Next Article in Journal
Are There Barriers Separating the Pink River Dolphin Populations (Inia boliviensis, Iniidae, Cetacea) within the Mamoré–Iténez River Basins (Bolivia)? An Analysis of Its Genetic Structure by Means of Mitochondrial and Nuclear DNA Markers
Previous Article in Journal
Changes in the Transcriptome and Long Non-Coding RNAs but Not the Methylome Occur in Human Cells Exposed to Borrelia burgdorferi
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Whole-Exome Analysis for Polish Caucasian Patients with Retinal Dystrophies and the Creation of a Reference Genomic Database for the Polish Population

by
Ewa Matczyńska
1,2,
Robert Szymańczak
1,
Katarzyna Stradomska
1,
Przemysław Łyszkiewicz
1,
Maria Jędrzejowska
1,
Karolina Kamińska
1,†,
Marta Beć-Gajowniczek
1,
Ewa Suchecka
1,
Marek Zagulski
1,
Marta Wiącek
3,
Edward Wylęgała
2,
Anna Machalińska
3,
Małgorzata Mossakowska
4,
Monika Puzianowska-Kuźnicka
5,6,
Sławomir Teper
2,7 and
Anna Boguszewska-Chachulska
1,*
1
Genomed S.A., 02-971 Warsaw, Poland
2
Chair and Clinical Department of Ophthalmology, Faculty of Medical Sciences in Zabrze, Medical University of Silesia, 40-055 Katowice, Poland
3
First Department of Ophthalmology, Pomeranian Medical University, 70-204 Szczecin, Poland
4
Study on Ageing and Longevity, International Institute of Molecular and Cell Biology, 02-109 Warsaw, Poland
5
Department of Human Epigenetics, Mossakowski Medical Research Institute, Polish Academy of Sciences, 02-106 Warsaw, Poland
6
Department of Geriatrics and Gerontology, Medical Centre of Postgraduate Education, 01-813 Warsaw, Poland
7
Department of Scientific Research, Branch in Bielsko-Biala, Medical University of Silesia, 43-300 Bielsko-Biała, Poland
*
Author to whom correspondence should be addressed.
Current address: Institute of Molecular and Clinical Ophthalmology Basel (IOB), 4031 Basel, Switzerland.
Genes 2024, 15(8), 1011; https://doi.org/10.3390/genes15081011
Submission received: 24 June 2024 / Revised: 25 July 2024 / Accepted: 29 July 2024 / Published: 1 August 2024
(This article belongs to the Special Issue Genetics of Eye Development and Diseases)

Abstract

:
We present the results of the first study of a large cohort of patients with inherited retinal dystrophies (IRD) performed for the Polish population using whole-exome sequencing (WES) in the years 2016–2019. Moreover, to facilitate such diagnostic analyses and enable future application of gene therapy and genome editing for IRD patients, a Polish genomic reference database (POLGENOM) was created based on whole-genome sequences of healthy Polish Caucasian nonagenarians and centenarians. The newly constructed database served as a control, providing a comparison for variant frequencies in the Polish population. The diagnostic yield for the selected group of IRD patients reached 64.9%. The study uncovered the most common pathogenic variants in ABCA4 and USH2A in the European population, along with several novel causative variants. A significant frequency of the ABCA4 complex haplotype p.(Leu541Pro; Ala1038Val) was observed, as well as that of the p.Gly1961Glu variant. The first VCAN causative variant NM_004385.5:c.4004-2A>G in Poland was found and described. Moreover, one of the first patients with the RPE65 causative variants was identified, and, in consequence, could receive the dedicated gene therapy. The availability of the reference POLGENOM database enabled comprehensive variant characterisation during the NGS data analysis, confirming the utility of a population-specific genomic database for enhancing diagnostic accuracy. Study findings suggest the significance of genetic testing in elder patients with unclear aetiology of eye diseases. The combined approach of NGS and the reference genomic database can improve the diagnosis, management, and future treatment of IRDs.

1. Introduction

Inherited retinal dystrophies (IRDs) represent the major group of congenital eye diseases, affecting 1 individual per 1500. They are a very heterogeneous group with at least 50 major types [1]. Similar clinical symptoms can therefore be due to variants in a large number of possible causative genes and, conversely, different genetic variants in the same gene can cause distinct disorders [2].
There is a wide spectrum of IRD phenotypes, from progressive conditions, including those affecting the macula (e.g., Stargardt disease), to stationary conditions such as achromatopsia and congenital stationary night blindness. Retinits pigmentosa (RP) is the most frequent IRD, with a prevalence close to 1:4000 in Europe and worldwide [2,3], encompassing progressive rod-cone dystrophies, characterised by the primary degeneration of rod photoreceptors followed by the cone photoreceptor loss [2]. A gradual loss of the visual field in a concentric pattern may ultimately result in the loss of vision at more advanced stages, with additional, extra-ocular symptoms in 20–30% of cases (syndromic RP) [2].
Most retinal dystrophies are inherited as an autosomal recessive trait, with RP being more frequently dominant [1,2]. There are, however, some genes in which both recessive and dominant inheritance is possible [1]. These dystrophy types can differ in their clinical symptoms. This makes the IRD diagnosis more complex. X-linked inheritance is relatively common, as shown by recent studies [4]. The involvement of the mitochondrial genome in the development of IRD should not be ignored [5]. The mtDNA analysis should therefore be included to search for causative variants in syndromic retinal dystrophies
Increasing application of next-generation sequencing (NGS) methods in genetic diagnostics and decreasing costs of such analyses have allowed for the performance of large studies using whole-exome sequencing (WES), whole-genome sequencing (WGS), and multigene panels. Results of these studies have been recently published and improve our knowledge and our ability to uncover the genetic background of IRD [4,6,7,8]. Over 300 genes are currently known as harbouring pathogenic variants that may cause retinal disease [1].
In response to the high incidence of IRDs and an urgent need for a cure, there are currently numerous gene therapies at various stages of clinical development.
Only one such treatment, Luxturna (voretigene neparvovec, Novartis), was authorised by the FDA (2017) and EMA (2018) to treat RP/Leber’s congenital amaurosis (LCA) patients with pathogenic variants in the RPE65 gene [9]. In spite of the hopes provoked by the introduction of this treatment, one should not forget that mutations of the RPE65 gene represent a small percentage of IRDs, while the efficacy of this treatment decreases with the age of the patient and a number of viable retinal cells [10]. Paradoxically, with the development of methods for both gene editing and introducing nucleic acids into the cell using appropriate vectors, it turned out that disease mechanisms are so complex that a simple insertion of the correct gene is an insufficient solution. Hence, an attempt to bypass many obstacles using gene-agnostic methods seems promising and may give hope to patients also at an advanced stage of the disease. Currently, we are observing a dynamic progress in both groups—specific gene-targeted and gene-independent methods [11]. Both will probably be necessary and, moreover, complementary to each other, depending on the type of mutation, retinal morphology, comorbidities, and financial constraints of healthcare systems.
Several genetic studies were carried out for small cohorts of Polish Caucasian IRD patients with Stargardt disease [12,13] and RP [14,15] clinical symptoms. These studies used mostly targeted approaches such as Sanger sequencing [15], SNP microarrays, or small NGS panels, e.g., in LCA [16,17] and cone-rode dystrophy [18] or broader NGS panels [12,13]. The WES approach has been applied in a few studies only [14,19]. Our group performed a large cohort study for another multifactorial retinal degeneration disorder, age-related macular degeneration (AMD), in the frame of the same National Centre for Research and Development grant [20].
Development of diagnostics based on the NGS method (especially using clinical WES and WGS) brings about a need for a well-curated database for the specific population to solve all the ambiguities connected with global databases. The combination of clinical (phenotypic) data with the genotypes is one of crucial requirements for such databases with the intended diagnostic use. Consideration of the best cohort for such purpose leads to the selection of healthy long-lived individuals as the source of genomic reference data. Analysis of the whole-genome data for a precisely selected group of people should also allow an indication of the genetic variants favouring the long-lived phenotype.
Having these two purposes in mind, we focused on the study of whole-genome sequence variation in the Polish population in combination with collected clinical data to create a Polish genomic reference database of extremely long-lived, healthy individuals who, compared to those not genetically predisposed to successful aging and longevity, probably have a lower number of pathogenic variants in genes affecting the risk of various diseases. Therefore, we assumed that such a database would serve as an excellent control for diagnostics of people affected with genetically determined diseases.
The goal of the current project was to identify the most frequent pathogenic variants which may serve as treatment targets in IRD patients. The specific approach adopted, based on WES, should allow to identify pathogenic variants not only in already known genes causing retinal diseases but also in new genes. Moreover, identification of common pathogenic variants would possibly allow for development of new therapies for patients with progressive, currently untreatable retinal dystrophies.

