this case, a role for complement in the pathogenesis of macular degeneration.

Duerr, R. H. et al. A genome-wide association study identifies IL23R as an inflammatory bowel disease gene. Science 314 , 1461–1463 (2006).

Wellcome Trust Case Control Consortium. Genome-wide association study of 14,000 cases of seven common diseases and 3,000 shared controls. Nature 447 , 661–678 (2007). This paper, which describes what was, at the time, the largest GWAS yet conducted, demonstrated the broad applicability of the approach, and set the scene for the decade of GWAS discovery that followed.

Mahajan, A. et al. Fine-mapping type 2 diabetes loci to single-variant resolution using high-density imputation and islet-specific epigenome maps. Nat. Genet . 50 , 1505–1513 (2018).

Astle, W. J. et al. The allelic landscape of human blood cell trait variation and links to common complex disease. Cell 167 , 1415–1429 (2016).

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

Buniello, A. et al. The NHGRI–EBI GWAS catalog of published genome-wide association studies, targeted arrays and summary statistics 2019. Nucleic Acids Res . 47 , D1005–D1012 (2019).

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

UK10K Consortium. The UK10K project identifies rare variants in health and disease. Nature 526 , 82–90 (2015).

Manolio, T. A. et al. Finding the missing heritability of complex diseases. Nature 461 , 747–753 (2009).

Wainschtein, P. et al. Recovery of trait heritability from whole genome sequence data. Preprint at bioRxiv https://doi.org/10.1101/588020 (2019).

Flannick, J. et al. Loss-of-function mutations in SLC30A8 protect against type 2 diabetes. Nat. Genet . 46 , 357–363 (2014).

Moltke, I. et al. A common Greenlandic TBC1D4 variant confers muscle insulin resistance and type 2 diabetes. Nature 512 , 190–193 (2014).

Nioi, P. et al. Variant ASGR1 associated with a reduced risk of coronary artery disease. N. Engl. J. Med . 374 , 2131–2141 (2016).

Pollin, T. I. et al. A null mutation in human APOC3 confers a favorable plasma lipid profile and apparent cardioprotection. Science 322 , 1702–1705 (2008).

Narasimhan, V. M. et al. Health and population effects of rare gene knockouts in adult humans with related parents. Science 352 , 474–477 (2016). A powerful illustration of the insights that can be gained from the study of individuals who are homozygous carriers of null alleles in genes of biomedical interest.

Saleheen, D. et al. Human knockouts and phenotypic analysis in a cohort with a high rate of consanguinity. Nature 544 , 235–239 (2017).

Martin, A. R. et al. Clinical use of current polygenic risk scores may exacerbate health disparities. Nat. Genet . 51 , 584–591 (2019). An elegant distillation of the ways in which the historical focus of genetic discovery among populations of European descent has adverse consequences for both discovery and translation.

Zeng, J. et al. Signatures of negative selection in the genetic architecture of human complex traits. Nat. Genet . 50 , 746–753 (2018).

Genovese, G. et al. Association of trypanolytic ApoL1 variants with kidney disease in African Americans. Science 329 , 841–845 (2010).

Luzzatto, L., Usanga, F. A. & Reddy, S. Glucose-6-phosphate dehydrogenase deficient red cells: resistance to infection by malarial parasites. Science 164 , 839–842 (1969).

Wojcik, G. L. et al. Genetic analyses of diverse populations improves discovery for complex traits. Nature 570 , 514–518 (2019).

Article   PubMed   CAS   PubMed Central   Google Scholar  

Farh, K. K. et al. Genetic and epigenetic fine mapping of causal autoimmune disease variants. Nature 518 , 337–343 (2015). An early demonstration of the value of overlapping genome-wide association data with tissue-specific regulatory maps to define the cell-types and tissues likely to be driving disease pathology.

Freund, M. K. et al. Phenotype-specific enrichment of Mendelian disorder genes near GWAS regions across 62 complex traits. Am. J. Hum. Genet . 103 , 535–552 (2018).

Boyle, E. A., Li, Y. I. & Pritchard, J. K. An expanded view of complex traits: from polygenic to omnigenic. Cell 169 , 1177–1186 (2017).

de Lange, K. M. et al. Genome-wide association study implicates immune activation of multiple integrin genes in inflammatory bowel disease. Nat. Genet . 49 , 256–261 (2017).

Sekar, A. et al. Schizophrenia risk from complex variation of complement component 4. Nature 530 , 177–183 (2016).

Jostins, L. et al. Host–microbe interactions have shaped the genetic architecture of inflammatory bowel disease. Nature 491 , 119–124 (2012).

Richardson, T. G., Harrison, S., Hemani, G. & Davey Smith, G. An atlas of polygenic risk score associations to highlight putative causal relationships across the human phenome. eLife 8 , e43657 (2019).

Dendrou, C. A. et al. Resolving TYK2 locus genotype-to-phenotype differences in autoimmunity. Sci. Transl. Med . 8 , 363ra149 (2016).

Voight, B. F. et al. Plasma HDL cholesterol and risk of myocardial infarction: a Mendelian randomisation study. Lancet 380 , 572–580 (2012).

Khera, A. V. et al. Genome-wide polygenic scores for common diseases identify individuals with risk equivalent to monogenic mutations. Nat. Genet . 50 , 1219–1224 (2018). This paper used large-scale genetic data to demonstrate the clinical potential of the polygenic scores that can be constructed for many common diseases, emphasizing that, in some situations, the lifetime risk of disease for those with the highest scores approaches that of established monogenic disease.

Mahmood, S. S., Levy, D., Vasan, R. S. & Wang, T. J. The Framingham Heart Study and the epidemiology of cardiovascular disease: a historical perspective. Lancet 383 , 999–1008 (2014).

Article   PubMed   Google Scholar  

Gudbjartsson, D. F. et al. Variants conferring risk of atrial fibrillation on chromosome 4q25. Nature 448 , 353–357 (2007).

Thorgeirsson, T. E. et al. A variant associated with nicotine dependence, lung cancer and peripheral arterial disease. Nature 452 , 638–642 (2008).

Collins, F. S. & Varmus, H. A new initiative on precision medicine. N. Engl. J. Med . 372 , 793–795 (2015).

Bycroft, C. et al. The UK Biobank resource with deep phenotyping and genomic data. Nature 562 , 203–209 (2018).

Nagai, A. et al. Overview of the BioBank Japan Project: study design and profile. J. Epidemiol . 27 , S2–S8 (2017).

Chen, Z. et al. China Kadoorie Biobank of 0.5 million people: survey methods, baseline characteristics and long-term follow-up. Int. J. Epidemiol . 40 , 1652–1666 (2011).

Kohane, I. S. Using electronic health records to drive discovery in disease genomics. Nat. Rev. Genet . 12 , 417–428 (2011).

Abul-Husn, N. S. & Kenny, E. E. Personalized medicine and the power of electronic health records. Cell 177 , 58–69 (2019).

Sun, B. B. et al. Genomic atlas of the human plasma proteome. Nature 558 , 73–79 (2018).

Shin, S. Y. et al. An atlas of genetic influences on human blood metabolites. Nat. Genet . 46 , 543–550 (2014).

Sung, Y. J. et al. A multi-ancestry genome-wide study incorporating gene-smoking interactions identifies multiple new loci for pulse pressure and mean arterial pressure. Hum. Mol. Genet . 28 , ddz070 (2019).

Pendergrass, S. A. & Crawford, D. C. Using electronic health records to generate phenotypes for research. Curr. Protoc. Hum. Genet . 100 , e80 (2019).

Denny, J. C. et al. PheWAS: demonstrating the feasibility of a phenome-wide scan to discover gene–disease associations. Bioinformatics 26 , 1205–1210 (2010).

Liu, Z. et al. Mendelian randomization analysis dissects the relationship between NAFLD, T2D and obesity and provides implications for precision medicine. Preprint at bioRxiv https://doi.org/10.1101/657734 (2019).

Cotsapas, C. et al. Pervasive sharing of genetic effects in autoimmune disease. PLoS Genet . 7 , e1002254 (2011).

Kanai, M. et al. Genetic analysis of quantitative traits in the Japanese population links cell types to complex human diseases. Nat. Genet . 50 , 390–400 (2018).

Gudmundsson, J. et al. Two variants on chromosome 17 confer prostate cancer risk, and the one in TCF2 protects against type 2 diabetes. Nat. Genet . 39 , 977–983 (2007).

Abul-Husn, N. S. et al. A protein-truncating HSD17B13 variant and protection from chronic liver disease. N. Engl. J. Med . 378 , 1096–1106 (2018).

Plenge, R. M., Scolnick, E. M. & Altshuler, D. Validating therapeutic targets through human genetics. Nat. Rev. Drug Discov . 12 , 581–594 (2013).

Nielsen, J. B. et al. Biobank-driven genomic discovery yields new insight into atrial fibrillation biology. Nat. Genet . 50 , 1234–1239 (2018).

Flannick, J. et al. Exome sequencing of 20,791 cases of type 2 diabetes and 24,440 controls. Nature 570 , 71–76 (2019).

Van Hout, C. V. et al. Whole exome sequencing and characterization of coding variation in 49,960 individuals in the UK Biobank. Preprint at bioRxiv https://doi.org/10.1101/572347 (2019).

Khera, A. V. et al. Whole-genome sequencing to characterize monogenic and polygenic contributions in patients hospitalized with early-onset myocardial infarction. Circulation 139 , 1593–1602 (2019).

Grarup, N. et al. Loss-of-function variants in ADCY3 increase risk of obesity and type 2 diabetes. Nat. Genet . 50 , 172–174 (2018).

Cohen, J. C., Boerwinkle, E., Mosley, T. H. Jr & Hobbs, H. H. Sequence variations in PCSK9 , low LDL, and protection against coronary heart disease. N. Engl. J. Med . 354 , 1264–1272 (2006).

Rivas, M. A. et al. Deep resequencing of GWAS loci identifies independent rare variants associated with inflammatory bowel disease. Nat. Genet . 43 , 1066–1073 (2011).

Chong, J. X. et al. The genetic basis of Mendelian phenotypes: discoveries, challenges, and opportunities. Am. J. Hum. Genet . 97 , 199–215 (2015).

Goldstein, J. L. & Brown, M. S. A century of cholesterol and coronaries: from plaques to genes to statins. Cell 161 , 161–172 (2015).

Peltonen, L., Perola, M., Naukkarinen, J. & Palotie, A. Lessons from studying monogenic disease for common disease. Hum. Mol. Genet . 15 , R67–R74 (2006).

Bulik-Sullivan, B. et al. An atlas of genetic correlations across human diseases and traits. Nat. Genet . 47 , 1236–1241 (2015).

Musunuru, K. et al. From noncoding variant to phenotype via SORT1 at the 1p13 cholesterol locus. Nature 466 , 714–719 (2010). Widely accepted as the first paper to describe the detailed mechanistic dissection of a complex trait locus and still one of relatively few examples of success in this endeavour.

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

Smemo, S. et al. Obesity-associated variants within FTO form long-range functional connections with IRX3 . Nature 507 , 371–375 (2014).

Gupta, R. M. et al. A genetic variant associated with five vascular diseases is a distal regulator of endothelin-1 gene expression. Cell 170 , 522–533 (2017).

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).

Lin, Y. T. et al. APOE4 causes widespread molecular and cellular alterations associated with Alzheimer’s disease phenotypes in human iPSC-derived brain cell types. Neuron 98 , 1294 (2018).

ENCODE Project Consortium. An integrated encyclopedia of DNA elements in the human genome. Nature 489 , 57–74 (2012).

Roadmap Epigenomics Consortium et al. Integrative analysis of 111 reference human epigenomes. Nature 518 , 317–330 (2015). The flagship paper from the NIH Roadmap Epigenomics Consortium, describing a detailed analysis of the epigenomic profiles of more than 100 human cell types and characterizing how these regulatory patterns relate to gene regulation, cellular differentiation and human disease.

Article   PubMed Central   CAS   Google Scholar  

Tansey, K. E., Cameron, D. & Hill, M. J. Genetic risk for Alzheimer’s disease is concentrated in specific macrophage and microglial transcriptional networks. Genome Med . 10 , 14 (2018).

Gamazon, E. R. et al. Using an atlas of gene regulation across 44 human tissues to inform complex disease- and trait-associated variation. Nat. Genet . 50 , 956–967 (2018).

Ongen, H. et al. Estimating the causal tissues for complex traits and diseases. Nat. Genet . 49 , 1676–1683 (2017).

GTEx Consortium. Genetic effects on gene expression across human tissues. Nature 550 , 204–213 (2017). One of several papers describing the results obtained from the Genotype-Tissue Expression (GTEx) project, which used RNA-sequence data from more than 40 human tissues in several hundred individuals to explore the relationship between DNA sequence variation and tissue-specific expression.

Article   PubMed Central   Google Scholar  

Chiou, J. et al. Single cell chromatin accessibility reveals pancreatic islet cell type- and state-specific regulatory programs of diabetes risk. Preprint at bioRxiv https://doi.org/10.1101/693671 (2019).

HuBMAP Consortium. The human body at cellular resolution: the NIH Human Biomolecular Atlas Program. Nature 574 , 187–192 (2019).

Regev, A. et al. The Human Cell Atlas. eLife 6 , e27041 (2017).

Lloyd, K. C. A knockout mouse resource for the biomedical research community. Ann. NY Acad. Sci . 1245 , 24–26 (2011).

Article   ADS   PubMed   Google Scholar  

Finucane, H. K. et al. Heritability enrichment of specifically expressed genes identifies disease-relevant tissues and cell types. Nat. Genet . 50 , 621–629 (2018).

Delaneau, O. et al. Chromatin three-dimensional interactions mediate genetic effects on gene expression. Science 364 , eaat8266 (2019).

Zhou, J. & Troyanskaya, O. G. Predicting effects of noncoding variants with deep learning-based sequence model. Nat. Methods 12 , 931–934 (2015).

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).

Patwardhan, R. P. et al. High-resolution analysis of DNA regulatory elements by synthetic saturation mutagenesis. Nat. Biotechnol . 27 , 1173–1175 (2009).

Melnikov, A. et al. Systematic dissection and optimization of inducible enhancers in human cells using a massively parallel reporter assay. Nat. Biotechnol . 30 , 271–277 (2012).

Wang, X. et al. High-resolution genome-wide functional dissection of transcriptional regulatory regions and nucleotides in human. Nat. Commun . 9 , 5380 (2018).

Potting, C. et al. Genome-wide CRISPR screen for PARKIN regulators reveals transcriptional repression as a determinant of mitophagy. Proc. Natl Acad. Sci. USA 115 , E180–E189 (2018).

Gasperini, M. et al. A genome-wide framework for mapping gene regulation via cellular genetic screens. Cell 176 , 377–390 (2019).

Fulco, C. P. et al. Activity-by-contact model of enhancer-promoter regulation from thousands of CRISPR perturbations. Nat. Genet . 51 , 1664–1669 (2019).

Komor, A. C., Kim, Y. B., Packer, M. S., Zuris, J. A. & Liu, D. R. Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage. Nature 533 , 420–424 (2016).

