Biallelic mutations in valyl-tRNA synthetase gene VARS are associated with a progressive neurodevelopmental epileptic encephalopathy

Jennifer Friedman, Desiree E Smith, Mahmoud Y Issa, Valentina Stanley, Rengang Wang, Marisa I Mendes, Meredith S Wright, Kristen Wigby, Amber Hildreth, John R Crawford, Alanna E Koehler, Shimul Chowdhury, Shareef Nahas, Liting Zhai, Zhiwen Xu, Wing-Sze Lo, Kiely N James, Damir Musaev, Andrea Accogli, Kether Guerrero, Luan T Tran, Tarek E I Omar, Tawfeg Ben-Omran, David Dimmock, Stephen F Kingsmore, Gajja S Salomons, Maha S Zaki, Geneviève Bernard, Joseph G Gleeson, Jennifer Friedman, Desiree E Smith, Mahmoud Y Issa, Valentina Stanley, Rengang Wang, Marisa I Mendes, Meredith S Wright, Kristen Wigby, Amber Hildreth, John R Crawford, Alanna E Koehler, Shimul Chowdhury, Shareef Nahas, Liting Zhai, Zhiwen Xu, Wing-Sze Lo, Kiely N James, Damir Musaev, Andrea Accogli, Kether Guerrero, Luan T Tran, Tarek E I Omar, Tawfeg Ben-Omran, David Dimmock, Stephen F Kingsmore, Gajja S Salomons, Maha S Zaki, Geneviève Bernard, Joseph G Gleeson

Abstract

Aminoacyl-tRNA synthetases (ARSs) function to transfer amino acids to cognate tRNA molecules, which are required for protein translation. To date, biallelic mutations in 31 ARS genes are known to cause recessive, early-onset severe multi-organ diseases. VARS encodes the only known valine cytoplasmic-localized aminoacyl-tRNA synthetase. Here, we report seven patients from five unrelated families with five different biallelic missense variants in VARS. Subjects present with a range of global developmental delay, epileptic encephalopathy and primary or progressive microcephaly. Longitudinal assessment demonstrates progressive cortical atrophy and white matter volume loss. Variants map to the VARS tRNA binding domain and adjacent to the anticodon domain, and disrupt highly conserved residues. Patient primary cells show intact VARS protein but reduced enzymatic activity, suggesting partial loss of function. The implication of VARS in pediatric neurodegeneration broadens the spectrum of human diseases due to mutations in tRNA synthetase genes.

Conflict of interest statement

D.D. is a consultant for Biomarin, Ichorian, and Complete Genomics. J.F. holds shares in Illumina and Spouse is Founder and Principal of Friedman Bioventure, which holds a variety of publicly traded and private biotechnology interests. D.D. is on the Scientific Advisory Board for Audentes Therapeutics. The remaining authors declare no competing interests.

