3,836
Views
17
CrossRef citations to date
0
Altmetric
Reviews

Atypical and ultra-rare Usher syndrome: a review

, , ORCID Icon, , , , , , , ORCID Icon & ORCID Icon show all
Pages 401-412 | Received 27 Dec 2019, Accepted 15 Mar 2020, Published online: 06 May 2020

ABSTRACT

Usher syndrome has classically been described as a combination of hearing loss and rod-cone dystrophy; vestibular dysfunction is present in many patients. Three distinct clinical subtypes were documented in the late 1970s. Genotyping efforts have led to the identification of several genes associated with the disease. Recent literature has seen multiple publications referring to “atypical” Usher syndrome presentations. This manuscript reviews the molecular etiology of Usher syndrome, highlighting rare presentations and molecular causes. Reports of “atypical” disease are summarized noting the wide discrepancy in the spectrum of phenotypic deviations from the classical presentation. Guidelines for establishing a clear nomenclature system are suggested.

Introduction

Usher syndrome (USH) is the most common hereditary form of deaf-blindness. It has an autosomal recessive inheritance pattern and is characterized by the combination of sensorineural hearing loss (SNHL), rod-cone dystrophy, and variable vestibular dysfunction (Citation1). Early sources reported a prevalence of 3.6 to 6.2 per 100,000 (Citation2Citation6). However, the prevalence of USH may be as high as 16.7 per 100,000 (Citation7).

Attempts at description of USH clinical subtypes date to the late 1970s (Citation8). Currently, USH has three recognized clinical subtypes distinguished by differences in severity and onset of hearing loss and the presence of vestibular dysfunction. The shared ophthalmic manifestation of all three subtypes is a retinal degeneration classified as rod-cone dystrophy or retinitis pigmentosa (RP). Type 1 (USH1) presents with profound congenital SNHL, peripheral vestibular areflexia, and infantile onset of retinitis pigmentosa (Citation9). Type 2 (USH2) results in mild to severe congenital SNHL and retinitis pigmentosa, but not vestibular dysfunction (Citation9). Type 3 (USH3) results in progressive SNHL, retinitis pigmentosa, and variable vestibular dysfunction (Citation10Citation12).

A poorly defined clinical subtype called atypical Usher syndrome (atypical USH) has emerged to include USH phenotypes that do not meet the canonical criteria for USH1, USH2, or USH3. Although reports of atypical USH vary, several of the cases described in the literature involve a combination of hearing loss and cone-rod dystrophy (CRD). Another commonly cited cause for atypical USH is related to the degree of severity of classical clinical findings, usually milder presentations. This paper reviews the literature on USH with a focus on summarizing data on atypical USH and the rare genetic causes putatively linked to USH (hitherto referred to as Ultra-rare USH).

Genetic heterogeneity

Usher syndrome is genetically heterogenous. USH1 can be caused by pathogenic variants in one of at least five genes: MYO7A (myosin VIIA) (Citation13), USH1 C (harmonin) (Citation14,Citation15), CDH23 (cadherin 23) (Citation16,Citation17), PCDH15 (protocadherin 15) (Citation18,Citation19), or USH1 G (sans) (Citation20). USH2 can be caused by pathogenic variants in one of the three genes: USH2A (usherin) (Citation21), ADGRV1 (adhesion G-protein coupled receptor V1) (Citation22), or WHRN (whirlin) (Citation23). USH3 can be caused by pathogenic variants in CLRN1 (clarin 1) (Citation24,Citation25).

USH genes encode proteins with a variety of molecular functions, including a motor protein (myosin VIIA), scaffold proteins (harmonin, sans, and whirlin), transmembrane proteins (usherin and clarin 1), adhesion proteins (cadherin 23 and protocadherin 15), and a G-protein coupled receptor (adhesion G-protein coupled receptor v1). USH proteins can interact to produce what is termed the Usher interactome (Citation26,Citation27). The Usher interactome is important for the function of inner ear hair cells and photoreceptors in the retina. Although a full discussion of the various functions of well-established USH proteins is beyond the scope of this manuscript, the reader is referred to a comprehensive review by Cosgrove and Zallocchi (Citation26).

The nine USH genes discussed above underlie the vast majority of molecularly diagnosed cases (Citation24,Citation28Citation38). However, current methods fail to yield a definitive molecular diagnosis in some USH patients, and bi-allelic pathogenic variants are generally discovered in less than 80% of patients using targeted or whole-exome sequencing methods (Citation39Citation43). Utilizing a combination of methods or a more comprehensive approach to genetic testing may increase the detection of bi-allelic pathogenic variants to above 90% (Citation44,Citation45). For example, including genes implicated in both non-syndromic hearing loss and retinal degenerations may reveal rare cases with a clinical presentation similar to that of USH but with bi-allelic variants in two different genes (Citation45,Citation46). Nonetheless, rare variants of hitherto unidentified genes for USH should be considered, noting that these likely account for a minority of patients.

Additional associations of variants in genes previously not described as causative of USH are continuously being proposed; we will refer to these as ultra-rare USH genes throughout this paper. In the past, several candidate genes, including MYO15A, SLC4A, and VEZT, have been proposed and then refuted as potential USH genes (Citation34,Citation40). Furthermore, several genes associated with USH have been proposed more recently. These include PDZD7 (Citation47), HARS (Citation48), ABHD12 (Citation49), CIB2 (Citation50), CEP250 (Citation51), CEP78 (Citation52), ARSG (Citation53), and ESPN (Citation54). Pathogenic variants in these genes likely represent rare causes of USH, and several of them were reported in association with atypical USH. Here, we review reports of these ultra-rare USH genes and atypical USH. For this review, “atypical USH” is based on the description of the authors of cited studies. This includes atypical USH clinical presentations with confirmed pathogenic variants in known USH causative genes. Between September 2018 and November 2019, we searched PubMed for terms including “Usher syndrome,” “Atypical Usher syndrome,” “Atypical USH,” and for each of the genes already identified as causative of USH. RetNet (https://sph.uth.edu/retnet/) and Human Genome Mutation Database (Citation55) were utilized to identify additional ultra-rare USH genes and relevant references. We also reviewed the pertinent references in identified papers and reviews on USH. Additional search methodology including accessing Scopus and Web of Science was used to identify relevant literature.

Ultra-rare USH genes

We identified eight putative ultra-rare USH genes through our literature search. The relevant references which associate each of these ultra-rare genes to USH are summarized in . Information on other important properties of the encoded proteins such as function, localization, and interactions with the USH interactome are summarized in .

Table 1. References for ultra-rare USH genes.

Table 2. Summary of ultra-rare USH proteins function, interactions, localization, and animal models.

One of these genes, PDZD7 (PDZ domain containing 7), is a 23.3-kb, 16-exon gene located on chromosome 10, encoding a scaffold protein (Citation62). The high sequence similarity of PDZD7 with USH1C and WHRN led Schneider et al. to first propose it as an USH gene (Citation62). Subsequently, in 2010, Ebermann et al. identified the first clinical cases of USH associated with pathogenic variants in PDZD7 (Citation47). They described two families in which there is evidence that PDZD7 contributes to the USH phenotype. In one family, they observed that the presence of a mono-allelic variant in PDZD7 appears to modify the severity of the retinal phenotype in the presence of homozygous pathogenic variants in USH2A. In the other family, mono-allelic variants in ADGRV1 and PDZD7 appear to contribute to digenic inheritance (Citation47). There is evidence that PDZD7, USH2A, ADGRV1, and WHRN interact to form an USH2 complex both in vitro and in vivo (Citation67,Citation68,Citation109). The localization and function of PDZD7 in the inner ear have been well characterized but are less well understood in the retina. For example, knockout or knockdown of gene function in animal models results in disorganization of the stereocilia, deafness in mice, and a reduced startle reflex and circling in zebrafish (Citation47,Citation68). However, there is evidence that the USH2 complex can assemble without PDZD7 in photoreceptors, unlike in inner ear hair cells where it is an obligate component (Citation68). A summary of the localization of the PDZD7 protein, interactions with other USH proteins, and animal models can be found in .