Feldman, D. et al. Optical pooled screens in human cells. Cell 179 , 787–799 (2019).

Dixit, A. et al. Perturb-Seq: dissecting molecular circuits with scalable single-cell RNA profiling of pooled genetic screens. Cell 167 , 1853–1866 (2016).

Datlinger, P. et al. Pooled CRISPR screening with single-cell transcriptome readout. Nat. Methods 14 , 297–301 (2017).

Ding, L. et al. Perspective on oncogenic processes at the end of the beginning of cancer genomics. Cell 173 , 305–320 (2018).

Chae, Y. K. et al. Path toward precision oncology: review of targeted therapy studies and tools to aid in defining “actionability” of a molecular lesion and patient management support. Mol. Cancer Ther . 16 , 2645–2655 (2017).

Shearer, A. E. & Smith, R. J. Massively parallel sequencing for genetic diagnosis of hearing loss: the new standard of care. Otolaryngol. Head Neck Surg . 153 , 175–182 (2015).

Yang, Y. et al. Molecular findings among patients referred for clinical whole-exome sequencing. J. Am. Med. Assoc . 312 , 1870–1879 (2014).

Friedman, J. M. et al. Genome-wide sequencing in acutely ill infants: genomic medicine’s critical application? Genet. Med . 21 , 498–504 (2019).

Clark, M. M. et al. Diagnosis of genetic diseases in seriously ill children by rapid whole-genome sequencing and automated phenotyping and interpretation. Sci. Transl. Med . 11 , eaat6177 (2019).

Eldomery, M. K. et al. Lessons learned from additional research analyses of unsolved clinical exome cases. Genome Med . 9 , 26 (2017).

Boycott, K. M. et al. A diagnosis for all rare genetic diseases: the horizon and the next frontiers. Cell 177 , 32–37 (2019).

Liu, S. et al. Genomic analyses from non-invasive prenatal testing reveal genetic associations, patterns of viral infections, and Chinese population history. Cell 175 , 347–359 (2018).

Thung, D. T., Beulen, L., Hehir-Kwa, J. & Faas, B. H. Implementation of whole genome massively parallel sequencing for noninvasive prenatal testing in laboratories. Expert Rev. Mol. Diagn . 15 , 111–124 (2015).

Lo, J. O., Cori D, F., Norton, M. E. & Caughey, A. B. Noninvasive prenatal testing. Obstet. Gynecol. Surv . 69 , 89–99 (2014).

Watson, M. S. et al. Newborn screening: toward a uniform screening panel and system—executive summary. Pediatrics 117 , S296–S307 (2006).

Article   Google Scholar  

Currier, R. J. et al. Genomic sequencing in cystic fibrosis newborn screening: what works best, two-tier predefined CFTR mutation panels or second-tier CFTR panel followed by third-tier sequencing? Genet. Med . 19 , 1159–1163 (2017).

Bauer, P. E. “Tell them it’s not so bad”: prenatal screening for Down syndrome and the bias toward abortion. Intellect. Dev. Disabil . 46 , 247–251 (2008).

Volpi, S. et al. Research directions in the clinical implementation of pharmacogenomics: an overview of US Programs and Projects. Clin. Pharmacol. Ther . 103 , 778–786 (2018).

International Schizophrenia Consortium. Common polygenic variation contributes to risk of schizophrenia and bipolar disorder. Nature 460 , 748–752 (2009).

Evans, D. M., Visscher, P. M. & Wray, N. R. Harnessing the information contained within genome-wide association studies to improve individual prediction of complex disease risk. Hum. Mol. Genet . 18 , 3525–3531 (2009).

Kurian, A. W. et al. Genetic testing and results in a population-based cohort of breast cancer patients and ovarian cancer patients. J. Clin. Oncol . 37 , 1305–1315 (2019).

Katz, S. J. et al. Association of attending surgeon with variation in the receipt of genetic testing after diagnosis of breast cancer. JAMA Surg . 153 , 909–916 (2018).

Kalokairinou, L. et al. Legislation of direct-to-consumer genetic testing in Europe: a fragmented regulatory landscape. J. Community Genet . 9 , 117–132 (2018).

Allyse, M. A., Robinson, D. H., Ferber, M. J. & Sharp, R. R. Direct-to-Consumer Testing 2.0: emerging models of direct-to-consumer genetic testing. Mayo Clin. Proc . 93 , 113–120 (2018).

Wright, C. F. et al. Assessing the pathogenicity, penetrance, and expressivity of putative disease-causing variants in a population setting. Am. J. Hum. Genet . 104 , 275–286 (2019).

ACMG Board of Directors. The use of ACMG secondary findings recommendations for general population screening: a policy statement of the American College of Medical Genetics and Genomics (ACMG). Genet. Med . 21 , 1467–1468 (2019).

George, A., Kaye, S. & Banerjee, S. Delivering widespread BRCA testing and PARP inhibition to patients with ovarian cancer. Nat. Rev. Clin. Oncol . 14 , 284–296 (2017).

CenterWatch. FDA Approved Drugs for Genetic Disease (CenterWatch); https://www.centerwatch.com/drug-information/fda-approved-drugs/therapeutic-area/34/genetic-disease

Wu, Y. et al. Highly efficient therapeutic gene editing of human hematopoietic stem cells. Nat. Med . 25 , 776–783 (2019).

Jaiswal, S. et al. Clonal hematopoiesis and risk of atherosclerotic cardiovascular disease. N. Engl. J. Med . 377 , 111–121 (2017).

Lee, M. H. et al. Somatic APP gene recombination in Alzheimer’s disease and normal neurons. Nature 563 , 639–645 (2018).

H3Africa Consortium. Enabling the genomic revolution in Africa. Science 344 , 1346–1348 (2014).

Dolman, L. et al. ClinGen advancing genomic data-sharing standards as a GA4GH driver project. Hum. Mutat . 39 , 1686–1689 (2018).

Fairfax, B. P. et al. Innate immune activity conditions the effect of regulatory variants upon monocyte gene expression. Science 343 , 1246949 (2014).

Nelson, M. R. et al. The support of human genetic evidence for approved drug indications. Nat. Genet . 47 , 856–860 (2015). An influential paper that demonstrates that the probability of therapeutic success is enhanced (by around a factor of 2) for drug targets with supportive evidence from human genetics.

Veitch, D. P. et al. Understanding disease progression and improving Alzheimer’s disease clinical trials: recent highlights from the Alzheimer’s Disease Neuroimaging Initiative. Alzheimers Dement . 15 , 106–152 (2019).

Ashley, E. A. Towards precision medicine. Nat. Rev. Genet . 17 , 507–522 (2016).

Franks, P. W. & McCarthy, M. I. Exposing the exposures responsible for type 2 diabetes and obesity. Science 354 , 69–73 (2016).

Sanna, S. et al. Causal relationships among the gut microbiome, short-chain fatty acids and metabolic diseases. Nat. Genet . 51 , 600–605 (2019).

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).

Rose, G. Sick individuals and sick populations. Int. J. Epidemiol . 14 , 32–38 (1985).

Hindorff, L. A. et al. Prioritizing diversity in human genomics research. Nat. Rev. Genet . 19 , 175–185 (2018).

Wittrup, A. & Lieberman, J. Knocking down disease: a progress report on siRNA therapeutics. Nat. Rev. Genet . 16 , 543–552 (2015).

Schuster, S. J. et al. Chimeric antigen receptor T cells in refractory B-cell lymphomas. N. Engl. J. Med . 377 , 2545–2554 (2017).

Ding, Q. et al. Permanent alteration of PCSK9 with in vivo CRISPR–Cas9 genome editing. Circ. Res . 115 , 488–492 (2014).

Risca, V. I. & Greenleaf, W. J. Unraveling the 3D genome: genomics tools for multiscale exploration. Trends Genet . 31 , 357–372 (2015).

Download references

Acknowledgements

We acknowledge grant funding from the following funders. J.H.C.: NIH (U01DK062429, U01DK062422 and R01DK106593); N.J.C.: NIH (U01HG009086, U54MD010722 and R01MH113362); E.T.D.: Swiss National Science Foundation, Louis Jeantet Foundation; E.E.K.: NIH (R01HG010297, R01HL104608, U01HG009610, R01DK110113, U01HG009080, X01HL1345 and UM1HG0089001); C.M.L.: NIHR Oxford Biomedical Research Centre and NIH (5P50HD028138-27); K.N.N.: NHMRC (APP1113531); S.E.P.: NIH (U01HG006485 and 1U41HG009649); C.N.R.: NIH Intramural Program at the Center for Research on Genomics and Global Health and National Human Genome Research Institute; and M.I.M.: Wellcome (090532, 098381, 106130, 203141 and 212259) and NIH (U01-DK105535, DK085545 and DK098032). Personal funding comes from the following sources. M.C.: Next Generation Fund at the Broad Institute of MIT and Harvard; J.H.C.: Sanford Grossman Charitable Trust and Helmsley Charitable Trust; R.C.: UKBiobank; C.M.L.: Li Ka Shing Foundation; J.S.: Howard Hughes Medical Institute; and M.I.M.: Wellcome and NIHR.

Author information

Authors and affiliations.

Beth Israel Deaconess Medical Center, Harvard Medical School, Boston, MA, USA

Melina Claussnitzer

Broad Institute of MIT and Harvard Cambridge, Cambridge, MA, USA

Melina Claussnitzer, Sekar Kathiresan, Cecilia M. Lindgren, Daniel G. MacArthur & Heidi L. Rehm

Institute of Nutritional Science, University of Hohenheim, Stuttgart, Germany

Charles Bronfman Institute for Personalized Medicine, Icahn School of Medicine at Mount Sinai, New York, NY, USA

Judy H. Cho & Eimear E. Kenny

Department of Medicine, Icahn School of Medicine at Mount Sinai, New York, NY, USA

Judy H. Cho

Department of Genetics and Genomic Sciences, Icahn School of Medicine at Mount Sinai, New York, NY, USA

Nuffield Department of Population Health (NDPH), University of Oxford, Oxford, UK

Rory Collins

UK Biobank, Stockport, UK

Vanderbilt Genetics Institute and Division of Genetic Medicine, Department of Medicine, Vanderbilt University Medical Center, Nashville, TN, USA

Nancy J. Cox

Department of Genetic Medicine and Development, University of Geneva Medical School, Geneva, Switzerland

Emmanouil T. Dermitzakis

Health 2030 Genome Center, Geneva, Switzerland

Wellcome Sanger Institute, Hinxton, UK

Matthew E. Hurles & Nicole Soranzo

Center for Genomic Medicine, Massachusetts General Hospital, Boston, MA, USA

Sekar Kathiresan, Daniel G. MacArthur & Heidi L. Rehm

Verve Therapeutics, Cambridge, MA, USA

Sekar Kathiresan

Center for Genomic Health, Icahn School of Medicine at Mount Sinai, New York, NY, USA

Eimear E. Kenny

Big Data Institute at the Li Ka Shing Centre for Health Information and Discovery, University of Oxford, Oxford, UK

Cecilia M. Lindgren

Wellcome Centre for Human Genetics, Nuffield Department of Medicine, University of Oxford, Oxford, UK

Cecilia M. Lindgren & Mark I. McCarthy

Analytic and Translational Genetics Unit, Massachusetts General Hospital, Boston, MA, USA

Daniel G. MacArthur & Heidi L. Rehm

Murdoch Children’s Research Institute, Parkville, Victoria, Australia

Kathryn N. North

University of Melbourne, Parkville, Victoria, Australia

Departments of Pediatrics and Molecular and Human Genetics, Baylor College of Medicine, Houston, TX, USA

Sharon E. Plon

Texas Children’s Cancer Center, Texas Children’s Hospital, Houston, TX, USA

Department of Pathology, Harvard Medical School, Boston, MA, USA

Heidi L. Rehm

Institute for Human Genetics, University of California San Francisco, San Francisco, CA, USA

Center for Research on Genomics and Global Health, National Human Genome Research Institute, Bethesda, MD, USA

Charles N. Rotimi

Department of Genome Sciences, University of Washington, Seattle, WA, USA

Jay Shendure

Brotman Baty Institute for Precision Medicine, Magnuson Health Sciences Building, Seattle, WA, USA

Howard Hughes Medical Institute, Seattle, WA, USA

Department of Haematology, University of Cambridge, Cambridge, UK

Nicole Soranzo

Oxford Centre for Diabetes, Endocrinology and Metabolism, Oxford, UK

Mark I. McCarthy

Oxford NIHR Biomedical Research Centre, Oxford University Hospitals NHS Foundation Trust, John Radcliffe Hospital, Oxford, UK

Human Genetics, Genentech, South San Francisco, CA, USA

You can also search for this author in PubMed   Google Scholar

Contributions

M.C. and M.I.M. coordinated the drafting of the Review. All other authors contributed to drafting of the text and approved the final version.

Corresponding author

Correspondence to Mark I. McCarthy .

Ethics declarations

Competing interests.

R.C. has received research grants from British Heart Foundation, Cancer Research UK, Medical Research Council, Merck & Co, UKBiobank, Wellcome and Medco; a Pfizer Prize Award (to NDPH) and is named on a patent for a statin-related myopathy genetic test. R.C. receives no personal remuneration from these: any share of royalties or other payments have been waived in favour of NDPH. E.T.D. is chairman and board member of Hybridstat and on the advisory board of DNAnexus. M.E.H. is a co-founder, shareholder and director of Congenica. S.K. has received research grants from Bayer and Novartis; is on Scientific Advisory Boards of Regeneron Genetics Center, Corvidia Therapeutics and Maze Therapeutics; has equity in San Therapeutics, Catabasis, Verve and Maze Therapeutics and is a consultant for Maze Therapeutics, Alynlam, ExpertConnect, Leerink Partners, Noble Insights, Bayer and Novo Ventures. E.E.K. receives honoraria from Illumina and Regeneron Pharmaceuticals. C.M.L. has research collaborations with Novo Nordisk and Bayer, receiving no personal payment. D.G.M. is co-founder and shareholder of Goldfinch Bio and has received research funding from AbbVie, Biogen, BioMarin, Merck, Pfizer and Sanofi-Genzyme. S.E.P. is a member of the Scientific Advisory Board of Baylor Genetics. J.S. is a member of the Scientific Advisory Board for Maze Therapeutics, Camp4 Therapeutics, Nanostring, Phase Genomics, Adaptive Biotechnology and Stratos Genomics, a founder of Phase Genomics and a consultant for Guardant Health. M.I.M. was a member of advisory panels for Pfizer, NovoNordisk and Zoe Global; received honoraria from Merck, Pfizer, NovoNordisk and Eli Lilly and research funding from Abbvie, Astra Zeneca, Boehringer Ingelheim, Eli Lilly, Janssen, Merck, NovoNordisk, Pfizer, Roche, Sanofi Aventis, Servier and Takeda. As of June 2019, M.I.M. is an employee of Genentech and a holder of Roche stock. The views expressed in this article are those of the author(s) and not necessarily those of the NHS, the NIHR, or the Department of Health.