2. Materials and Methods

2.1. IRD Group and Genetic Testing

This study involved 77 patients with clinical symptoms of retinal dystrophies, recruited by several centres: the Medical University of Silesia in Katowice, Genomed S.A. in Warsaw, and the Pomeranian Medical University in Szczecin (Table S1), and who were included into the cohort between the years 2016 and 2018. The mean age at the time of referral was 40.9 (SD = 16.7).
Patients were subjected to detailed ophthalmological examinations, which involved visual acuity and colour vision testing, autorefractometry, tonometry, perimetry, and dilated fundus examinations, including fundus autofluorescence, optical coherence tomography (OCT), fluorescein angiography, and ERG/EOG, depending on the clinical symptoms. The referring ophthalmologist selected the type of examination for each patient. Results of ophthalmological examinations for a subgroup of retinitis pigmentosa patients were described by Wiacek et al. [19].
Genetic testing was carried out in the laboratory of Genomed S.A., Warsaw, Poland. Genomic DNA was isolated from the peripheral blood using a total DNA isolation kit (Blood Mini kit, A&A Biotechnology, Gdańsk, Poland) according to the manufacturer’s protocol or as described in [20]. Genomic DNA was used as input for constructing an exome-enriched library. Two types of the WES enrichment were used during the study: SureSelectXT Clinical Research Exome or Human Exome V5 (Agilent Technologies, Santa Clara, CA, US) and ClinicalExome (Roche, Basel, Switzerland).
The WES libraries were prepared according to the Agilent or Roche protocols using 100 ng of genomic DNA for each sample.
All the libraries were sequenced using the NextSeq500 or HiSeq4000 (Illumina Inc., San Diego, CA, USA) in the PE150 mode, aiming at the mean target coverage above 100×.
Initial processing of BCL files and demultiplexing was done using the Illumina bcl2fastq. Trimmomatic [21] was applied to trim raw FASTQ files from adapter sequences and low-quality bases. Read mapping to the GRCh37 (hg19) reference genome was performed using the Burrows–Wheeler Alignment tool [22]. Duplicate read pairs were removed with MarkDuplicates (Picard tools package [23]). Alignment files were further processed in accordance with the Genome Analysis Toolkit v3 best practice pipeline [24], and Haplotype Caller was used for variant identification. The quality metrics for the alignment and variants were examined, and all variants within the targeted regions, above the default variant quality threshold, were annotated with Annovar [25]. Further variant filtration and interpretation involved the Gemini framework [26], enabling an analysis of multiple samples in the search for rare pathogenic variants in the whole exome, as well as an in-house variant analysis software (BroVar v2), using the ACMG variant classification guidelines [27,28].
In the first step, targeted exome sequencing data analysis was carried out to identify pathogenic variants in 267 genes associated with retinal dystrophies (based on the RetNet database [1]), and subsequently a search for loss-of-function and likely pathogenic variants of the entire exome sequence was performed for samples with negative results. The result set has been limited to rare variants based on the allele frequencies from 1000 Genomes [29], as well as newly constructed POLGENOM [30] databases. Each rare variant (MAF < 0.01 for 1000 Genomes), with the exception of the known pathogenic variants, was assessed for pathogenicity using multiple in silico predictors, including Alamut ver. 2.9.0 (SOPHiA GENETICS SA, Rolle, Switzerland) with all the incorporated predictors, such as SIFT, Mutation Taster, and Polyphen2. Its presence in variant databases such as ClinVar [31] and HGMD Professional [32] and in the in-house database was checked. Finally, the variant frequency was compared to the expected frequency for the considered disease. Segregation was assessed when possible. Pathogenic and likely pathogenic variants were confirmed using Sanger sequencing if diagnostic results were issued or the quality of NGS data was below expected. Mutation Surveyor V 5.0.1 (Softgenetics, State College, PA, USA) was employed for the Sanger data analysis. Copy number variants were searched for using XHMM [33], and an attempt of copy number variation (CNV) analysis was performed with GermlineCNVCaller from GATKv4 package [34].
A patient case was considered likely solved if one of the confirmatory variants for an autosomal recessive disease was classified as an uncertain significance variant and genotyping results were consistent with clinical data. If such variants were the only ones identified and there were additional data supporting its pathogenicity, a case was also considered likely solved. A partial re-analysis of data was recently performed using updated information on the pathogenicity of presumably causative variants—mainly ClinVar and HGMD Professional—and including re-mapping selected exome datasets to the GRCh38 (hg38) reference genome.
This study was approved by the Ethics Committee of Medical University of Silesia (resolution no. KNW/0022/KB1/105/13, KB 31/2012, and KB/36/A/2013) and adhered to the tenets of The Declaration of Helsinki. Informed written consent was obtained from all the participants.

2.2. POLGENOM Group and Genetic Analysis

Recruitment and selection of nonagenarians and centenarians to the Polish Reference Genome for Genomic Diagnostics and Personalised Medicine research project (POLGENOM, alternative acronym PlGen) was described in Skubiszewska et al. [35]. In addition, 23 participants were selected from the POLSTU Project [36] and 58 from the POLSENIOR project [37]. The criteria for successful aging were as follows: age ≥ 90 years; independence or slight dependence in the activities of daily living (the ADL score ≥ 4); no moderate or severe dementia (the MMSE score ≥ 20); no diagnosis of either diabetes, stroke, myocardial infarction, or cancer (basal cell carcinoma and squamous cell carcinoma were allowed); and diagnosis after the age of 80 years. Only 2 cases of myocardial infarction, 2 cases of stroke, 2 cases of diabetes, and 3 cases of cancers (urothelial carcinoma, breast cancer, uterine cancer) were diagnosed at old age. The POLGENOM study participants were not specifically tested for the presence of ophthalmological conditions. The questionnaire checked for the presence or absence of cataracts. Nevertheless, their visual acuity had to be good enough to perform the MMSE test, as well as to warrant independence in everyday living.
In total, 126 genomes (74 women and 52 men) of successfully aging individuals aged ≥ 90 years were sequenced and analysed. The general characteristics of the study group are shown in Table S2.
Genomic DNA was isolated using a salting-out procedure [38]. Whole-genome sequences were obtained using HiSeq 2000 and HiSeq X Ten (Illumina Inc., San Diego, CA, USA) in the PE150 mode, with a mean 30× coverage. An in-house-developed bioinformatic pipeline was used that started by the initial processing of BCL files, along with demultiplexing with the Illumina bcl2fastq, followed by raw FASTQ file trimming from adapter sequences and low-quality bases using Trimmomatic [21] and ending by read mapping to GRCh38 (hg38) with a Burrows–Wheeler Aligner [22]. Alignment files were further processed accordingly to the Genome Analysis Toolkit v3 best practice pipeline [24], with the use of HaplotypeCaller for variant identification, while VEP Ensembl release 83 [39] was applied to annotate sequence variants. Additionally, LOFTEE (version 0.2; Loss-of-function Transcript Effect Estimator), a plugin to VEP, was applied to identify high confidence loss-of-function variants. Variants that had not passed the following filters: coverage sufficient for variant calling in more than 33 samples, location out of low complexity regions, successful GATK variant recalibration, were excluded from further analysis. Selected variants were confirmed using the classical (Sanger sequencing) method.
Mitochondrial genome sequences were also analysed using CLC Genomics Workbench 7 (Qiagen, Qiagen, Hilden, Germany), Mitomaster, MITOMAP, and mtDB Genebank databases [40,41,42]. Identified and categorised variants were added to the POLGENOM database.

2.3. POLGENOM Reference Database

The genomic database (https://polgenom.pl/) is supported by one dedicated server, containing a database and application. The Ubuntu LTS version of the Linux operating system was installed on specially selected computer hardware. The database is served by the PostgreSQL engine. The application was written in JavaScript, using the Express/Node.js framework.

3. Results

3.1. IRD Group

Seventy-seven patients diagnosed with retinal dystrophies underwent genetic testing using the WES approach. The group included 34 (44.2%) women and 43 (55.8%) men, unrelated except for two families, with parent–child duos (Table 1). Patients were divided into two phenotypic groups: the main group included 47 patients with clinical symptoms of retinitis pigmentosa (RP) (61.0%), and the other one included those with unspecific IRD symptoms (30, 39.0%).

3.2. POLGENOM Group and Database

The POLGENOM group served as the Polish Caucasian-specific control group for this study. The inclusion criteria into the successfully aging group were fulfilled by 126 individuals (Table S2), for whom the whole-genome sequencing and bioinformatic analysis were carried out, and whose genomic variants were used to create the database records.

3.2.1. Statistics

Almost 16.7 million variants in the nuclear genome of the study participants were found (22 million before filtering), including 15.3 million SNVs and 1.4 million small insertions and deletions (InDels). A total of 13.6 million (81%) variants have been already present in the dbSNP database (build 147 and 148), 12.3 million (73.4%) in the 1000 Genomes Project (1000G, Phase 3). Almost 2.76 million novel SNVs and 0.34 million novel indels were detected (database record on 18 February 2019).
The POLGENOM project results were also compared with the current version of the GnomAD database (v4.1.0) (https://gnomad.broadinstitute.org), revealing 1,197,199 novel variants found in the Polish population, of which 1,149,439 (96%) are SNVs, 13,631 small insertions, and 34,129 small deletions.