Figures

Fig. 1
Fig. 1
VARS mutations lead to a neurodevelopmental disorder with microcephaly, seizures, and cortical atrophy. a Pedigrees showing consanguinity in three families (double bars). Family 3007 showed remote consanguinity. Seven affected individuals were identified (black), and one individual was likely affected but DNA was not available for genotyping (3439-IV-1). Multiple miscarriages in family GB31 suggested fetal demise. b Available axial or midline sagittal brain MRIs demonstrated diffuse cortical atrophy (yellow star) and thinned corpus callosum (yellow arrow) compared to control. The brain MRI in 3007-II-1 was acquired at 6 months of age, and thus the patient was probably too young to observe these features. Two individuals from family 3439 had follow-up MRIs that demonstrated progressive cortical atrophy. D—days; M—months; Y—years. c Major domains of the 1264 amino acid protein, showing the GST, aminoacylation (including HIGH, Editing and KMSKS domains) and anticodon-binding domains. Disease-associated variants described in this study (red) and those from Karaca et al. (black) are noted. Variants were in the aminoacylation or in or near the anticodon-binding domains, suggesting altered function. d Sequence conservation of mutated amino acids. The variants were either precisely conserved or maintain charge/polarity, with the exception of NM_006295.2:c.2653C>T p.(Leu885Phe), which is absent in T. thermophiles
Fig. 2
Fig. 2
Predicted mutational impact on VARS structure and tRNA binding. a Locations of the mutational sites on the structural model of human VARS-tRNAVal complex (VARS green, tRNAVal orange, mutational sites magenta). Residues Pro661, Arg947, and Arg1119 are closer to the tRNA molecule (<6 Å), while Ala692 and Arg1058 have a longer distance of >15 Å from tRNA. bf Pair-wise comparisons between the wild-type (left) and mutant (right) residues for predicted changes in local contacts with tRNA or other amino acids. Amino acids are indicated by letter and number (e.g., Pro661), tRNA bases are indicated by letter and dash and number (e.g., A-75). Hydrogen bonds are presented as yellow dashed lines. b Residue Pro661 forms a hydrogen bond with Val670 and has a distance of ~4.5 Å from A-75 of the tRNA. c Residue Ala692 forms hydrogen bonds with adjacent residues Ala689, Arg695, Gln696, and Asp697. Mutation to proline predicts abolished contacts with Ala689 and disrupted helical structure. d Residue Arg947 is close to C-37 and predicts hydrogen-bonding interaction that is abolished by the Arg947H mutation. This mutation does not predict altered hydrogen-bonding with Ile943 and Asn951. e Residue Arg1058 forms hydrogen bonds with Asp1054, Ser1061 and Pro1090. The NM_006295.2:c.3173G>A p.(Arg1058Gln) mutation predicts elimination one of the two possible bonds with Asp1054. f The side chain of Arg1119 forms a hydrogen bond with C-39, predicted to be abolished by the mutation
Fig. 3
Fig. 3
VARS patient cells show normal VARS protein level, but reduced enzymatic ARS activity. a Western blot of patient or control cells showing retained VARS protein level in affected individuals (2937-III-1A and 3308-III-1A), carrier parents (2937-II-1F and 3308-II-2M), and controls (3311-III-2C, 3838-III-1C, and 4936-III-1C). b VARS, LARS (leucine ARS) and RARS (arginine ARS) aminoacylation activity. VARS activity was specifically reduced in patient cells, normalized to control cells at 100% (from n = 3 replicates, error bars = S.E.M.). c VARS activity in patients compared with carrier parents and controls. Black: patients, light gray: parents of affecteds (obligate carriers of the VARS mutation), dark gray: healthy controls (from n = 3 replicates, error bars = S.E.M). Differences between patients and controls or carriers were calculated based upon the average of the group (P < 0.0001 for both comparisons, One tailed Student’s t-test). Full gel in Supplementary Fig. 2

References

    1. Sobreira NLM, et al. Matchmaker exchange. Curr. Protoc. Human. Genet. 2017;95:9.31.1–9.31.15. doi: 10.1002/cphg.50.
    1. Boycott KM, et al. International cooperation to enable the diagnosis of all rare genetic diseases. Am. J. Hum. Genet. 2017;100:695–705. doi: 10.1016/j.ajhg.2017.04.003.
    1. Karaca E, et al. Genes that affect brain structure and function identified by rare variant analyses of Mendelian neurologic disease. Neuron. 2015;88:499–513. doi: 10.1016/j.neuron.2015.09.048.
    1. Alazami AM, et al. Accelerating novel candidate gene discovery in neurogenetic disorders via whole-exome sequencing of prescreened multiplex consanguineous families. Cell Rep. 2015;10:148–161. doi: 10.1016/j.celrep.2014.12.015.
    1. Najmabadi H, et al. Deep sequencing reveals 50 novel genes for recessive cognitive disorders. Nature. 2011;478:57–63. doi: 10.1038/nature10423.
    1. Schimmel P. The emerging complexity of the tRNA world: mammalian tRNAs beyond protein synthesis. Nat. Rev. Mol. Cell Biol. 2018;19:45–58. doi: 10.1038/nrm.2017.77.
    1. Antonellis A, Green ED. The role of aminoacyl-tRNA synthetases in genetic diseases. Annu. Rev. Genom. Hum. Genet. 2008;9:87–107. doi: 10.1146/annurev.genom.9.081307.164204.
    1. Santos-Cortez RL, et al. Mutations in KARS, encoding lysyl-tRNA synthetase, cause autosomal-recessive nonsyndromic hearing impairment DFNB89. Am. J. Hum. Genet. 2013;93:132–140. doi: 10.1016/j.ajhg.2013.05.018.
    1. Puffenberger EG, et al. Genetic mapping and exome sequencing identify variants associated with five novel diseases. PLoS. One. 2012;7:e28936. doi: 10.1371/journal.pone.0028936.
    1. Pierce SB, et al. Mutations in mitochondrial histidyl tRNA synthetase HARS2 cause ovarian dysgenesis and sensorineural hearing loss of Perrault syndrome. Proc. Natl Acad. Sci. USA. 2011;108:6543–6548. doi: 10.1073/pnas.1103471108.
    1. Baertling F, et al. Neonatal encephalocardiomyopathy caused by mutations in VARS2. Metab. Brain. Dis. 2017;32:267–270. doi: 10.1007/s11011-016-9890-2.
    1. Taylor RW, et al. Use of whole-exome sequencing to determine the genetic basis of multiple mitochondrial respiratory chain complex deficiencies. JAMA. 2014;312:68–77. doi: 10.1001/jama.2014.7184.
    1. Diodato D, et al. VARS2 and TARS2 mutations in patients with mitochondrial encephalomyopathies. Hum. Mutat. 2014;35:983–989. doi: 10.1002/humu.22590.
    1. MacArthur DG, et al. Guidelines for investigating causality of sequence variants in human disease. Nature. 2014;508:469–476. doi: 10.1038/nature13127.
    1. Akizu N, et al. Biallelic mutations in SNX14 cause a syndromic form of cerebellar atrophy and lysosome-autophagosome dysfunction. Nat. Genet. 2015;47:528–534. doi: 10.1038/ng.3256.
    1. Thiffault I, et al. Recessive mutations in POLR1C cause a leukodystrophy by impairing biogenesis of RNA polymerase III. Nat. Commun. 2015;6:7623. doi: 10.1038/ncomms8623.
    1. Sherlin LD, Perona JJ. tRNA-dependent active site assembly in a class I aminoacyl-tRNA synthetase. Structure. 2003;11:591–603. doi: 10.1016/S0969-2126(03)00074-1.
    1. Fukai S, et al. Mechanism of molecular interactions for tRNA(Val) recognition by valyl-tRNA synthetase. RNA. 2003;9:100–111. doi: 10.1261/rna.2760703.
    1. Fukai S, et al. Structural basis for double-sieve discrimination of L-valine from L-isoleucine and L-threonine by the complex of tRNA(Val) and valyl-tRNA synthetase. Cell. 2000;103:793–803. doi: 10.1016/S0092-8674(00)00182-3.
    1. Nakayama T, et al. Deficient activity of alanyl-tRNA synthetase underlies an autosomal recessive syndrome of progressive microcephaly, hypomyelination, and epileptic encephalopathy. Hum. Mutat. 2017;38:1348–1354. doi: 10.1002/humu.23250.
    1. Oliveira R, et al. Lethal neonatal LTBL associated with biallelic EARS2 variants: case report and review of the reported neuroradiological features. JIMD Rep. 2017;33:61–68. doi: 10.1007/8904_2016_581.
    1. Coughlin CR, II, et al. Mutations in the mitochondrial cysteinyl-tRNA synthase gene, CARS2, lead to a severe epileptic encephalopathy and complex movement disorder. J. Med. Genet. 2015;52:532–540. doi: 10.1136/jmedgenet-2015-103049.
    1. Vanlander AV, et al. Two siblings with homozygous pathogenic splice-site variant in mitochondrial asparaginyl-tRNA synthetase (NARS2) Hum. Mutat. 2015;36:222–231. doi: 10.1002/humu.22728.
    1. Guo M, Schimmel P. Essential nontranslational functions of tRNA synthetases. Nat. Chem. Biol. 2013;9:145–153. doi: 10.1038/nchembio.1158.
    1. Bec G, Kerjan P, Waller JP. Reconstitution in vitro of the valyl-tRNA synthetase-elongation factor (EF) 1 beta gamma delta complex. Essential roles of the NH2-terminal extension of valyl-tRNA synthetase and of the EF-1 delta subunit in complex formation. J. Biol. Chem. 1994;269:2086–2092.