Additional information

Peer review information Nature thanks Ewan Birney, Jeff Schloss, Richard Trembath and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Rights and permissions

Reprints and permissions

About this article

Cite this article.

Claussnitzer, M., Cho, J.H., Collins, R. et al. A brief history of human disease genetics. Nature 577 , 179–189 (2020). https://doi.org/10.1038/s41586-019-1879-7

Download citation

Received : 16 July 2019

Accepted : 13 November 2019

Published : 08 January 2020

Issue Date : 09 January 2020

DOI : https://doi.org/10.1038/s41586-019-1879-7

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Genetic effects of sequence-conserved enhancer-like elements on human complex traits.

  • Wing Hung Wong

Genome Biology (2024)

The interplay of ethics and genetic technologies in balancing the social valuation of the human genome in UNESCO declarations

  • Hristina Gaydarska
  • Kayo Takashima
  • Jusaku Minari

European Journal of Human Genetics (2024)

Five steps to connect genetic risk variants to disease

Nature (2024)

Improved phenotypic classification of male infertility to promote discovery of genetic causes

  • Margot J. Wyrwoll
  • Godfried W. van der Heijden
  • Frank Tüttelmann

Nature Reviews Urology (2024)

Variational autoencoder for design of synthetic viral vector serotypes

  • Shahin Sowlati-Hashjin
  • Michael Garton

Nature Machine Intelligence (2024)

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.

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

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

gene disorder research paper

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed.

  • Search in PubMed
  • Search in NLM Catalog
  • Add to Search

The genetics of bipolar disorder

Affiliations.

  • 1 Human Genetics Branch, National Institute of Mental Health Intramural Research Program, Department of Health and Human Services, National Institutes of Health, Bethesda, MD, USA.
  • 2 College of Medicine, University of the Philippines Manila, 1000, Ermita, Manila, Philippines.
  • 3 Human Genetics Branch, National Institute of Mental Health Intramural Research Program, Department of Health and Human Services, National Institutes of Health, Bethesda, MD, USA. [email protected].
  • PMID: 31907381
  • DOI: 10.1038/s41380-019-0634-7

Bipolar disorder (BD) is one of the most heritable mental illnesses, but the elucidation of its genetic basis has proven to be a very challenging endeavor. Genome-Wide Association Studies (GWAS) have transformed our understanding of BD, providing the first reproducible evidence of specific genetic markers and a highly polygenic architecture that overlaps with that of schizophrenia, major depression, and other disorders. Individual GWAS markers appear to confer little risk, but common variants together account for about 25% of the heritability of BD. A few higher-risk associations have also been identified, such as a rare copy number variant on chromosome 16p11.2. Large scale next-generation sequencing studies are actively searching for other alleles that confer substantial risk. As our understanding of the genetics of BD improves, there is growing optimism that some clear biological pathways will emerge, providing a basis for future studies aimed at molecular diagnosis and novel therapeutics.

PubMed Disclaimer

Similar articles

  • Contribution of Rare Copy Number Variants to Bipolar Disorder Risk Is Limited to Schizoaffective Cases. Charney AW, Stahl EA, Green EK, Chen CY, Moran JL, Chambert K, Belliveau RA Jr, Forty L, Gordon-Smith K, Lee PH, Bromet EJ, Buckley PF, Escamilla MA, Fanous AH, Fochtmann LJ, Lehrer DS, Malaspina D, Marder SR, Morley CP, Nicolini H, Perkins DO, Rakofsky JJ, Rapaport MH, Medeiros H, Sobell JL, Backlund L, Bergen SE, Juréus A, Schalling M, Lichtenstein P, Knowles JA, Burdick KE, Jones I, Jones LA, Hultman CM, Perlis R, Purcell SM, McCarroll SA, Pato CN, Pato MT, Di Florio A, Craddock N, Landén M, Smoller JW, Ruderfer DM, Sklar P. Charney AW, et al. Biol Psychiatry. 2019 Jul 15;86(2):110-119. doi: 10.1016/j.biopsych.2018.12.009. Epub 2018 Dec 20. Biol Psychiatry. 2019. PMID: 30686506 Free PMC article.
  • Association Between Schizophrenia-Related Polygenic Liability and the Occurrence and Level of Mood-Incongruent Psychotic Symptoms in Bipolar Disorder. Allardyce J, Leonenko G, Hamshere M, Pardiñas AF, Forty L, Knott S, Gordon-Smith K, Porteous DJ, Haywood C, Di Florio A, Jones L, McIntosh AM, Owen MJ, Holmans P, Walters JTR, Craddock N, Jones I, O'Donovan MC, Escott-Price V. Allardyce J, et al. JAMA Psychiatry. 2018 Jan 1;75(1):28-35. doi: 10.1001/jamapsychiatry.2017.3485. JAMA Psychiatry. 2018. PMID: 29167880 Free PMC article.
  • After GWAS: searching for genetic risk for schizophrenia and bipolar disorder. Gershon ES, Alliey-Rodriguez N, Liu C. Gershon ES, et al. Am J Psychiatry. 2011 Mar;168(3):253-6. doi: 10.1176/appi.ajp.2010.10091340. Epub 2011 Feb 1. Am J Psychiatry. 2011. PMID: 21285144 Free PMC article. Review.
  • Cross-disorder genomewide analysis of schizophrenia, bipolar disorder, and depression. Huang J, Perlis RH, Lee PH, Rush AJ, Fava M, Sachs GS, Lieberman J, Hamilton SP, Sullivan P, Sklar P, Purcell S, Smoller JW. Huang J, et al. Am J Psychiatry. 2010 Oct;167(10):1254-63. doi: 10.1176/appi.ajp.2010.09091335. Epub 2010 Aug 16. Am J Psychiatry. 2010. PMID: 20713499 Free PMC article.
  • New technologies provide insights into genetic basis of psychiatric disorders and explain their co-morbidity. Rudan I. Rudan I. Psychiatr Danub. 2010 Jun;22(2):190-2. Psychiatr Danub. 2010. PMID: 20562745 Review.
  • Key Roles of CACNA1C/Cav1.2 and CALB1/Calbindin in Prefrontal Neurons Altered in Cognitive Disorders. Datta D, Yang S, Joyce MKP, Woo E, McCarroll SA, Gonzalez-Burgos G, Perone I, Uchendu S, Ling E, Goldman M, Berretta S, Murray J, Morozov Y, Arellano J, Duque A, Rakic P, O'Dell R, van Dyck CH, Lewis DA, Wang M, Krienen FM, Arnsten AFT. Datta D, et al. JAMA Psychiatry. 2024 May 22:e241112. doi: 10.1001/jamapsychiatry.2024.1112. Online ahead of print. JAMA Psychiatry. 2024. PMID: 38776078 Free PMC article.
  • Heterogeneity analysis provides evidence for a genetically homogeneous subtype of bipolar-disorder. McGrouther CC, Rangan AV, Di Florio A, Elman JA, Schork NJ, Kelsoe J; Bipolar Disorder Working Group of the Psychiatric Genomics Consortium. McGrouther CC, et al. ArXiv [Preprint]. 2024 Apr 30:arXiv:2405.00159v1. ArXiv. 2024. PMID: 38745705 Free PMC article. Preprint.
  • Genetic regulation of human brain proteome reveals proteins implicated in psychiatric disorders. Luo J, Li L, Niu M, Kong D, Jiang Y, Poudel S, Shieh AW, Cheng L, Giase G, Grennan K, White KP, Chen C, Wang SH, Pinto D, Wang Y, Liu C, Peng J, Wang X. Luo J, et al. Mol Psychiatry. 2024 May 9. doi: 10.1038/s41380-024-02576-8. Online ahead of print. Mol Psychiatry. 2024. PMID: 38724566
  • Bipolar Patient-Specific In Vitro Diagnostic Test Reveals Underlying Cardiac Arrhythmia Phenotype Caused by Calcium Channel Genetic Risk Factor. Dow R, DeLong C, Jiang G, Attili D, Creech J, Kraan R, Campbell K, Saraithong P, O'Shea S, Monteiro da Rocha A, McInnis MG, Herron TJ. Dow R, et al. Biol Psychiatry Glob Open Sci. 2024 Feb 20;4(3):100296. doi: 10.1016/j.bpsgos.2024.100296. eCollection 2024 May. Biol Psychiatry Glob Open Sci. 2024. PMID: 38560725 Free PMC article.
  • Associations of altered leukocyte DDR1 promoter methylation and childhood trauma with bipolar disorder and suicidal behavior in euthymic patients. Garcia-Ruiz B, Jiménez E, Aranda S, Verdolini N, Gutiérrez-Zotes A, Sáez C, Losantos E, Alonso-Lana S, Fatjó-Vilas M, Sarró S, Torres L, Panicalli F, Bonnin CDM, Pomarol-Clotet E, Vieta E, Vilella E. Garcia-Ruiz B, et al. Mol Psychiatry. 2024 Mar 19. doi: 10.1038/s41380-024-02522-8. Online ahead of print. Mol Psychiatry. 2024. PMID: 38503928
  • Merikangas KR, Jin R, He J-P, Kessler RC, Lee S, Sampson NA, et al. Prevalence and correlates of bipolar spectrum disorder in the world mental health survey initiative. Arch Gen Psychiatry. 2011;68:241–51. - PubMed - PMC
  • Oxford University Press. Manic-Depressive illness: bipolar disorders and recurrent depression. 2nd ed. Oxford, New York: Oxford University Press; 2007.
  • McMahon FJ, Stine OC, Chase GA, Meyers DA, Simpson SG, DePaulo JRJ. Influence of clinical subtype, sex, and lineality on age at onset of major affective disorder in a family sample. Am J Psychiatry. 1994;151:210–5. - PubMed
  • Perlis RH, Ostacher MJ, Goldberg JF, Miklowitz DJ, Friedman E, Calabrese J, et al. Transition to mania during treatment of bipolar depression. Neuropsychopharmacol. 2010;35:2545–52.
  • Simon NM, Otto MW, Wisniewski SR, Fossey M, Sagduyu K, Frank E, et al. Anxiety disorder comorbidity in bipolar disorder patients: data from the first 500 participants in the Systematic Treatment Enhancement Program for Bipolar Disorder (STEP-BD). Am J Psychiatry. 2004;161:2222–9. - PubMed

Publication types

  • Search in MeSH

Related information

Grants and funding.

  • ZIA-MH002843/U.S. Department of Health & Human Services | NIH | National Institute of Mental Health (NIMH)/International

LinkOut - more resources

Full text sources.

  • Nature Publishing Group
  • Genetic Alliance
  • MedlinePlus Consumer Health Information
  • MedlinePlus Health Information

Research Materials

  • NCI CPTC Antibody Characterization Program

full text provider logo

  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

NIMH Logo

Transforming the understanding and treatment of mental illnesses.

Información en español

Celebrating 75 Years! Learn More >>

  • Research Funded by NIMH

Research Conducted at NIMH (Intramural Research Program)

  • Priority Research Areas
  • Research Resources

Labs landing banner

The Division of Intramural Research Programs (IRP) at the National Institute of Mental Health (NIMH) is the internal research division of the NIMH. The division plans and conducts basic, clinical, and translational research to advance understanding of the diagnosis, causes, treatment, and prevention of psychiatric disorders. NIMH IRP conducts state-of-the-art research that utilizes the unique resources of the National Institutes of Health (NIH), provides an environment conducive to the training and development of clinical and basic scientists, and in part, complements extramural research activities.

Arash Afraz

Investigators

Gree Blue Neurons

Research Groups

Group looking at charts

About the Office of Fellowship Training

Circle and bar chart

Office of the Scientific Director (OSD)

Hands shaking with blue background

Collaborations and Partnerships

Man brain scan

Join a Study

Screenshot from NIMH video

News and Multimedia Featuring IRP

  • Basic Research Powers the First Medication for Postpartum Depression | May 14, 2024 ● Science News
  • Novel Treatment Helps Children With Severe Irritability | April 5, 2024 ● Science News
  • NIH Researchers Identify Brain Connections Associated With ADHD in Youth | March 13, 2024 ● Science News
  • Patient Care & Health Information
  • Diseases & Conditions
  • Neurofibromatosis type 1

Neurofibromatosis type 1 (NF1) is a genetic condition that causes changes in skin pigment and tumors on nerve tissue. Skin changes include flat, light brown spots and freckles in the armpits and groin. Tumors can grow anywhere in the nervous system, including the brain, spinal cord and nerves. NF1 is rare. About 1 in 2,500 is affected by NF1.

The tumors often are not cancerous, known as benign tumors. But sometimes they can become cancerous. Symptoms often are mild. But complications can occur and may include trouble with learning, heart and blood vessel conditions, vision loss, and pain.

Treatment focuses on supporting healthy growth and development in children and early management of complications. If NF1 causes large tumors or tumors that press on a nerve, surgery can reduce symptoms. A newer medicine is available to treat tumors in children, and other new treatments are being developed.

Neurofibromatosis type 1 (NF1) usually is diagnosed during childhood. Symptoms are seen at birth or shortly afterward and almost always by age 10. Symptoms tend to be mild to moderate, but they can vary from person to person.

Symptoms include:

  • Flat, light brown spots on the skin, known as cafe au lait spots. These harmless spots are common in many people. But having more than six cafe au lait spots suggests NF1. They often are present at birth or appear during the first years of life. After childhood, new spots stop appearing.
  • Freckling in the armpits or groin area. Freckling often appears by ages 3 to 5. Freckles are smaller than cafe au lait spots and tend to occur in clusters in skin folds.
  • Tiny bumps on the iris of the eye, known as Lisch nodules. These nodules can't easily be seen and don't affect vision.
  • Soft, pea-sized bumps on or under the skin called neurofibromas. These benign tumors usually grow in or under the skin but can also grow inside the body. A growth that involves many nerves is called a plexiform neurofibroma. Plexiform neurofibromas, when located on the face, can cause disfigurement. Neurofibromas may increase in number with age.
  • Bone changes. Changes in bone development and low bone mineral density can cause bones to form in an irregular way. People with NF1 may have a curved spine, known as scoliosis, or a bowed lower leg.
  • Tumor on the nerve that connects the eye to the brain, called an optic pathway glioma. This tumor usually appears by age 3. The tumor rarely appears in late childhood and among teenagers, and almost never in adults.
  • Learning disabilities. It's common for children with NF1 to have some trouble with learning. Often there is a specific learning disability, such as trouble with reading or math. Attention-deficit/hyperactivity disorder (ADHD) and speech delay also are common.
  • Larger than average head size. Children with NF1 tend to have a larger than average head size due to increased brain volume.
  • Short stature. Children who have NF1 often are below average in height.

When to see a doctor

See a healthcare professional if your child has symptoms of neurofibromatosis type 1. The tumors are often not cancerous and are slow growing, but complications can be managed. If your child has a plexiform neurofibroma, a medicine is available to treat it.

Neurofibromatosis type 1 is caused by an altered gene that either is passed down by a parent or occurs at conception.

The NF1 gene is located on chromosome 17. This gene produces a protein called neurofibromin that helps regulate cell growth. When the gene is altered, it causes a loss of neurofibromin. This allows cells to grow without control.