The next ultra-rare USH gene to be proposed was HARS (histidyl-tRNA synthetase), a 13-exon, 17.4-kb gene located on chromosome 5 (Citation110,Citation111). The HARS gene product charges tRNA molecules with histidine amino acids for protein translation (Citation75). In 2012, Puffenberger et al. identified a homozygous pathogenic variant in HARS associated with an USH3 phenotype (Citation48). Tiwari et al. also described compound heterozygous pathogenic variants in HARS in a patient with an unspecified type of Usher syndrome in 2016 (Citation56). The specific variant described by Puffenberger et al. was further characterized in vitro by Abbott et al., who observed that it caused a reduction in thermal stability of the protein (Citation78). The localization of HARS in the inner ear and retina and interactions with other USH proteins are unknown and require further investigation.

ABHD12 (α/β-hydrolase domain containing 12) is a membrane-embedded serine hydrolase (Citation113,Citation114). Its substrates include endocannabinoid transmitter 2-arachidonylglycerol (2-AG) and signaling lipids (lysophosphatidylserines, LPS) (Citation83,Citation114). Both Eisenberger et al. and Yoshimura et al. identified bi-allelic ABHD12 pathogenic variants in patients initially presenting with USH3-like symptoms (Citation49,Citation57). However, it is important to note that both sets of authors ultimately diagnosed the patients with a different condition comprising polyneuropathy, hearing loss, ataxia, retinitis pigmentosa, and early-onset cataract (PHARC). Nonetheless, ABHD12 has been considered by some to be an USH gene, leading it to be included on panels of genes utilized for Next Generation Sequencing (NGS) specifically targeted at USH. For example, in 2018, Sun et al. performed a screening of 119 clinically diagnosed USH patients and identified bi-allelic pathogenic variants in ABHD12 in one individual (Citation42). Information about whether this patient had undergone a neurological exam was not presented. Abhd12-null mice display many of the phenotypic characteristics of PHARC (Citation83). The localization of ABHD12 in the inner ear and retina and possible interactions with USH proteins require further investigation.

CIB2 (calcium and integrin-binding family member 2) was proposed as another potential USH gene. Previously known as KIP2 (kinase interacting protein 2), CIB2 is located on chromosome 15 and encodes an EF-domain containing protein that binds both calcium and integrin α7b (Citation115,Citation116). In 2012, Riazuddin et al. reported a homozygous pathogenic variant of CIB2 in USH1 patients (Citation50) which affected the function of CIB2 as a calcium sensor. Vallone et al. later demonstrated that, in physiological conditions, CIB2 can function as a magnesium sensor and the variant described by Riazuddin et al. affects this capacity (Citation88). Studies show that knockdown of Cib2 in mice eliminates mechanotransduction in cochlear hair cells (Citation89,Citation90).

CEP250 encodes CEP250 (also known as centrosomal protein 250 or CNAP1), a member of the CEP family of centrosome-associated proteins (Citation92). There have been three separate papers reporting bi-allelic variants in CEP250 linked to retinal degeneration and hearing loss. In 2014, Khateb et al. identified homozygous pathogenic variants in both CEP250 and PCARE in a family diagnosed with atypical USH (Citation51). This family had multiple-affected individuals, and individuals with bi-allelic pathogenic variants in both genes had more severe phenotypes (Citation51). In 2018, Kubota et al. identified compound heterozygous pathogenic variants in CEP250 in patients diagnosed with mild CRD and SNHL (Citation58). Lastly, Fuster-Garcia et al. recently identified compound heterozygous CEP250 pathogenic variants in a patient presenting with similar ocular symptoms and progressive hearing loss (Citation59). CEP250 is important for centriole-centriole cohesion of chromosomes and has been shown to promote cilia formation (Citation91,Citation117). Its localization is described in . Interactions with other USH proteins and the function of CEP250 in the ear and eye are areas for further research.

Another member of the CEP family, CEP78 (centrosomal protein 78), has also been linked to USH. As a ciliary/centrosomal protein, it is important for the cell cycle (Citation118). Three different reports have associated CEP78 with retinal degeneration and hearing loss (Citation52,Citation60,Citation61). In 2016, both Namburi et al. and Nikopolous et al. identified bi-allelic CEP78 pathogenic variants in patients clinically diagnosed with CRD and hearing loss (Citation52,Citation60). The following year, Fu et al. reported bi-allelic CEP78 pathogenic variants in patients with CRD and progressive hearing loss who were clinically diagnosed with atypical USH (Citation61). While a full description of localization patterns can be found in , work performed by both Namburi et al. and Nikopolous et al. indicates that, in the retina, CEP78 is primarily localized to the photoreceptor connecting cilium, and this localization is stronger in cones than rods (Citation52,Citation60). Studies involving adult human tissue samples indicate that CEP78 mRNA is also present in the inner ear (Citation60,Citation119). Interestingly, tandem affinity purification has been used to show that CEP78 interacts with CEP250 (Citation93).

Another putative USH gene is ARSG (arylsulfatase G), which encodes a member of a class of enzymes called sulfatases, which are responsible for hydrolyzing ester sulfate bonds and which are implicated in a wide variety of biochemical processes (Citation101). It catalyzes an important step in the breakdown of heparan sulfate (Citation97). Khateb et al. identified a homozygous pathogenic variant in ARSG in patients clinically diagnosed with atypical USH presenting as a retinal degeneration involving a ring scotoma and moderate to severe SNHL (Citation53). It is noteworthy that the patients were also found to have a normal metabolic and neurological function. The authors also demonstrated that the ARSG variant they identified nearly abolishes all enzyme function and that ARSG is expressed in the human retina. Homozygous Arsg knockout mice displays retinal degeneration as well as behavioral dysfunction indicative of systemic effects (Citation97,Citation98).

Lastly, ESPN (espin) has recently been associated with USH. ESPN, which has several known alternative splice isoforms, encodes an actin-bundling protein important in various neurosensory cell types, including those in the retina and inner ear (Citation106). Ahmed et al. identified an in-frame deletion in ESPN as the cause of a phenotype described as atypical USH1 in a Pakistani family (Citation54). The phenotype included delayed dark adaptation with irregular retinal contour, temporal flecks, and optic nerve pallor. Electroretinography showed a mild reduction in scotopic responses with preserved photopic responses. Upon detailed review of the presented data, we felt that retinal clinical findings and electroretinography are atypical for USH. ESPN is an established cause of non-syndromic hearing loss DFNB36 (Citation104), but this was the first time it has been associated with USH.

Atypical USH

Our literature search identified a variety of articles describing atypical USH, the findings of which are summarized in . These papers generally either describe variants in well-characterized USH genes associated with atypical clinical presentations or they describe atypical USH associated with ultra-rare USH genes, such as the ones discussed above.

Table 3. Atypical USH references.

MYO7A is one of the well-characterized USH1 genes which has been linked to atypical USH. Although all these cases involved RP and SNHL, they were classified as atypical USH because the symptoms deviated from the expected clinical presentation of USH1. Many involve mild RP or progressive SNHL. For example, Liu et al. reported two siblings with bilateral progressive hearing loss and mild RP who were compound heterozygous for pathogenic variants in MYO7A (Citation120). Additional examples of atypical USH are included in . Of note, not all instances of unusual clinical presentations are defined as atypical USH by the authors (Citation132,Citation133).