3.2.2. Mitochondrial Data

For each long-lived individual, a list of mtDNA variants with their respective heteroplasmy level and haplogroup assignment was obtained. The identified variants were classified based on their frequency in the GenBank and mtDB databases, localisation, functional effect, and clinical data (MitoMaster, MITOMAP). In total, there were 215 variants with <0.5% frequency in mtDB or newly identified that were included into the database.

3.3. Molecular Diagnosis of IRD

In total, a definitive molecular diagnosis was reached in 45 (58.44%) cases, while 5 cases could be considered as likely solved, which resulted in total of 50 solved and likely solved cases (64,94%). For seven patients, results were inconclusive. For 20 patients, causative variants could not be found, and only carriership for pathogenic or likely pathogenic variants was detected. Positive cases among RP patients accounted for 63.8%, and for the unspecific IRD group, 66.6%.
Confirmatory variants were found in 22 distinct genes (Table 2). The largest group of patients had causative variants in the ABCA4 gene (10, 20%), followed by USH2A (7, 14%), EYS (4, 8%), and RS1 (4, 8%) (Figure 1). A total of 11 (22%) patients presented causative variants in genes that were identified only once in our group. Looking at the distributions of confirmatory genes for both groups, namely, patients with indication of RP and the unspecific IRD group (Figure 2), a predominance of patients with pathogenic variants identified in ABCA4 was clearly visible in the unspecific IRD group, whereas USH2A, EYS, and PRPF31 genes contributed the most to the genetic diagnosis of patients in the RP group.
Pathogenic or likely pathogenic variants were identified predominantly in genes already known to cause retinal degenerations (with a significant number of ABCA4 variants), although variants in genes newly identified as involved in retinal disorders were also uncovered (such as CFAP410 and GUCA1A). A new clinical diagnosis could be suggested in three cases with the initial indication of RP (CHM, AHI1, RS1—single case each), as well as in five cases from the Unspecific IRD group, where a new diagnosis was found due to causal variants in the VCAN and RS1 genes (two cases and three cases, respectively) (Table 1).
The ABCA4 (NM_000350.3, NP_000341.2) complex haplotype p.(Leu541Pro; Ala1038Val) was observed in three patients, while the hypomorphic variant (c.5603A>T p.Asn1868Ile) was observed in two patients. All these patients were referred for genetic diagnostics at an older age (45 years and more, Table S1), confirming the correlation of these variants with a milder form of late-onset macular dystrophy.
Missense variants corresponded to 49% of the identified confirmatory variants, frameshift variants to 19%, stopgain variants to 18%, and splicing ones to 11%. One of the few splicing variants identified in the study group was a canonical splice site variant NM_004385.5:c.4004-2A>G in VCAN. This transition presumably modifies the 3′ acceptor splice site of intron 7 and activates a cryptic exon 8 splice site, 39 bp downstream from the authentic one [43,44]. In one patient, two missense variants (NM_000329.3:c.131G>A p.Arg44Gln and NM_000329.3:c.709C>T p.Pro237Ser) in the RPE65 gene were uncovered, the first one being a well-characterised pathogenic variant while the second one a likely pathogenic variant with only one record in ClinVar but with strong predictions of pathogenicity, obtained using a set of prediction programmes. Both variants change highly conserved amino acids, affecting the RPE65 protein function. In one patient, two missense variants in the AHI1 gene (NM_017651.5, NP_060121.3) were indicated as the most probable causative variants (c.1828C>T p.Arg610Ter and c.1829G>C p.Arg610Pro), indicating a possible genetic diagnosis of Joubert Syndrome type 3.
Seven novel variants identified in our cohort are still not present in any databases (Table 2). They include missense variants in PRPF31 (NM_015629.4:c.831T>A p.Ser277Arg) and in USH2A (NM_206933.4:c.10732A>C p.Ser3578Arg); frameshift variants in ABCA4 (c.2548dup p.Tyr850LeufsTer9), in EYS (NM_001142800.2:c.4462_4469dup p.Met1491AlafsTer12), and PRPF31 (NM_015629.4:c.428_430delins22ACAAGTGCAAGGCTGTTCTTGC p.Gly143AspfsTer17); and a splicing variant in CHM (NM_000390.4:c.49+2T>A). Of those, the ABCA4 and EYS variants have been subsequently uncovered in a single patient each, while the USH2A variant was found twice in our further IRD panel studies (Matczyska et al. [45]).
The POLGENOM database served as a reference database in the bioinformatic analysis of data. Data on the POLGENOM frequency of IRD causative variants, identified in this study, are shown in Table 2. Interestingly, only the predominant ABCA4 variant in our study group (found in five patients, always in compound heterozygosity with another variant), c.5882G>A p.Gly1961Glu, and the hypomorphic variant, c.5603A>T, p.Asn1868Ile (found in two patients, always in compound heterozygosity with a severe pathogenic variant), were present.
Among the most interesting unsolved cases was a family duo in which two VUS variants of a possible AD effect were uncovered (in AIPL1 and RP1) with the second variant that could possibly influence the clinical phenotype (Table S1). This seems to be a case that cannot be solved without a deeper analysis of all the family members, including a search for CNVs.

3.4. Selected Genetic Diagnosis Cases

Using the WES approach for the recruited group of the IRD patients, a few rare causative genes and treatable disease cases were uncovered. VUS variants corresponding to clinical symptoms were reported as likely causative due to their absence from population databases and indications of prediction programmes suggesting their strong pathogenicity. As the diagnostic process is always based on collaboration between a molecular geneticist and a referring physician, the final diagnostic decision is made by an ophthalmogenetic specialist.

3.4.1. Retinitis Pigmentosa Case (Compound Heterozygote in RPE65)

The young (14-year-old) male patient, for whom two missense variants in the RPE65 gene were uncovered, reported the following symptoms: narrow field of vision, decreased visual acuity, and poor night vision. The patient also had myopic astigmatism. Best corrected visual acuity: OD 0.1; OS: 0.3 by Snellen. In May 2023, he received treatment with Luxturna (voretigene neparvovec, Spark Therapeutics/Novartis) in the eye with lower visual acuity, which allowed for the stabilisation of retinal morphology, improvement in night vision, and reduced photophobia.

3.4.2. Wagner Syndrome - Family Case (Heterozygote in VCAN)

Wagner syndrome was detected for the first time in Poland in a family previously diagnosed as having dominant retinitis pigmentosa (Figure 3). Wagner syndrome, also known as dominant hyaloideoretinal dystrophy or VCAN-related vitreoretinopathy, is a hereditary eye disorder that leads to progressive vision loss. The vitreous becomes thin and watery, appearing empty. Retinal tractions and tears can result in the retinal detachment. Wagner syndrome is associated by ophthalmologists primarily with the above-mentioned structural changes of the vitreous body. Therefore, in patients who do not present with early retinal detachment, diagnosis may be difficult. Other signs and symptoms are decreased visual acuity (often associated with lens opacities), progressive night blindness, and a narrowing visual field, as well as thinned retina and choroid, pigment changes in the periphery of the retina, and slightly reduced thickness of retinal vessels, which all may suggest retinitis pigmentosa.
Wagner syndrome is caused by genetic changes in the VCAN gene (previously named CSPG2), which encodes Versican protein, a chondroitin sulphate proteoglycan-2 found in the vitreous among other tissues. It is an extremely rare disease, with not much more than 250 patients having been described to date (together with another disease due to mutations in the VCAN gene-erosive vitreoretinopathy) [44]. The canonical splice site variant NM_004385.5:c.4004-2A>G in VCAN, found in this study, was previously reported only in two families [44,45], while, based on the HGMD records, splicing mutations in intron 7 and 8 as well as deletions of the entire exon 8 or of its part are the main causes of Wagner syndrome.