Risk factors

Autosomal dominant inheritance pattern

Autosomal dominant inheritance pattern

In an autosomal dominant disorder, the changed gene is a dominant gene. It's located on one of the nonsex chromosomes, called autosomes. Only one changed gene is needed for someone to be affected by this type of condition. A person with an autosomal dominant condition — in this example, the father — has a 50% chance of having an affected child with one changed gene and a 50% chance of having an unaffected child.

The biggest risk factor for neurofibromatosis type 1 (NF1) is a family history. For about half of people who have NF1, the disease was passed down from a parent. People who have NF1 and whose relatives aren't affected are likely to have a new change to a gene.

NF1 has an autosomal dominant inheritance pattern. This means that any child of a parent who is affected by the disease has a 50% chance of having the altered gene.

Complications

Complications of neurofibromatosis type 1 (NF1) vary, even within the same family. Generally, complications occur when tumors affect nerve tissue or press on internal organs.

Complications of NF1 include:

  • Neurological symptoms. Trouble with learning and thinking are the most common neurological symptoms associated with NF1. Less common complications include epilepsy and the buildup of excess fluid in the brain.
  • Concerns with appearance. Visible signs of NF1 can include widespread cafe au lait spots, many neurofibromas in the facial area or large neurofibromas. In some people this can cause anxiety and emotional distress, even if they're not medically serious.
  • Skeletal symptoms. Some children have bones that didn't form as usual. This can cause bowing of the legs and fractures that sometimes don't heal. NF1 can cause curvature of the spine, known as scoliosis, that may need bracing or surgery. NF1 also is associated with lower bone mineral density, which increases the risk of weak bones, known as osteoporosis.
  • Changes in vision. Sometimes a tumor called an optic pathway glioma develops on the optic nerve. When this happens, it can affect vision.
  • Increase in symptoms during times of hormonal change. Hormonal changes associated with puberty or pregnancy might cause an increase in neurofibromas. Most people who have NF1 have healthy pregnancies but will likely need monitoring by an obstetrician who is familiar with NF1.
  • Cardiovascular symptoms. People who have NF1 have an increased risk of high blood pressure and may develop blood vessel conditions.
  • Trouble breathing. Rarely, plexiform neurofibromas can put pressure on the airway.
  • Cancer. Some people who have NF1 develop cancerous tumors. These usually arise from neurofibromas under the skin or from plexiform neurofibromas. People who have NF1 also have a higher risk of other forms of cancer. They include breast cancer, leukemia, colorectal cancer, brain tumors and some types of soft tissue cancer. Screening for breast cancer should begin earlier, at age 30, for women with NF1 compared to the general population.
  • Benign adrenal gland tumor, known as a pheochromocytoma. This noncancerous tumor produces hormones that raise your blood pressure. Surgery often is needed to remove it.

Neurofibromatosis type 1 care at Mayo Clinic

  • Ferri FF. Neurofibromatosis. In: Ferri's Clinical Advisor 2024. Elsevier; 2024. https://www.clinicalkey.com. Accessed Feb. 21, 2024.
  • Neurofibromatosis. National Institute of Neurological Disorders and Stroke. https://www.ninds.nih.gov/Disorders/Patient-Caregiver-Education/Fact-Sheets/Neurofibromatosis-Fact-Sheet. Accessed Feb. 21, 2024.
  • Korf BR, et al. Neurofibromatosis type 1 (NF1): Pathogenesis, clinical features, and diagnosis. https://www.uptodate.com/contents/search. Accessed Feb. 21, 2024.
  • Saleh M, et al. Neurofibromatosis type 1 system-based manifestations and treatments: A review. Neurological Sciences. 2023; doi:10.1007/s10072-023-06680-5.
  • Neurofibromatosis. American Association of Neurological Surgeons. https://www.aans.org/en/Patients/Neurosurgical-Conditions-and-Treatments/Neurofibromatosis. Accessed Feb. 21, 2024.
  • Neurofibromatosis. Merck Manual Professional Version. https://www.merckmanuals.com/professional/pediatrics/neurocutaneous-syndromes/neurofibromatosis. Accessed Feb. 21, 2024.
  • Jankovic J, et al., eds. Neurocutaneous syndromes. In: Bradley and Daroff's Neurology in Clinical Practice. 8th ed. Elsevier; 2022. https://www.clinicalkey.com. Accessed Feb. 21, 2024.
  • Armstrong AE, et al. Treatment decisions and the use of the MEK inhibitors for children with neurofibromatosis type 1-related plexiform neurofibromas. BMC Cancer. 2023; doi:10.1186/s12885-023-10996-y.
  • Zitelli BJ, et al., eds. Neurology. In: Zitelli and Davis' Atlas of Pediatric Physical Diagnoses. 8th ed. Elsevier; 2023. https://www.clinicalkey.com. Accessed Feb. 21, 2024.
  • Kellerman RD, et al. Neurofibromatosis (type 1). In: Conn's Current Therapy 2024. Elsevier; 2024. https://www.clinicalkey.com. Accessed Feb. 21, 2024.
  • Babovic-Vuksanovic D (expert opinion). Mayo Clinic. March 26, 2024.
  • Tamura R. Current understanding of neurofibromatosis type 1, 2 and schwannomatosis. International Journal of Molecular Sciences. 2021; doi:10.3390/ijms22115850.
  • Legius E, et al. Revised diagnostic criteria for neurofibromatosis type 1 and Legius syndrome: An international consensus recommendation. Genetics in Medicine. 2021; doi:10.1038/s41436-021-01170-5.
  • Find a doctor. Children's Tumor Foundation. https://www.ctf.org/understanding-nf/find-a-doctor/. Accessed Feb. 26, 2024.
  • Ami TR. Allscripts EPSi. Mayo Clinic. April 18, 2024.

Mayo Clinic in Rochester, Minnesota, Mayo Clinic in Phoenix/Scottsdale, Arizona, and Mayo Clinic in Jacksonville, Florida, have been ranked among the best Neurology & Neurosurgery hospitals in the nation for 2023-2024 by U.S. News & World Report.

  • Symptoms & causes
  • Diagnosis & treatment
  • Doctors & departments
  • Care at Mayo Clinic

Mayo Clinic does not endorse companies or products. Advertising revenue supports our not-for-profit mission.

  • Opportunities

Mayo Clinic Press

Check out these best-sellers and special offers on books and newsletters from Mayo Clinic Press .

  • Mayo Clinic on Incontinence - Mayo Clinic Press Mayo Clinic on Incontinence
  • The Essential Diabetes Book - Mayo Clinic Press The Essential Diabetes Book
  • Mayo Clinic on Hearing and Balance - Mayo Clinic Press Mayo Clinic on Hearing and Balance
  • FREE Mayo Clinic Diet Assessment - Mayo Clinic Press FREE Mayo Clinic Diet Assessment
  • Mayo Clinic Health Letter - FREE book - Mayo Clinic Press Mayo Clinic Health Letter - FREE book

We’re transforming healthcare

Make a gift now and help create new and better solutions for more than 1.3 million patients who turn to Mayo Clinic each year.

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • HHS Author Manuscripts

Logo of nihpa

Genetics of eating disorders in the genome-wide era

Hunna j. watson.

1 Department of Psychiatry, The University of North Carolina at Chapel Hill, Chapel Hill, NC, USA

2 Division of Paediatrics, School of Medicine, The University of Western Australia, Perth, Australia

3 School of Psychology, Curtin University, Perth, Australia

Alish B. Palmos

4 Social, Genetic & Developmental Psychiatry Centre, Institute of Psychiatry, Psychology & Neuroscience, King’s College London, London, United Kingdom

Avina Hunjan

5 National Institute for Health Research (NIHR) Biomedical Research Centre for Mental Health, South London and Maudsley National Health Service (NHS) Trust, London, United Kingdom

Jessica H Baker

Zeynep yilmaz.

6 National Centre for Register-based Research, Aarhus BSS, Aarhus University, Aarhus, Denmark

7 Department of Genetics, The University of North Carolina at Chapel Hill, Chapel Hill, NC, USA

Helena L. Davies

Associated data.

Enabled by advances in high throughput genomic sequencing and an unprecedented level of global data sharing, molecular genetic research is beginning to unlock the biological basis of eating disorders. This invited review provides an overview of genetic discoveries in eating disorders in the genome-wide era. To date, five genome-wide association studies (GWAS) on eating disorders have been conducted—all on anorexia nervosa (AN). For AN, several risk loci have been detected, and ~11–17% of the heritability has been accounted for by common genetic variants. There is extensive genetic overlap between AN and psychological traits, especially obsessive-compulsive disorder, and intriguingly, with metabolic phenotypes even after adjusting for BMI risk variants. Further, genetic risk variants predisposing to lower BMI may be causal risk factors for AN. Causal genes and biological pathways of eating disorders have yet to be elucidated and will require greater sample sizes and statistical power, and functional follow-up studies. Several studies are underway to recruit individuals with bulimia nervosa and binge-eating disorder to enable further genome-wide studies. Data collections and research labs focused on the genetics of eating disorders have joined together in a global effort with the Psychiatric Genomics Consortium. In sum, molecular genetics research in the genome-wide era is improving knowledge about the biology behind the established heritability of eating disorders. This has the potential to offer new hope for understanding eating disorder etiology and for overcoming the therapeutic challenges that confront the eating disorder field.

Introduction

Cross-disciplinary efforts spanning the behavioral sciences, medicine, and genomics are furthering progress toward unlocking the biological basis of eating disorders. These efforts to develop insights into pathophysiology are intended to encompass major translational areas for the personalized care of patients, including screening, risk assessment, and treatment. Genes and the biological pathways underpinning eating disorders could provide novel targets for the development of safe, effective treatments and improve diagnostic nosology and classification of these disorders. This review provides an overview of genetic discoveries in eating disorders from human genomics research in the genome-wide era. Human genomics research in psychiatry has accelerated significantly in the past decade due to advances in high-throughput genomic sequencing and large-scale genomic data sharing and collaboration. A timeline of key achievements in the genetics of eating disorders is shown in Figure 1 . Table 1 overviews the topics covered in this paper and synthesizes the current state of knowledge.

An external file that holds a picture, illustration, etc.
Object name is nihms-1771245-f0001.jpg

Timeline outlining the history of our understanding of the genetics of eating disorders. Boxes with a heavy outline indicate the dates different types of studies were first undertaken within the eating disorders field. Boxes with a faded outline represent landmark achievements in genetics more broadly. The timeline is not drawn to scale and contains only a small portion of genetic studies in the field. References can be found in the Supplementary Material . AN = anorexia nervosa; BED = binge-eating disorder; BN = bulimia nervosa; CNV = copy-number variation; GWAS = genome-wide association study; h 2 = narrow-sense heritability; NNAT = neuronatin (protein coding gene), SNP = single nucleotide polymorphism.

Summary of the state of knowledge on the genetics of eating disorders

Genetic Epidemiology
Molecular GeneticsGWAS
Genetic correlations
Genetic risk scores
Cross-disorder GWAS
Mendelian randomization
Rare and structural variants , , and variants in AN, linked to the estrogen system , , , , , , and variants in AN
Gene expression and expression in AN is positively associated with leptin, a hormone linking nutritional status and the immune response shows expression in induced pluripotent stem cells from AN patients; the gene was previously associated with anxiety, bipolar disorder, and ADHD
Epigenetic and gene-environment mechanisms hypermethylation has been replicated in multiple studies
Clinical Implications
Future Directions

ADHD = attention-deficit/hyperactivity disorder, AN = anorexia nervosa, BED = binge-eating disorder, BMI = body mass index, BN = bulimia nervosa, GWAS = genome-wide association study, MDD = major depressive disorder, OCD = obsessive-compulsive disorder, SCZ = schizophrenia. See the manuscript text for citations.

Eating Disorders

Eating disorders are mental health conditions characterized by disordered eating behaviors and carry serious physical and mental health morbidity, and elevated mortality ( Arcelus et al., 2011 ). The Diagnostic and Statistical Manual of Mental Disorders , fifth edition (DSM-5), defines the most widely recognized eating disorder, anorexia nervosa (AN), as maintenance of a significantly low body weight through restrictive eating behavior, acute fear of gaining weight, and body image disturbance ( American Psychiatric Association, 2013 ). AN is subdivided into two subtypes: restrictive and binge-eating/purging. Binge eating is a key feature of two other DSM-5 defined eating disorders: bulimia nervosa (BN) and binge-eating disorder (BED) ( American Psychiatric Association, 2013 ). BN is accompanied by recurrent compensatory behaviors such as self-induced vomiting (i.e., a purging behavior) and/or fasting (i.e., a non-purging behavior), whereas BED is defined by binge eating in the absence of such behaviors. Unlike with AN, there are no weight criteria for BN and BED. Persons with BN are typically of normal weight and overweight or obese for BED. Whilst AN and BN have been recognized in the DSM since 1980 ( American Psychiatric Association, 1980 ), BED was only formally recognized as of the DSM-5 in 2013 ( American Psychiatric Association, 2013 ). Lifetime prevalence estimates for AN, BN, and BED range from 0.9–1.4%, 0.5–1.5%, and 1.2–3.5% for females and 0–0.3%, 0.1–0.5%, and 0.3–2.0% in males, respectively ( Hudson et al., 2007 ; Preti et al., 2009 ; Udo & Grilo, 2018 ). The DSM-5 defines additional eating and feeding disorders; however, this review will focus on AN, BN and BED, due to a current lack of genetic and epidemiological research on other eating disorders.

Genetic Epidemiology

Twin studies were among the first lines of evidence to suggest a genetic component for eating disorders. They revealed substantial heritability ( h 2 twin ), which ranges from 16–74% for AN ( Klump et al., 2001 ; Kortegaard et al., 2001 ; Wade et al., 2000 ; Walters & Kendler, 1995 ), 28–83% for BN ( Bulik et al., 1998 ; Bulik et al., 2000 ; Kendler et al., 1995 ), and 39–45% for BED ( Javaras et al., 2008 ; Mitchell et al., 2010 ). The estimates vary, partly, based on whether studies use threshold or relaxed DSM criteria. For example, the heritability of AN is higher when the definition of AN is broadened to include subsyndromal cases ( Dellava et al., 2011 ).

Twin and sibling studies are also a valuable tool to assess the genetic overlap between phenotypes, and have shown that roughly 60% of the genetic effects of AN and BN may be shared ( Bulik et al., 2010 ; Yao et al., in press). Additionally, twin studies have suggested shared genetic risk between eating disorders and alcohol and substance use disorders ( Baker et al., 2010 ; Munn-Chernoff & Baker, 2016 ), obsessive-compulsive disorder (OCD) ( Cederlof et al., 2015 ), and major depressive disorder (MDD) ( Wade et al., 2000 ).

Molecular Genetics

Twin studies do not illuminate the biological and molecular mechanisms involved in risk for which molecular genetics is instrumental. Genetics research in eating disorders has transitioned to the genome-wide era, further, genome-wide association studies (GWAS) have become the dominant approach to identify genetic risk variants associated with complex traits and disorders. GWAS findings can facilitate gene and biological pathway discovery, polygenic risk prediction, and can illuminate causal risk factors and cross-disorder relationships. In this section, we focus predominantly on molecular genetic discoveries for AN, since such genetic studies of BN and BED are limited.