USH2A has also been associated with atypical USH. Many of these cases involve progressive hearing loss. For example, Liu et al. identified the common c.2299delG USH2A pathogenic variant in four unrelated patients with atypical USH (Citation122). This was in a homozygous state in three of the patients and a heterozygous state in one, with the second disease-causing allele variant unknown. These patients had progressive SNHL and RP, and one patient had vestibular dysfunction as well. Additional examples are described in . As with MYO7A, not all descriptions of similarly unusual phenotypes are classified as atypical USH by the authors (Citation134). There has also been discussion in the literature as to whether progressive hearing loss is necessarily an atypical natural history for some patients with Usher syndrome caused by USH2A (Citation135–138).

Other well-established USH genes described in cases of atypical USH include CDH23 and SANS. These cases, which involved unusual severity or progression of symptoms, also varied in terms of classification as atypical USH. For example, each of the two papers which first identified pathogenic variants of CDH23 as a cause of USH also described a family with a mild form of RP (Citation16,Citation17), but only one of the papers specifically defined this phenotype as atypical USH (Citation16). Similarly, bi-allelic pathogenic variants in SANS have been reported in conjunction with unusual phenotypes characterized by mild RP with or without vestibular dysfunction (Citation130,Citation131), but only the former was described as atypical USH (Citation130).

Several of the previously discussed ultra-rare causes of USH have also been described in association with atypical USH. All of these reports are summarized in . Unlike well-established genetic phenotypes, these cases sometimes present as non-RP retinal degenerations, including CRD. For example, Fu et al. reported either of two different homozygous variants in CEP78 in association with CRD and SNHL in patients clinically diagnosed with atypical USH (Citation61). Authors also differ on the use of the atypical USH diagnosis in these cases, with some simply describing the clinical presentation as CRD and SNHL (Citation52,Citation58,Citation60) or an “Usher-like” phenotype (Citation59).

Lastly, it has become more common for authors to include patients clinically diagnosed with atypical USH as a small subset of a larger cohort of USH patients screened for USH pathogenic variants. Pathogenic variants in many USH genes associated with atypical USH have been reported, including MYO7A, USH2A, CDH23, ADGRV1, and ABHD12. However, there is little clinical information available for these patients. The results of these studies are also included in .

Discussion

The literature shows great variability in the USH phenotype. Although most USH phenotypes meet the current criteria for USH1, USH2, or USH3, a significant number of phenotypes do not meet criteria for any of them, whether in terms of their phenotype or genotype. Clear and established nomenclature guidelines are now needed for patients with these atypical genotypes and phenotypes. Additionally, the routine implementation of massively parallel sequencing has greatly increased the number of variants being discovered. As a result, it is imperative for the community of clinicians and scientists treating and investigating USH to establish clear criteria for the diagnosis of USH and characterization of potential USH genes.

In pursuit of this goal, we raise a few points of consideration. First, it is important to consider the typical clinical definitions for each subtype of USH. Currently, it is not uncommon within the literature for a similar clinical presentation described as USH1 or USH2 in one case to be described as atypical USH in another case. For example, this frequently occurs with cases of progressive hearing loss in USH2. It is important to establish clear criteria for which signs fall within the spectrum of disease for each clinical subtype, and which are truly atypical.

Second, vestibular function is often not evaluated or well characterized in persons with USH. Note that many of the manuscripts in have “not available” in the vestibular column. Furthermore, the methods used to determine vestibular function vary considerably from patient self-report to complex diagnostic assessments. This makes it difficult to determine if the vestibular presentation is typical or atypical and to compare cases across studies. Criteria for normal vestibular function should be defined and based upon accessible testing.

Third, the presence or absence of a retinal degeneration consistent with RP should be a defining criterion for USH. This is a point on which there is no consensus in the literature; however, establishing clearer criteria will allow for more consistency in diagnosis and reporting. We believe that a clinical diagnosis of USH should involve a retinal degeneration categorized as a rod-cone dystrophy.

Finally, it is important to carefully consider whether additional symptoms are present that would be indicative of a disorder affecting additional organ systems other than the eye and ear. This is particularly relevant given that several of the ultra-rare USH genes have been linked to systemic disorders in humans or animal models, such as lysosomal, peroxisomal, or mitochondrial disorders with associated neurological findings ().

We have searched the literature and observed that greater consistency is necessary in the description of USH. Establishing concrete criteria for diagnosis and the characterization of genetic causes will benefit the clinicians and scientists who are working to understand and treat this condition. Furthermore, it will provide clarity in diagnosis to patients and establish a common foundation on which future research can be built.

Declaration of interest statement

The authors report no conflicts of interest. The authors alone are responsible for the content and writing of this article.

Additional information

Funding

This study was supported by the Intramural Research Program of the National Institutes of Health.