4. Discussion

We attempted one of the first and the widest WES-based characterisations of variants causing inherited retinal dystrophies in the Polish Caucasian population.
To facilitate all types of high-throughput genetic testing for the Polish Caucasian population, a Polish Genomic Database was created, including WGS sequences together with biochemical data and health records of 126 health nonagenarians and centenarians.
Such databases have been created for other populations, beginning from the Icelandic one [46], and currently the TopMed [47] and the UK project [48], but rarely for a selected group of healthy, long-lived people, gathering additional, easily accessible data on their individual health status. The first data from a larger Polish whole-genome project were published in 2022 [49]. Its participants were recruited by several medical centres to study the genetic susceptibility to COVID-19 infections, and therefore the cohort is not representative of the entire Polish population. Moreover, these data are currently not available for other projects and diagnostic purposes while the POLGENOM database, a homogenous whole-genome database of healthy extremely age-advanced individuals, remains the only accessible reference genomic database for the Polish Caucasian population. It is available at https://polgenom.pl/ for any type of project requiring assembled genetic and health status records for the Caucasian population.
This combined approach allowed us to provide molecular confirmation of clinical diagnosis in 50 (64.94%) cases and establish possible genetic aetiology in 7 cases out of 77 WES-analysed individuals. This result was an improvement in regard to the initial conclusions, owing to the partial re-analysis of the variant lists, taking into account new data on the pathogenicity of variants, especially those in the ABCA4 gene. The data re-analysis led us also to the re-classification of three variants from VUS to likely benign and removal of these cases from the list of possibly solved cases. The final diagnostic yield, however, is comparable to the results of other NGS-based studies of IRD in different populations. The diagnostic success rate varied between 51 and 76% [4,6,7,8], depending on the approach adopted (NGS panel/WES/WGS), availability of CNV analysis, and the study group composition. Interestingly, the highest success rate reported for WES was 89% for 19 well-characterised non-syndromic Polish RP patients, owing to a large set of clinical/family data; extensive feedback from an ophthalmogenetic specialist; and an improved CNV analysis based on consistent data from a newer, fine-tuned exome enrichment [14].
The prevalence of confirmatory ABCA4 variants, noted in the unspecific IRD group, correlated with the higher referral age of patients (Table S3), is of importance for diagnostics and medical care for elder patients with eye disease problems of unclear aetiology, suggesting a need for genetic testing even at a more advanced age. A diagnosis of a late-onset genetic disorder may allow for the avoidance of unnecessary treatment with a risk of serious adverse effects, especially since differential diagnosis often includes inflammatory diseases requiring immunosuppressive or biological agents.
This was the first WES-based study of a large group of patients with inherited eye diseases initiated in Poland. The introduction of a new diagnostic method, based on the whole-exome analysis, encouraged people who had not previously decided to use genetic diagnostics, especially in diseases considered difficult to confirm using classical methods. Interestingly, this may apply to situations where a specific gene (or set of genes) is strongly suspected as the cause of the disease, but a slight change in the phenotype compared to the descriptions of the disease in scientific publications may mislead the initial diagnosis. This way, for the first time in Poland, Wagner syndrome was detected and genetically confirmed in a family previously diagnosed as having dominant retinitis pigmentosa. The identification of the genetic background of the disease allowed for the provision of specific lifestyle advice, a warning against the symptoms of retinal detachment, and a discontinuation of unnecessary medications.
For the patient with the RPE65 mutations, a gene therapy could be started with positive results two months after its application. Other patients with positive genetic results, including pathogenic variants in most frequently mutated genes (such as ABCA4, EYS, USH2A), new genetic counselling options and new therapies will be available soon, with genome editing as one of the options.
Among future gene therapies for RP, there is a combination of gene therapy and so-called optogenetics, with two complementary components: a construct, encoding a photoactivatable channel rhodopsin protein, delivered via a modified adenovirus vector, and biomimetic goggles that stimulate the engineered retinal cells [50,51]. A phase I/II trial was successfully ended for intravitreal antisense oligonucleotide sepofarsen [52] to treat LCA type 10, caused by CEP290 mutations. Another advanced and positive gene therapy clinical trial is that for patients with LCA type 1 caused by mutations in GUCY2D [53]. Treatment of LCA5 gene-dependent LCA (type 5) is similarly advanced [54], while a so-called “molecular photoswitch”, conferring light-sensing capabilities to special types of retinal neuron cells, could potentially become a new treatment in RP and choroideremia [55]. Future treatment of IRD will, however, probably rely on CRISPR-Cas9-mediated gene editing either in human-induced pluripotent stem cells derived from patients and subsequent subretinal transplantation of differentiated retinal cells, or in vivo, by subretinal injection of a modified adenovirus vector carrying a CRISPR-Cas9-guide RNA-expressing construct. Successful results in cell cultures were achieved for several types of retinal dystrophies such as Stargardt disease [56] and LCA [57]. For patients with the most common CEP290 mutation, a gene editing treatment is already available in frame of a clinical trial [58]. A gene-agnostic treatment—an antioxidant nanotherapeutic—could also be available soon [59]. It has been proposed for diseases affecting the posterior segment of the eye and it consists of polycaprolactone nanoparticles delivering anti-oxidative drugs to the retina with a durable effect.
Possible reasons of unsolved cases or nonconclusive results in our study are the following, based on our current experience: exome enrichments still under development—in consequence, GC-rich regions, regions with tandem repeats, and genes with pseudogenes have not been efficiently sequenced; bioinformatic pipelines were still under development during the study period; inefficient identification of CNV and of deep intronic variants; not enough data on variant pathogenicity, with databases and tools still under development; data missing on new genes involved in retinal degeneration; and the study group was too small to identify new IRD genes [6,7,45].
Taking into account the presented results, we focused our efforts on developing an optimised NGS panel, because it allowed for an overall increase in uniformity and level of coverage at a lower cost (along with an improved coverage in cumbersome regions, such as the RPGR ORF15), supporting a successful CNV analysis [45]. This approach enabled us to reach the final molecular diagnosis in one of the patients of this cohort (Table S1, patient S42). A customised exome enrichment with additional probes improving CNV identification and sequencing of deep intron variants could be an option with a similar diagnostic yield. Surely, the WGS approach should allow for up to 10% increase in molecular diagnoses when compared to WES and targeted approaches as a consequence of a more homogeneous coverage. It has not only a much higher potential for identifying structural variants such as translocations and inversions due to significantly more complete analysis of breakpoints but also allows for a more comprehensive sequencing of variants in non-coding regions [6].

5. Conclusions

Our results confirm that over 60% IRD patients referred for genetic testing by ophthalmologists and clinical geneticists with experience in IRD diagnosis can expect a positive result, based on WES analysis limited to the RetNet scope. Causative variants for an extremely rare disease (Wagner syndrome) as well as those in the gene being a target of one of the first gene therapies (RPE65 and Luxturna) were able to be uncovered using this WES approach, still under development a few years ago, supported by the availability of a reference population genomic database.
The results of this study supported, however, the development of a targeted retinal panel that would cover most pathogenic variants occurring in the Polish population, and would allow for a fast and low-cost genetic analysis. Along the diagnostic algorithm proposed by our group (Matczyńska et al. [45]), such a targeted approach can be followed by a wider dystrophy panel, based on WES or WGS for unsolved cases, eventually leading to the final diagnosis and selection of a personalised therapy. The POLGENOM database remains a constant support for such types of analysis.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/genes15081011/s1. Table S1: List of 77 IRD patients along with the confirmatory/VUS variants. Table S2: General characteristics of the POLGENOM group. Table S3: Genes with confirmatory variants grouped by patients’ age quartile.

Author Contributions

Conceptualisation, M.P.-K., M.M., E.M., S.T. and A.B.-C.; methodology, E.M., R.S. and A.B.-C.; software, E.M. and P.Ł.; validation, E.M. and A.B.-C.; formal analysis, E.M.; investigation, E.M., R.S., K.S., K.K., M.B.-G., M.J., E.S., P.Ł., M.W. and A.B.-C.; resources, S.T., M.Z. and A.B.-C.; data curation, E.M. and A.B.-C.; writing—original draft preparation, E.M., S.T. and A.B.-C.; writing—review and editing, M.P.-K., A.M., M.J., R.S., E.M., S.T., M.Z., E.W. and A.B.-C.; visualisation, E.M.; supervision, A.B.-C.; project administration, M.P.-K., M.M., E.W., S.T. and A.B.-C.; funding acquisition, M.P.-K., M.M., A.M., E.W., S.T., M.Z. and A.B.-C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Centre for Research and Development grant STRATEGMED1/234261/2/NCBR/2014 and INNOTECH-K2/IN2/10/181852/NCBR/13. E.M. was partially supported by the Ministry of Science and Higher Education grant PCTT/1874/2019.

Institutional Review Board Statement

The study was conducted in accordance with the Declaration of Helsinki and approved by the Ethics Committee of Medical University of Silesia (resolution no. KNW/0022/KB1/105/13, 8 October 2013) and Ethics Committee of the Medical University of Warsaw (KB80/2012, 14 February 2012).

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

All necessary data are included in the Supplementary Materials. Please contact the corresponding author to discuss the possibility of obtaining additional data or detailed information.

Acknowledgments

We thank the patients for their participation in this study. We also thank A. Piotrowska-Nowak for mtDNA analysis of the Polgenom group.