Genome-wide association studies

To date, there have been several eating disorder GWASs, all on AN ( Boraska et al., 2014 ; Duncan et al., 2017 ; Nakabayashi et al., 2009 ; Wang et al., 2011 ; Watson et al., 2019 ). Early GWAS for AN have been subject to important criticisms. With the benefit of hindsight, sample sizes have been too small and underpowered. The first GWAS (320 cases, 341 controls) identified 10 associated microsatellite markers, two of which remained associated in fine-mapping analysis (331 cases, 872 controls). However, the study used DNA pooling which is prone to errors, included only 23,000 markers, and did not safeguard against false positives from multiple testing or population stratification ( Nakabayashi et al., 2009 ). Neither Wang et al.’s (2011) study among 1,033 cases and 3,733 controls or the Genetics Consortium of Anorexia Nervosa and Wellcome Trust Case Control Consortium 3 GWAS among 2,907 cases and 14,860 controls ( Boraska et al., 2014 ) identified any single-nucleotide polymorphisms (SNPs) of genome-wide significance, but this is typical of small sample sizes.

The Psychiatric Genomics Consortium’s (PGC) first GWAS for AN, amongst 3,495 cases and 10,982 controls, identified one genome-wide significant locus on chromosome 12 (rs4622308) that also tags genes implicated in type 1 diabetes and other autoimmune diseases ( Duncan et al., 2017 ). SNP-based heritability ( h 2 snp ) was 20% (s.e. = 2%), indicating that common genetic variants accounted for a large proportion of the twin-based heritability of AN (i.e., at least 27%).

The second PGC GWAS with 16,992 AN cases and 55,525 controls identified eight genome-wide significant loci ( Watson et al., 2019 ) ( Figure 2 ). The locus in the first PGC GWAS did not replicate in the second ( P = 7.02 × 10 −5 ). The risk effect was stronger in the first (i.e., C-allele: OR = 1.20, s.e. = 0.03 vs OR = 1.06, s.e. = 0.01) and showed between-cohort heterogeneity in the second GWAS. Note, we expect that if the effect is in the more modest realm observed in the second GWAS, which is likely given it was initially large, or if the source of heterogeneity is accounted for, the locus will re-emerge in subsequent higher-powered data freezes. Gene-wise association tests, eQTL analyses, chromatin interaction analyses, and base pair coordinates implicated 133 genes. The h 2 snp ranged from 11–17% (s.e. = 1%). The odds ratios (ORs) of SNPs with the largest effects were in the same realm (OR ~1.08–1.17) as observed for psychiatric disorders with more advanced molecular genetic research (OR ~ 1.05–1.15), such as schizophrenia ( Smoller, 2019 ). The results of these GWAS confirm that AN is highly polygenic and suggest that as sample sizes continue to grow, the field will discover more novel risk variants associated with eating disorders.

An external file that holds a picture, illustration, etc.
Object name is nihms-1771245-f0002.jpg

A Manhattan plot depicting eight genome-wide significant loci associated with AN in the second Psychiatric Genomics Consortium genome-wide association study of AN ( Watson et al., 2019 ). In single-gene loci, the gene is annotated. The genome-wide significance threshold ( P < 5 × 10 −8 ) is represented by the horizontal line.

SNP-based genetic correlations

SNP-based genetic correlations ( r g ) provide insight into the overlapping genetics of traits and give further clues to their biological basis. The second PGC GWAS estimated SNP- r g of AN with 447 phenotypes. Statistically significant results fell into six categories: psychiatric, personality, educational attainment, physical activity, metabolic, and anthropometric ( Figure 3 ) ( Watson et al., 2019 ). AN had significant positive SNP- r g with other psychiatric disorders, corresponding with many of the comorbidities observed in clinical and epidemiological studies ( Hudson et al., 2007 ), and with physical activity, which is compelling since compulsive exercise can be a clinical feature of AN. There were negative genetic correlations with anthropometric and metabolic traits, such as body mass index (BMI), leptin, and fasting insulin. The authors investigated whether the SNP- r g between AN and such traits was confounded by low body weight being a diagnostic criterion of AN, by exploring whether SNP- r g s remained significant after removing variance in AN genetic risk accounted for by BMI risk variants. There were modest and statistically non-significant SNP- r g attenuations with metabolic, glycemic, and anthropometric traits, suggesting that AN may be driven, at least in part, by metabolic mechanisms ( Watson et al., 2019 ). This biological clue could explain why those with AN are able to sustain extremely low weights and long-term caloric restriction in the face of strong evolutionary and metabolic forces in the human population toward body fat retention. In a separate study examining the genetic architecture of substance-use-related traits with phenotype sample sizes ranging from 2,400 to 537,000, a significant positive genetic correlation was found for AN with cannabis initiation (SNP- r g = 0.23) and alcohol use disorder (SNP- r g = 0.18), but the latter was no longer significant after co-varying for MDD loci (Munn-Chernoff et al., in press). Additionally, while a positive genetic correlation was observed with cannabis initiation and AN with binge eating (SNP- r g = 0.27), significant negative genetic correlations emerged for smoking phenotypes and AN without binge eating (SNP- r g = −0.19 to −0.23), providing evidence for potential differences in the genetic architecture of AN subtypes (Munn-Chernoff et al., in press).

An external file that holds a picture, illustration, etc.
Object name is nihms-1771245-f0003.jpg

Error bars show the standard error of the r g . Correlations with 447 phenotypes were tested. Only significant correlations surpassing a Bonferroni-corrected P value threshold (P < 1.11 × 10 –4 ) are shown. Complete results are in Supplementary Table 10 of Watson et al. 2019 .

Genetic risk scores

Similar to other psychiatric disorders, genetic risk score (GRS) analyses on GWAS results have suggested that thousands of genetic variants are associated with AN disease risk. GRS, also known as polygenic risk score, is the sum of risk alleles weighted by their effect sizes from GWAS (for a primer, see Wray et al., in press). GRSs for eating disorders are in an early stage of use, given they rely on large, well-powered GWAS. In the second PGC GWAS of AN with 16,992 cases and 55,525 controls, individuals in the highest decile of AN GRS had more than four times the odds of lifetime AN than those in the lowest decile ( Watson et al., 2019 ). GRS for psychiatric traits (such as OCD, schizophrenia, and bipolar disorder) and anthropometric traits have shown significant associations with eating disorder diagnosis and symptom phenotypes in population-based samples ( Abdulkadir et al., 2020 ; Nagata et al., 2019 ; Solmi et al., 2019 ). Many uses for GRS in eating disorder research are on the horizon. GRS can be used to investigate premorbid developmental trajectory, endophenotypes, and course of illness, such as disease severity, relapse, and treatment response. Other possible uses include shedding light on diagnostic nosology and classification, modeling gene-environment etiology, and evaluating whether GWAS findings generalize to multi-ethnic populations for which GWAS may not yet be available. Larger GWAS and more powerful GRS will improve prediction accuracy and advance scientific discovery.

Cross-disorder GWAS

Mental health conditions are highly comorbid with one another, and at times differentiation of a diagnosis based on symptoms may be complex, suggesting shared risk. Cross-disorder GWAS have begun to investigate common genetic pathways in etiology ( Cross-Disorder Group of the Psychiatric Genomics Consortium, 2013 ; Cross-Disorder Phenotype Group of the Psychiatric GWAS Consortium et al., 2009 ). One such effort that included AN and seven other disorders (schizophrenia, bipolar disorder, MDD, attention-deficit/hyperactivity disorder (ADHD), autism spectrum disorder, OCD, and Tourette syndrome) in 232,964 cases and 494,162 controls identified 109 pleiotropic loci, with the 18q21.2 region showing pleiotropic association with all eight disorders ( Cross-Disorder Group of the Psychiatric Genomics Consortium, 2019 ). AN, OCD, and—to a smaller extent—Tourette syndrome clustered together at a genetic level ( Cross-Disorder Group of the Psychiatric Genomics Consortium, 2019 ). A cross-disorder GWAS of AN and OCD (combined sample of 3,495 AN cases, 2,688 OCD cases, and 18,013 controls) found that the SNP- r g s of the AN-OCD cross-disorder phenotype resembled the SNP- r g patterns of both disorders, but the GWAS did not detect any genome-wide significant variants ( Yilmaz et al., 2020 ). When the unique contributions by AN and OCD were examined, the metabolic and anthropometric correlations observed were driven by AN and not OCD. A gene-set enrichment analysis using AN and OCD GWAS summary data (among 16,992 AN cases and 55,525 controls and 2,688 OCD cases and 7,037 controls) also revealed an overlap in many common features of brain regions and developmental stages for AN and OCD in mRNA expression profile, but not as much in DNA-level transcriptome-wide association studies ( Cheng et al., 2020 ), suggesting a role for future mRNA sequencing efforts alongside GWAS to better understand the biological nature of the shared risk between AN and OCD.

Mendelian randomization

Although follow-up analyses using GWAS data (such as genetic correlations and GRSs) have been instrumental in identifying risk factors associated with eating disorders, they do not provide evidence on causality. Mendelian randomization analysis uses linkage-disequilibrium independent genome-wide significant SNPs identified in GWAS as instrumental variables for a given exposure, and measures the degree to which the exposure is causally associated with the outcome ( Davey Smith & Ebrahim, 2003 ). Since genotypes are transmitted randomly from parents to offspring during meiosis, the genotype distribution should be unrelated to confounding factors that are often present in observational studies ( Burgess et al., 2020 ; Teumer, 2018 ). For this reason, Mendelian randomization is often referred to as a natural randomized controlled trial. In the second PGC GWAS of AN, a Mendelian randomization analysis revealed a causal bi-directional relationship between AN and BMI, whereby genetic risk variants for AN led to lower BMI and genetic variants for lower BMI led to an increased risk of lifetime AN ( Watson et al., 2019 ). This finding is complemented by findings that lower BMI predicts the onset of AN ( Stice et al., 2017 ; Yilmaz et al., 2019 ). In general population samples, Mendelian randomization results point to a positive causal association between higher BMI and eating disorder behaviors and symptoms ( Reed et al., 2017 ). Similarly, epidemiological results link higher BMI to the onset of body dissatisfaction and disordered eating in the general population ( Stice et al., 2002 ).

Adiponectin is a fat-derived hormone that plays a key role in energy homeostasis and appetite regulation ( Coll et al., 2007 ; Steinberg & Kemp, 2007 ). Altered adiponectin levels have been observed in patients with AN and BN. Awofala and colleagues (2019) performed a Mendelian randomization study using GWAS summary data for adiponectin (29,347 samples) and eating behavior disinhibition (898 samples) and found that higher blood adiponectin was causally associated with eating behavior disinhibition. These findings support previous studies which showed an association between effect allele carriers in the adiponectin gene and increased frequency of overeating, which may consequently lead to symptoms of eating disorders ( Rohde et al., 2015 ). These studies point towards innate biological drivers that may lead toward symptoms of eating disorders.

Mendelian randomization for eating disorder research is in its infancy, since the strength of genetic instruments in Mendelian randomization is determined by well-powered GWAS ( Zhu et al., 2018 ). Recent Mendelian randomization approaches allow for nonlinear-associations, which may significantly advance the application of Mendelian randomization in eating disorder research in the future, particularly with respect to anthropometric risk factors ( Burgess et al., 2014 ).

Rare and structural variants

Genetic studies have started to explore the role of rare and structural variants in eating disorders. Regarding studies of copy number variants (CNVs), Wang et al. (2011) examined 2,158 AN cases and 15,458 controls and found no evidence that cases had a significantly higher burden of CNVs. However, they identified a recurrent 13q12 deletion (1.5 Mb) in two cases, and CNVs disrupting the CNTN6 / CNTN4 region in several cases. Similarly, in a sample of 1,983 cases with AN, Yilmaz et al. (2017) found another AN case with a deletion in the 13q12 region. The authors also observed two instances of CNVs with at least 50% reciprocal overlap with regions associated with psychiatric and neurodevelopmental disorders ( Yilmaz et al., 2017 ). Alongside these well-established neuropsychiatric CNVs, instances of rare and large CNVs in AN cases were also observed ( Yilmaz et al., 2017 ). In addition, mixed results have been found for microduplications at 15q11.2 ( Chang et al., 2019 ; Wang et al., 2011 ; Yilmaz et al., 2017 ).

Whole-exome and whole-genome analyses have also provided evidence for an enrichment of rare variants in AN ( Bienvenu et al., 2019 ; Cui et al., 2013 ; Iacobellis & Barbaro, 2019 ; Lombardi et al., 2019 ; Lutter et al., 2017 ). A whole‐exome analysis in two independent families with males with AN found variants in the neuronatin ( NNAT ) gene in both probands: one nonsense variant (p.Trp33*) and one rare variant in the 5′UTR ( Lombardi et al., 2019 ). Eleven additional NNAT variants were found in a follow-up cohort of eight male and 144 females with AN. Another study combined exome sequencing, whole‐genome sequencing, and linkage analysis to examine two families with recurrence of AN ( Cui et al., 2013 ). In the first pedigree, they found a missense variant co‐segregating with the affected family members in the ESRRA (estrogen-related receptor alpha), and a potentially damaging mutation in the HDAC4 (histone deacetylase 4) in the second pedigree ( Cui et al., 2013 ). These genes are linked to the estrogen system.

Whole-genome sequencing analyses in eating disorders are less common. Bergen et al. (2019) performed a whole-genome sequencing analysis in six individuals—two maternally-linked cousins with severe AN and their parents—and found, of the approximately 5.3 million variants per individual analyzed, that 494,712 variants were shared identical-by-descent by the cousin pair based on maternally derived haplotypes ( Bergen et al., 2019 ). They identified novel variants in seven genes: TTC22 , MRPS9 , DNAJC30 , HEPACAM2 , USP20 , ESF1 , and CDK5RAP1 . These findings suggest that there may be utility in whole-genome sequencing of families with affected individuals to detect rare variants that may influence AN ( Bergen et al., 2019 ).

Despite strong evidence for the heritable polygenic risk of AN, rare variant contributions of large effect have not yet been identified. Early studies show promise and larger-scale studies with well-matched control groups and replication studies will be necessary for illuminating whether rare and structural variants contribute to eating disorders.

Gene expression

Gene expression offers insight into the genes and molecular mechanisms that influence phenotypes. Note, that whereas GWASs identify inherited genetic variants associated with disease risk and epigenetic studies investigate changes to the physical structure of DNA, gene expression studies measure messenger RNA expression levels in any given tissue, thus capturing the degree to which a gene is being expressed. Howard et al. (2020) investigated brain regions enriched for gene expression to understand the molecular neuroanatomy of AN. The authors combined the gene lists from two common variant, a rare variant, and a stem-cell study ( Duncan et al., 2017 ; Lutter et al., 2017 ; Negraes et al., 2017 ; Watson et al., 2019 ), and used genetic and transcriptomic resources spanning human fetal and adult and mouse gene expression data. Genes associated with AN resided in subcortical feeding and reward circuits; and furthermore, they implicated microglia genes and genes responding to fasting in mice hypothalami (i.e., RPS26 and DALRD3 ). Likewise, the PGCs recent GWAS of AN (2019) found an enrichment of gene expression in CNS brain tissues and striatal and hippocampus neurons linked to feeding and reward.