References

  • Friedman TB, Schultz JM, Ahmed ZM, Tsilou ET, Brewer CC. Usher syndrome: hearing loss with vision loss. In: Alford RL, Sutton V, editors. Medical genetics in the clinical practice of ORL. Advances in oto-rhino-laryngology. Vol. 70. Basel (Switzerland): Karger Publishers; 2011. p. 56–65. doi:10.1159/000322473.
  • Hope C, Bundey S, Proops D, Fielder A. 1997. Usher syndrome in the city of Birmingham—prevalence and clinical classification. Br J of Ophthalmol. 81(1):46–53. doi:10.1136/bjo.81.1.46.
  • Grøndahl J. 1987. Estimation of prognosis and prevalence of retinitis pigmentosa and Usher syndrome in Norway. Clin Genet. 31(4):255–64. doi:10.1111/j.1399-0004.1987.tb02804.x.
  • Rosenberg T, Haim M, Hauch AM, Parving A. 1997. The prevalence of Usher syndrome and other retinal dystrophy-hearing impairment associations. Clin Genet. 51(5):314–21. doi:10.1111/j.1399-0004.1997.tb02480.x.
  • Spandau UH, Rohrschneider K. 2002. Prevalence and geographical distribution of Usher syndrome in Germany. Graefes Arch Clin Exp Ophthalmol. 240(6):495–98. doi:10.1007/s00417-002-0485-8.
  • Boughman JA, Vernon M, Shaver KA. 1983. Usher syndrome: definition and estimate of prevalence from two high-risk populations. J Chronic Dis. 36(8):595–603. doi:10.1016/0021-9681(83)90147-9.
  • Kimberling WJ, Hildebrand MS, Shearer AE, Jensen ML, Halder JA, Trzupek K. Frequency of Usher syndrome in two pediatric populations: implications for genetic screening of deaf and hard of hearing children. Genet Med. 2010;12(8):512–16. doi:10.1097/GIM.0b013e3181e5afb8.
  • Davenport S, Omenn G The heterogeneity of Usher syndrome. Vth Int. Conf. Birth Defects, Montreal. 1977.
  • Smith RJ, Berlin CI, Hejtmancik JF, Keats BJ, Kimberling WJ, Lewis RA. Clinical diagnosis of the Usher syndromes. Usher Syndrome Consortium. Am J Med Genet. 1994;50(1):32–38. doi:10.1002/ajmg.1320500107.
  • Gorlin RJ, Tilsner TJ, Feinstein S, Duvall AJ 3rd. 1979. Usher’s syndrome type III. Arch Otolaryngol. 105(6):353–54. doi:10.1001/archotol.1979.00790180051011.
  • Karjalainen S, Terasvirta M, Karja J, Kaariainen H. 1983. An unusual otological manifestation of Usher’s syndrome in four siblings. Clin Genet. 24(4):273–79. doi:10.1111/j.1399-0004.1983.tb00082.x.
  • Smith RJH, Pelias MZ, Daiger SP, Keats B, Kimberling W, Hejtmancik JF. 1992. Clinical variability and genetic-heterogeneity within the Acadian Usher population. Am J Med Genet. 43(6):964–69. doi:10.1002/ajmg.1320430612.
  • Weil D, Blanchard S, Kaplan J, Guilford P, Gibson F, Walsh J. Defective myosin VIIA gene responsible for Usher syndrome type 1B. Nature. 1995;374(6517):60–61. doi:10.1038/374060a0.
  • Verpy E, Leibovici M, Zwaenepoel I, Liu XZ, Gal A, Salem N. A defect in harmonin, a PDZ domain-containing protein expressed in the inner ear sensory hair cells, underlies Usher syndrome type 1C. Nat Genet. 2000;26(1):51–55. doi:10.1038/79171.
  • Bitner-Glindzicz M, Lindley KJ, Rutland P, Blaydon D, Smith VV, Milla PJ. A recessive contiguous gene deletion causing infantile hyperinsulinism, enteropathy and deafness identifies the Usher type 1C gene. Nat Genet. 2000;26(1):56–60. doi:10.1038/79178.
  • Bork JM, Peters LM, Riazuddin S, Bernstein SL, Ahmed ZM, Ness SL. Usher syndrome 1D and nonsyndromic autosomal recessive deafness DFNB12 are caused by allelic mutations of the novel cadherin-like gene CDH23. Am J Hum Genet. 2001;68(1):26–37. doi:10.1086/316954.
  • Bolz H, von Brederlow B, Ramirez A, Bryda EC, Kutsche K, Nothwang HG. Mutation of CDH23, encoding a new member of the cadherin gene family, causes Usher syndrome type 1D. Nat Genet. 2001;27(1):108–12. doi:10.1038/83667.
  • Ahmed ZM, Riazuddin S, Ahmad J, Bernstein SL, Guo Y, Sabar MF. PCDH15 is expressed in the neurosensory epithelium of the eye and ear and mutant alleles are responsible for both USH1F and DFNB23. Hum Mol Genet. 2003;12(24):3215–23. doi:10.1093/hmg/ddg358.
  • Alagramam KN, Yuan H, Kuehn MH, Murcia CL, Wayne S, Srisailpathy CR. Mutations in the novel protocadherin PCDH15 cause Usher syndrome type 1F. Hum Mol Genet. 2001;10(16):1709–18. doi:10.1093/hmg/10.16.1709.
  • Weil D, El-Amraoui A, Masmoudi S, Mustapha M, Kikkawa Y, Lainé S. Usher syndrome type IG (USH1G) is caused by mutations in the gene encoding SANS, a protein that associates with the USH1C protein, harmonin. Hum Mol Genet. 2003;12(5):463–71. doi:10.1093/hmg/ddg051.
  • Eudy JD, Weston MD, Yao S, Hoover DM, Rehm HL, Ma-Edmonds M. Mutation of a gene encoding a protein with extracellular matrix motifs in Usher syndrome type IIa. Science. 1998;280(5370):1753–57. doi:10.1126/science.280.5370.1753.
  • Weston MD, Luijendijk MW, Humphrey KD, Moller C, Kimberling WJ. 2004. Mutations in the VLGR1 gene implicate G-protein signaling in the pathogenesis of Usher syndrome type II. Am J Hum Genet. 74(2):357–66. doi:10.1086/381685.
  • Ebermann I, Scholl HP, Charbel Issa P, Becirovic E, Lamprecht J, Jurklies B. A novel gene for Usher syndrome type 2: mutations in the long isoform of whirlin are associated with retinitis pigmentosa and sensorineural hearing loss. Hum Genet. 2007;121(2):203–11. doi:10.1007/s00439-006-0304-0.
  • Joensuu T, Hamalainen R, Yuan B, Johnson C, Tegelberg S, Gasparini P. Mutations in a novel gene with transmembrane domains underlie Usher syndrome type 3. Am J Hum Genet. 2001;69(4):673–84. doi:10.1086/323610.
  • Adato A, Vreugde S, Joensuu T, Avidan N, Hamalainen R, Belenkiy O. USH3A transcripts encode clarin-1, a four-transmembrane-domain protein with a possible role in sensory synapses. Eur J Hum Genet. 2002;10(6):339–50. doi:10.1038/sj.ejhg.5200831.
  • Cosgrove D, Zallocchi M. Usher protein functions in hair cells and photoreceptors. Int J Biochem Cell Biol. 2014;46:80–89. doi:10.1016/j.biocel.2013.11.001.
  • Kremer H, van Wijk E, Marker T, Wolfrum U, Roepman R. 2006. Usher syndrome: molecular links of pathogenesis, proteins and pathways. Hum Mol Genet. 15(Issue suppl_2):R262–70. doi:10.1093/hmg/ddl205.
  • Bharadwaj AK, Kasztejna JP, Huq S, Berson EL, Dryja TP. 2000. Evaluation of the myosin VIIA gene and visual function in patients with Usher syndrome type I. Exp Eye Res. 71(2):173–81. doi:10.1006/exer.2000.0863.
  • Ouyang XM, Yan D, Du LL, Hejtmancik JF, Jacobson SG, Nance WE. Characterization of Usher syndrome type I gene mutations in an Usher syndrome patient population. Hum Genet. 2005;116(4):292–99. doi:10.1007/s00439-004-1227-2.
  • Jaijo T, Aller E, Oltra S, Beneyto M, Najera C, Ayuso C. Mutation profile of the MYO7A gene in Spanish patients with Usher syndrome type I. Hum Mutat. 2006;27(3):290–91. doi:10.1002/humu.9404.
  • Roux AF, Faugere V, Le Guedard S, Pallares-Ruiz N, Vielle A, Chambert S. Survey of the frequency of USH1 gene mutations in a cohort of Usher patients shows the importance of cadherin 23 and protocadherin 15 genes and establishes a detection rate of above 90%. J Med Genet. 2006;43(9):763–68. doi:10.1136/jmg.2006.041954.
  • Pennings RJ, Te Brinke H, Weston MD, Claassen A, Orten DJ, Weekamp H. USH2A mutation analysis in 70 Dutch families with Usher syndrome type II. Hum Mutat. 2004;24(2):185. doi:10.1002/humu.9259.
  • Baux D, Larrieu L, Blanchet C, Hamel C, Ben Salah S, Vielle A. Molecular and in silico analyses of the full-length isoform of usherin identify new pathogenic alleles in Usher type II patients. Hum Mutat. 2007;28(8):781–89. doi:10.1002/humu.20513.
  • Stabej PL, Saihan Z, Rangesh N, Steele-Stallard HB, Ambrose J, Coffey A. Comprehensive sequence analysis of nine Usher syndrome genes in the UK national collaborative Usher study. J Med Genet. 2012;49(1):27–36. doi:10.1136/jmedgenet-2011-100468.
  • Garcia-Garcia G, Besnard T, Baux D, Vache C, Aller E, Malcolm S. The contribution of GPR98 and DFNB31 genes to a Spanish Usher syndrome type 2 cohort. Mol Vis. 2013;19:367–73.
  • Oshima A, Jaijo T, Aller E, Millan JM, Carney C, Usami S. Mutation profile of the CDH23 gene in 56 probands with Usher syndrome type I. Hum Mutat. 2008;29(6):E37–46. doi:10.1002/humu.20761.
  • Ness SL, Ben-Yosef T, Bar-Lev A, Madeo AC, Brewer CC, Avraham KB. Genetic homogeneity and phenotypic variability among Ashkenazi Jews with Usher syndrome type III. J Med Genet. 2003;40(10):767–72. doi:10.1136/jmg.40.10.767.
  • Pakarinen L, Karjalainen S, Simola KO, Laippala P, Kaitalo H. 1995. Usher’s syndrome type 3 in Finland. Laryngoscope. 105(6):613–17. doi:10.1288/00005537-199506000-00010.
  • Bonnet C, Grati M, Marlin S, Levilliers J, Hardelin JP, Parodi M. Complete exon sequencing of all known Usher syndrome genes greatly improves molecular diagnosis. Orphanet J Rare Dis. 2011;6(1):21. doi:10.1186/1750-1172-6-21.
  • Aparisi MJ, Aller E, Fuster-Garcia C, Garcia-Garcia G, Rodrigo R, Vazquez-Manrique RP. Targeted next generation sequencing for molecular diagnosis of Usher syndrome. Orphanet J Rare Dis. 2014;9(1):168. doi:10.1186/s13023-014-0168-7.
  • Besnard T, Garcia-Garcia G, Baux D, Vache C, Faugere V, Larrieu L. Experience of targeted Usher exome sequencing as a clinical test. Mol Genet Genomic Med. 2014;2(1):30–43. doi:10.1002/mgg3.25.
  • Sun T, Xu K, Ren Y, Xie Y, Zhang X, Tian L. Comprehensive molecular screening in Chinese Usher syndrome patients. Invest Ophthalmol Vis Sci. 2018;59(3):1229–37. doi:10.1167/iovs.17-23312.
  • Khalaileh A, Abu-Diab A, Ben-Yosef T, Raas-Rothschild A, Lerer I, Alswaiti Y. The genetics of Usher syndrome in the Israeli and Palestinian populations. Invest Ophthalmol Vis Sci. 2018;59(2):1095–104. doi:10.1167/iovs.17-22817.
  • Bonnet C, Riahi Z, Chantot-Bastaraud S, Smagghe L, Letexier M, Marcaillou C. An innovative strategy for the molecular diagnosis of Usher syndrome identifies causal biallelic mutations in 93% of European patients. Eur J Hum Genet. 2016;24(12):1730–38. doi:10.1038/ejhg.2016.99.
  • Neuhaus C, Eisenberger T, Decker C, Nagl S, Blank C, Pfister M. Next‐generation sequencing reveals the mutational landscape of clinically diagnosed Usher syndrome: copy number variations, phenocopies, a predominant target for translational read‐through, and PEX 26 mutated in Heimler syndrome. Mol Genet Genomic Med. 2017;5(5):531–52. doi:10.1002/mgg3.312.
  • Fuster-García C, García-García G, Jaijo T, Blanco-Kelly F, Tian L, Hakonarson H. Expanding the genetic landscape of Usher-like phenotypes. Invest Ophthalmol Vis Sci. 2019;60(14):4701–10. doi:10.1167/iovs.19-27470.
  • Ebermann I, Phillips JB, Liebau MC, Koenekoop RK, Schermer B, Lopez I. PDZD7 is a modifier of retinal disease and a contributor to digenic Usher syndrome. J Clin Invest. 2010;120(6):1812–23. doi:10.1172/Jci39715.
  • Puffenberger EG, Jinks RN, Sougnez C, Cibulskis K, Willert RA, Achilly NP. Genetic mapping and exome sequencing identify variants associated with five novel diseases. PLoS One. 2012;7(1):e28936. doi:10.1371/journal.pone.0028936.
  • Eisenberger T, Slim R, Mansour A, Nauck M, Nurnberg G, Nurnberg P. Targeted next-generation sequencing identifies a homozygous nonsense mutation in ABHD12, the gene underlying PHARC, in a family clinically diagnosed with Usher syndrome type 3. Orphanet J Rare Dis. 2012;7(1):59. doi:10.1186/1750-1172-7-59.
  • Riazuddin S, Belyantseva IA, Giese AP, Lee K, Indzhykulian AA, Nandamuri SP. Alterations of the CIB2 calcium- and integrin-binding protein cause Usher syndrome type 1J and nonsyndromic deafness DFNB48. Nat Genet. 2012;44(11):1265–71. doi:10.1038/ng.2426.
  • Khateb S, Zelinger L, Mizrahi-Meissonnier L, Ayuso C, Koenekoop RK, Laxer U. A homozygous nonsense CEP250 mutation combined with a heterozygous nonsense C2orf71 mutation is associated with atypical Usher syndrome. J Med Genet. 2014;51(7):460–69. doi:10.1136/jmedgenet-2014-102287.
  • Namburi P, Ratnapriya R, Khateb S, Lazar CH, Kinarty Y, Obolensky A. Bi-allelic truncating mutations in CEP78, encoding centrosomal protein 78, cause cone-rod degeneration with sensorineural hearing loss. Am J Hum Genet. 2016;99(3):777–84. doi:10.1016/j.ajhg.2016.07.010.
  • Khateb S, Kowalewski B, Bedoni N, Damme M, Pollack N, Saada A. A homozygous founder missense variant in arylsulfatase G abolishes its enzymatic activity causing atypical Usher syndrome in humans. Genet Med. 2018;20(9):1004–12. doi:10.1038/gim.2017.227.
  • Ahmed ZM, Jaworek TJ, Sarangdhar GN, Zheng LL, Gul K, Khan SN. Inframe deletion of human ESPN is associated with deafness, vestibulopathy and vision impairment. J Med Genet. 2018;55(7):479–88. doi:10.1136/jmedgenet-2017-105221.
  • Stenson PD, Mort M, Ball EV, Shaw K, Phillips AD, Cooper DN. The human gene mutation database: building a comprehensive mutation repository for clinical and molecular genetics, diagnostic testing and personalized genomic medicine. Hum Genet. 2014;133(1):1–9.