Conflicts of Interest

E.M., M.B.-G., K.S., E.G., P.Ł., R.S., M.J., M.Z., and A.B.-C. are employees of Genomed S.A. A.B.-C. and M.Z. keep stocks in Genomed S.A. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. RetNet. The Retinal Information Network. Available online: https://web.sph.uth.edu/RetNet/ (accessed on 2 April 2024).
  2. Verbakel, S.K.; van Huet, R.A.C.; Boon, C.J.F.; den Hollander, A.I.; Collin, R.W.J.; Klaver, C.C.W.; Hoyng, C.B.; Roepman, R.; Klevering, B.J. Non-syndromic retinitis pigmentosa. Prog. Retin. Eye Res. 2018, 66, 157–186. [Google Scholar] [CrossRef] [PubMed]
  3. Haim, M. Epidemiology of retinitis pigmentosa in Denmark. Acta Ophthalmol. Scand. 2002, s233, 1–34. [Google Scholar] [CrossRef] [PubMed]
  4. Tuupanen, S.; Gall, K.; Sistonen, J.; Saarinen, I.; Kämpjärvi, K.; Wells, K.; Merkkiniemi, K.; von Nandelstadh, P.; Sarantaus, L.; Känsäkoski, J.; et al. Prevalence of RPGR-Mediated Retinal Dystrophy in an Unselected Cohort of Over 5000 Patients. Transl. Vis. Sci. Technol. 2022, 11, 6. [Google Scholar] [CrossRef] [PubMed]
  5. Zeviani, M.; Carelli, V. Mitochondrial Retinopathies. Int. J. Mol. Sci. 2021, 23, 210. [Google Scholar] [CrossRef] [PubMed]
  6. Weisschuh, N.; Mazzola, P.; Zuleger, T.; Schaeferhoff, K.; Kühlewein, L.; Kortüm, F.; Witt, D.; Liebmann, A.; Falb, R.; Pohl, L.; et al. Diagnostic Genome Sequencing Improves Diagnostic Yield: A Prospective Single-Centre Study in 1000 Patients with Inherited Eye Diseases. J. Med. Genet. 2023, 61, 186–195. [Google Scholar] [CrossRef] [PubMed]
  7. Karali, M.; Testa, F.; Di Iorio, V.; Torella, A.; Zeuli, R.; Scarpato, M.; Romano, F.; Onore, M.E.; Pizzo, M.; Melillo, P.; et al. Genetic Epidemiology of Inherited Retinal Diseases in a Large Patient Cohort Followed at a Single Center in Italy. Sci. Rep. 2022, 12, 20815. [Google Scholar] [CrossRef] [PubMed]
  8. Peter, V.G.; Kaminska, K.; Santos, C.; Quinodoz, M.; Cancellieri, F.; Cisarova, K.; Gobert, R.P.; Rodrigues, R.; Custódio, S.; Paris, L.P.; et al. The First Genetic Landscape of Inherited Retinal Dystrophies in Portuguese Patients Identifies Recurrent Homozygous Mutations as a Frequent Cause of Pathogenesis. PNAS Nexus 2023, 2, pgad043. [Google Scholar] [CrossRef] [PubMed]
  9. Available online: https://www.ema.europa.eu/en/medicines/human/EPAR/luxturna#ema-inpage-item-authorisation-details (accessed on 11 February 2024).
  10. Fischer, M.D.; Simonelli, F.; Sahni, J.; Holz, F.G.; Maier, R.; Fasser, C.; Suhner, A.; Stiehl, D.P.; Chen, B.; Audo, I.; et al. Real-World Safety and Effectiveness of Voretigene Neparvovec: Results up to 2 Years from the Prospective, Registry-Based PERCEIVE Study. Biomolecules 2024, 14, 122. [Google Scholar] [CrossRef] [PubMed]
  11. Chew, L.A.; Iannaccone, A. Gene-agnostic approaches to treating inherited retinal degenerations. Front. Cell. Dev. Biol. 2023, 11, 1177838. [Google Scholar] [CrossRef]
  12. Tracewska, A.M.; Kocyła-Karczmarewicz, B.; Rafalska, A.; Murawska, J.; Jakubaszko-Jablonska, J.; Rydzanicz, M.; Stawiński, P.; Ciara, E.; Khan, M.I.; Henkes, A.; et al. Genetic Spectrum of ABCA4-Associated Retinal Degeneration in Poland. Genes 2019, 10, 959. [Google Scholar] [CrossRef]
  13. Tracewska, A.M.; Kocyla-Karczmarewicz, B.; Rafalska, A.; Murawska, J.; Jakubaszko-Jablonska, J.; Rydzanicz, M.; Stawinski, P.; Ciara, E.; Lipska-Zitkiewicz, B.S.; Khan, M.I.; et al. Non-Syndromic Inherited Retinal Diseases in Poland: Genes, Mutations, and Phenotypes. Mol. Vis. 2021, 27, 457–465. [Google Scholar]
  14. Wawrocka, A.; Socha, M.; Walczak-Sztulpa, J.; Koczyk, G.; Skorczyk-Werner, A.; Krawczyński, M.R. Molecular Re-Diagnosis with Whole-Exome Sequencing Increases the Diagnostic Yield in Patients with Non-Syndromic Retinitis Pigmentosa. Diagnostics 2023, 13, 730. [Google Scholar] [CrossRef] [PubMed]
  15. Nowomiejska, K.; Baltaziak, K.; Całka, P.; Ciesielka, M.; Teresiński, G.; Rejdak, R. Identification of the RPGR Gene Pathogenic Variants in a Cohort of Polish Male Patients with Retinitis Pigmentosa Phenotype. Genes 2023, 14, 1950. [Google Scholar] [CrossRef]
  16. Skorczyk-Werner, A.; Niedziela, Z.; Stopa, M.; Krawczyński, M.R. Novel Gene Variants in Polish Patients with Leber Congenital Amaurosis (LCA). Orphanet. J. Rare Dis. 2020, 15, 345. [Google Scholar] [CrossRef]
  17. Skorczyk-Werner, A.; Sowińska-Seidler, A.; Wawrocka, A.; Walczak-Sztulpa, J.; Krawczyński, M.R. Molecular Background of Leber Congenital Amaurosis in a Polish Cohort of Patients—Novel Variants Discovered by NGS. J. Appl. Genet. 2023, 64, 89–104. [Google Scholar] [CrossRef]
  18. Wawrocka, A.; Skorczyk-Werner, A.; Wicher, K.; Niedziela, Z.; Ploski, R.; Rydzanicz, M.; Sykulski, M.; Kociecki, J.; Weisschuh, N.; Kohl, S.; et al. Novel Variants Identified with Next-Generation Sequencing in Polish Patients with Cone-Rod Dystrophy. Mol. Vis. 2018, 24, 326–339. [Google Scholar]
  19. Wiącek, M.P.; Gosławski, W.; Grabowicz, A.; Sobuś, A.; Kawa, M.P.; Baumert, B.; Paczkowska, E.; Milczarek, S.; Osȩkowska, B.; Safranow, K.; et al. Long-Term Effects of Adjuvant Intravitreal Treatment with Autologous Bone Marrow-Derived Lineage-Negative Cells in Retinitis Pigmentosa. Stem Cells Int. 2021, 2021, 6631921. [Google Scholar] [CrossRef]
  20. Ulańczyk, Z.; Grabowicz, A.; Mozolewska-Piotrowska, K.; Safranow, K.; Kawa, M.P.; Pałucha, A.; Krawczyk, M.; Sikora, P.; Matczyńska, E.; Machaliński, B.; et al. Genetic Factors Associated with Age-Related Macular Degeneration: Identification of a Novel PRPH2 Single Nucleotide Polymorphism Associated with Increased Risk of the Disease. Acta Ophthalmol. 2021, 99, 739–749. [Google Scholar] [CrossRef]
  21. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A Flexible Trimmer for Illumina Sequence Data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [PubMed]
  22. Li, H.; Durbin, R. Fast and Accurate Short Read Alignment with Burrows-Wheeler Transform. Bioinformatics 2009, 25, 1754–1760. [Google Scholar] [CrossRef]
  23. Broad Institute. Picard Tools. Available online: http://broadinstitute.github.io/picard (accessed on 2 April 2024).
  24. Van der Auwera, G.A.; O’Connor, B.D. Genomics in the Cloud: Using Docker, GATK, and WDL in Terra, 1st ed.; O’Reilly Media: Sebastopol, CA, USA, 2020. [Google Scholar]
  25. Wang, K.; Li, M.; Hakonarson, H. ANNOVAR: Functional Annotation of Genetic Variants from High-Throughput Sequencing Data. Nucleic Acids Res. 2010, 38, e164. [Google Scholar] [CrossRef]
  26. Paila, U.; Chapman, B.A.; Kirchner, R.; Quinlan, A.R. GEMINI: Integrative exploration of genetic variation and genome annotations. PLoS Comput. Biol. 2013, 9, e1003153. [Google Scholar] [CrossRef]
  27. Richards, S.; Aziz, N.; Bale, S.; Bick, D.; Das, S.; Gastier-Foster, J.; Grody, W.W.; Hegde, M.; Lyon, E.; Spector, E.; et al. Standards and Guidelines for the Interpretation of Sequence Variants: A Joint Consensus Recommendation of the American College of Medical Genetics and Genomics and the Association for Molecular Pathology. Genet. Med. 2015, 17, 405–424. [Google Scholar] [CrossRef]
  28. Ellard, S.; Baple, E.L.; Callaway, A.; Berry, I.; Forrester, N.; Turnbull, C.; Owens, M.; Eccles, D.M.; Abbs, S.; Scott, R.; et al. ACGS Best Practice Guidelines for Variant Classification in Rare Disease 2020. v4.01; Association for Clinical Genomic Science: London, UK, 2020. [Google Scholar]
  29. 1000 Genomes Project Consortium; Auton, A.; Brooks, L.D.; Durbin, R.M.; Garrison, E.P.; Kang, H.M.; Korbel, J.O.; Marchini, J.L.; McCarthy, S.; McVean, G.A.; et al. A global reference for human genetic variation. Nature 2015, 526, 68–74. [Google Scholar] [CrossRef]
  30. POLGENOM. Available online: https://polgenom.pl/ (accessed on 2 April 2024).
  31. 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]
  32. Stenson, P.D.; Mort, M.; Ball, E.V.; Howells, K.; Phillips, A.D.; Cooper, D.N.; Thomas, N.S.T. The Human Gene Mutation Database: 2008 Update. Genome Med. 2009, 1, 13. [Google Scholar] [CrossRef]
  33. Fromer, M.; Purcell, S.M. Using XHMM Software to Detect Copy Number Variation in Whole-Exome Sequencing Data. Curr. Protoc. Hum. Genet. 2014, 81, 7.23.1–7.23.21. [Google Scholar] [CrossRef] [PubMed]
  34. Babadi, M.; Fu, J.M.; Lee, S.K.; Smirnov, A.N.; Gauthier, L.D.; Walker, M.; Benjamin, D.I.; Zhao, X.; Karczewski, K.J.; Wong, I.; et al. GATK-GCNV Enables the Discovery of Rare Copy Number Variants from Exome Sequencing Data. Nat. Genet. 2023, 55, 1589–1597. [Google Scholar] [CrossRef] [PubMed]