Another set of studies applied transcriptome expression profiling to assess gene expression changes in six individuals with AN before and after inpatient weight restoration. Among the top 20 genes, was down-regulation of genes encoding for a cholesterol side-chain cleavage enzyme (CYPP450scc) and up-regulation of genes related to protein secretion, protein signaling, defense response to bacterial regulation, and olfactory receptor regulation ( Kim et al., 2013 ). Of the top differentially expressed genes, CPA3 and GATA2 expression were positively associated with levels of leptin, a hormone linked to nutritional status and the immune response ( Baker et al., 2019 ). This aligns with studies suggesting a genetic overlap between AN, autoimmune disease, and metabolic function ( Baker et al., 2019 ).

In a study with seven females with AN and four healthy controls, Negraes and colleagues (2017) modelled AN using induced pluripotent stem cells, with their transcriptomic analyses revealing a novel gene, TACR1 , that may contribute to AN pathophysiology. The TACR1 gene encodes the tachykinin (or neurokinin) 1 receptor which is involved in a range of biological processes, interacts with several neurotransmitters, and has previously been associated with anxiety disorders, bipolar disorder, and ADHD, suggesting a novel system that might contribute to AN symptoms ( Schank, 2014 ; Sharp et al., 2014 ). Although several studies on gene expression in eating disorders exist, there are not many and most have small samples, limiting the conclusions that can be drawn ( Kim et al., 2013 ).

Epigenetic and gene-environment mechanisms

Epigenetics refers to chemical modifications to DNA and chromatin proteins that control gene expression but do not change the underlying base-pair sequence of the DNA ( Ryan et al., 2018 ). Epigenetic changes are typically measured via global methylation (amount of methylated cytosine compared to total cytosine), via a candidate gene approach, or more recently, via epigenome-wide association studies (EWAS), which have gained popularity and are carried out in a similar way to GWAS. Global methylation and candidate gene methylation study results in eating disorders, specifically AN, have been mixed, with largely inconsistent findings and opposite effects ( Booij et al., 2015 ; Frieling et al., 2007 ; Hübel et al., 2019 ; Saffrey et al., 2014 ; Tremolizzo et al., 2014 ). EWAS using a hypothesis-free driven method have however identified multiple differentially methylated sites associated with AN ( Booij et al., 2015 ; Kesselmeier et al., 2018 ; Ramoz et al., 2017 ). Samples in these studies have ranged from 29 to 47 AN cases and 15 to 147 controls. TNXB hypermethylation has been replicated, although the significance level was nominal and therefore a false positive finding cannot be ruled out ( Kesselmeier et al., 2018 ). TNXB plays a role in maintaining muscles, joints, organs and skin and regulates the production of collagen. Future studies would need to replicate this finding using larger sample sizes, but these early findings could indicate epigenetic changes in people with eating disorders. Notably, future studies need to be well-designed in order to disentangle epigenetic differences in eating disorder patients by disorder type, tissue type, cell type, and take into account large numbers of environmental factors such as diet, binge eating and purging behaviors, and medication ( Horvath & Raj, 2018 ; Kubota et al., 2012 ; Moore et al., 2013 ).

Gene-environment interaction studies in eating disorders have predominantly focused on candidate genes relating to behavior, emotion, and the stress response, such as serotonin and glucocorticoid genes. However, candidate gene studies are subject to false-positive results. Another avenue to study gene-environment interaction is via the use of GRS to capture ‘G’. Recent studies have begun modelling GRS by environment interactions in various psychiatric disorders with some interesting findings ( Mullins et al., 2016 ). However, this methodology is in its infancy and has not yet been applied to eating disorders due to a lack of well-powered GWAS needed to calculate GRS. In the future, we are likely to see great advancements in this area of study.

Clinical Implications

Integrating genomics into clinical practice.

Translatability is a key goal for genomics research in eating disorders. Through continued research with larger sample sizes, the era in which personal genomic information—in combination with other known risk factors—occupies a potential role in forecasting eating disorder risk, outcome, and clinical decision-making will emerge. The most immediate benefit of current genomics research is an improved scientific understanding of the role genomics plays in eating disorder risk. For instance, genomics can be integrated into clinical settings in the context of psychoeducation. As part of standard care, providers give patients and families psychoeducation about eating disorders, which ought to include up-to-date information about heritability and genetic risk. Communicating information to patients and families is complex for it can arouse emotions such as guilt and fear (i.e., “passing on ‘bad’ genes) and unhelpful cognitions, such as reduced self-efficacy and fatalism, the belief that little can be done to reduce risk. There are few empirical studies on genetic counseling for eating disorders, or mental health conditions in general, to date, but guidance for clinicians is available in a synthesis from related literature (see Bulik et al., 2019 ). When communicating such information with patients, it will be important for clinicians to recognize the possibility of such unhelpful cognitions and beliefs and be able to orient the patient to the meaning of genetic findings in the context of the totality of what is known about the risk for eating disorders. As genetics is just one, non-deterministic piece of the puzzle, it is essential that clinicians convey this. Evidence-based, user-friendly resources for treatment providers, patients, and families, continually updated as genomics research evolves, will help patients benefit from cutting-edge research ( Bulik et al., 2019 ). In the future when personalized genetic testing may become available, clinicians may want to collaborate with genetic counselors with expertise in communicating test results.

Bridging the therapeutic impasse

Genomic research progress is taking place contextually in a field with a long-standing clinical plateau in therapeutics, similar to other mental health disorders. The hope is that insights gained through genomics will break through the therapeutic impasse and speed up the search for novel, effective treatments.

Leading, empirically-supported treatments for eating disorders are a half-century old and their efficacy is limited: for example, cognitive-behavioral therapy (CBT), developed first for BN in the early 1980s ( Fairburn, 1981 ) was informed by depression theory and treatment pioneered in the 1960s ( Beck, 2019 ); interpersonal psychotherapy treatment (IPT) was first applied to BN in clinical trials in the 1990s ( Fairburn et al., 1991 ); family-based treatment (FBT), also known as the Maudsley approach and the most recent addition to the therapeutic repertoire, was developed at the Maudsley Hospital in London in the 1970s ( Lock & Le Grange, 2013 ); and pharmacological agents to treat BN and BED, most notably selective serotonin reuptake inhibitors, were an extension of serendipitous discoveries related to depression from the 1950s (an exception is a recent treatment advance for BED in the form of stimulant, lisdexamfetamine: Hudson et al., 2017 ). The long-term success rates of these treatments are around 25% to 65% ( Agras, 1997 ; Hilbert et al., 2012 ; Lock & Le Grange, 2019 ), with many individuals having chronic, partially recovered, or relapsing courses of illness, and no empirically-supported treatment has yet been established for AN in adults.

Understanding the biological pathways involved and potential drug target genes may be fruitful for developing the next generation of interventions. Schizophrenia GWAS results, for example, are associated with antipsychotic drugs, which are already used in treatment, and selective calcium channel blockers and antiepileptics, therapeutic classes that present repurposing opportunities ( Gaspar & Breen, 2017 ). Similarly, it is hoped that genomics findings will provide leads for novel eating disorder treatments. For example, pharmacological agents that address metabolic processes may represent pharmacotherapeutic targets for AN. Drug targets based on GWAS findings are more likely to be successful in phase II and III clinical trials and to make it to market ( King et al., 2019 ; Nelson et al., 2015 ).

Future Directions

The immediate priority is to increase the statistical power of analyses by increasing sample sizes, a pursuit underway within the Eating Disorders Workgroup of the PGC for AN and other eating disorders. Twin and SNP-heritability estimates imply that with increased sample size it is a matter of time before more risk loci are identified, for example, as has occurred for schizophrenia from 7 loci with ~18k cases ( Schizophrenia Psychiatric Genome-Wide Association Study (GWAS) Consortium, 2011 ) to 108 loci with ~37k cases ( Schizophrenia Working Group of the Psychiatric Genomics Consortium, 2014 ), and for MDD from 0 loci with ~9k cases ( Ripke et al., 2013 ), 44 loci with ~135k cases ( Wray et al., 2018 ), and 102 loci with ~246k cases ( Howard et al., 2019 ). The genetic architecture of complex traits and diseases means that individual risk loci account for only a very small fraction of the heritability, but they may tag causal genes and cooperatively make contributions to disease pathways. Further, statistical power will facilitate systems analyses that interpret the data in the context of how gene and biological pathways influence the phenotype, and other types of analyses discussed throughout this paper.

Expansion of eating disorder phenotypes

Similarly, expansion of eating disorder phenotypes to eating disorders beyond AN is needed. This includes gaining an understanding of how phenotype measurement—from gold standard (i.e., yielded through clinical interview) to other means (i.e., register diagnoses, electronic health record data, questionnaire-based algorithms, self-report diagnosis history)—affects genomic findings. Interestingly, the MDD field found that similar GWAS-related results were yielded under detailed interviewer-based versus self-reported depression diagnosis and that continuous measures of non-pathological depressive symptoms yielded substantial genetic correlations with MDD ( McIntosh et al., 2019 ). This tapped more samples globally for use and accelerated statistical power and discovery.

Functional genomics

As GWAS and rare and structural variant studies correlate genetic variants with eating disorder phenotypes, functional genomic analyses will be needed to convert these insights into an understanding of the underlying biological mechanisms. These analyses are important for identifying the causal variant tagged by the genome-wide significant SNP, the biological function of the causal variant, and the gene/s involved in its association. Many computational (i.e., statistical fine mapping, eQTL analysis, TWAS, gene-set enrichment analysis) and molecular biology (i.e., RNA-seq datasets, ChIP-seq studies, Hi-C analysis, chromatin accessibility assays, knock-out animal models) approaches are being used for functional follow-up. A challenge is that many significant SNP signals in mental health disorders, including AN, are falling in non-protein coding regions.

Conclusions

Genomics is a rapidly-evolving research area in eating disorders. Eating disorders aggregate in families, are moderately heritable, and (at least for AN so far) some of the risk is attributed to genetic variants commonly found in the population ( Duncan et al., 2017 ; Watson et al., 2019 ). Genome-wide era work is beginning to unravel the genetic architecture of AN. Several loci and over 130 genes have shown associations with AN, and genetic overlap between AN and psychiatric (especially OCD), personality and behavioral, physical activity, cognitive, metabolic, and anthropometric traits has been revealed. Now, efforts are needed to elaborate on the functional context of these genes. Genomic approaches such as Mendelian randomization are helping to identify causal risk factors and have so far highlighted the importance of metabolic traits such as BMI, and adiponectin, in AN. Further research is needed to disentangle metabolic genomic factors from low weight in AN. A peek over the horizon into clinical management suggests that patient screening, care, and outcomes may improve from advances in molecular genetics. Genomic discovery depends on very large sample sizes and large-scale collaborations. In the next few years, the Eating Disorders Working Group of the PGC and large-scale studies such as the Eating Disorders Genetics Initiative (EDGI) and Binge Eating Genetics INitiative (BEGIN) ( Bulik et al., 2020 ) will be important to watch for advances in progress.

Supplementary Material

Funding statement:.

AH and AP are supported by the National Institute for Health Research (NIHR) Biomedical Research Centre at the South London and Maudsley National Health Service (NHS) Foundation Trust. HD is supported by an Economic and Social Research Council studentship. HW is supported by the National Institute of Mental Health (NIMH) (U01MH109528, R01MH120170). JB is supported by the NIMH (K01MH106675). ZY is supported by the NIMH (K01MH109782; R01MH105500; R01MH120170) and a Brain and Behavior Research Foundation NARSAD Young Investigator Award (grant # 28799). The content is solely the responsibility of the authors and does not necessarily represent the official views of the funding agencies.

Conflicts of interest: All authors declare that they have no conflicts of interest.