  • Tiwari A, Bahr A, Bähr L, Fleischhauer J, Zinkernagel MS, Winkler N. Next generation sequencing based identification of disease-associated mutations in Swiss patients with retinal dystrophies. Sci Rep. 2016;6:1–11.
  • Yoshimura H, Hashimoto T, Murata T, Fukushima K, Sugaya A, Nishio SY. Novel ABHD12 mutations in PHARC patients: the differential diagnosis of deaf-blindness. Ann Otol Rhinol Laryngol. 2015;(124 Suppl 1(1_suppl)):77S–83S. doi:10.1177/0003489415574513.
  • Kubota D, Gocho K, Kikuchi S, Akeo K, Miura M, Yamaki K. CEP250 mutations associated with mild cone-rod dystrophy and sensorineural hearing loss in a Japanese family. Ophthalmic Genet. 2018;39(4):500–07. doi:10.1080/13816810.2018.1466338.
  • Fuster-Garcia C, Garcia-Garcia G, Jaijo T, Fornes N, Ayuso C, Fernandez-Burriel M. High-throughput sequencing for the molecular diagnosis of Usher syndrome reveals 42 novel mutations and consolidates CEP250 as Usher-like disease causative. Sci Rep. 2018;8(1):17113. doi:10.1038/s41598-018-35085-0.
  • Nikopoulos K, Farinelli P, Giangreco B, Tsika C, Royer-Bertrand B, Mbefo MK. Mutations in CEP78 cause cone-rod dystrophy and hearing loss associated with primary-cilia defects. Am J Hum Genet. 2016;99(3):770–76. doi:10.1016/j.ajhg.2016.07.009.
  • Fu Q, Xu MC, Chen X, Sheng XL, Yuan ZS, Liu YN. CEP78 is mutated in a distinct type of Usher syndrome. J Med Genet. 2017;54(3):190–95. doi:10.1136/jmedgenet-2016-104166.
  • Schneider E, Marker T, Daser A, Frey-Mahn G, Beyer V, Farcas R. Homozygous disruption of PDZD7 by reciprocal translocation in a consanguineous family: a new member of the Usher syndrome protein interactome causing congenital hearing impairment. Hum Mol Genet. 2009;18(4):655–66. doi:10.1093/hmg/ddn395.
  • Booth KT, Azaiez H, Kahrizi K, Simpson AC, Tollefson WTA, Sloan CM, Meyer NC, Babanejad M, Ardalani F, Arzhangi S, et al. PDZD7 and hearing loss: more than just a modifier. Am J Med Genet Part A. 2015;167(12):2957–65. doi:10.1002/ajmg.a.37274.
  • Vona B, Lechno S, Hofrichter MAH, Hopf S, Lassig AK, Haaf T. Confirmation of PDZD7 as a nonsyndromic hearing loss gene. Ear Hear. 2016;37(4):E238–E46. doi:10.1097/Aud.0000000000000278.
  • Stabej PL, James C, Ocaka L, Tekman M, Grunewald S, Clement E. An example of the utility of genomic analysis for fast and accurate clinical diagnosis of complex rare phenotypes. Orphanet J Rare Dis. 2017;12(1):24. doi:10.1186/s13023-017-0582-8.
  • Guan J, Wang H, Lan L, Wang L, Yang J, Xie L. Novel recessive PDZD7 biallelic mutations in two Chinese families with non-syndromic hearing loss. Am J Med Genet A. 2018;176(1):99–106. doi:10.1002/ajmg.a.38477.
  • Chen Q, Zou JH, Shen ZL, Zhang WP, Yang J. 2014. Whirlin and PDZ Domain-containing 7 (PDZD7) proteins are both required to form the quaternary protein complex associated with Usher syndrome type 2. J Biol Chem. 289(52):36070–88. doi:10.1074/jbc.M114.610535.
  • Zou J, Zheng T, Ren C, Askew C, Liu XP, Pan B. Deletion of PDZD7 disrupts the Usher syndrome type 2 protein complex in cochlear hair cells and causes hearing loss in mice. Hum Mol Genet. 2014;23(9):2374–90. doi:10.1093/hmg/ddt629.
  • Morgan CP, Krey JF, Grati M, Zhao B, Fallen S, Kannan-Sundhari A. PDZD7-MYO7A complex identified in enriched stereocilia membranes. Elife. 2016;5:e18312.
  • Zou J, Chen Q, Almishaal A, Mathur PD, Zheng T, Tian C. The roles of USH1 proteins and PDZ domain-containing USH proteins in USH2 complex integrity in cochlear hair cells. Hum Mol Genet. 2017;26(3):624–36. doi:10.1093/hmg/ddw421.
  • Grati M, Shin JB, Weston MD, Green J, Bhat MA, Gillespie PG. Localization of PDZD7 to the stereocilia ankle-link associates this scaffolding protein with the Usher syndrome protein network. J Neurosci. 2012;32(41):14288–93. doi:10.1523/Jneurosci.3071-12.2012.
  • Vester A, Velez‐Ruiz G, McLaughlin HM, Program NCS, Lupski JR, Talbot K. A loss‐of‐function variant in the human histidyl‐t RNA synthetase (HARS) gene is neurotoxic in vivo. Human Mutat. 2013;34(1):191–99. doi:10.1002/humu.22210.
  • Abbott JA, Meyer-Schuman R, Lupo V, Feely S, Mademan I, Oprescu SN. Substrate interaction defects in histidyl-tRNA synthetase linked to dominant axonal peripheral neuropathy. Hum Mutat. 2018;39(3):415–32. doi:10.1002/humu.23380.
  • Safka Brozkova D, Deconinck T, Griffin LB, Ferbert A, Haberlova J, Mazanec R. Loss of function mutations in HARS cause a spectrum of inherited peripheral neuropathies. Brain. 2015;138(Pt 8):2161–72. doi:10.1093/brain/awv158.
  • Antonellis A, Green ED. The role of aminoacyl-tRNA synthetases in genetic diseases. Annu Rev Genomics Hum Genet. 2008;9:87–107. doi:10.1146/annurev.genom.9.081307.164204.
  • Fiskerstrand T, H’Mida-Ben Brahim D, Johansson S, M’Zahem A, Haukanes BI, Drouot N. Mutations in ABHD12 cause the neurodegenerative disease PHARC: an inborn error of endocannabinoid metabolism. Am J Hum Genet. 2010;87(3):410–17. doi:10.1016/j.ajhg.2010.08.002.
  • Chen DH, Naydenov A, Blankman JL, Mefford HC, Davis M, Sul Y. Two novel mutations in ABHD 12: expansion of the mutation spectrum in PHARC and assessment of their functional effects. Hum Mutat. 2013;34(12):1672–78. doi:10.1002/humu.22437.
  • Nishiguchi KM, Avila-Fernandez A, van Huet RA, Corton M, Perez-Carro R, Martin-Garrido E. Exome sequencing extends the phenotypic spectrum for ABHD12 mutations: from syndromic to nonsyndromic retinal degeneration. Ophthalmology. 2014;121(8):1620–27. doi:10.1016/j.ophtha.2014.02.008.
  • Lerat J, Cintas P, Beauvais-Dzugan H, Magdelaine C, Sturtz F, Lia AS. 2017. A complex homozygous mutation in ABHD12 responsible for PHARC syndrome discovered with NGS and review of the literature. J Peripher Nerv Syst. 22(2):77–84. doi:10.1111/jns.12216.
  • Tingaud-Sequeira A, Raldua D, Lavie J, Mathieu G, Bordier M, Knoll-Gellida A. Functional validation of ABHD12 mutations in the neurodegenerative disease PHARC. Neurobiol Dis. 2017;98:36–51.
  • Frasquet M, Lupo V, Chumillas MJ, Vazquez-Costa JF, Espinos C, Sevilla T. Phenotypical features of two patients diagnosed with PHARC syndrome and carriers of a new homozygous mutation in the ABHD12 gene. J Neurol Sci. 2018;387:134–38. doi:10.1016/j.jns.2018.02.021.
  • Simon GM, Cravatt BF. 2010. Activity-based proteomics of enzyme superfamilies: serine hydrolases as a case study. J Biol Chem. 285(15):11051–55. doi:10.1074/jbc.R109.097600.
  • Blankman JL, Long JZ, Trauger SA, Siuzdak G, Cravatt BF. 2013. ABHD12 controls brain lysophosphatidylserine pathways that are deregulated in a murine model of the neurodegenerative disease PHARC. P Natl Acad Sci USA. 110(4):1500–05. doi:10.1073/pnas.1217121110.
  • Patel K, Giese AP, Grossheim J, Hegde RS, Delio M, Samanich J. A novel C-terminal CIB2 (calcium and integrin binding protein 2) mutation associated with non-syndromic hearing loss in a hispanic family. PLoS One. 2015;10(10):e0133082. doi:10.1371/journal.pone.0133082.
  • Seco CZ, Giese AP, Shafique S, Schraders M, Oonk AM, Grossheim M. Novel and recurrent CIB2 variants, associated with nonsyndromic deafness, do not affect calcium buffering and localization in hair cells. Eur J Hum Genet. 2016;24(4):542. doi:10.1038/ejhg.2015.157.
  • Booth K, Kahrizi K, Babanejad M, Daghagh H, Bademci G, Arzhangi S. Variants in CIB2 cause DFNB48 and not USH1J. Clin Genet. 2018;93(4):812–21. doi:10.1111/cge.13170.
  • Talbi S, Bonnet C, Riahi Z, Boudjenah F, Dahmani M, Hardelin J-P. Genetic heterogeneity of congenital hearing impairment in Algerians from the Ghardaïa province. Int J Pediatr Otorhinolaryngol. 2018;112:1–5.
  • Vallone R, Dal Cortivo G, D’Onofrio M, Dell’Orco D. Preferential binding of Mg(2+) over Ca(2+) to CIB2 triggers an allosteric switch impaired in Usher syndrome type 1J. Front Mol Neurosci. 2018;11:274. doi:10.3389/fnmol.2018.00274.
  • Giese APJ, Tang YQ, Sinha GP, Bowl MR, Goldring AC, Parker A. CIB2 interacts with TMC1 and TMC2 and is essential for mechanotransduction in auditory hair cells. Nat Commun. 2017;8(1):43. doi:10.1038/s41467-017-00061-1.
  • Wang Y, Li J, Yao X, Li W, Du H, Tang M. Loss of CIB2 causes profound hearing loss and abolishes mechanoelectrical transduction in mice. Front Mol Neurosci. 2017;10:401.
  • de Castro-miró M, Tonda R, Escudero-Ferruz P, Andrés R, Mayor-Lorenzo A, Castro J. Novel candidate genes and a wide spectrum of structural and point mutations responsible for inherited retinal dystrophies revealed by exome sequencing. PLoS One. 2016;11(12):e0168966. doi:10.1371/journal.pone.0168966.
  • Kumar A, Rajendran V, Sethumadhavan R, Purohit R. 2013. CEP proteins: the knights of centrosome dynasty. Protoplasma. 250(5):965–83. doi:10.1007/s00709-013-0488-9.
  • Fogeron ML, Muller H, Schade S, Dreher F, Lehmann V, Kuhnel A. LGALS3BP regulates centriole biogenesis and centrosome hypertrophy in cancer cells. Nat Commun. 2013;4:1531.
  • Floriot S, Vesque C, Rodriguez S, Bourgain-Guglielmetti F, Karaiskou A, Gautier M. C-Nap1 mutation affects centriole cohesion and is associated with a Seckel-like syndrome in cattle. Nat Commun. 2015;6:6894.
  • Andersen JS, Wilkinson CJ, Mayor T, Mortensen P, Nigg EA, Mann M. 2003. Proteomic characterization of the human centrosome by protein correlation profiling. Nature. 426(6966):570–74. doi:10.1038/nature02166.
  • Azimzadeh J, Wong ML, Downhour DM, Sanchez Alvarado A, Marshall WF. 2012. Centrosome loss in the evolution of planarians. Science. 335(6067):461–63. doi:10.1126/science.1214457.
  • Kowalewski B, Lamanna WC, Lawrence R, Damme M, Stroobants S, Padva M. Arylsulfatase G inactivation causes loss of heparan sulfate 3-O-sulfatase activity and mucopolysaccharidosis in mice. Proc Natl Acad Sci U S A. 2012;109(26):10310–15. doi:10.1073/pnas.1202071109.
  • Kruszewski K, Lullmann-Rauch R, Dierks T, Bartsch U, Damme M. 2016. Degeneration of photoreceptor cells in arylsulfatase G-deficient mice. Invest Ophthalmol Vis Sci. 57(3):1120–31. doi:10.1167/iovs.15-17645.
  • Kowalewski B, Heimann P, Ortkras T, Lüllmann-Rauch R, Sawada T, Walkley SU. Ataxia is the major neuropathological finding in arylsulfatase G-deficient mice: similarities and dissimilarities to Sanfilippo disease (mucopolysaccharidosis type III). Hum Mol Genet. 2014;24(7):1856–68. doi:10.1093/hmg/ddu603.
  • Abitbol M, Thibaud JL, Olby NJ, Hitte C, Puech JP, Maurer M. A canine Arylsulfatase G (ARSG) mutation leading to a sulfatase deficiency is associated with neuronal ceroid lipofuscinosis. P Natl Acad Sci USA. 2010;107(33):14775–80. doi:10.1073/pnas.0914206107.
  • Ferrante P, Messali S, Meroni G, Ballabio A. 2002. Molecular and biochemical characterisation of a novel sulphatase gene: arylsulfatase G (ARSG). Eur J Hum Genet. 10(12):813–18. doi:10.1038/sj.ejhg.5200887.
  • Frese MA, Schulz S, Dierks T, Arylsulfatase G. 2008. a novel lysosomal sulfatase. J Biol Chem. 283(17):11388–95. doi:10.1074/jbc.M709917200.
  • Ratzka A, Mundlos S, Vortkamp A. 2010. Expression patterns of sulfatase genes in the developing mouse embryo. Dev Dyn. 239(6):1779–88. doi:10.1002/dvdy.22294.
  • Naz S, Griffith AJ, Riazuddin S, Hampton LL, Battey JF Jr., Khan SN. Mutations of ESPN cause autosomal recessive deafness and vestibular dysfunction. J Med Genet. 2004;41(8):591–95. doi:10.1136/jmg.2004.018523.
  • Donaudy F, Zheng L, Ficarella R, Ballana E, Carella M, Melchionda S. Espin gene (ESPN) mutations associated with autosomal dominant hearing loss cause defects in microvillar elongation or organisation. J Med Genet. 2006;43(2):157–61. doi:10.1136/jmg.2005.032086.
  • Sekerkova G, Zheng L, Loomis PA, Mugnaini E, Bartles JR. 2006. Espins and the actin cytoskeleton of hair cell stereocilia and sensory cell microvilli. Cell Mol Life Sci. 63(19–20):2329–41. doi:10.1007/s00018-006-6148-x.
  • Wang L, Zou J, Shen Z, Song E, Yang J. 2011. Whirlin interacts with espin and modulates its actin-regulatory function: an insight into the mechanism of Usher syndrome type II. Hum Mol Genet. 21(3):692–710. doi:10.1093/hmg/ddr503.
  • Zheng LL, Sekerkova G, Vranich K, Tilney LG, Mugnaini E, Bartles JR. 2000. The deaf jerker mouse has a mutation in the gene encoding the espin actin-bundling proteins of hair cell stereocilia and lacks espins. Cell. 102(3):377–85. doi:10.1016/S0092-8674(00)00042-8.
  • Zou JH, Mathur PD, Zheng TH, Wang Y, Almishaal A, Park AH. Individual USH2 proteins make distinct contributions to the ankle link complex during development of the mouse cochlear stereociliary bundle. Hum Mole Genet. 2015;24(24):6944–57. doi:10.1093/hmg/ddv398.
  • O’Hanlon TP, Miller FW. 2002. Genomic organization, transcriptional mapping, and evolutionary implications of the human bi-directional histidyl-tRNA synthetase locus (HARS/HARSL). Biochem Biophys Res Commun. 294(3):609–14. doi:10.1016/S0006-291X(02)00525-9.
  • Wasmuth JJ, Carlock LR. 1986. Chromosomal localization of human gene for histidyl-tRNA synthetase: clustering of genes encoding aminoacyl-tRNA synthetases on human chromosome 5. Somat Cell Mol Genet. 12(5):513–17. doi:10.1007/bf01539922.
  • Abbott JA, Guth E, Kim C, Regan C, Siu VM, Rupar CA. The usher syndrome type IIIB histidyl-tRNA synthetase mutation confers temperature sensitivity. Biochemistry. 2017;56(28):3619–31. doi:10.1021/acs.biochem.7b00114.
  • Navia-Paldanius D, Savinainen JR, Laitinen JT. Biochemical and pharmacological characterization of human α/β-hydrolase domain containing 6 (ABHD6) and 12 (ABHD12). J Lipid Res. 2012;53:2413–24. doi:10.1194/jlr.M030411.
  • Blankman JL, Simon GM, Cravatt BF. 2007. A comprehensive profile of brain enzymes that hydrolyze the endocannabinoid 2-arachidonoylglycerol. Chem Biol. 14(12):1347–56. doi:10.1016/j.chembiol.2007.11.006.
  • Seki N, Hattori A, Hayashi A, Kozuma S, Ohira M, Hori T-A. Structure, expression profile and chromosomal location of an isolog of DNA-PKcs interacting protein (KIP) gene1. Biochimica Et Biophysica Acta (Bba)-gene Structure and Expression. 1999;1444(1):143–47. doi:10.1016/S0167-4781(98)00253-X.
  • Häger M, Bigotti MG, Meszaros R, Carmignac V, Holmberg J, Allamand V. Cib2 binds integrin α7Bβ1D and is reduced in laminin α2 chain-deficient muscular dystrophy. J Biol Chem. 2008;283(36):24760–69. doi:10.1074/jbc.M801166200.
  • Mayor T, Stierhof Y-D, Tanaka K, Fry AM, Nigg EA. 2000. The centrosomal protein C-Nap1 is required for cell cycle–regulated centrosome cohesion. J Cell Biol. 151(4):837–46. doi:10.1083/jcb.151.4.837.
  • Brunk K, Zhu M, Kratz AS, Haselmann-Weiss U, Claude A, Hoffmann I. Cep78 is a new centriolar protein involved in Plk4-induced centriole overduplication. J Cell Sci. 2016;129:2713–18. doi:10.1242/jcs.184093.
  • Schrauwen I, Hasin-Brumshtein Y, Corneveaux JJ, Ohmen J, White C, Allen AN. A comprehensive catalogue of the coding and non-coding transcripts of the human inner ear. Hear Res. 2016;333:266–74.
  • Liu XZ, Hope C, Walsh J, Newton V, Ke XM, Liang CY. Mutations in the myosin VIIA gene cause a wide phenotypic spectrum, including atypical Usher syndrome. Am J Hum Genet. 1998;63(3):909–12. doi:10.1086/302026.
  • Cremers FP, Kimberling WJ, Kulm M, de Brouwer AP, van Wijk E, Te Brinke H. Development of a genotyping microarray for Usher syndrome. J Med Genet. 2007;44(2):153–60. doi:10.1136/jmg.2006.044784.
  • Liu XZ, Hope C, Liang CY, Zou JM, Xu LR, Cole T. A mutation (2314delG) in the Usher syndrome type IIA gene: high prevalence and phenotypic variation. Am J Hum Genet. 1999;64(4):1221–25. doi:10.1086/302332.
  • Blanco-Kelly F, Jaijo T, Aller E, Avila-Fernandez A, Lopez-Molina MI, Gimenez A. Clinical aspects of Usher syndrome and the USH2A gene in a cohort of 433 patients. JAMA Ophthalmol. 2015;133(2):157–64. doi:10.1001/jamaophthalmol.2014.4498.
  • Aller E, Nájera C, Millán JM, Oltra JS, Pérez-Garrigues H, Vilela C. Genetic analysis of 2299delG and C759F mutations (USH2A) in patients with visual and/or auditory impairments. Eur J Hum Genet. 2004;12(5):407. doi:10.1038/sj.ejhg.5201138.
  • Nájera C, Beneyto M, Blanca J, Aller E, Fontcuberta A, Millán JM. Mutations in myosin VIIA (MYO7A) and usherin (USH2A) in Spanish patients with Usher syndrome types I and II, respectively. Hum Mutat. 2002;20(1):76–77. doi:10.1002/humu.9042.
  • Garcia-Garcia G, Aparisi MJ, Jaijo T, Rodrigo R, Leon AM, Avila-Fernandez A. Mutational screening of the USH2A gene in Spanish USH patients reveals 23 novel pathogenic mutations. Orphanet J Rare Dis. 2011;6(1):65. doi:10.1186/1750-1172-6-65.
  • Steele-Stallard HB, Le Quesne Stabej P, Lenassi E, Luxon LM, Claustres M, Roux AF. Screening for duplications, deletions and a common intronic mutation detects 35% of second mutations in patients with USH2A monoallelic mutations on Sanger sequencing. Orphanet J Rare Dis. 2013;8(1):122. doi:10.1186/1750-1172-8-122.
  • Astuto LM, Bork JM, Weston MD, Askew JW, Fields RR, Orten DJ. CDH23 mutation and phenotype heterogeneity: a profile of 107 diverse families with Usher syndrome and nonsyndromic deafness. Am J Hum Genet. 2002;71(2):262–75. doi:10.1086/341558.
  • Valero R, de Castro-miro M, Jimenez-Ochoa S, Rodriguez-Ezcurra JJ, Marfany G, Gonzalez-Duarte R. 2019. Aberrant splicing events associated to CDH23 noncanonical splice site mutations in a proband with atypical Usher syndrome 1. Genes (Basel). 10(10):732. doi:10.3390/genes10100732.
  • Kalay E, de Brouwer AP, Caylan R, Nabuurs SB, Wollnik B, Karaguzel A. A novel D458V mutation in the SANS PDZ binding motif causes atypical Usher syndrome. J Mol Med (Berl). 2005;83(12):1025–32. doi:10.1007/s00109-005-0719-4.
  • Bashir R, Fatima A, Naz S. 2010. A frameshift mutation in SANS results in atypical Usher syndrome. Clin Genet. 78(6):601–03. doi:10.1111/j.1399-0004.2010.01500.x.
  • Ben Rebeh I, Morinière M, Ayadi L, Benzina Z, Charfedine I, Feki J. Reinforcement of a minor alternative splicing event in MYO7A due to a missense mutation results in a mild form of retinopathy and deafness. Mol Vis. 2010;16:1898–906. PMID: 21031134.
  • Zina ZB, Masmoudi S, Ayadi H, Chaker F, Ghorbel AM, Drira M. From DFNB2 to Usher syndrome: variable expressivity of the same disease. Am J Med Genet. 2001;101(2):181–83. doi:10.1002/ajmg.1335.
  • Bernal S, Meda C, Solans T, Ayuso C, Garcia‐Sandoval B, Valverde D. Clinical and genetic studies in Spanish patients with Usher syndrome type II: description of new mutations and evidence for a lack of genotype–phenotype correlation. Clin Genet. 2005;68(3):204–14. doi:10.1111/j.1399-0004.2005.00481.x.
  • Pennings RJ, Huygen PL, Cremers WR. 2003. Hearing impairment in Usher syndrome type II. Ann Otol Rhinol Laryngol. 112(9 Pt 1):825. doi:10.1177/000348940311200915.
  • Pennings RJ, Huygen PL, Weston MD, van Aarem A, Wagenaar M, Kimberling WJ. Pure tone hearing thresholds and speech recognition scores in Dutch patients carrying mutations in the USH2A gene. Otol Neurotol. 2003;24(1):58–63. doi:10.1097/00129492-200301000-00013.
  • Wagenaar M, van Aarem A, Huygen P, Pieke-Dahl S, Kimberling W, Cremers C. 1999. Hearing impairment related to age in Usher syndrome types 1B and 2A. Arch Otolaryngol Head Neck Surg. 125(4):441–45. doi:10.1001/archotol.125.4.441.
  • Sadeghi M, Cohn ES, Kelly WJ, Kimberling WJ, Tranebjoerg L, Moller C. 2004. Audiological findings in Usher syndrome types IIa and II (non-IIa). Int J Audiol. 43(3):136–43. doi:10.1080/14992020400050019.