  35. Skubiszewska, A.; Broczek, K.; Maruniak-Chudek, I.; Oledzka, G.; Jonas, M.I.; Puzianowska-Kuznicka, M.; Mossakowska, M. Frailty and Survivability of Polish Caucasian Nonagenarians and Centenarians. Geriatrics 2024, 9, 14. [Google Scholar] [CrossRef] [PubMed]
  36. Mossakowska, M.; Barcikowska, M.; Broczek, K.; Grodzicki, T.; Klich-Raczka, A.; Kupisz-Urbanska, M.; Podsiadly-Marczykowska, T.; Sikora, E.; Szybinska, A.; Wieczorowska-Tobis, K.; et al. Polish Centenarians Programme. Multidisciplinary studies of successful ageing: Aims, methods, and preliminary results. Exp. Gerontol. 2008, 43, 238–244. [Google Scholar] [CrossRef]
  37. Bledowski, P.; Mossakowska, M.; Chudek, J.; Grodzicki, T.; Milewicz, A.; Szybalska, A.; Wieczorowska-Tobis, K.; Wiecek, A.; Bartoszek, A.; Dabrowski, A.; et al. Medical, psychological and socioeconomic aspects of aging in Poland: Assumptions and objectives of the PolSenior project. Exp. Gerontol. 2011, 46, 1003–1009. [Google Scholar] [CrossRef] [PubMed]
  38. Miller, S.A.; Dykes, D.D.; Polesky, H.F. A simple salting out procedure for extracting DNA from human nucleated cells. Nucleic Acids Res. 1988, 16, 1215. [Google Scholar] [CrossRef] [PubMed]
  39. McLaren, W.; Gil, L.; Hunt, S.E.; Riat, H.S.; Ritchie, G.R.S.; Thormann, A.; Flicek, P.; Cunningham, F. The Ensembl Variant Effect Predictor. Genome Biol. 2016, 17, 122. [Google Scholar] [CrossRef] [PubMed]
  40. Brandon, M.C.; Ruiz-Pesini, E.; Mishmar, D.; Procaccio, V.; Lott, M.T.; Nguyen, K.C.; Spolim, S.; Patil, U.; Baldi, P.; Wallace, D.C. MITOMASTER: A bioinformatics tool for the analysis of mitochondrial DNA sequences. Hum. Mutat. 2009, 30, 1–6. [Google Scholar] [CrossRef] [PubMed]
  41. Brandon, M.C.; Lott, M.T.; Nguyen, K.C.; Spolim, S.; Navathe, S.B.; Baldi, P.; Wallace, D.C. MITOMAP: A human mitochondrial genome database—2004 update. Nucleic Acids Res. 2005, 33, D611–D613. [Google Scholar] [CrossRef] [PubMed]
  42. Ingman, M.; Gyllensten, U. mtDB: Human Mitochondrial Genome Database, a resource for population genetics and medical sciences. Nucleic Acids Res. 2006, 34, D749–D751. [Google Scholar] [CrossRef] [PubMed]
  43. Tang, P.H.; Velez, G.; Tsang, S.H.; Bassuk, A.G.; Mahajan, V.B. VCAN Canonical Splice Site Mutation is Associated with Vitreoretinal Degeneration and Disrupts an MMP Proteolytic Site. Investig. Ophthalmol. Vis. Sci. 2019, 60, 282–293. [Google Scholar] [CrossRef] [PubMed]
  44. Miyamoto, T.; Inoue, H.; Sakamoto, Y.; Kudo, E.; Naito, T.; Mikawa, T.; Mikawa, Y.; Isashiki, Y.; Osabe, D.; Shinohara, S.; et al. Identification of a novel splice site mutation of the CSPG2 gene in a Japanese family with Wagner syndrome. Investig. Ophthalmol. Vis. Sci. 2005, 46, 2726–2735. [Google Scholar] [CrossRef] [PubMed]
  45. Matczyńska, E.; Beć-Gajowniczek, M.; Sivitskaya, L.; Gregorczyk, E.; Łyszkiewicz, P.; Szymańczak, R.; Jędrzejowska, M.; Wylęgała, E.; Krawczyński, M.R.; Teper, S.; et al. Optimised, Broad NGS Panel for Inherited Eye Diseases to Diagnose 1000 Patients in Poland. Biomedicines 2024, 12, 1355. [Google Scholar] [CrossRef]
  46. Gudbjartsson, D.F.; Helgason, H.; Gudjonsson, S.A.; Zink, F.; Oddson, A.; Gylfason, A.; Besenbacher, S.; Magnusson, G.; Halldorsson, B.V.; Hjartarson, E.; et al. Large-scale whole-genome sequencing of the Icelandic population. Nat. Genet. 2015, 47, 435–444. [Google Scholar] [CrossRef]
  47. Taliun, D.; Harris, D.N.; Kessler, M.D.; Carlson, J.; Szpiech, Z.A.; Torres, R.; Taliun, S.A.G.; Corvelo, A.; Gogarten, S.; Kang, H.M.; et al. Sequencing of 53,831 diverse genomes from the NHLBI TOPMed Program. Nature 2021, 590, 290–299. [Google Scholar] [CrossRef]
  48. Halldorsson, B.V.; Eggertsson, H.P.; Moore, K.H.S.; Hauswedell, H.; Eiriksson, O.; Ulfarsson, M.O.; Palsson, G.; Hardarson, M.T.; Oddsson, A.; Jensson, B.O.; et al. The sequences of 150,119 genomes in the UK Biobank. Nature 2022, 607, 732–740. [Google Scholar] [CrossRef] [PubMed]
  49. Kaja, E.; Lejman, A.; Sielski, D.; Sypniewski, M.; Gambin, T.; Dawidziuk, M.; Suchocki, T.; Golik, P.; Wojtaszewska, M.; Mroczek, M.; et al. The Thousand Polish Genomes-A Database of Polish Variant Allele Frequencies. Int. J. Mol. Sci. 2022, 23, 4532. [Google Scholar] [CrossRef] [PubMed]
  50. GenSight Biologics. Available online: https://www.gensight-biologics.com/clinical-development-summary (accessed on 2 April 2024).
  51. Mansouri, V. X-Linked Retinitis Pigmentosa Gene Therapy: Preclinical Aspects. Ophthalmol. Ther. 2022, 12, 7–34. [Google Scholar] [CrossRef] [PubMed]
  52. Russell, S.R.; Drack, A.V.; Cideciyan, A.V.; Jacobson, S.G.; Leroy, B.P.; Van Cauwenbergh, C.; Ho, A.C.; Dumitrescu, A.V.; Han, I.C.; Martin, M.; et al. Intravitreal Antisense Oligonucleotide Sepofarsen in Leber Congenital Amaurosis Type 10: A Phase 1b/2 Trial. Nat. Med. 2022, 28, 1014–1021. [Google Scholar] [CrossRef]
  53. Jacobson, S.G.; Cideciyan, A.V.; Ho, A.C.; Roman, A.J.; Wu, V.; Garafalo, A.V.; Sumaroka, A.; Krishnan, A.K.; Swider, M.; Mascio, A.A.; et al. Night Vision Restored in Days after Decades of Congenital Blindness. iScience 2022, 25, 105274. [Google Scholar] [CrossRef]
  54. Opus Genetics, Inc. NCT05616793. Available online: https://classic.clinicaltrials.gov/ct2/show/study/NCT05616793 (accessed on 11 February 2024).
  55. Kiora Pharmaceuticals, Inc. NCT05282953. Available online: https://classic.clinicaltrials.gov/ct2/show/NCT05282953 (accessed on 11 February 2024).
  56. Siles, L.; Ruiz-Nogales, S.; Navinés-Ferrer, A.; Méndez-Vendrell, P.; Pomares, E. Efficient Correction of ABCA4 Variants by CRISPR-Cas9 in HiPSCs Derived from Stargardt Disease Patients. Mol. Ther. Nucleic Acids 2023, 32, 64–79. [Google Scholar] [CrossRef]
  57. Maeder, M.L.; Stefanidakis, M.; Wilson, C.J.; Baral, R.; Barrera, L.A.; Bounoutas, G.S.; Bumcrot, D.; Chao, H.; Ciulla, D.M.; DaSilva, J.A.; et al. Development of a Gene-Editing Approach to Restore Vision Loss in Leber Congenital Amaurosis Type 10. Nat. Med. 2019, 25, 229–233. [Google Scholar] [CrossRef]
  58. Pierce, E.A.; Aleman, T.S.; Jayasundera, K.T.; Ashimatey, B.S.; Kim, K.; Rashid, A.; Jaskolka, M.C.; Myers, R.L.; Lam, B.L.; Bailey, S.T.; et al. Gene Editing for CEP290-Associated Retinal Degeneration. N. Engl. J. Med. 2024, 390, 1972–1984. [Google Scholar] [CrossRef]
  59. Nguyen, D.D.; Luo, L.J.; Yang, C.J.; Lai, J.Y. Highly Retina-Permeating and Long-Acting Resveratrol/Metformin Nanotherapeutics for Enhanced Treatment of Macular Degeneration. ACS Nano 2023, 17, 168–183. [Google Scholar] [CrossRef]
Figure 1. Distribution of genes with confirmatory variants among 50 distinct solved patients (total number of patients with confirmatory variants in the denoted gene in brackets).
Figure 1. Distribution of genes with confirmatory variants among 50 distinct solved patients (total number of patients with confirmatory variants in the denoted gene in brackets).
Genes 15 01011 g001
Figure 2. Distribution of genes with confirmatory variants among distinct solved patients in the two groups: (a) unspecific IRD group; (b) RP group. Total number of patients with confirmatory variants in the denoted gene in brackets.
Figure 2. Distribution of genes with confirmatory variants among distinct solved patients in the two groups: (a) unspecific IRD group; (b) RP group. Total number of patients with confirmatory variants in the denoted gene in brackets.
Genes 15 01011 g002
Figure 3. Ultrawidefield eye fundus images (Optos) - VCAN-related Wagner syndrome family: patient S48, father, 52 years old: (a) (right eye, RE), (b) (left eye, LE); patient S46, son, 20 years old (c) (LE), (d) (RE); (e) Sanger confirmation of the NM_004385.5:c.4004-2A>G variant in VCAN for patient S46 (top), patient S48 (middle), and wild-type control sample (bottom). Due to changes involving the anterior segment of the eye in the father, with extensive phimosis of the lens capsule after cataract surgery, fundus imaging is difficult. In this case, widefield images cover a limited area of the retina; in the right eye, only the nasal part of the retina is visible (a). Pigmentary changes located mainly on the periphery of the retina resemble retinitis pigmentosa, which was the first, initial diagnosis in both father and son. Visual acuity was low at 0.01 (RE), 0.05 (LE), but stable for the last 10 years. In his son, pigmentary and atrophic changes were less severe. The boundary line of the optically empty part of the vitreous body was clearly marked, from which vitreoretinal tractions extend towards the periphery. Best corrected visual acuity is 1.0 (RE) and 0.9 (LE), but with a narrowed field of vision. Snellen charts were used for visual acuity examination.