  • Abdulkadir M, Herle M, De Stavola BL, Hübel C, Santos Ferreira DL, Loos RJF, … Micali N (2020). Polygenic score for body mass index is associated with disordered eating in a general population cohort . Journal of Clinical Medicine , 9 , 1187. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Agras WS (1997). Pharmacotherapy of bulimia nervosa and binge eating disorder: Longer-term outcomes . Psychopharmacology Bulletin , 33 , 433–436. [ PubMed ] [ Google Scholar ]
  • American Psychiatric Association. (1980). Diagnostic and statistical manual of mental disorders (3rd ed.). Washington, DC: American Psychiatric Association. [ Google Scholar ]
  • American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders (5th ed.). Washington, DC: American Psychiatric Association. [ Google Scholar ]
  • Arcelus J, Mitchell AJ, Wales J, & Nielsen S (2011). Mortality rates in patients with anorexia nervosa and other eating disorders: A meta-analysis of 36 studies . Archives of General Psychiatry , 68 , 724–731. [ PubMed ] [ Google Scholar ]
  • Awofala AA, Ogundele OE, Adekoya KO, & Osundina SA (2019). Adiponectin and human eating behaviour: A Mendelian randomization study . Egyptian Journal of Medical Human Genetics , 20 , 17. [ Google Scholar ]
  • Baker J, Kim Y, Crowley J, Trace S, Brownley K, Pisetsky D, & Bulik C (2019). SU44 - Correlation between gene expression change and hormone change during refeeding in anorexia nervosa . European Neuropsychopharmacology , 29 , S911. [ Google Scholar ]
  • Baker JH, Mitchell KS, Neale MC, & Kendler KS (2010). Eating disorder symptomatology and substance use disorders: Prevalence and shared risk in a population based twin sample . International Journal of Eating Disorders , 43 , 648–658. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Beck AT (2019). A 60-year evolution of cognitive theory and therapy . Perspectives on Psychologica Science , 14 , 16–20. [ PubMed ] [ Google Scholar ]
  • Bergen A, Shih P.-a. B., Zeeland AS-V, Carland T, Bansal V, Magistretti P, … Schork N (2019). M42 - Whole genome sequence analysis of a cousin pair with restricting anorexia nervosa . European Neuropsychopharmacology , 29 , S977–S978. [ Google Scholar ]
  • Bienvenu T, Lebrun N, Clarke J, Duriez P, Gorwood P, & Ramoz N (2019). Exome sequencing in a familial form of anorexia nervosa supports multigenic etiology . Journal of Neural Transmission , 126 , 1505–1511. [ PubMed ] [ Google Scholar ]
  • Booij L, Casey KF, Antunes JM, Szyf M, Joober R, Israël M, & Steiger H (2015). DNA methylation in individuals with anorexia nervosa and in matched normal-eater controls: A genome-wide study . International Journal of Eating Disorders , 48 , 874–882. [ PubMed ] [ Google Scholar ]
  • Boraska V, Franklin CS, Floyd JA, Thornton LM, Huckins LM, Southam L, … Bulik CM (2014). A genome-wide association study of anorexia nervosa . Molecular Psychiatry , 19 , 1085–1094. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Bulik C, Thornton L, Root T, Pisetsky E, Lichtenstein P, & Pedersen N (2010). Understanding the relation between anorexia nervosa and bulimia nervosa in a Swedish national twin sample . Biological Psychiatry , 67 , 71–77. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Bulik CM, Blake L, & Austin J (2019). Genetics of Eating Disorders: What the Clinician Needs to Know . Psychiatric Clinics of North America , 42 , 59–73. [ PubMed ] [ Google Scholar ]
  • Bulik CM, Butner JE, Tregarthen J, Thornton LM, Flatt RE, Smith T, … Deboeck PR (2020). The Binge Eating Genetics Initiative (BEGIN): Study protocol . BMC Psychiatry , 20 , 1–9. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Bulik CM, Sullivan PF, & Kendler KS (1998). Heritability of binge-eating and broadly defined bulimia nervosa . Biological Psychiatry , 44 , 1210–1208. [ PubMed ] [ Google Scholar ]
  • Bulik CM, Sullivan PF, Wade TD, & Kendler KS (2000). Twin studies of eating disorders: A review . International Journal of Eating Disorders , 27 , 1–20. [ PubMed ] [ Google Scholar ]
  • Burgess S, Davey Smith G, Davies NM, Dudbridge F, Gill D, Glymour MM, … Theodoratou E (2020). Guidelines for performing Mendelian randomization investigations . Wellcome Open Research , 4 , 186. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Burgess S, Davies NM, Thompson SG, & Consortium EP-I (2014). Instrumental variable analysis with a nonlinear exposure-outcome relationship . Epidemiology , 25 , 877–885. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Cederlof M, Thornton LM, Baker J, Lichtenstein P, Larsson H, Ruck C, … Mataix-Cols D (2015). Etiological overlap between obsessive-compulsive disorder and anorexia nervosa: a longitudinal cohort, multigenerational family and twin study . World Psychiatry , 14 , 333–338. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Chang X, Qu H, Liu Y, Glessner J, Hou C, Wang F, … Hakonarson H (2019). Microduplications at the 15q11.2 BP1-BP2 locus are enriched in patients with anorexia nervosa . Journal of Psychiatric Research , 113 , 34–38. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Cheng B, Qi X, Liang C, Zhang L, Ma M, Li P, … Zhang F (2020). Integrative genomic enrichment analysis identified the brain regions and development stages related to anorexia nervosa and obsessive-compulsive disorder . Cerebral Cortex , 30 , 6481–6489. [ PubMed ] [ Google Scholar ]
  • Coll AP, Farooqi IS, & O’Rahilly S (2007). The hormonal control of food intake . Cell , 129 , 251–262. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Cross-Disorder Group of the Psychiatric Genomics Consortium. (2013). Identification of risk loci with shared effects on five major psychiatric disorders: a genome-wide analysis . Lancet , 381 , 1371–1379. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Cross-Disorder Group of the Psychiatric Genomics Consortium. (2019). Genomic Relationships, Novel Loci, and Pleiotropic Mechanisms across Eight Psychiatric Disorders . Cell , 179 , 1469–1482.e1411. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Cross-Disorder Phenotype Group of the Psychiatric GWAS Consortium, Craddock N, Kendler K, Neale M, Nurnberger J, Purcell S, … Thapar A (2009). Dissecting the phenotype in genome-wide association studies of psychiatric illness . British Journal of Psychiatry , 195 , 97–99. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Cui H, Moore J, Ashimi SS, Mason BL, Drawbridge JN, Han S, … Lutter M (2013). Eating disorder predisposition is associated with ESRRA and HDAC4 mutations . Journal of Clinical Investigation , 123 , 4706–4713. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Davey Smith G, & Ebrahim S (2003). ‘Mendelian randomization’: Can genetic epidemiology contribute to understanding environmental determinants of disease? International Journal of Epidemiology , 32 , 1–22. [ PubMed ] [ Google Scholar ]
  • Dellava JE, Thornton LM, Lichtenstein P, Pedersen NL, & Bulik CM (2011). Impact of broadening definitions of anorexia nervosa on sample characteristics . Journal of Psychiatric Research , 45 , 691–698. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Duncan L, Yilmaz Z, Gaspar H, Walters R, Goldstein J, Anttila V, … Bulik CM (2017). Significant locus and metabolic genetic correlations revealed in genome-wide association study of anorexia nervosa . American Journal of Psychiatry , 174 , 850–858. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Fairburn C (1981). A cognitive behavioural approach to the treatment of bulimia . Psychological Medicine , 11 , 707–711. [ PubMed ] [ Google Scholar ]
  • Fairburn CG, Jones R, Peveler RC, Carr SJ, Solomon RA, O’Connor ME, … Hope RA (1991). Three psychological treatments for bulimia nervosa. A comparative trial . Archives of General Psychiatry , 48 , 463–469. [ PubMed ] [ Google Scholar ]
  • Frieling H, Gozner A, Römer KD, Lenz B, Bönsch D, Wilhelm J, … Bleich S (2007). Global DNA hypomethylation and DNA hypermethylation of the alpha synuclein promoter in females with anorexia nervosa . Molecular Psychiatry , 12 , 229–230. [ PubMed ] [ Google Scholar ]
  • Gaspar HA, & Breen G (2017). Drug enrichment and discovery from schizophrenia genome-wide association results: An analysis and visualisation approach . Scientific Reports , 7 , 12460. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Hilbert A, Bishop ME, Stein RI, Tanofsky-Kraff M, Swenson AK, Welch RR, & Wilfley DE (2012). Long-term efficacy of psychological treatments for binge eating disorder . British Journal of Psychiatry , 200 , 232–237. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Horvath S, & Raj K (2018). DNA methylation-based biomarkers and the epigenetic clock theory of ageing . Nature Reviews Genetics , 19 , 371–384. [ PubMed ] [ Google Scholar ]
  • Howard D, Negraes P, Voineskos AN, Kaplan AS, Muotri AR, Duvvuri V, & French L (2020). Molecular neuroanatomy of anorexia nervosa . Scientific Reports , 10 , 11411. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Howard DM, Adams MJ, Clarke TK, Hafferty JD, Gibson J, Shirali M, … McIntosh AM (2019). Genome-wide meta-analysis of depression identifies 102 independent variants and highlights the importance of the prefrontal brain regions . Nature Neuroscience , 22 , 343–352. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Hübel C, Marzi SJ, Breen G, & Bulik CM (2019). Epigenetics in eating disorders: A systematic review . Molecular Psychiatry , 24 , 901–915. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Hudson JI, Hiripi E, Pope HG Jr., & Kessler RC (2007). The prevalence and correlates of eating disorders in the National Comorbidity Survey Replication . Biological Psychiatry , 61 , 348–358. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Hudson JI, McElroy SL, Ferreira-Cornwell MC, Radewonuk J, & Gasior M (2017). Efficacy of lisdexamfetamine in adults with moderate to severe binge-eating disorder: A randomized clinical trial . JAMA Psychiatry , 74 , 903–910. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Iacobellis G, & Barbaro G (2019). Targeting the organ-specific adiposity . Eating and Weight Disorders , 24 , 1–2. [ PubMed ] [ Google Scholar ]
  • Javaras KN, Laird NM, Reichborn-Kjennerud T, Bulik CM, Pope HG Jr., & Hudson JI (2008). Familiality and heritability of binge eating disorder: results of a case-control family study and a twin study . International Journal of Eating Disorders , 41 , 174–179. [ PubMed ] [ Google Scholar ]
  • Kendler KS, Walters EE, Neale MC, Kessler RC, Heath AC, & Eaves LJ (1995). The structure of the genetic and environmental risk factors for six major psychiatric disorders in women: Phobia, generalized anxiety disorder, panic disorder, bulimia, major depression and alcoholism . Archives of General Psychiatry , 52 , 374–383. [ PubMed ] [ Google Scholar ]
  • Kesselmeier M, Pütter C, Volckmar AL, Baurecht H, Grallert H, Illig T, … Hinney A (2018). High-throughput DNA methylation analysis in anorexia nervosa confirms TNXB hypermethylation . World Journal of Biological Psychiatry , 19 , 187–199. [ PubMed ] [ Google Scholar ]
  • Kim Y, Trace SE, Crowley JJ, Brownley KA, Hamer RM, Pisetsky DS, … Bulik CM (2013). Assessment of gene expression in peripheral blood using RNAseq before and after weight restoration in anorexia nervosa . Psychiatry Research , 210 , 287–293. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • King EA, Davis JW, & Degner JF (2019). Are drug targets with genetic support twice as likely to be approved? Revised estimates of the impact of genetic support for drug mechanisms on the probability of drug approval . PLoS Genetics , 15 , e1008489. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Klump KL, Miller KB, Keel PK, McGue M, & Iacono WG (2001). Genetic and environmental influences on anorexia nervosa syndromes in a population-based twin sample . Psychological Medicine , 31 , 737–740. [ PubMed ] [ Google Scholar ]
  • Kortegaard LS, Hoerder K, Joergensen J, Gillberg C, & Kyvik KO (2001). A preliminary population-based twin study of self-reported eating disorder . Psychological Medicine , 31 , 361–365. [ PubMed ] [ Google Scholar ]
  • Kubota T, Miyake K, & Hirasawa T (2012). Epigenetic understanding of gene-environment interactions in psychiatric disorders: A new concept of clinical genetics . Clinical Epigenetics , 4 , 1. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Lock J, & Le Grange D (2013). Treatment manual for anorexia nervosa: A family-based approach (2nd ed.). New York: Guilford Press. [ Google Scholar ]
  • Lock J, & Le Grange D (2019). Family-based treatment: Where are we and where should we be going to improve recovery in child and adolescent eating disorders . International Journal of Eating Disorders , 52 , 481–487. [ PubMed ] [ Google Scholar ]
  • Lombardi L, Blanchet C, Poirier K, Lebrun N, Ramoz N, Rose Moro M, … Bienvenu T (2019). Anorexia nervosa is associated with neuronatin variants . Psychiatric Genetics , 29 , 103–110. [ PubMed ] [ Google Scholar ]
  • Lutter M, Bahl E, Hannah C, Hofammann D, Acevedo S, Cui H, … Michaelson JJ (2017). Novel and ultra-rare damaging variants in neuropeptide signaling are associated with disordered eating behaviors . PLoS One , 12 , e0181556. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • McIntosh AM, Sullivan PF, & Lewis CM (2019). Uncovering the genetic architecture of major depression . Neuron , 102 , 91–103. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Mitchell KS, Neale MC, Bulik CM, Aggen SH, Kendler KS, & Mazzeo SE (2010). Binge eating disorder: a symptom-level investigation of genetic and environmental influences on liability . Psychological Medicine , 40 , 1899–1906. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Moore LD, Le T, & Fan G (2013). DNA methylation and its basic function . Neuropsychopharmacology , 38 , 23–38. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Mullins N, Power RA, Fisher HL, Hanscombe KB, Euesden J, Iniesta R, … Lewis CM (2016). Polygenic interactions with environmental adversity in the aetiology of major depressive disorder . Psychological Medicine , 46 , 759–770. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Munn-Chernoff MA, & Baker JH (2016). A primer on the genetics of comorbid eating disorders and substance use disorders . European Eating Disorders Review , 24 , 91–100. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Munn-Chernoff MA, Johnson EC, Chou YL, Coleman JRI, Thornton LM, Walters RK, … Agrawal A (2021). Shared genetic risk between eating disorder- and substance-use-related phenotypes: Evidence from genome-wide association studies . Addiction Biology , 26 , e128880. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Nagata JM, Braudt DB, Domingue BW, Bibbins-Domingo K, Garber AK, Griffiths S, & Murray SB (2019). Genetic risk, body mass index, and weight control behaviors: Unlocking the triad . International Journal of Eating Disorders , 52 , 825–833. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Nakabayashi K, Komaki G, Tajima A, Ando T, Ishikawa M, Nomoto J, … Shirasawa S (2009). Identification of novel candidate loci for anorexia nervosa at 1q41 and 11q22 in Japanese by a genome-wide association analysis with microsatellite markers . Journal of Human Genetics , 54 , 531–537. [ PubMed ] [ Google Scholar ]
  • Negraes PD, Cugola FR, Herai RH, Trujillo CA, Cristino AS, Chailangkarn T, … Duvvuri V (2017). Modeling anorexia nervosa: transcriptional insights from human iPSC-derived neurons . Translational Psychiatry , 7 , e1060. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Nelson MR, Tipney H, Painter JL, Shen J, Nicoletti P, Shen Y, … Sanseau P (2015). The support of human genetic evidence for approved drug indications . Nature Genetics , 47 , 856–860. [ PubMed ] [ Google Scholar ]
  • Preti A, Girolamo G, Vilagut G, Alonso J, Graaf R, Bruffaerts R, … Morosini P (2009). The epidemiology of eating disorders in six European countries: Results of the ESEMeD-WMH project . Journal of Psychiatric Research , 43 , 1125–1132. [ PubMed ] [ Google Scholar ]
  • Ramoz N, Guillaume S, Courtet P, & Gorwood P (2017). Epigenetics in the remission of anorexia nervosa: A follow-up study of whole-genome methylation profiles . European Psychiatry , 41 , S102. [ Google Scholar ]
  • Reed ZE, Micali N, Bulik CM, Davey Smith G, & Wade KH (2017). Assessing the causal role of adiposity on disordered eating in childhood, adolescence, and adulthood: a Mendelian randomization analysis . American Journal of Clinical Nutrition , 106 , 764–772. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Ripke S, Wray NR, Lewis CM, Hamilton SP, Weissman MM, Breen G, … Sullivan PF (2013). A mega-analysis of genome-wide association studies for major depressive disorder . Molecular Psychiatry , 18 , 497–511. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Rohde K, Keller M, Horstmann A, Liu X, Eichelmann F, Stumvoll M, … Böttcher Y (2015). Role of genetic variants in ADIPOQ in human eating behavior . Genes and Nutrition , 10 , 449. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Ryan J, Saffery R, & Patton G (2018). Epigenetics: a missing link in understanding psychiatric disorders? Lancet Psychiatry , 5 , 8–9. [ PubMed ] [ Google Scholar ]
  • Saffrey R, Novakovic B, & Wade TD (2014). Assessing global and gene specific DNA methylation in anorexia nervosa: A pilot study . International Journal of Eating Disorders , 47 , 206–210. [ PubMed ] [ Google Scholar ]
  • Schank JR (2014). The neurokinin-1 receptor in addictive processes . Journal of Pharmacology and Experimental Therapeutics , 351 , 2–8. [ PubMed ] [ Google Scholar ]
  • Schizophrenia Psychiatric Genome-Wide Association Study (GWAS) Consortium. (2011). Genome-wide association study identifies five new schizophrenia loci . Nature Genetics , 43 , 969–976. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Schizophrenia Working Group of the Psychiatric Genomics Consortium. (2014). Biological insights from 108 schizophrenia-associated genetic loci . Nature , 511 , 421–427. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Sharp SI, McQuillin A, Marks M, Hunt SP, Stanford SC, Lydall GJ, … Gurling HM (2014). Genetic association of the tachykinin receptor 1 TACR1 gene in bipolar disorder, attention deficit hyperactivity disorder, and the alcohol dependence syndrome . American Journal of Medical Genetics, Part B, Neuropsychiatr Genetics, 165b , 373–380. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Smoller JW (2019). Psychiatric genetics begins to find its footing . American Journal of Psychiatry , 176 , 609–614. [ PubMed ] [ Google Scholar ]
  • Solmi F, Mascarell MC, Zammit S, Kirkbride JB, & Lewis G (2019). Polygenic risk for schizophrenia, disordered eating behaviours and body mass index in adolescents . British Journal of Psychiatry , 215 , 428–433. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Steinberg GR, & Kemp BE (2007). Adiponectin: starving for attention . Cell Metabolism , 6 , 3–4. [ PubMed ] [ Google Scholar ]
  • Stice E, Gau JM, Rohde P, & Shaw H (2017). Risk factors that predict future onset of each DSM-5 eating disorder: Predictive specificity in high-risk adolescent females . Journal of Abnormal Psychology , 126 , 38–51. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Stice E, Presnell K, & Spangler D (2002). Risk factors for binge eating onset in adolescent girls: A 2-year prospective investigation . Health Psychology , 21 , 131–138. [ PubMed ] [ Google Scholar ]
  • Teumer A (2018). Common methods for performing Mendelian randomization . Frontiers in Cardiovascular Medicine , 5 , 51. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Tremolizzo L, Conti E, Bomba M, Uccellini O, Rossi MS, Marfone M, … Nacinovich R (2014). Decreased whole-blood global DNA methylation is related to serum hormones in anorexia nervosa adolescents . World Journal of Biological Psychiatry , 15 , 327–333. [ PubMed ] [ Google Scholar ]
  • Udo T, & Grilo CM (2018). Prevalence and correlates of DSM-5-defined eating disorders in a nationally representative sample of U.S. adults . Biological Psychiatry , 84 , 345–354. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Wade TD, Bulik CM, Neale M, & Kendler KS (2000). Anorexia nervosa and major depression: Shared genetic and environmental risk factors . American Journal of Psychiatry , 157 , 469–471. [ PubMed ] [ Google Scholar ]
  • Walters EE, & Kendler KS (1995). Anorexia nervosa and anorexic-like syndromes in a population-based female twin sample . American Journal of Psychiatry , 152 , 64–71. [ PubMed ] [ Google Scholar ]
  • Wang K, Zhang H, Bloss CS, Duvvuri V, Kaye W, Schork NJ, … Hakonarson H (2011). A genome-wide association study on common SNPs and rare CNVs in anorexia nervosa . Molecular Psychiatry , 16 , 949–959. [ PubMed ] [ Google Scholar ]
  • Watson HJ, Yilmaz Z, Thornton LM, Hübel C, Coleman JRI, Gaspar HA, … Bulik CM (2019). Genome-wide association study identifies eight risk loci and implicates metabo-psychiatric origins for anorexia nervosa . Nature Genetics , 51 , 1207–1214. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Wray NR, Lin T, Austin J, McGrath JJ, Hickie IB, Murray GK, & Visscher PM (2021). From basic science to clinical application of polygenic risk scores: A primer . JAMA Psychiatry , 78 , 101–109. [ PubMed ] [ Google Scholar ]
  • Wray NR, Ripke S, Mattheisen M, Trzaskowski M, Byrne EM, Abdellaoui A, … Sullivan PF (2018). Genome-wide association analyses identify 44 risk variants and refine the genetic architecture of major depression . Nature Genetics , 50 , 668–681. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Yao S, Larsson H, Norring C, Birgegard A, Lichtenstein P, D’Onofrio BM, … Kuja-Halkola R (2021). Genetic and environmental contributions to diagnostic fluctuation in anorexia nervosa and bulimia nervosa . Psychological Medicine , 51 , 62–69. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Yilmaz Z, Halvorsen M, Bryois J, Yu D, Thornton LM, Zerwas S, … Crowley JJ (2020). Examination of the shared genetic basis of anorexia nervosa and obsessive-compulsive disorder . Molecular Psychiatry , 25 , 2036–2046. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Yilmaz Z, Schaumberg K, Crowley J, Breen G, Bulik C, Micali N, & Zerwas S (2019). 64 - Anorexia nervosa and obsessive compulsive polygenic risk score: associations with adolescent eating disorder phenotypes . European Neuropsychopharmacology , 29 , S818–S819. [ Google Scholar ]
  • Yilmaz Z, Szatkiewicz JP, Crowley JJ, Ancalade N, Brandys MK, van Elburg A, … Bulik CM (2017). Exploration of large, rare copy number variants associated with psychiatric and neurodevelopmental disorders in individuals with anorexia nervosa . Psychiatric Genetics , 27 , 152–158. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Zhu Z, Zheng Z, Zhang F, Wu Y, Trzaskowski M, Maier R, … Yang J (2018). Causal associations between risk factors and common diseases inferred from GWAS summary data . Nature Communications , 9 , 224. [ PMC free article ] [ PubMed ] [ Google Scholar ]