    1. Negrutskii BS, Shalak VF, Kerjan P, El’skaya AV, Mirande M. Functional interaction of mammalian valyl-tRNA synthetase with elongation factor EF-1ɑ in the complex with EF-1H. J. Biol. Chem. 1999;274:4545–4550. doi: 10.1074/jbc.274.8.4545.
    1. Lo WS, et al. Human tRNA synthetase catalytic nulls with diverse functions. Science. 2014;345:328–332. doi: 10.1126/science.1252943.
    1. Meyer-Schuman R, Antonellis A. Emerging mechanisms of aminoacyl-tRNA synthetase mutations in recessive and dominant human disease. Hum. Mol. Genet. 2017;26:R114–R127. doi: 10.1093/hmg/ddx231.
    1. Wolf NI, et al. Mutations in RARS cause hypomyelination. Ann. Neurol. 2014;76:134–139. doi: 10.1002/ana.24167.
    1. Taft RJ, et al. Mutations in DARS cause hypomyelination with brain stem and spinal cord involvement and leg spasticity. Am. J. Hum. Genet. 2013;92:774–780. doi: 10.1016/j.ajhg.2013.04.006.
    1. Mendes MI, et al. Bi-allelic mutations in EPRS, encoding the glutamyl-prolyl-aminoacyl-tRNA synthetase, cause a hypomyelinating leukodystrophy. Am. J. Hum. Genet. 2018;102:1–9. doi: 10.1016/j.ajhg.2018.02.011.
    1. Musante L. Mutations of the aminoacyl-tRNA-synthetases SARS and WARS2 are implicated in the etiology of autosomal recessive intellectual disability. Hum. Mutat. 2017;38:621–636. doi: 10.1002/humu.23205.
    1. Safka Brozkova D, et al. Loss of function mutations in HARS cause a spectrum of inherited peripheral neuropathies. Brain. 2015;138:2161–2172. doi: 10.1093/brain/awv158.
    1. Vester A, et al. A loss-of-function variant in the human histidyl-tRNA synthetase (HARS) gene is neurotoxic in vivo. Hum. Mutat. 2013;34:191–199. doi: 10.1002/humu.22210.
    1. McLaughlin HM, et al. A recurrent loss-of-function alanyl-tRNA synthetase (AARS) mutation in patients with Charcot-Marie-Tooth disease type 2 N (CMT2N) Hum. Mutat. 2012;33:244–253. doi: 10.1002/humu.21635.
    1. Latour P, et al. A major determinant for binding and aminoacylation of tRNA(Ala) in cytoplasmic Alanyl-tRNA synthetase is mutated in dominant axonal Charcot-Marie-Tooth disease. Am. J. Hum. Genet. 2010;86:77–82. doi: 10.1016/j.ajhg.2009.12.005.
    1. Jordanova A, et al. Disrupted function and axonal distribution of mutant tyrosyl-tRNA synthetase in dominant intermediate Charcot-Marie-Tooth neuropathy. Nat. Genet. 2006;38:197–202. doi: 10.1038/ng1727.
    1. Antonellis A, et al. Glycyl tRNA synthetase mutations in Charcot-Marie-Tooth disease type 2D and distal spinal muscular atrophy type V. Am. J. Hum. Genet. 2003;72:1293–1299. doi: 10.1086/375039.
    1. van Meel E, et al. Rare recessive loss-of-function methionyl-tRNA synthetase mutations presenting as a multi-organ phenotype. Bmc. Med. Genet. 2013;14:106. doi: 10.1186/1471-2350-14-106.
    1. Casey JP, et al. Identification of a mutation in LARS as a novel cause of infantile hepatopathy. Mol. Genet. Metab. 2012;106:351–358. doi: 10.1016/j.ymgme.2012.04.017.
    1. Feinstein M, et al. Pelizaeus-Merzbacher-like disease caused by AIMP1/p43 homozygous mutation. Am. J. Hum. Genet. 2010;87:820–828. doi: 10.1016/j.ajhg.2010.10.016.
    1. Armstrong L, et al. AIMP1 deficiency presents as a cortical neurodegenerative disease with infantile onset. Neurogenetics. 2014;15:157–159. doi: 10.1007/s10048-014-0411-3.
    1. Schaffer AE, et al. CLP1 founder mutation links tRNA splicing and maturation to cerebellar development and neurodegeneration. Cell. 2014;157:651–663. doi: 10.1016/j.cell.2014.03.049.
    1. Karaca E, et al. Human CLP1 mutations alter tRNA biogenesis, affecting both peripheral and central nervous system function. Cell. 2014;157:636–650. doi: 10.1016/j.cell.2014.02.058.
    1. Cassandrini D, et al. Pontocerebellar hypoplasia: clinical, pathologic, and genetic studies. Neurology. 2010;75:1459–1464. doi: 10.1212/WNL.0b013e3181f88173.
    1. Budde BS, et al. tRNA splicing endonuclease mutations cause pontocerebellar hypoplasia. Nat. Genet. 2008;40:1113–1118. doi: 10.1038/ng.204.
    1. Novarino G, et al. Exome sequencing links corticospinal motor neuron disease to common neurodegenerative disorders. Science. 2014;343:506–511. doi: 10.1126/science.1247363.
    1. Farnaes L. et al. Rapid whole-genome sequencing identifies a novel GABRA1 variant associated with West syndrome. Cold Spring Harb. Mol. Case Stud. 3, a001776 (2017).
    1. Kelley LA, Mezulis S, Yates CM, Wass MN, Sternberg MJ. The Phyre2 web portal for protein modeling, prediction and analysis. Nat. Protoc. 2015;10:845–858. doi: 10.1038/nprot.2015.053.
    1. Schneidman-Duhovny D, Inbar Y, Nussinov R, Wolfson HJ. PatchDock and SymmDock: servers for rigid and symmetric docking. Nucleic Acids Res. 2005;33:W363–W367. doi: 10.1093/nar/gki481.

Source: PubMed

3
Abonneren