Figure 3. Ultrawidefield eye fundus images (Optos) - VCAN-related Wagner syndrome family: patient S48, father, 52 years old: (a) (right eye, RE), (b) (left eye, LE); patient S46, son, 20 years old (c) (LE), (d) (RE); (e) Sanger confirmation of the NM_004385.5:c.4004-2A>G variant in VCAN for patient S46 (top), patient S48 (middle), and wild-type control sample (bottom). Due to changes involving the anterior segment of the eye in the father, with extensive phimosis of the lens capsule after cataract surgery, fundus imaging is difficult. In this case, widefield images cover a limited area of the retina; in the right eye, only the nasal part of the retina is visible (a). Pigmentary changes located mainly on the periphery of the retina resemble retinitis pigmentosa, which was the first, initial diagnosis in both father and son. Visual acuity was low at 0.01 (RE), 0.05 (LE), but stable for the last 10 years. In his son, pigmentary and atrophic changes were less severe. The boundary line of the optically empty part of the vitreous body was clearly marked, from which vitreoretinal tractions extend towards the periphery. Best corrected visual acuity is 1.0 (RE) and 0.9 (LE), but with a narrowed field of vision. Snellen charts were used for visual acuity examination.
Genes 15 01011 g003
Table 1. IRD group characteristics.
Table 1. IRD group characteristics.
Initial IndicationTotal CasesFemalesMalesMean Age (SD)Solved and Likely Solved CasesNew Clinical Diagnosis
RP4719 (40.4%)28 (59.6%)38.26 (SD = 14.55)30 (63.8%)3
Unspecific IRD3015 (50.0%)15 (50.0%)46.04 (SD = 19.56)20 (66.7%)5
Total77344340.93 (SD = 16.71)50 (64.9%)8
Table 2. List of causative and possibly causative variants found in the analysed cohort of 77 IRD patients (LP—likely pathogenic).
Table 2. List of causative and possibly causative variants found in the analysed cohort of 77 IRD patients (LP—likely pathogenic).
GeneVariant ClassHGVScHGVSpNo PatientsNo AllelesPOLGENOM FrequencygnomAD_2.1.1.AF TotaldbSNPClinvar IDConsequenceNovelP/LP/VUS
ABCA4SNVNM_000350.3:c.214G>ANP_000341.2:p.Gly72Arg11 0.00001rs6175141299120missense Pathogenic
ABCA4SNVNM_000350.3:c.1622T>CNP_000341.2: p.Leu541Pro45-0.00016rs6175139299067missense Pathogenic
ABCA4SNVNM_000350.3:c.3113C>TNP_000341.2:p.Ala1038Val45-0.00175rs617513747894missense Pathogenic
ABCA4InDelNM_000350.3:c.2548dupNP_000341.2:p.Tyr850LeufsTer911--- frameshiftyesPathogenic
ABCA4SNVNM_000350.2:c.2588G>CNP_000341.2:p.Gly863Ala22-0.00430rs761576387879missense Pathogenic
ABCA4SNVNM_000350.3:c.4234C>TNP_000341.2:p.Gln1412Ter11-0.00001rs6175013799263stop_gained Pathogenic
ABCA4InDelNM_000350.3:c.4537dupNP_000341.2:p.Gln1513ProfsTer4211-0.00002rs28186537799292frameshift Pathogenic
ABCA4SNVNM_000350.3:c.5189G>ANP_000341.2:p.Trp1730Ter11-0.00000rs886044747236125stop_gained Pathogenic
ABCA4SNVNM_000350.3:c.5318C>TNP_000341.2:p.Ala1773Val11-0.00001rs760549861236129missense Pathogenic
ABCA4SNVNM_000350.3:c.5603A>TNP_000341.2:p.Asn1868Ile220.06350.04219rs180146699390missense Hypomorphic
ABCA4SNVNM_000350.3:c.5882G>ANP_000341.2:p.Gly1961Glu550.01190.00456rs18005537888missense Pathogenic
AHI1SNVNM_001134831.2:c.1828C>TNP_001128303.1:p.Arg610Ter11-0.00001rs751734985984718stop_gained Pathogenic
AHI1SNVNM_001134831.2:c.1829G>CNP_001128303.1:p.Arg610Pro11--rs374009466802272missense VUS
CEP290SNVNM_025114.4:c.4882C>TNP_079390.3:p.Gln1628Ter11-0.00007rs376493409217635stop_gained Pathogenic
CEP290SNVNM_025114.4:c.5254C>TNP_079390.3:p.Arg1752Trp11-0.00002rs748471942866339missense LP
CFAP410InDelNM_004928.3:NP_004919.1:p.Ala12SerfsTer6011-0.00008rs748531024438159frameshift Pathogenic
c.33_34insAGCTGCACAGCGTGCA
CFAP410SNVNM_004928.2:c.218G>CNP_004919.1:p.Arg73Pro11-0.00033rs140451304428573missense LP
CHMSNVNM_000390.4:c.49+2T>A 11--- splice_donoryesPathogenic
CNGB3SNVNM_019098.5:c.1578+1G>A 12-0.00002rs372006750189031splice_donor Pathogenic
CRB1SNVNM_201253.3:c.2042G>ANP_957705.1:p.Cys681Tyr11-0.00000rs6263626699874missense Pathogenic
CRB1SNVNM_201253.3:c.2171A>GNP_957705.1:p.Tyr724Cys22-0.00001rs765676754872033missense Pathogenic
CRB1SNVNM_201253.3:c.2843G>ANP_957705.1:p.Cys948Tyr11-0.00013rs6264574839614missense Pathogenic
CRXSNVNM_000554.6:c.766C>TNP_000545.1:p.Gln256Ter11--rs1968173024861103missense LP
EYSSNVNM_001142800.2:c.2055T>ANP_001136272.1:p.Cys685Ter11--rs372354156194194stop_gained Pathogenic
EYSSNVNM_001142800.2:c.2259+1G>A 11-0.00004rs752736741558511splice_donor Pathogenic
EYSInDelNM_001142800.2:c.4350_4356delNP_001136272.1:pIle1451ProfsTer312-0.00008rs761238771195936frameshift Pathogenic
EYSInDelNM_001142800.2:c.4462_4469dupNP_001136272.1:p.Met1491AlafsTer1211--- frameshiftyesPathogenic
EYSSNVNM_001142800.2:c.4955C>GNP_001136272.1:p.Ser1652Ter11--rs909730457867110stop_gained Pathogenic
EYSSNVNM_001142800.2:c.8816G>ANP_001136272.1:p.Cys2939Tyr11--rs15821399651039939missense VUS
EYSInDelNM_001142800.2:c.9079_9082delNP_001136272.1:p.Arg3027SerfsTer511--rs1427770112556940frameshift LP
GUCA1ASNVNM_001384910.1:c.320T>CNP_000400.2:p.Ile107Thr22--rs869320710225233missense LP
GUCY2DInDelNM_000180.4:c.2944+1del 11-0.00036rs6175018598582splice_donor Pathogenic
IMPDH1SNVNM_000883.4:c.730A>CNP_000874.2:p.Thr244Pro11-0.00000rs1253235754845744missense VUS
IMPDH1SNVNM_000883.4:c.931G>ANP_000874.2:p.Asp311Asn11--rs12191255014834missense Pathogenic
MERTKSNVNM_006343.2:c.79G>TNP_006334.2:p.Glu27Ter12--rs1223798126 stop_gained Pathogenic
NR2E3InDelNM_014249.4:c.481delNP_055064.1:p.Thr161HisfsTer1812-0.00007rs759339012636047frameshift Pathogenic
PRPF3SNVNM_004698.2:c.1481C>TNP_004689.1:p.Thr494Met11--rs1214342413352missense Pathogenic
PRPF31InDelNM_015629.4: c.428_430delins22AC-NP_056444.3: p.Gly143AspfsTer1711--- frameshiftyesLP
AAGTGCAAGGCTGTTCTTGC
PRPF31InDelNM_015629.4:c.808dupNP_056444.3:p.His270ProfsTer911--rs2073875313865752frameshift LP
PRPF31SNVNM_015629.4:c.831T>ANP_056444.3:p.Ser277Arg11--- missenseyesVUS
REEP6SNVNM_138393.4:c.349-1G>T 11--rs2085005383958464splice_acceptor LP
REEP6SNVNM_001329556.3:c.367T>CNP_001316485.1:p.Cys123Arg11-0.00001rs1341620671 missense VUS
RHOSNVNM_000539.3:c.560G>ANP_000530.1:p.Cys187Tyr11--rs1578280588812397missense Pathogenic
RHOSNVNM_000539.3:c.1040C>TNP_000530.1:p.Pro347Leu11-0.00000rs2900156613014missense Pathogenic
RP2InDelNM_006915.3:c.14_16delNP_008846.2:p.Phe5del11--rs1556313414437942inframe deletion Pathogenic
RPE65SNVNM_000329.3:c.131G>ANP_000320.1:p.Arg44Gln11-0.00004rs6175128298840missense Pathogenic
RPE65SNVNM_000329.3:c.709C>TNP_000320.1:p.Pro237Ser11---3028862missense LP
RS1SNVNM_000330.4:c.214G>ANP_000321.1:p.Glu72Lys11-0.00001rs1048949289888missense Pathogenic
RS1SNVNM_000330.4:c.440G>ANP_000321.1:p.Trp147Ter11--- missenseyesLP
RS1InDelNM_000330.4:c.522+1G>T 11--rs28186534898980splice_donor Pathogenic
RS1SNVNM_000330.4:c.625C>TNP_000321.1:p.Arg209Cys11--rs28186536199006missense Pathogenic
USH2AInDelNM_206933.4:c.775_776delNP_996816.3:p.Ser259PhefsTer6311--rs2038566220866168frameshift Pathogenic
USH2AInDelNM_206933.4:c.1836_1839dupNP_996816.3:p.Gly614TyrfsTer611--rs21026106071075936frameshift Pathogenic
USH2ASNVNM_206933.4:c.2276G>TNP_996816.3:p.Cys759Phe11-0.00097rs803389022356missense Pathogenic
USH2ASNVNM_206933.4:c.4210G>TNP_996816.3:p.Glu1404Ter11--rs2034849647813119stop_gained Pathogenic
USH2ASNVNM_206933.4:c.5530C>TNP_996816.3:p.Gln1844Ter11--rs7610753031065688stop_gained Pathogenic
USH2AInDelNM_206933.4:c.5865_5866delNP_996816.3:p.Ser1956CysfsTer1611--rs2818653613028715frameshift LP
USH2ASNVNM_206933.4:c.9424G>TNP_996816.3:p.Gly3142Ter11-0.000024rs39751804848626stop_gained Pathogenic
USH2ASNVNM_206933.4:c.10732A>CNP_996816.3:p.Ser3578Arg11--- missenseyesLP
USH2ASNVNM_206933.4:c.11864G>ANP_996816.3:p.Trp3955Ter23-0.00012rs1110333642357stop_gained Pathogenic
USH2ASNVNM_206933.4:c.12284G>ANP_996816.2:p.Gly4095Asp11--rs759898765378847missense Pathogenic
USH2AInDelNM_206933.4:NP_996816.3:22--rs1553252388438013inframe_indel Pathogenic
c.13335_13347delinsCTTGp.Glu4445_Ser4449delinsAspLeu
VCANSNVNM_004385.5:c.4004-2A>G 22--rs8035655517494splice_acceptor Pathogenic
Variants in inconclusive results
AIPL1SNVNM_014336.5: c.737A>CNP_055151.3: p.Tyr246Ser22-0.00021rs138585919324614missense VUS
CAPN5SNVNM_004055.5:c.64C>TNP_004046.2:p.Arg22Cys11 0.00007rs7818376401064369missense VUS
HMCN1SNVNM_031935.3: c.16034A>GNP_114141.2: pGln5345Arg11 0.00071rs1214343822205missense VUS/Risk factor
PRPH2SNVNM_000322.5:c.94A>GNP_000313.2: p.Ile32Val11-0.00081rs6175576798722missense VUS
RP1SNVNM_006269.2:c.6338C>ANP_006260.1: p.The2113Asn22-0.00022rs137887415847529missense VUS
SEMA4ASNVNM_022367.4:c.241C>GNP_071762.2: p.Arg81Gly11--rs745715951 missense VUS
Disclaimer/Publisher’s Note: 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