IMAGES

  1. Genetic Disorder / Abnormality Paper Research Paper

    gene disorder research paper

  2. Genetic Disorders Research Project

    gene disorder research paper

  3. Genetic Disorder Research Project Packet by Beth Steckelberg

    gene disorder research paper

  4. Genetic Disorder Research and Doctor Report by Middle School Frizz

    gene disorder research paper

  5. Genetic Disorders Research Project

    gene disorder research paper

  6. GENETICS DISORDER RESEARCH PROJECT

    gene disorder research paper

VIDEO

  1. BIOLOGY INVESTIGATORY PROJECT ON GENETIC DISORDERS (class XII )🦠🧫🧬🔬

  2. Highlights from ASH with Dr. Akshat Jain

  3. Genetic Disorder

  4. அப்பா வயசு அதிகமாக அதிகமாக Single Gene Disorder

  5. Gene editing technology 🧬🤯

  6. Higher lithium levels in drinking water may raise autism risk

COMMENTS

  1. The genetic basis of disease

    The age at which symptoms first appear and the severity, varies considerably from one individual to another. Rett syndrome affects approximately 1 in 10000 females and is a single gene disorder involving the X-linked MECP2 gene. Due to the severity of symptoms, this usually arises as a de novo mutation. Boys with a similar mutation have a more ...

  2. Rare Genetic Diseases: Nature's Experiments on Human Development

    Rare genetic diseases are the result of a continuous forward genetic screen that nature is conducting on humans. Here, we present epistemological and systems biology arguments highlighting the importance of studying these rare genetic diseases. We contend that the expanding catalog of mutations in ∼4,000 genes, which cause ∼6,500 diseases ...

  3. The genetic basis of disease

    Genetics plays a role, to a greater or lesser extent, in all diseases. ... Essays Biochem. 2018 Dec 2;62(5):643-723. doi: 10.1042/EBC20170053. ... contribute to disease processes. This review explores the genetic basis of human disease, including single gene disorders, chromosomal imbalances, epigenetics, cancer and complex disorders, and ...

  4. Predicting Genetic Disorder and Types of Disorder Using Chain

    The analysis demonstrates that there are high chances of genetic disorders when the mother's age is between 20 and 60 years. When the mother's age is less than 20 years, the probability of a genetic disorder is low. A high chance of genetic disorder is associated with the age of the father begins between 20 and 70 years.

  5. Genome Sequencing for Diagnosing Rare Diseases

    We sequenced the genomes of 822 families (744 in the initial cohort and 78 in the replication cohort) and made a molecular diagnosis in 218 of 744 families (29.3%). Of the 218 families, 61 (28.0% ...

  6. Evidence for 28 genetic disorders discovered by combining ...

    It has previously been estimated that around 42-48% of patients with a severe developmental disorder (DD) have a pathogenic de novo mutation (DNM) in a protein-coding gene 1, 2. However, most of ...

  7. Evidence for 28 genetic disorders discovered by combining ...

    De novo mutations in protein-coding genes are a well-established cause of developmental disorders 1.However, genes known to be associated with developmental disorders account for only a minority of the observed excess of such de novo mutations 1,2.Here, to identify previously undescribed genes associated with developmental disorders, we integrate healthcare and research exome-sequence data ...

  8. Disease genetics

    DYRK1A gene linked to heart defects in Down syndrome. A study shows that congenital heart defects in Down syndrome are in part caused by increased dosage of the DYRK1A gene, which lies on ...

  9. An online compendium of treatable genetic disorders

    An online compendium of treatable genetic disorders. Am J Med Genet C Semin Med Genet. 2021 Mar;187 (1):48-54. doi: 10.1002/ajmg.c.31874. Epub 2020 Dec 22.

  10. Rare diseases, common challenges

    The genetics community has a particularly important part to play in accelerating rare disease research and contributing to improving diagnosis and treatment. Innovations in sequencing technology ...

  11. (PDF) The genetic basis of disease

    This review explores. the genetic basis of human disease, including single gene disorders, chromosomal imbal-. ances, epigenetics, cancer and complex disorders, and considers how our understanding ...

  12. Network analysis of genes and their association with diseases

    1. Introduction. The majority of human genetic disorders are rarely attributed to the activity of a single gene, but rather to the combinatorial activity of more than one gene (Cordell and Clayton, 2005, Goldstein, 2009).Disease-associated genes have been traditionally studied using family-based linkage methods (Oti et al., 2006).However, during the last decade, the deep-sequencing ...

  13. Human Molecular Genetics and Genomics

    Genomic research has evolved from seeking to understand the fundamentals of the human genetic code to examining the ways in which this code varies among people, and then applying this knowledge to ...

  14. Rare and undiagnosed diseases: From disease-causing gene identification

    The advent of NGS has changed the landscape of rare disease research [10]. Since the completion of the Human Genome Project, the cost of WGS has decreased, allowing researchers to conduct WGS studies. ... (an autosomal linear genetic disorder caused by inactivation of the transmembrane regulator CTFR gene mutation) can differentiate into ...

  15. Identification and bioinformatics analysis of a novel variant in the

    Abstract. HERC2-associated neurodevelopmental-disorders(NDD) encompass a cluster of medical conditions that arise from genetic mutations occurring within the HERC2 gene. These disorders can manifest a spectrum of symptoms that impact the brain and nervous system, including delayed psychomotor development, severe mental retardation, seizures and autistic features.

  16. Snijders Blok-Campeau syndrome: a novel neurodevelopmental genetic disorder

    A comprehensive review of the features of Snijders Blok-Campeau syndrome is provided to facilitate identification and genetic diagnostics of the syndrome. Snijders Blok-Campeau syndrome is a recently discovered genetic disorder characterized by childhood apraxia of speech, delays in intellectual development, and a plethora of other neurodevelopmental disorders (e.g., vision disorders, muscle ...

  17. Unraveling the Genetic Landscape of Neurological Disorders ...

    Genetic abnormalities play a crucial role in the development of neurodegenerative disorders (NDDs). Genetic exploration has indeed contributed to unraveling the molecular complexities responsible for the etiology and progression of various NDDs. The intricate nature of rare and common variants in NDDs contributes to a limited understanding of the genetic risk factors associated with them ...

  18. Genetics of neurodegenerative diseases: an overview

    Genetics plays an essential role in translational research, ultimately aiming to develop novel disease-modifying therapies for neurodegenerative disorders. We anticipate that individual genetic profiling will also be increasingly relevant in a clinical context, with implications for patient care in line with the proposed ideal of personalized ...

  19. Genetics

    Genome Sequencing for Diagnosing Rare Diseases. M.H. Wojcik and OthersN Engl J Med 2024;390:1985-1997. Genetic diagnosis of rare diseases is made through a variety of methods. This study gauged ...

  20. Genetic contributions to autism spectrum disorder

    Abstract. Autism spectrum disorder (autism) is a heterogeneous group of neurodevelopmental conditions characterized by early childhood-onset impairments in communication and social interaction alongside restricted and repetitive behaviors and interests. This review summarizes recent developments in human genetics research in autism ...

  21. PLOS Genetics

    Genomic analyses of Symbiomonas scintillans show no evidence for endosymbiotic bacteria but does reveal the presence of giant viruses. A multi-gene tree showed the three SsV genome types branched within highly supported clades with each of BpV2, OlVs, and MpVs, respectively. Image credit: pgen.1011218. 03/28/2024. Research Article.

  22. Reversible Nucleic Acid Storage in Deconstructable Glassy Polymer

    The rapid decline in DNA sequencing costs has fueled the demand for nucleic acid collection to unravel genomic information, develop treatments for genetic diseases, and track emerging biological threats. Current approaches to maintaining these nucleic acid collections hinge on continuous electricity for maintaining low-temperature and intricate cold-chain logistics. Inspired by the millennia ...

  23. A brief history of human disease genetics

    A primary goal of human genetics is to identify DNA sequence variants that influence biomedical traits, particularly those related to the onset and progression of human disease. Over the past 25 ...

  24. The genetics of bipolar disorder

    Bipolar disorder (BD) is one of the most heritable mental illnesses, but the elucidation of its genetic basis has proven to be a very challenging endeavor. Genome-Wide Association Studies (GWAS) have transformed our understanding of BD, providing the first reproducible evidence of specific genetic markers and a highly polygenic architecture ...

  25. What is a Genetic Mutation? Definition & Types

    A genetic mutation is a change to a gene's DNA sequence to produce something different. It creates a permanent change to that gene's DNA sequence. Genetic variations are important for humans to evolve, which is the process of change over generations. A sporadic genetic mutation occurs in one person.

  26. Gene therapy: advances, challenges and perspectives

    Currently, gene therapy is an area that exists predominantly in research laboratories, and its application is still experimental. Most trials are conducted in the United States, Europe, and Australia. The approach is broad, with potential treatment of diseases caused by recessive gene disorders (cystic fibrosis, hemophilia, muscular dystrophy ...

  27. Researchers find a single, surprising gene behind a disorder that

    Scientists have found the genetic root of a disorder that causes intellectual disability, which they estimate affects as many as one in 20,000 young people. ... research director of the South West ...

  28. Research Conducted at NIMH (Intramural Research Program)

    The Division of Intramural Research Programs (IRP) at the National Institute of Mental Health (NIMH) is the internal research division of the NIMH. The division plans and conducts basic, clinical, and translational research to advance understanding of the diagnosis, causes, treatment, and prevention of psychiatric disorders.

  29. Neurofibromatosis type 1

    Neurofibromatosis type 1 (NF1) is a genetic condition that causes changes in skin pigment and tumors on nerve tissue. Skin changes include flat, light brown spots and freckles in the armpits and groin. Tumors can grow anywhere in the nervous system, including the brain, spinal cord and nerves. NF1 is rare. About 1 in 2,500 is affected by NF1.

  30. Genetics of eating disorders in the genome-wide era

    Table 1 overviews the topics covered in this paper and synthesizes the current state of knowledge. Open in a separate window. Figure 1. ... Genetics research in eating disorders has transitioned to the genome-wide era, further, genome-wide association studies (GWAS) have become the dominant approach to identify genetic risk variants associated ...