MDPI and ACS Style

Matczyńska, E.; Szymańczak, R.; Stradomska, K.; Łyszkiewicz, P.; Jędrzejowska, M.; Kamińska, K.; Beć-Gajowniczek, M.; Suchecka, E.; Zagulski, M.; Wiącek, M.; et al. Whole-Exome Analysis for Polish Caucasian Patients with Retinal Dystrophies and the Creation of a Reference Genomic Database for the Polish Population. Genes 2024, 15, 1011. https://doi.org/10.3390/genes15081011

AMA Style

Matczyńska E, Szymańczak R, Stradomska K, Łyszkiewicz P, Jędrzejowska M, Kamińska K, Beć-Gajowniczek M, Suchecka E, Zagulski M, Wiącek M, et al. Whole-Exome Analysis for Polish Caucasian Patients with Retinal Dystrophies and the Creation of a Reference Genomic Database for the Polish Population. Genes. 2024; 15(8):1011. https://doi.org/10.3390/genes15081011

Chicago/Turabian Style

Matczyńska, Ewa, Robert Szymańczak, Katarzyna Stradomska, Przemysław Łyszkiewicz, Maria Jędrzejowska, Karolina Kamińska, Marta Beć-Gajowniczek, Ewa Suchecka, Marek Zagulski, Marta Wiącek, and et al. 2024. "Whole-Exome Analysis for Polish Caucasian Patients with Retinal Dystrophies and the Creation of a Reference Genomic Database for the Polish Population" Genes 15, no. 8: 1011. https://doi.org/10.3390/genes15081011

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop