[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (74)

Search Parameters:
Keywords = Stargardt disease

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 1830 KiB  
Review
Emerging Therapeutic Approaches and Genetic Insights in Stargardt Disease: A Comprehensive Review
by Laura Andreea Ghenciu, Ovidiu Alin Hațegan, Emil Robert Stoicescu, Roxana Iacob and Alina Maria Șișu
Int. J. Mol. Sci. 2024, 25(16), 8859; https://doi.org/10.3390/ijms25168859 - 14 Aug 2024
Cited by 4 | Viewed by 1279
Abstract
Stargardt disease, one of the most common forms of inherited retinal diseases, affects individuals worldwide. The primary cause is mutations in the ABCA4 gene, leading to the accumulation of toxic byproducts in the retinal pigment epithelium (RPE) and subsequent photoreceptor cell degeneration. Over [...] Read more.
Stargardt disease, one of the most common forms of inherited retinal diseases, affects individuals worldwide. The primary cause is mutations in the ABCA4 gene, leading to the accumulation of toxic byproducts in the retinal pigment epithelium (RPE) and subsequent photoreceptor cell degeneration. Over the past few years, research on Stargardt disease has advanced significantly, focusing on clinical and molecular genetics. Recent studies have explored various innovative therapeutic approaches, including gene therapy, stem cell therapy, and pharmacological interventions. Gene therapy has shown promise, particularly with adeno-associated viral (AAV) vectors capable of delivering the ABCA4 gene to retinal cells. However, challenges remain due to the gene’s large size. Stem cell therapy aims to replace degenerated RPE and photoreceptor cells, with several clinical trials demonstrating safety and preliminary efficacy. Pharmacological approaches focus on reducing toxic byproduct accumulation and modulating the visual cycle. Precision medicine, targeting specific genetic mutations and pathways, is becoming increasingly important. Novel techniques such as clustered regularly interspaced palindromic repeats (CRISPR)/Cas9 offer potential for directly correcting genetic defects. This review aims to synthesize recent advancements in understanding and treating Stargardt disease. By highlighting breakthroughs in genetic therapies, stem cell treatments, and novel pharmacological strategies, it provides a comprehensive overview of emerging therapeutic options. Full article
Show Figures

Figure 1

Figure 1
<p>Left eye color fundus image of a 52-year-old patient with Stargardt disease. Personal casuistry.</p>
Full article ">Figure 2
<p>Left eye fundus image of the same patient. (<b>a</b>) Infrared image; (<b>b</b>) Red-free image.</p>
Full article ">Figure 3
<p>Presentation of the retinoid cycle. When rhodopsin (11-cis retinal) becomes stimulated by a photon, it undergoes a molecular conversion into all-trans-retinal, a process that occurs in the disk membrane of photoreceptor outer segments. The recently produced all-trans retinal must next be transferred over the disk membrane and into the cytoplasm. Some of the all-trans retinal diffuses straight to that subcellular region, but a large portion diffuses into the lumen of the disk membrane, where it interacts with phosphatidylethanolamine (PE) to generate N-retinylidene-PE (retPE), which remains trapped due to its protonated condition. <span class="html-italic">ABCA4</span> actively moves ret-PE across the disk membrane. On the cytoplasmic side, ret-PE separates into all-trans retinal and PE. All-trans retinal is then converted into all-trans retinol by all-trans retinol dehydrogenase (RDH). Afterward, all-trans retinol is delivered to the retinal pigmented epithelium (RPE), where the cycle continues, as shown. Created with Biorender.</p>
Full article ">Figure 4
<p>Gene therapy for <span class="html-italic">ABCA4</span> mutations: gene supplementation using viral vectors to deliver functional <span class="html-italic">ABCA4</span> genes, and gene correction through CRISPR and AONs to precisely edit or correct the existing mutations in retinal cells. Created with Biorender.</p>
Full article ">Figure 5
<p>Therapeutic strategies targeting retinal pigment epithelium (RPE) and photoreceptors in Stargardt disease. This figure illustrates key therapies: gene and stem cell therapy for <span class="html-italic">ABCA4</span> mutation correction and RPE regeneration, deuterated vitamin A and RBP4 antagonists to reduce toxic retinoid byproducts, emixustat to inhibit harmful visual cycle enzymes, metformin and avacincaptad pegol to decrease lipofuscin and inhibit the complement cascade, and DHA/EPA to reduce oxidative stress and inflammation. Created with Draw.io (<a href="https://app.diagrams.net/" target="_blank">https://app.diagrams.net/</a> (accessed on 15 June 2024).</p>
Full article ">
17 pages, 13894 KiB  
Article
The Surviving, Not Thriving, Photoreceptors in Patients with ABCA4 Stargardt Disease
by Hanna De Bruyn, Megan Johnson, Madelyn Moretti, Saleh Ahmed, Mircea Mujat, James D. Akula, Tomislav Glavan, Ivana Mihalek, Sigrid Aslaksen, Laurie L. Molday, Robert S. Molday, Bruce A. Berkowitz and Anne B. Fulton
Diagnostics 2024, 14(14), 1545; https://doi.org/10.3390/diagnostics14141545 - 17 Jul 2024
Viewed by 878
Abstract
Stargardt disease (STGD1), associated with biallelic variants in the ABCA4 gene, is the most common heritable macular dystrophy and is currently untreatable. To identify potential treatment targets, we characterized surviving STGD1 photoreceptors. We used clinical data to identify macular regions with surviving STGD1 [...] Read more.
Stargardt disease (STGD1), associated with biallelic variants in the ABCA4 gene, is the most common heritable macular dystrophy and is currently untreatable. To identify potential treatment targets, we characterized surviving STGD1 photoreceptors. We used clinical data to identify macular regions with surviving STGD1 photoreceptors. We compared the hyperreflective bands in the optical coherence tomographic (OCT) images that correspond to structures in the STGD1 photoreceptor inner segments to those in controls. We used adaptive optics scanning light ophthalmoscopy (AO-SLO) to study the distribution of cones and AO-OCT to evaluate the interface of photoreceptors and retinal pigment epithelium (RPE). We found that the profile of the hyperreflective bands differed dramatically between patients with STGD1 and controls. AO-SLOs showed patches in which cone densities were similar to those in healthy retinas and others in which the cone population was sparse. In regions replete with cones, there was no debris at the photoreceptor-RPE interface. In regions with sparse cones, there was abundant debris. Our results raise the possibility that pharmaceutical means may protect surviving photoreceptors and so mitigate vision loss in patients with STGD1. Full article
(This article belongs to the Special Issue High-Resolution Retinal Imaging: Hot Topics and Recent Developments)
Show Figures

Figure 1

Figure 1
<p>Photoreceptor (PR), retinal pigment epithelium (RPE), and <span class="html-italic">ABCA4</span> protein. (<b>A</b>) The PR and RPE have close structural and functional relationships. <span class="html-italic">ABCA4</span> protein is located in the rim of the discs of the photoreceptor outer segment. In the photoreceptor’s inner segment is the ellipsoid zone (EZ), which has abundant mitochondria (Mito), which are needed to support the high energy demands of the photoreceptor. Also, at the level of the PR inner segment, adherens junctions form the external limiting membrane (ELM), one of the OCT hyperreflective bands. Images were adapted from Scortecci et al. [<a href="#B14-diagnostics-14-01545" class="html-bibr">14</a>] and Steinberg et al. [<a href="#B15-diagnostics-14-01545" class="html-bibr">15</a>]. (<b>B</b>) This diagram of <span class="html-italic">ABCA4</span> protein highlights its functional domains and indicates the site of variants found in our patients. The extracellular domains 1 and 2 (ECD1, ECD2) reside in the lumen, while the transmembrane domains 1 and 2 (TMD1, TMD2) are embedded in the lipid bilayer of the disc. The nucleotide-binding domains 1 and 2 (NBD1, NBD2) are in the cytosol [<a href="#B16-diagnostics-14-01545" class="html-bibr">16</a>]. In Patient 1, R2038W and Q2190R are on the same allele; on the other allele, there is a deep intronic variant resulting in complete protein loss. Patient 2 and Patient 3 have C1490Y on one allele; their other allele contains a deep intronic variant. In Patient 4, the two variants are V989A and E2096K, and in Patient 5, the two variants are G863A and G1961E, which are located as indicated. See <a href="#diagnostics-14-01545-t001" class="html-table">Table 1</a>, as well. For details about variants and genotypes, please see <a href="#app1-diagnostics-14-01545" class="html-app">Appendix A</a>.</p>
Full article ">Figure 2
<p>Retinal images of right eye. Left column: healthy Control 1. Right column: Patient 1. (<b>A</b>) Fundus autofluorescence (200°; California; Optos, Dunfermline, Scotland, UK); (<b>B</b>) color fundus photograph (45°; TRC-NW8F; Topcon Corporation, Tokyo, Japan); (<b>C</b>) blue autofluorescent image (30°; HRA + OCT Spectralis; Heidelberg Engineering, Heidelberg, Germany); (<b>D</b>) horizontal OCT showing the 30/61 b-scan slice (30°; HRA + OCT Spectralis; Heidelberg Engineering, Heidelberg, Germany). The dotted box in A indicates the region shown in (<b>B</b>). The dotted box in (<b>B</b>) indicates the region imaged in (<b>C</b>). The box in (<b>C</b>) indicates the region shown in (<b>D</b>). In (<b>C</b>), the dashed red line indicates the site of the OCT slice, as shown in (<b>D</b>). In the control, the OCT slice, bound by the red rectangle, has a dark band, the outer nuclear layer (ONL), representing photoreceptor nuclei; the ONL normally widens in subfoveal retina (white arrow). In the patient with STGD1, the ONL is absent in subfoveal retina, and there is debris at the retina–pigment epithelium (RPE) interface.</p>
Full article ">Figure 3
<p>Fundus photograph and horizontal transfoveal OCT image of asymptomatic patient (Patient 2, <a href="#diagnostics-14-01545-t001" class="html-table">Table 1</a>) with biallelic pathogenic changes in <span class="html-italic">ABCA4</span>. The photograph (<b>left panel</b>), as well as ophthalmoscopy, showed no signs of maculopathy. The OCT (<b>right panel</b>) shows thickened and hyperreflective ELM and indistinct EZ similar to that reported by others [<a href="#B19-diagnostics-14-01545" class="html-bibr">19</a>,<a href="#B20-diagnostics-14-01545" class="html-bibr">20</a>].</p>
Full article ">Figure 4
<p>For each of the 3 patients (Patient 3, Patient 4, &amp; Patient 5), from left to right we present, Fundus photograph, Flattened OCT images of retina nasal to the fovea for STGD1 and age- and sex- matched controls (See <a href="#diagnostics-14-01545-t001" class="html-table">Table 1</a>); the yellow boxes indicate the region of interest (ROI). The yellow arrow indicates the direction of the A-scan. In the right most panel the mitochondrial configuration within photoreceptors/aspect ratio (MCP/AR). Patient characteristics are shown in <a href="#diagnostics-14-01545-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 5
<p>Healthy control 1. Far-left fundus photograph of a healthy control (right eye); location and size of AO Scan 1 and Scan 2 are as indicated. Left column—cone density heat maps. Center column—AO-SLO cone images with identified cones (green dots). Right column—OCT B-scan at the location indicated by the green line on the AO-SLO. AO imaging details are shown in <a href="#diagnostics-14-01545-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 6
<p>Patient 4, right eye. Format is similar to that of <a href="#diagnostics-14-01545-f005" class="html-fig">Figure 5</a>, with the zoom-in (yellow square on the AO-SLO) column added. * in the superior retina, at ~10° is the patient’s preferred retinal locus for fixation (PRL). AO imaging details are shown in <a href="#diagnostics-14-01545-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 7
<p>Patient 5, both eyes. Top-right eye scan (<b>D</b>–<b>F</b>). Bottom-left eye scan (<b>G</b>–<b>J</b>). Format is similar to that of <a href="#diagnostics-14-01545-f006" class="html-fig">Figure 6</a>. AO imaging details are shown in <a href="#diagnostics-14-01545-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure A1
<p>Biochemical characterization of the Val989Ala variant associated with Stargardt disease. WT <span class="html-italic">ABCA4</span> and the Val989Ala variant were expressed in HEK293T cells. (<b>A</b>) Protein expression levels were determined on Western blots labeled for <span class="html-italic">ABCA4</span>. The Val989Ala expression was 70% that of WT <span class="html-italic">ABCA4</span>. (<b>B</b>) The ATPase activity of purified WT and the Val989Ala variant was determined at the same protein concentration in the absence (−) and presence (+) of all-trans-retinal (ATR). Data are expressed relative to WT <span class="html-italic">ABCA4</span> in the absence of ATR. A small but significant (<span class="html-italic">p</span> &lt; 0.05) activation by ATR was observed for the Val989Ala variant. Data show SD for <span class="html-italic">n</span> = 3 independent experiments.</p>
Full article ">Figure A2
<p>Fovea OCT image from the right eye from a healthy normal control overlaid on their fundus photo. Note the intact layers of the retina and how their fixation in the fixation plot is centered in their foveola, and the Bivariate Contour Ellipse (BCEA) is small (63% BCEA: 0.4° × 0.3°, area = 0.1°<sup>2</sup>, angle = −6.5°).</p>
Full article ">Figure A3
<p>Fovea OCT image from the right eye from Patient 4 overlaid on fundus photo at age 16.8 years old. * marks the patient’s fixation during adaptive optics testing session. The fixation plot taken from their MAIA exam at 25 years confirms this fixation location (63% BCEA: 0.5° × 3.9°, area = 1.7°<sup>2</sup>, angle = 87.3°). + on the fundus marks the fovea as shown in OCT slice.</p>
Full article ">Figure A4
<p>Fovea OCT image from the left eye from Patient 5 overlaid on fundus photo at age 58 years old. The fixation plot taken from their MAIA exam at 60 years confirms this fixation location (63% BCEA: 1.2° × 4.1°, area = 3.8°<sup>2</sup>, angle = 77.7°). Note how the patient is able to fixate centrally due to foveal sparing.</p>
Full article ">
10 pages, 792 KiB  
Article
Genetic Characterization of 191 Probands with Inherited Retinal Dystrophy by Targeted NGS Analysis
by Alessandra Mihalich, Gabriella Cammarata, Gemma Tremolada, Emanuela Manfredini, Stefania Bianchi Marzoli and Anna Maria Di Blasio
Genes 2024, 15(6), 766; https://doi.org/10.3390/genes15060766 - 12 Jun 2024
Viewed by 798
Abstract
Inherited retinal diseases (IRDs) represent a frequent cause of blindness in children and adults. As a consequence of the phenotype and genotype heterogeneity of the disease, it is difficult to have a specific diagnosis without molecular testing. To date, over 340 genes and [...] Read more.
Inherited retinal diseases (IRDs) represent a frequent cause of blindness in children and adults. As a consequence of the phenotype and genotype heterogeneity of the disease, it is difficult to have a specific diagnosis without molecular testing. To date, over 340 genes and loci have been associated with IRDs. We present the molecular finding of 191 individuals with IRD, analyzed by targeted next-generation sequencing (NGS). For 67 of them, we performed a family segregation study, considering a total of 126 relatives. A total of 359 variants were identified, 44 of which were novel. Genetic diagnostic yield was 41%. However, after stratifying the patients according to their clinical suspicion, diagnostic yield was higher for well-characterized diseases such as Stargardt disease (STGD), at 65%, and for congenital stationary night blindness 2 (CSNB2), at 64%. Diagnostic yield was higher in the patient group where family segregation analysis was possible (68%) and it was higher in younger (55%) than in older patients (33%). The results of this analysis demonstrated that targeted NGS is an effective method for establishing a molecular genetic diagnosis of IRDs. Furthermore, this study underlines the importance of segregation studies to understand the role of genetic variants with unknow pathogenic role. Full article
(This article belongs to the Section Genetic Diagnosis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Distribution of patients according to their clinical diagnosis. “rod–cone dystrophy” (RCD), “cone and cone–rod dystrophy” (CRD), “retinitis pigmentosa” (RT), “Stargardt disease” (STGD), “best vitelliform macular dystrophy” (BVMD), “Usher syndrome” (USH), “generalized retinal dystrophy” (GRD), “congenital stationary night blindness 2” (CSNB2), “Leber congenital amaurosis” (LCA), “achromatopsia” (ACHM) and “retinoschisis” (XLRS).</p>
Full article ">Figure 2
<p>Genetic diagnosis results: (<b>A</b>) diagnostic yield in total cohort; (<b>B</b>) diagnostic yield according to the disease type “rod–cone dystrophy” (RCD), “cone and cone–rod dystrophy” (CRD), “retinitis pigmentosa” (RT), “Stargardt disease” (STGD), “best vitelliform macular dystrophy” (BVMD), “Usher syndrome” (USH) (8.9%), “generalized retinal dystrophy” (GRD), “congenital stationary night blindness 2” (CSNB2), “Leber congenital amaurosis” (LCA), “achromatopsia” (ACHM) and “retinoschisis” (XLRS) and (<b>C</b>) diagnostic yield according to the patients’ ages.</p>
Full article ">Figure 3
<p>Distribution of patients with genetic diagnosis according to the gene where variants were localized.</p>
Full article ">
14 pages, 1256 KiB  
Article
Novel and Recurrent Copy Number Variants in ABCA4-Associated Retinopathy
by Zelia Corradi, Claire-Marie Dhaenens, Olivier Grunewald, Ipek Selen Kocabaş, Isabelle Meunier, Sandro Banfi, Marianthi Karali, Frans P. M. Cremers and Rebekkah J. Hitti-Malin
Int. J. Mol. Sci. 2024, 25(11), 5940; https://doi.org/10.3390/ijms25115940 - 29 May 2024
Viewed by 865
Abstract
ABCA4 is the most frequently mutated gene leading to inherited retinal disease (IRD) with over 2200 pathogenic variants reported to date. Of these, ~1% are copy number variants (CNVs) involving the deletion or duplication of genomic regions, typically >50 nucleotides in length. An [...] Read more.
ABCA4 is the most frequently mutated gene leading to inherited retinal disease (IRD) with over 2200 pathogenic variants reported to date. Of these, ~1% are copy number variants (CNVs) involving the deletion or duplication of genomic regions, typically >50 nucleotides in length. An in-depth assessment of the current literature based on the public database LOVD, regarding the presence of known CNVs and structural variants in ABCA4, and additional sequencing analysis of ABCA4 using single-molecule Molecular Inversion Probes (smMIPs) for 148 probands highlighted recurrent and novel CNVs associated with ABCA4-associated retinopathies. An analysis of the coverage depth in the sequencing data led to the identification of eleven deletions (six novel and five recurrent), three duplications (one novel and two recurrent) and one complex CNV. Of particular interest was the identification of a complex defect, i.e., a 15.3 kb duplicated segment encompassing exon 31 through intron 41 that was inserted at the junction of a downstream 2.7 kb deletion encompassing intron 44 through intron 47. In addition, we identified a 7.0 kb tandem duplication of intron 1 in three cases. The identification of CNVs in ABCA4 can provide patients and their families with a genetic diagnosis whilst expanding our understanding of the complexity of diseases caused by ABCA4 variants. Full article
(This article belongs to the Special Issue Genetics and Epigenetics of Eye Diseases: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Deletions identified in probands sequenced by smMIPs. In each schematic, intronic regions are represented as horizontal lines and exons are as black boxes. In light grey are represented deleted regions, including partial exon deletions. Black box frames highlight novel deletions. Question marks indicate uncharacterized breakpoints. CNV 1: c.699_768+341del, CNV 2: c.1239+303_1555-5571del, CNV 3: c.2160+531_2569del, CNV 4a: c.(2918+757_2918+981)_(3329-420_3329-644)del, CNV 4b: c.2918+533_3329-622del, CNV 5: g.(?_94458389)_(94505684_94506764)del, CNV 6: c.3863-553_4539+578, CNV 7: c.4254-197_4672delinsGCTTTTT, CNV 8: c.4352+123_4540-585del, CNV 9: c.4540-1000_4635-389delinsTGCCCG, CNV 10: c.4539+872_4635-565delins28 and CNV 11: c.5864_6085del.</p>
Full article ">Figure 2
<p>Duplications and complex rearrangement identified in probands sequenced by smMIPs. In each schematic, intronic regions are represented as horizontal lines and exons are as black boxes. Duplicated regions are depicted as bold, white-filled sections. Black box frames highlight the novel duplication. The stapled line arrow in CNV 15 highlights the downstream insertion of the duplicated region. CNV 12: c.66+520_67-389dup, CNV 13: c.768+6839_858+66dup, CNV 14: c.6006-29_6370dup and CNV 15: c.6147+411_c.6479+293delins4583_5715-778.</p>
Full article ">Figure 3
<p>The landscape of the reported structural variants in <span class="html-italic">ABCA4.</span> All the CNVs submitted to LOVD and identified in this study are depicted as triangles encompassing the affected region of <span class="html-italic">ABCA4</span>. The <span class="html-italic">ABCA4</span> gene sequence has been divided into three sections (top: exon 1 to exon 9, middle: exon 10 to exon 26 and bottom: exon 27 to exon 50), and the numbered vertical lines represent the exons. For each section, the peaks above and below represent the deletions and duplications, respectively. The height of the peaks corresponds to the number of alleles reported in LOVD. Below each section, the color legends for all the CNVs show the details of the intronic or exonic region of the breakpoints. Complex defects, such as deletion-inversions and deletion-insertions, are annotated as del/inv and delins, respectively, in the legend. Defects with one or both breakpoints in the intergenic regions are not included.</p>
Full article ">
19 pages, 326 KiB  
Review
Exploring Stem-Cell-Based Therapies for Retinal Regeneration
by Madalina Radu, Daniel Constantin Brănișteanu, Ruxandra Angela Pirvulescu, Otilia Maria Dumitrescu, Mihai Alexandru Ionescu and Mihail Zemba
Life 2024, 14(6), 668; https://doi.org/10.3390/life14060668 - 23 May 2024
Cited by 2 | Viewed by 2186
Abstract
The escalating prevalence of retinal diseases—notably, age-related macular degeneration and hereditary retinal disorders—poses an intimidating challenge to ophthalmic medicine, often culminating in irreversible vision loss. Current treatments are limited and often fail to address the underlying loss of retinal cells. This paper explores [...] Read more.
The escalating prevalence of retinal diseases—notably, age-related macular degeneration and hereditary retinal disorders—poses an intimidating challenge to ophthalmic medicine, often culminating in irreversible vision loss. Current treatments are limited and often fail to address the underlying loss of retinal cells. This paper explores the potential of stem-cell-based therapies as a promising avenue for retinal regeneration. We review the latest advancements in stem cell technology, focusing on embryonic stem cells (ESCs), pluripotent stem cells (PSCs), and mesenchymal stem cells (MSCs), and their ability to differentiate into retinal cell types. We discuss the challenges in stem cell transplantation, such as immune rejection, integration into the host retina, and functional recovery. Previous and ongoing clinical trials are examined to highlight the therapeutic efficacy and safety of these novel treatments. Additionally, we address the ethical considerations and regulatory frameworks governing stem cell research. Our analysis suggests that while stem-cell-based therapies offer a groundbreaking approach to treating retinal diseases, further research is needed to ensure long-term safety and to optimize therapeutic outcomes. This review summarizes the clinical evidence of stem cell therapy and current limitations in utilizing stem cells for retinal degeneration, such as age-related macular degeneration, retinitis pigmentosa, and Stargardt’s disease. Full article
(This article belongs to the Special Issue Retinal Diseases: From Molecular Mechanisms to Therapeutics)
20 pages, 6680 KiB  
Article
A Proximity Complementation Assay to Identify Small Molecules That Enhance the Traffic of ABCA4 Misfolding Variants
by Davide Piccolo, Christina Zarouchlioti, James Bellingham, Rosellina Guarascio, Kalliopi Ziaka, Robert S. Molday and Michael E. Cheetham
Int. J. Mol. Sci. 2024, 25(8), 4521; https://doi.org/10.3390/ijms25084521 - 20 Apr 2024
Viewed by 1055
Abstract
ABCA4-related retinopathy is the most common inherited Mendelian eye disorder worldwide, caused by biallelic variants in the ATP-binding cassette transporter ABCA4. To date, over 2200 ABCA4 variants have been identified, including missense, nonsense, indels, splice site and deep intronic defects. Notably, more than [...] Read more.
ABCA4-related retinopathy is the most common inherited Mendelian eye disorder worldwide, caused by biallelic variants in the ATP-binding cassette transporter ABCA4. To date, over 2200 ABCA4 variants have been identified, including missense, nonsense, indels, splice site and deep intronic defects. Notably, more than 60% are missense variants that can lead to protein misfolding, mistrafficking and degradation. Currently no approved therapies target ABCA4. In this study, we demonstrate that ABCA4 misfolding variants are temperature-sensitive and reduced temperature growth (30 °C) improves their traffic to the plasma membrane, suggesting the folding of these variants could be rescuable. Consequently, an in vitro platform was developed for the rapid and robust detection of ABCA4 traffic to the plasma membrane in transiently transfected cells. The system was used to assess selected candidate small molecules that were reported to improve the folding or traffic of other ABC transporters. Two candidates, 4-PBA and AICAR, were identified and validated for their ability to enhance both wild-type ABCA4 and variant trafficking to the cell surface in cell culture. We envision that this platform could serve as a primary screen for more sophisticated in vitro testing, enabling the discovery of breakthrough agents to rescue ABCA4 protein defects and mitigate ABCA4-related retinopathy. Full article
(This article belongs to the Special Issue Molecular Advances in Retinal Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><b>Structure of ABCA4 in the unbound state.</b> The structure is represented as a cartoon with the N- and C-terminal halves coloured in silver and gold, respectively. The selected residues are represented as spheres and colour coded. The structure has been modelled in Pymol (The PyMOL Molecular Graphics System, Version 2.5.5 Schrödinger LLC, New York, NY, USA) using the cryo-EM map of substrate-free ABCA4 [<a href="#B42-ijms-25-04521" class="html-bibr">42</a>]. The α-ABCA4 Abbexa antibody antigen is highlighted with a black curve.</p>
Full article ">Figure 2
<p><b>Localisation of WT ABCA4 and disease-associated variants.</b> CHO cells were transfected with plasmids expressing WT-ABCA4 protein and the indicated variants. At 48 h post-transfection, cells were fixed in 4% PFA and permeabilised with 0.2% Triton X-100 and labelled with 1D4 antibody (yellow) and CNX antibody (magenta). Nuclei were stained with DAPI (cyan). WT-ABCA4 mostly localised in large CNX-positive intracellular vesicle-like structures, whereas the indicated mutants localised in a reticular pattern characteristic of the ER (insets). Scale bars = 10 μm.</p>
Full article ">Figure 3
<p><b>Missense ABCA4 variants present partially impaired trafficking and maturation in cultured cells.</b> (<b>A</b>) CHO cells were transfected with plasmids expressing WT-ABCA4 and selected mutants. At 48 h post-transfection, cells were double-labelled with the 1D4 antibody to stain the intracellular ABCA4 (yellow) and with the ABCA4 Abbexa antibody (magenta, upper panels; white, lower panels) to stain the membrane-exposed extracellular epitope. Nuclei were stained with DAPI (cyan). WT-ABCA4 was localised both in the ER and at the plasma membrane, whereas missense variants were mainly trapped in the ER. Scale bars = 10 μm. (<b>B</b>) HEK293T cells were transfected with plasmids expressing WT-ABCA4, T983A and R2077W. At 48 h post-transfection, cell lysates were collected, and 6 µg of protein lysate was treated with glycosidases Endo H and PNGase-F, or buffer and temperature-only protocol control (T°). The untreated sample migrated as two species, band A and band B. The treated samples only showed band A. (<b>C</b>) The B/A band ratio was quantified using ImageJ (Version 2.14.0/1.54f). Fold changes relative to WT are shown. T983A and R2077W band ratios were reduced with respect to WT protein. Error bars are mean of fold change ± SD. n = 3 independent experiments. One-way ANOVAs and post-hoc analysis comparisons were performed only against WT ABCA4 sample.</p>
Full article ">Figure 4
<p><b>The effect of temperature on ABCA4 protein level and localisation.</b> (<b>A</b>) HEK293T cells were transfected with plasmids expressing WT-ABCA4 and the indicated variants and subsequently incubated at 37 °C or 30 °C. WT ABCA4 and variant protein levels detected by western blot (10 µg of protein lysate). Mock represents a non-transfected control. (<b>B</b>) Quantification of band intensity using ImageJ. Data were normalised to GAPDH reference protein intensity and 37 °C level (dashed line, 1.00). Error bars are mean ± SD. n = 3 independent experiments. Significant differences are displayed. (<b>C</b>) CHO cells transfected with WT-ABCA4 and variants incubated at 37 °C or 30 °C and stained for cell surface accessible ABCA4 (Abbexa) and total ABCA4 (1D4). Confocal images showing the localisation of ABCA4 in vesicle structures (blue arrows) and plasma membrane (red arrows). Cells treated at 30 °C show an increase in the plasma membrane and vesicle-positive cells. (<b>D</b>) ABCA4 vesicle-positive cells and plasma membrane-positive cells were scored among all transfected cells and expressed as a percentage of total transfected cells. n = 10 fields of view (×40 objective). Mean ± SD. Scale bars = 10 μm. Two-tailed Student’s <span class="html-italic">t</span>-test for 37 °C vs. 30 °C.</p>
Full article ">Figure 5
<p><b>Split NanoBiT proximity complementation assay</b>. (<b>A</b>) Schematic showing the design of the system. Misfolded membrane protein (ABCA4; blue) is tagged at the cytoplasmic C-terminus with SmBiT part of NanoLuc (dark red) and the LgBiT subunit (light red) is targeted to the plasma membrane by the N-terminus of RP2 (magenta). When misfolded protein is retained in the ER the SmBiT (dark red) and LgBiT (light red) cannot complement (<b>upper panel</b>). Small molecule rescue of ABCA4 plasma membrane traffic enables the SmBiT–LgBiT system to produce luminescence through proximity induced complementation of the NanoLuc (<b>lower panel</b>). (<b>B</b>) WT-RHO fused with SmBiT localises to the plasma membrane. The staining was performed with antibody 4D2 without a permeabilization step, as the 4D2 epitope is located on the extracellular N-terminus the protein. (<b>C</b>) P23H fused to SmBiT staining is observed in a perinuclear and reticular pattern, consistent with the ER, following a detergent permeabilization step to reveal 4D2 immunoreactivity. (<b>D</b>) P23H-SmBiT is not detected with 4D2 using a no-permeabilization protocol. (<b>E</b>) The RP2-LgBiT localised to the plasma membrane, when stained with anti-LgBiT antibody in permeabilized cells. (<b>F</b>) Intracellular WT-ABCA4-SmBiT was detected with the ABCA4 3F4 ABCAM antibody (yellow) and WT-ABCA4-SmBiT localising to the plasma membrane was detected using an Abbexa antibody (magenta) 48 h post-transfection analysis. Scale bars = 10 μm. (<b>G</b>) 24 h post-transfection with RHO-SmBiT + RP2-LgBiT and P23H-SmBiT + RP2-LgBiT plasmids luminescence signal was measured. Raw luminescence values are shown. Error bars are ± SD. n = 3. two-tailed Student’s <span class="html-italic">t</span>-test for RHO vs. P23H. (<b>H</b>) Luminescence signal was analysed 48 h post-transfection with ABCA4 variants + RP2-LgBiT plasmids in live HEK293T cells. Mean of fold change relative to WT ± SD. n = 4 independent experiments. One-way ANOVA and post-hoc analysis comparisons were performed only against WT-ABCA4-SmBit + RP2-LgBit sample. Significant differences are displayed.</p>
Full article ">Figure 6
<p><b>Cell-based complementation assay reveals drug potency for cell surface traffic.</b> (<b>A</b>) HEK293T cells were treated with 9-cis-retinal for 24 h, and luminescence was analysed 48 h post-transfection. P23H-SmBiT + RP2-LgBiT treated with 10 µM and 20 µM of pharmacological chaperone showed a significant increase in luminescence. Raw values are shown ± SD, n = 3. One-way ANOVAs and post-hoc analysis comparisons were conducted only against vehicle. (<b>B</b>,<b>C</b>) At 48 h post transfection, cells were treated with different concentrations of 4-PBA (<b>B</b>) and AICAR (<b>C</b>) for 24 h. Luminescence signal was measured. Fold change relative to the vehicle ± SD. n = 3. Two-way ANOVA against WT-ABCA4-SmBit + RP2-LgBit sample. Significant differences after post-hoc correction analysis are displayed.</p>
Full article ">Figure 7
<p><b>The effect of 4-PBA on ABCA4 protein expression and traffic.</b> (<b>A</b>,<b>B</b>) Western blot analysis and quantification of ABCA4 variants protein level (10 µg protein lysate), mean fold change ± SD compared to vehicle after normalisation to vinculin, n = 3. Black dashed line represents baseline vehicle value, 1.00. Two-tailed Student’s <span class="html-italic">t</span>-test for treated vs. vehicle. Significant differences are displayed. (<b>C</b>) CHO cells transfected with ABCA4 variants were treated with 5 mM 4-PBA for 24 h. The sub-cellular localisation was analysed by confocal microscopy using two different antibodies recognising ABCA4 intra-(1D4) and extra-(Abbexa) cellular epitopes. Blue arrows indicate vesicles, red arrows indicate ABCA4 at the plasma membrane, and, for T983A and R2077W, the insets show cells with vesicular staining. Scale bars = 10 μm.</p>
Full article ">Figure 8
<p><b>The effect of AICAR on ABCA4 protein.</b> (<b>A</b>,<b>B</b>) Western blot analysis and respective quantification of 10 µg of protein lysate from HEK293T cells expressing ABCA4 variants showing that AICAR treatment increases the steady state of the ABCA4 protein. n = 3. Error bars are mean of fold change ± SD. Black dashed line represents baseline vehicle value, 1.00. Two-tailed Student’s <span class="html-italic">t</span>-test for treated vs. vehicle. Significant differences are displayed. (<b>C</b>) CHO cells transfected with ABCA4 variants were treated with 0.5 mM AICAR for 24 h. The sub-cellular localisation was analysed by confocal microscopy using two different antibodies recognising ABCA4 intra-(1D4) and extra-(Abbexa) cellular epitopes. Blue arrows indicate vesicles, red arrows indicate ABCA4 at the plasma membrane. Scale bars = 10 μm.</p>
Full article ">
21 pages, 2265 KiB  
Article
Preclinical Development of Antisense Oligonucleotides to Rescue Aberrant Splicing Caused by an Ultrarare ABCA4 Variant in a Child with Early-Onset Stargardt Disease
by Nuria Suárez-Herrera, Catherina H. Z. Li, Nico Leijsten, Dyah W. Karjosukarso, Zelia Corradi, Femke Bukkems, Lonneke Duijkers, Frans P. M. Cremers, Carel B. Hoyng, Alejandro Garanto and Rob W. J. Collin
Cells 2024, 13(7), 601; https://doi.org/10.3390/cells13070601 - 29 Mar 2024
Cited by 1 | Viewed by 1370
Abstract
Precision medicine is rapidly gaining recognition in the field of (ultra)rare conditions, where only a few individuals in the world are affected. Clinical trial design for a small number of patients is extremely challenging, and for this reason, the development of N-of-1 strategies [...] Read more.
Precision medicine is rapidly gaining recognition in the field of (ultra)rare conditions, where only a few individuals in the world are affected. Clinical trial design for a small number of patients is extremely challenging, and for this reason, the development of N-of-1 strategies is explored to accelerate customized therapy design for rare cases. A strong candidate for this approach is Stargardt disease (STGD1), an autosomal recessive macular degeneration characterized by high genetic and phenotypic heterogeneity. STGD1 is caused by pathogenic variants in ABCA4, and amongst them, several deep-intronic variants alter the pre-mRNA splicing process, generally resulting in the insertion of pseudoexons (PEs) into the final transcript. In this study, we describe a 10-year-old girl harboring the unique deep-intronic ABCA4 variant c.6817-713A>G. Clinically, she presents with typical early-onset STGD1 with a high disease symmetry between her two eyes. Molecularly, we designed antisense oligonucleotides (AONs) to block the produced PE insertion. Splicing rescue was assessed in three different in vitro models: HEK293T cells, fibroblasts, and photoreceptor precursor cells, the last two being derived from the patient. Overall, our research is intended to serve as the basis for a personalized N-of-1 AON-based treatment to stop early vision loss in this patient. Full article
(This article belongs to the Special Issue Nucleic Acid Therapeutics (NATs): Advances and Perspectives)
Show Figures

Figure 1

Figure 1
<p><b>Mutation analysis and clinical description of the patient and her family.</b> (<b>A</b>) Pedigree showing the segregation of disease. Parents (I:1, I:2) are each heterozygous for one of the two variants. The affected daughter (II:2) is compound heterozygous, harboring both variants. The sibling (II:1) was not analyzed and did not present any symptoms. (<b>B</b>) Sanger sequencing traces of the two <span class="html-italic">ABCA4</span> regions encompassing the respective variants. (<b>C</b>) Imaging of the left and right eye of the patient demonstrates an identical phenotype, with a visual acuity of 0.16 (Snellen decimals). Color fundus photography reveals macular alterations indicated by the white arrows (L1, R1). Central retinal pigment epithelium (RPE) atrophy is evident in both eyes as not well-demarcated hypoautofluorescent areas, encircled by a hyperfluorescent ring, as indicated by the white arrows in the fundus autofluorescence images (L2, R2). Transfoveal OCT scans (L3, R3) show corresponding retinal layer loss of the ellipsoid zone (EZ), external limiting membrane (ELM), and the RPE, along with lipofuscin depositions.</p>
Full article ">Figure 2
<p><b>Schematic representation of the splicing defect caused by variant c.6817-713A&gt;G and initial screening of four antisense oligonucleotides (AONs) in minigene splice assays.</b> (<b>A</b>) Boundaries of the pseudoexon inclusion of 122 nt and flanking exons (GRCh37/hg19 genomic positions). Dashed lines represent the correct splicing process between the canonical splice sites and aberrant splicing between canonical and cryptic splice sites due to the presence of variant c.6817-713A&gt;G (highlighted in orange color). Scores for strength of the different splice donor sites (SDS, in blue color) and splice acceptor sites (SAS, in green color) were predicted with AlamutVisual Plus 1.7.1 software and SpliceAI, indicating that the substitution creates an SDS (scores highlighted in orange color). The score range from each prediction tool is indicated between brackets. (<b>B</b>) Depiction of the minigene construct used in splice assays and relative position of the designed AONs (A1, A2, A3, and A4) along the PE sequence. A1-A3 are targeting two high-scored exonic splicing enhancer (ESE) motifs, whereas A4 is targeting the cryptic SAS used for the PE inclusion. (<b>C</b>) AON-mediated rescue in minigene-transfected HEK293T cells. Analysis of correct (Correct) and pseudoexon (PE)-including <span class="html-italic">ABCA4</span> transcripts by RT-PCR. wild-type (WT) minigene and the respective mutant (MUT) minigene harboring variant c.6817-713A&gt;G were transfected in HEK293T cells. Non-transfected cells were used as endogenous expression control (HEK). The four AONs were then delivered at 0.5 µM, except for the non-treated MUT lane. Scrambled oligonucleotide (SON) delivery at 0.5 µM was used as negative control. Below the representative gel image: semi-quantification analysis graph of the different RT-PCR products is represented, indicating the percentages of the observed <span class="html-italic">ABCA4</span> transcripts in each condition. Amplification of β-actin (<span class="html-italic">ACTB</span>) gene was used as loading control, and exon 5 of the rhodopsin (<span class="html-italic">RHO</span>) gene was used as a minigene transfection control. MQ is used as negative control of all reactions. Data (n = 3) are presented as mean ± SD. Statistical significance is indicated as * <span class="html-italic">p</span> &lt; 0.05 using one-way ANOVA test with Dunnett’s multiple comparison analysis, in which non-treated MUT column was the reference condition for correct transcript levels.</p>
Full article ">Figure 3
<p><b>AON-mediated rescue of aberrant splicing in patient-derived fibroblasts.</b> Initial testing of the four designed AONs (<b>A</b>) and further -dose-response curve of the best-performing AONs (<b>B</b>) in control-individual- and patient-derived fibroblasts carrying variant c.6817-713A&gt;G in heterozygosity. Analysis of correct (Correct) and pseudoexon (PE)-including <span class="html-italic">ABCA4</span> transcripts by RT-PCR. The four designed AONs (A1–A4) were first tested at 0.5 µM, and increasing concentrations (0.1, 0.25, and 0.5 µM) of the lead candidates (A1–A3) were then delivered in a follow-up assay. Scrambled oligonucleotide (SON) delivery at 0.5 µM was used as negative control. Non-treated (NT) fibroblasts were used as endogenous <span class="html-italic">ABCA4</span> expression control. Below the representative gel images, semi-quantification analysis graphs of the different RT-PCR products are represented, indicating the percentages of the observed <span class="html-italic">ABCA4</span> transcripts in each condition. Amplification of β-actin (<span class="html-italic">ACTB</span>) gene was used as the loading control. MQ is used as the negative control of all reactions. Data (n = 3) are presented as mean ± SD. Statistical significance is indicated as ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 using one-way ANOVA test with Dunnett’s multiple comparison analysis, in which non-treated (NT) column was the reference condition for correct transcript levels.</p>
Full article ">Figure 4
<p><b>AON-mediated rescue of aberrant splicing in patient-derived photoreceptor precursor cells (PPCs).</b> (<b>A</b>) Schematic representation of the 30-day protocol used to obtain PPCs from patient-derived induced pluripotent stem cells (iPSCs). Essential 8 Flex (E8F) medium was used for iPSC seeding and at Day 0 was changed to Essential 6 (E6) medium. On Day 2, medium was replaced by neural induction medium (NIM), and later on Day 6, BMP4 pulse was performed. NIM was then refreshed every other day until day 30. (<b>B</b>) -Dose-response curve of the best-performing AONs in control-individual- and patient-derived PPCs carrying variant c.6817-713A&gt;G in heterozygosity. Analysis of correct (Correct) and pseudoexon (PE)-including <span class="html-italic">ABCA4</span> transcripts via RT-PCR. Increasing concentrations (0.25, 0.5, and 1 µM) of the lead candidates (A1-A3) were delivered on Day 20 of differentiation. PPCs were treated with cycloheximide (CHX) to assess the accumulation of total aberrant transcript or left untreated as endogenous <span class="html-italic">ABCA4</span> expression control (NT). Scrambled oligonucleotide (SON) delivery at 1 µM was used as negative control. Semi-quantification analysis of the different RT-PCR products are represented in the graph below the representative gel image. Amplification of β-actin (<span class="html-italic">ACTB</span>) gene was used as loading control. MQ is used as negative control of all reactions. Data (n = 2) are presented as mean ± SD. Statistical significance is indicated as * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 using one-way ANOVA test with Dunnett’s multiple comparison analysis, in which the CHX-treated column was the reference condition for correct transcript levels.</p>
Full article ">Figure 5
<p><b>Assessment of PE effect at the protein level.</b> (<b>A</b>) Open reading frame prediction of the correct and aberrant transcripts by ORF finder. Correct amino acids are depicted in blue color in both correct and aberrant ABCA4 proteins, whereas the extra 10 amino acids added by the PE sequence are depicted in orange color in the aberrant protein. Termination or stop codons are shown in red color. Depicted below, the Gateway p3xHA_CMV/DEST destination vector expressing either the wild-type (WT) <span class="html-italic">ABCA4</span> cDNA or presenting the 33-bp insertion between exons 49 and 50 (PE), and the corresponding vector transfection. Western blot analysis (n = 2) of non-transfected (NT) or transfected HEK293T cells with WT or PE <span class="html-italic">ABCA4</span> cDNA-expressing vector. HA-Tag was used as transfection control, and β-tubulin was used as a loading control. (<b>B</b>) AON-mediated protein rescue in PPCs by western blot analysis. On Day 20 of differentiation, control-individual-derived PPCs were treated with A3 at 1 µM, whereas increasing concentrations (0.25, 0.5, and 1 µM) were delivered to patient-derived PPCs carrying variant c.6817-713A&gt;G in heterozygosity. Both PPC lines were left as non-treated (NT) control, and scrambled oligonucleotide (SON) delivery at 1 µM was used as negative control. β-tubulin was used as a loading control and to normalize ABCA4 levels. Data (n = 2) are presented as mean ± SD in the semi-quantification graph as % of normalized ABCA4 levels to non-treated control. Statistical significance is indicated as * <span class="html-italic">p</span> &lt; 0.05 using one-way ANOVA test with Dunnett’s multiple comparison analysis, in which non-treated patient-derived PPCs were the reference condition.</p>
Full article ">
15 pages, 358 KiB  
Review
Artificial Intelligence (AI) for Early Diagnosis of Retinal Diseases
by Uday Pratap Singh Parmar, Pier Luigi Surico, Rohan Bir Singh, Francesco Romano, Carlo Salati, Leopoldo Spadea, Mutali Musa, Caterina Gagliano, Tommaso Mori and Marco Zeppieri
Medicina 2024, 60(4), 527; https://doi.org/10.3390/medicina60040527 - 23 Mar 2024
Cited by 6 | Viewed by 4337
Abstract
Artificial intelligence (AI) has emerged as a transformative tool in the field of ophthalmology, revolutionizing disease diagnosis and management. This paper provides a comprehensive overview of AI applications in various retinal diseases, highlighting its potential to enhance screening efficiency, facilitate early diagnosis, and [...] Read more.
Artificial intelligence (AI) has emerged as a transformative tool in the field of ophthalmology, revolutionizing disease diagnosis and management. This paper provides a comprehensive overview of AI applications in various retinal diseases, highlighting its potential to enhance screening efficiency, facilitate early diagnosis, and improve patient outcomes. Herein, we elucidate the fundamental concepts of AI, including machine learning (ML) and deep learning (DL), and their application in ophthalmology, underscoring the significance of AI-driven solutions in addressing the complexity and variability of retinal diseases. Furthermore, we delve into the specific applications of AI in retinal diseases such as diabetic retinopathy (DR), age-related macular degeneration (AMD), Macular Neovascularization, retinopathy of prematurity (ROP), retinal vein occlusion (RVO), hypertensive retinopathy (HR), Retinitis Pigmentosa, Stargardt disease, best vitelliform macular dystrophy, and sickle cell retinopathy. We focus on the current landscape of AI technologies, including various AI models, their performance metrics, and clinical implications. Furthermore, we aim to address challenges and pitfalls associated with the integration of AI in clinical practice, including the “black box phenomenon”, biases in data representation, and limitations in comprehensive patient assessment. In conclusion, this review emphasizes the collaborative role of AI alongside healthcare professionals, advocating for a synergistic approach to healthcare delivery. It highlights the importance of leveraging AI to augment, rather than replace, human expertise, thereby maximizing its potential to revolutionize healthcare delivery, mitigate healthcare disparities, and improve patient outcomes in the evolving landscape of medicine. Full article
29 pages, 5422 KiB  
Article
Scavenging of Cation Radicals of the Visual Cycle Retinoids by Lutein, Zeaxanthin, Taurine, and Melanin
by Malgorzata Rozanowska, Ruth Edge, Edward J. Land, Suppiah Navaratnam, Tadeusz Sarna and T. George Truscott
Int. J. Mol. Sci. 2024, 25(1), 506; https://doi.org/10.3390/ijms25010506 - 29 Dec 2023
Cited by 1 | Viewed by 1270
Abstract
In the retina, retinoids involved in vision are under constant threat of oxidation, and their oxidation products exhibit deleterious properties. Using pulse radiolysis, this study determined that the bimolecular rate constants of scavenging cation radicals of retinoids by taurine are smaller than 2 [...] Read more.
In the retina, retinoids involved in vision are under constant threat of oxidation, and their oxidation products exhibit deleterious properties. Using pulse radiolysis, this study determined that the bimolecular rate constants of scavenging cation radicals of retinoids by taurine are smaller than 2 × 107 M−1s−1 whereas lutein scavenges cation radicals of all three retinoids with the bimolecular rate constants approach the diffusion-controlled limits, while zeaxanthin is only 1.4–1.6-fold less effective. Despite that lutein exhibits greater scavenging rate constants of retinoid cation radicals than other antioxidants, the greater concentrations of ascorbate in the retina suggest that ascorbate may be the main protectant of all visual cycle retinoids from oxidative degradation, while α-tocopherol may play a substantial role in the protection of retinaldehyde but is relatively inefficient in the protection of retinol or retinyl palmitate. While the protection of retinoids by lutein and zeaxanthin appears inefficient in the retinal periphery, it can be quite substantial in the macula. Although the determined rate constants of scavenging the cation radicals of retinol and retinaldehyde by dopa-melanin are relatively small, the high concentration of melanin in the RPE melanosomes suggests they can be scavenged if they are in proximity to melanin-containing pigment granules. Full article
(This article belongs to the Special Issue The Role of Carotenoids in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Representative kinetics of the formation and decay of the transient species monitored at 610 nm after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinyl palmitate (<b>A</b>) and transient absorption spectra at indicated times after the pulse radiolysis of that solution (<b>B</b>). (<b>C</b>) Representative kinetics of the formation and decay of the transient species monitored at 610 nm after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinyl palmitate in the presence of 0.1 mM α-tocopherol. (<b>D</b>) Transient absorption spectra at indicated times after the pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinyl palmitate and 0.1 mM zeaxanthin, and representative kinetics of the formation and decay of transient species monitored at 610 nm (<b>E</b>) and 1000 nm (<b>F</b>) after pulse radiolysis of that solution. (<b>G</b>,<b>H</b>) Representative kinetics of the formation and decay of transient species monitored at 610 nm (<b>G</b>) and 950 nm (<b>H</b>) after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinyl palmitate in the presence of 0.1 mM lutein.</p>
Full article ">Figure 2
<p>Transient absorption spectra after pulse radiolysis of aqueous solution saturated with N<sub>2</sub>O and containing 10 mM phosphate, pH 7, 0.1 M KBr, and 1 mM retinyl palmitate incorporated in 2% Triton X-100 micelles (<b>A</b>) and a representative kinetics of the formation and decay of retinyl palmitate cation radicals monitored at 590 nm after pulse radiolysis of that solution in the absence (<b>B</b>) and presence of 0.1 mM of ascorbate (<b>C</b>).</p>
Full article ">Figure 3
<p>Representative kinetics of the formation and decay of retinyl palmitate cation radicals monitored at 590 nm after pulse radiolysis of aqueous solution saturated with N<sub>2</sub>O and containing 10 mM phosphate, pH 7, 0.1 M KBr, and 1 mM retinyl palmitate incorporated in 2% Triton X-100 micelles, in the absence (<b>A</b>) and presence of 0.1 mM of taurine (<b>B</b>) or 0.1 mg/mL (equivalent to 0.67 mM monomers) dopa-melanin (<b>C</b>).</p>
Full article ">Figure 4
<p>(<b>A</b>) Representative kinetics of the formation and decay of the transient species monitored at 610 nm after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinaldehyde. (<b>B</b>–<b>D</b>): Transient absorption spectra at indicated times after the pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinaldehyde and 0.1 mM lutein (<b>B</b>) and representative kinetics of the of the formation and decay of the transient species after the pulse radiolysis of that solution monitored at 610 nm (<b>C</b>) and 950 nm (<b>D</b>) after pulse radiolysis of that solution. (<b>E</b>,<b>F</b>) Representative kinetics of the formation and decay of transient species monitored at 610 nm (<b>E</b>) and 1000 nm (<b>F</b>) after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinaldehyde in the presence of 0.1 mM zeaxanthin.</p>
Full article ">Figure 5
<p>Representative kinetics of the formation and decay of retinaldehyde cation radicals monitored at 590 nm after pulse radiolysis of aqueous solution saturated with N<sub>2</sub>O and containing 10 mM phosphate, pH 7, 0.1 M KBr, and 1 mM retinaldehyde incorporated in 2% Triton X-100 micelles, in the absence (<b>A</b>) and presence of 0.1 mM of taurine (<b>B</b>) or 0.1 mg/mL (equivalent to 0.67 mM monomers) dopa-melanin (<b>C</b>).</p>
Full article ">Figure 6
<p>(<b>A</b>) Representative kinetics of the formation and decay of the transient species monitored at 610 nm after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinol. (<b>B</b>,<b>C</b>) Representative kinetics of the formation and decay of transient species monitored at 610 nm (<b>B</b>) and 950 nm (<b>C</b>) after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinol in the presence of 0.1 mM lutein. (<b>D</b>,<b>E</b>) Representative kinetics of the formation and decay of transient species monitored at 610 nm (<b>D</b>) and 1000 nm (<b>E</b>) after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinaldehyde in the presence of 0.1 mM zeaxanthin.</p>
Full article ">Figure 7
<p>Representative kinetics of the formation and decay of retinol cation radicals monitored at 590 nm after pulse radiolysis of aqueous solution saturated with N<sub>2</sub>O and containing 10 mM phosphate, pH 7, 0.1 M KBr, and 1 mM retinol incorporated in 2% Triton X-100 micelles, in the absence (<b>A</b>) and presence of 0.1 mM of taurine (<b>B</b>) or 0.1 mg/mL (equivalent to 0.67 mM monomers) dopa-melanin (<b>C</b>).</p>
Full article ">
15 pages, 22246 KiB  
Article
Diagnostic Challenges in ABCA4-Associated Retinal Degeneration: One Gene, Many Phenotypes
by Tien-En Tan, Rachael Wei Chao Tang, Choi Mun Chan, Ranjana S. Mathur and Beau J. Fenner
Diagnostics 2023, 13(23), 3530; https://doi.org/10.3390/diagnostics13233530 - 25 Nov 2023
Cited by 1 | Viewed by 1216
Abstract
(1) Purpose: ABCA4-associated retinal degeneration (ABCA4-RD) is a phenotypically diverse disease that often evades diagnosis, even by experienced retinal specialists. This may lead to inappropriate management, delayed genetic testing, or inaccurate interpretation of genetic testing results. Here, we illustrate the [...] Read more.
(1) Purpose: ABCA4-associated retinal degeneration (ABCA4-RD) is a phenotypically diverse disease that often evades diagnosis, even by experienced retinal specialists. This may lead to inappropriate management, delayed genetic testing, or inaccurate interpretation of genetic testing results. Here, we illustrate the phenotypic diversity of ABCA4-RD using a series of representative cases and compare these to other conditions that closely mimic ABCA4-RD. (2) Methods: Genetically confirmed ABCA4-RD cases with representative phenotypes were selected from an inherited retinal disease cohort in Singapore and compared to phenocopies involving other retinal diseases. (3) Results: ABCA4-RD phenotypes in this series included typical adolescent-onset Stargardt disease with flecks, bull’s eye maculopathy without flecks, fundus flavimaculatus, late-onset Stargardt disease, and severe early-onset Stargardt disease. Phenocopies of ABCA4-RD in this series included macular dystrophy, pattern dystrophy, cone dystrophy, advanced retinitis pigmentosa, Leber congenital amaurosis, drug toxicity, and age-related macular degeneration. Key distinguishing features that often suggested a diagnosis of ABCA4-RD were the presence of peripapillary sparing, macular involvement and centrifugal distribution, and a recessive pedigree. (4) Conclusions: ABCA4-RD demonstrates a remarkable phenotypic spectrum that makes diagnosis challenging. Awareness of the clinical spectrum of disease can facilitate prompt recognition and accurate diagnostic testing. Full article
(This article belongs to the Section Pathology and Molecular Diagnostics)
Show Figures

Figure 1

Figure 1
<p>Cases 1a versus 1b: Typical adolescent-onset Stargardt disease with macular atrophy, flecks and peripapillary sparing (<b>a</b>) versus <span class="html-italic">PRPH2</span>-associated macular dystrophy (<b>b</b>). Note the similar appearance on autofluorescence imaging, except for the peripapillary sparing, which is present in Case 1a, but absent in Case 1b. These cases can also be distinguished clinically based on age of onset and inheritance patterns.</p>
Full article ">Figure 2
<p>Cases 2a versus 2b versus 2c: Stargardt disease without flecks (<b>a</b>) versus <span class="html-italic">GUCY2D</span>-associated cone dystrophy (<b>b</b>) versus hydroxychloroquine toxicity (<b>c</b>). Note the similar appearance between Cases 2a and Case 2b, particularly on autofluorescence and optical coherence tomography imaging. The slight metallic sheen within the area of atrophy on the color photograph in Case 2a points towards <span class="html-italic">ABCA4</span>-associated Stargardt disease. These cases can also be distinguished based on clinical symptoms, electroretinography, and inheritance patterns. Case 2c shows that systemic drug toxicity can present with similar imaging features that mimic inherited retinal disease.</p>
Full article ">Figure 3
<p>Cases 3a versus 3a: Fundus flavimaculatus (<b>a</b>) versus autosomal-dominant pattern dystrophy (<b>b</b>). These cases both presented with late onset, yellow flecks, good vision, and minimal macular atrophy. Note the peripapillary sparing on autofluorescence imaging in Case 3a that points towards <span class="html-italic">ABCA4</span>-associated retinal degeneration. Also note the prominent pigment clumps within the yellow flecks in Case 3b, which are more characteristic of <span class="html-italic">PRPH2</span>-associated pattern dystrophy, although genetic testing results were not available for this case. These cases can also be distinguished based on the inheritance patterns.</p>
Full article ">Figure 4
<p>Cases 4a versus 4b: Late-onset Stargardt disease (<b>a</b>) versus dry age-related macular degeneration (AMD) with geographic atrophy (<b>b</b>). In Case 4a, the yellow flecks are more elongated and pisciform, and many more flecks are apparent on autofluorescence imaging where they are both hyper- and hypo-autofluorescent. There is also peripapillary sparing, with the nasal edge of the atrophy concave towards the disc as a result. In Case 4b, the yellow drusen are rounder, and do not have a significant autofluorescence signal. The nasal edge of the atrophy is convex towards the disc. There is a shallow irregular pigment epithelial detachment temporal to the atrophy on optical coherence tomography that is associated with AMD.</p>
Full article ">Figure 5
<p>Cases 5a versus 5b: Severe, early-onset Stargardt disease (<b>a</b>) versus late-stage retinitis pigmentosa (<b>b</b>). Note the similar appearances of the posterior pole on color photography, with widespread diffuse atrophy, arteriolar attenuation, and pigment clumps. Ultrawidefield imaging shows that, in Case 5a, the distribution of disease is more central, with extensive macular involvement and centrifugal disease progression, which is more characteristic of <span class="html-italic">ABCA4</span>-associated retinal degeneration. In contrast, with retinitis pigmentosa the disease progression is centripetal. These cases can also be distinguished in terms of clinical history.</p>
Full article ">Figure 6
<p>Cases 6a versus 6b: Severe, early-onset <span class="html-italic">ABCA4</span>-associated retinal degeneration (<b>a</b>) versus <span class="html-italic">CRB1</span>-associated Leber congenital amaurosis (LCA) (<b>b</b>). Note the similar appearances on color photography, with widespread diffuse atrophy and scattered pigment clumps. The cases also look similar on optical coherence tomography with diffuse outer retinal loss, inner retinal thickening, and loss of retinal laminations. The centrifugal distribution and peripapillary sparing on autofluorescence imaging in Case 6a points towards <span class="html-italic">ABCA4</span>-related retinal degeneration, while the relative lack of arteriolar attenuation in Case 6b points towards <span class="html-italic">CRB1</span>-associated LCA. These cases can also be distinguished clinically based on age of onset and presence or absence of nystagmus.</p>
Full article ">
Communication
Major Contribution of c.[1622T>C;3113C>T] Complex Allele and c.5882G>A Variant in ABCA4-Related Retinal Dystrophy in an Eastern European Population
by Vitaly V. Kadyshev, Ekaterina A. Alekseeva, Vladimir V. Strelnikov, Anna A. Stepanova, Alexander V. Polyakov, Andrey V. Marakhonov, Sergey I. Kutsev and Rena A. Zinchenko
Int. J. Mol. Sci. 2023, 24(22), 16231; https://doi.org/10.3390/ijms242216231 - 12 Nov 2023
Cited by 1 | Viewed by 1068
Abstract
Inherited retinal diseases (IRDs) constitute a prevalent group of inherited ocular disorders characterized by marked genetic diversity alongside moderate clinical variability. Among these, ABCA4-related eye pathology stands as a prominent form affecting the retina. In this study, we conducted an in-depth analysis [...] Read more.
Inherited retinal diseases (IRDs) constitute a prevalent group of inherited ocular disorders characterized by marked genetic diversity alongside moderate clinical variability. Among these, ABCA4-related eye pathology stands as a prominent form affecting the retina. In this study, we conducted an in-depth analysis of 96 patients harboring ABCA4 variants in the European part of Russia. Notably, the complex allele c.[1622T>C;3113C>T] (p.Leu541Pro;Ala1038Val, or L541P;A1038V) and the variant c.5882G>A (p.Gly1961Glu or G1961E) emerged as primary contributors to this ocular pathology within this population. Additionally, we elucidated distinct disease progression characteristics associated with the G1961E variant. Furthermore, our investigation revealed that patients with loss-of-function variants in ABCA4 were more inclined to develop phenotypes distinct from Stargardt disease. These findings provide crucial insights into the genetic and clinical landscape of ABCA4-related retinal dystrophies in this specific population. Full article
(This article belongs to the Special Issue Molecular Research of Ocular Pathology)
Show Figures

Figure 1

Figure 1
<p>Distribution of ABCA4 variants by type revealed in Russian cohort of IRD patients.</p>
Full article ">Figure 2
<p>Optical coherence tomograms of the right (<b>a</b>) and left (<b>b</b>) eyes in a patient (ID-16) with Stargardt’s disease 1 (STGD1).</p>
Full article ">Figure 3
<p>Optical coherence tomograms and image of the fundus of the right (<b>a</b>,<b>c</b>) and left (<b>b</b>,<b>d</b>) eyes of a patient (ID-76) with Retinitis pigmentosa 19 (RP19).</p>
Full article ">Figure 4
<p>Optical coherence tomograms of the right (<b>a</b>) and left (<b>b</b>) eyes in a patient (ID-31) with cone-rod dystrophy 3 (CORD3).</p>
Full article ">
19 pages, 4746 KiB  
Review
Updates on Emerging Interventions for Autosomal Recessive ABCA4-Associated Stargardt Disease
by Liang Wang, Serena M. Shah, Simran Mangwani-Mordani and Ninel Z. Gregori
J. Clin. Med. 2023, 12(19), 6229; https://doi.org/10.3390/jcm12196229 - 27 Sep 2023
Cited by 2 | Viewed by 2643
Abstract
Autosomal recessive Stargardt disease (STGD1) is an inherited retinal degenerative disease associated with a mutated ATP-binding cassette, subfamily A, member 4 (ABCA4) gene. STGD1 is the most common form of juvenile macular degeneration with onset in late childhood to early or [...] Read more.
Autosomal recessive Stargardt disease (STGD1) is an inherited retinal degenerative disease associated with a mutated ATP-binding cassette, subfamily A, member 4 (ABCA4) gene. STGD1 is the most common form of juvenile macular degeneration with onset in late childhood to early or middle adulthood and causes progressive, irreversible visual impairment and blindness. No effective treatment is currently available. In the present article, we review the most recent updates in clinical trials targeting the management of STGD1, including gene therapy, small molecule therapy, and stem cell therapy. In gene therapy, dual adeno-associated virus and non-viral vectors have been successful in delivering the human ABCA4 gene in preclinical studies. For pharmaceutical therapies ALK-001, deuterated vitamin A shows promise with preliminary data for phase 2 trial, demonstrating a decreased atrophy growth rate after two years. Stem cell therapy using human pluripotent stem cell-derived retinal pigment epithelium cells demonstrated long-term safety three years after implantation and visual acuity improvements in the first two years after initiation of therapy. Many other treatment options have ongoing investigations and clinical trials. While multiple potential interventions have shown promise in attenuating disease progression, further exploration is necessary to demonstrate treatment safety and efficacy. Full article
(This article belongs to the Special Issue Clinical Diagnosis and Treatment of Retinal Degeneration)
Show Figures

Figure 1

Figure 1
<p>Typical presentation of Stargardt Type 1 at diagnosis (<b>A<sub>1</sub></b>–<b>A<sub>3</sub></b>) and 3-year follow-up (<b>B<sub>1</sub>–B<sub>3</sub></b>). Color fundus (<b>A<sub>1</sub></b>,<b>B<sub>1</sub></b>) shows macular atrophy with yellow-white retinal flecks. Fundus autofluorescence (<b>A<sub>2</sub></b>,<b>B<sub>2</sub></b>) shows patches of hypoautofluorescence surrounded by an increased signal with flecks of both increased and decreased autofluorescence. Optical coherence tomography B-scans (<b>A<sub>3</sub></b>,<b>B<sub>3</sub></b>) show outer retinal and retinal pigment epithelium loss with hypertransmission defects corresponding to atrophy. Atrophy growth is significant at 3-year follow-up (<b>B<sub>1</sub></b>–<b>B<sub>3</sub></b>) compared to presentation at diagnosis (<b>A<sub>1</sub></b>–<b>A<sub>3</sub></b>). Best corrected visual acuity (BCVA) remains stable at 0.9 Snellen decimals (20/160). Reprinted from [<a href="#B1-jcm-12-06229" class="html-bibr">1</a>], used under an open-access license agreement distributed under the terms of the Creative Commons CC-BY license.</p>
Full article ">Figure 2
<p>Potential pharmaceutical interventions for Stargardt disease (STGD1). Treatment options (blue “x”) are shown in a summarized representation of the visual cycle with mutated ATP-binding cassette, subfamily A, member 4 (<span class="html-italic">ABCA4</span>) gene resulting in the accumulation of lipofuscin in the retinal pigment epithelium (RPE). Treatment for STGD1 includes several strategies. Gene therapy includes vector delivery of human <span class="html-italic">ABCA4</span> gene. Pharmaceutical therapies include visual cycle modulators (VCM), metformin, avacincaptad pegol, and docosahexaenoic acid (DHA) and eicosapentaenoic acid (EPA). VCM represented are retinol binding protein 4 (RBP4) antagonists, deuterated vitamin A, and retinal pigment epithelium-specific 65 kDa protein (RPE65) inhibitor (emixustat). Stem cell therapy includes human pluripotent stem cell-derived retinal pigment epithelium (hESC-RPE) transplantation for regenerating RPE layer. Created with BioRender.com.</p>
Full article ">
18 pages, 5011 KiB  
Article
Progression of Rare Inherited Retinal Dystrophies May Be Monitored by Adaptive Optics Imaging
by Katarzyna Samelska, Jacek Paweł Szaflik, Barbara Śmigielska and Anna Zaleska-Żmijewska
Life 2023, 13(9), 1871; https://doi.org/10.3390/life13091871 - 5 Sep 2023
Cited by 2 | Viewed by 1176
Abstract
Inherited retinal dystrophies (IRDs) are bilateral genetic conditions of the retina, leading to irreversible vision loss. This study included 55 eyes afflicted with IRDs affecting the macula. The diseases examined encompassed Stargardt disease (STGD), cone dystrophy (CD), and cone–rod dystrophy (CRD) using adaptive [...] Read more.
Inherited retinal dystrophies (IRDs) are bilateral genetic conditions of the retina, leading to irreversible vision loss. This study included 55 eyes afflicted with IRDs affecting the macula. The diseases examined encompassed Stargardt disease (STGD), cone dystrophy (CD), and cone–rod dystrophy (CRD) using adaptive optics (Rtx1™; Imagine Eyes, Orsay, France). Adaptive optics facilitate high-quality visualisation of retinal microstructures, including cones. Cone parameters, such as cone density (DM), cone spacing (SM), and regularity (REG), were analysed. The best corrected visual acuity (BCVA) was assessed as well. Examinations were performed twice over a 6-year observation period. A significant change was observed in DM (1282.73/mm2 vs. 10,073.42/mm2, p< 0.001) and SM (9.83 μm vs. 12.16 μm, p< 0.001) during the follow-up. BCVA deterioration was also significant (0.16 vs. 0.12, p = 0.001), albeit uncorrelated with the change in cone parameters. No significant difference in REG was detected between the initial examination and the follow-up (p = 0.089). Full article
(This article belongs to the Collection New Diagnostic and Therapeutic Developments in Eye Diseases)
Show Figures

Figure 1

Figure 1
<p>An adaptive optics image showing photoreceptors in a healthy eye (Rtx1™, Imagine Eyes, France). The photoreceptor mosaic appears intact (not disrupted) with individual photoreceptors visible as white and greyish spots.</p>
Full article ">Figure 2
<p>An adaptive optics image of the photoreceptors of an eye afflicted by cone dystrophy (Rtx1™, Imagine Eyes, France). Observe the cone disruption throughout the image with “dark spaces” apparent within the cone mosaic across different areas of the image.</p>
Full article ">Figure 3
<p>An adaptive optics image of photoreceptors in an eye affected by cone–rod dystrophy (Rtx1™, Imagine Eyes, France). Throughout the image, the cones are not clearly visible. Observe the “dark spaces” scattered within the cone mosaic across various regions of the picture.</p>
Full article ">Figure 4
<p>An adaptive optics image of the photoreceptors of the eye with Stargardt disease (Rtx1™; Imagine Eyes, France). The photoreceptor mosaic is disrupted, note the appearance “dark spaces” among the cone mosaic in various regions of the picture.</p>
Full article ">Figure 5
<p>Difference in DM change over a 6-year observation period with respect to sex. DM: cone density (1/<math display="inline"><semantics> <msup> <mi>mm</mi> <mn>2</mn> </msup> </semantics></math>).</p>
Full article ">Figure 6
<p>Difference in SM change over a 6-year observation with respect to sex. SM: cone spacing (<math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m).</p>
Full article ">
12 pages, 3365 KiB  
Case Report
A Splicing Variant in RDH8 Is Associated with Autosomal Recessive Stargardt Macular Dystrophy
by Stefania Zampatti, Cristina Peconi, Giulia Calvino, Rosangela Ferese, Stefano Gambardella, Raffaella Cascella, Jacopo Sebastiani, Benedetto Falsini, Andrea Cusumano and Emiliano Giardina
Genes 2023, 14(8), 1659; https://doi.org/10.3390/genes14081659 - 21 Aug 2023
Cited by 1 | Viewed by 1159
Abstract
Stargardt macular dystrophy is a genetic disorder, but in many cases, the causative gene remains unrevealed. Through a combined approach (whole-exome sequencing and phenotype/family-driven filtering algorithm) and a multilevel validation (international database searching, prediction scores calculation, splicing analysis assay, segregation analyses), a biallelic [...] Read more.
Stargardt macular dystrophy is a genetic disorder, but in many cases, the causative gene remains unrevealed. Through a combined approach (whole-exome sequencing and phenotype/family-driven filtering algorithm) and a multilevel validation (international database searching, prediction scores calculation, splicing analysis assay, segregation analyses), a biallelic mutation in the RDH8 gene was identified to be responsible for Stargardt macular dystrophy in a consanguineous Italian family. This paper is a report on the first family in which a biallelic deleterious mutation in RDH8 is detected. The disease phenotype is consistent with the expected phenotype hypothesized in previous studies on murine models. The application of the combined approach to genetic data and the multilevel validation allowed the identification of a splicing mutation in a gene that has never been reported before in human disorders. Full article
(This article belongs to the Collection Genetics and Genomics of Rare Disorders)
Show Figures

Figure 1

Figure 1
<p>Pedigree of the family. Genotypes for c.262+1G&gt;A in <span class="html-italic">RDH8</span> gene are reported below each tested individual. The affected sisters (II:2 and II:4) received a clinical diagnosis of Stargardt macular dystrophy. Their parents (I:1 and I:2) were reported as sighted but never evaluated. The healthy brother (II:5), son (III:4), and daughters (III:1, III:2, and III:3) of the affected sisters were evaluated by ophthalmologists who confirmed the absence of signs of macular disease.</p>
Full article ">Figure 2
<p>(<b>A</b>,<b>B</b>) OCT imaging of the macular region showing subretinal deposits in the foveal region and thinning of the outer nuclear layers ((<b>A</b>): Right eye; (<b>B</b>): Left eye). (<b>C</b>) OCT imaging of the normal macular region.</p>
Full article ">Figure 3
<p>Fundus autofluorescence showing perifoveal and peripheral flecks as well retinal pigment epithelium atrophy in the foveal and perifoveal region. RE: Right Eye, LE: Left Eye.</p>
Full article ">Figure 4
<p>(<b>A</b>) Ganzfeld mixed rod–cone electroretinograms in both eyes. (<b>B</b>) Ganzfeld cone electroretinograms in both eyes. RE: Right eye, LE: left eye.</p>
Full article ">Figure 5
<p>Minigene assay, Sanger sequencing of RHD8. (<b>A</b>) Agarose gel shows RT-PCR results of minigene assay for variant c.262+1G&gt;A. Lane 1 shows an amplicon of 423 bp corresponding to a wild-type genotype (159 bp of normal splicing of exons 2 + 264 bp of pSPL3 exon); lane 2 shows an amplicon of 492 bp corresponding to abnormal splicing produced by mutation c.262+1G&gt;A (492 bp [159 bp of normal splicing (exon 2) + 69 bp of partial retention of intron 2] + 264 bp of pSPL3 exon); lane 3 shows the amplification of pSPL3 without RHD8 cloning; lane 4 shows the amplification of HEK 293 T cDNA without transfection of pSPL3; and lane 5 shows negative control of PCR amplification. (<b>B</b>) Sanger sequence shows the normal sequence and the partial retention of intron 2. (<b>C</b>) Genomic sequence of RHD8 exon 2 is in blue capital letters, underlined in yellow is c.262+1G&gt;A variant, and underlined in green is the partial retention of intron 2.</p>
Full article ">
23 pages, 5727 KiB  
Article
Characteristics of Rare Inherited Retinal Dystrophies in Adaptive Optics—A Study on 53 Eyes
by Katarzyna Samelska, Jacek Paweł Szaflik, Maria Guszkowska, Anna Katarzyna Kurowska and Anna Zaleska-Żmijewska
Diagnostics 2023, 13(15), 2472; https://doi.org/10.3390/diagnostics13152472 - 25 Jul 2023
Cited by 4 | Viewed by 1231
Abstract
Inherited retinal dystrophies (IRDs) are genetic disorders that lead to the bilateral degeneration of the retina, causing irreversible vision loss. These conditions often manifest during the first and second decades of life, and their primary symptoms can be non-specific. Diagnostic processes encompass assessments [...] Read more.
Inherited retinal dystrophies (IRDs) are genetic disorders that lead to the bilateral degeneration of the retina, causing irreversible vision loss. These conditions often manifest during the first and second decades of life, and their primary symptoms can be non-specific. Diagnostic processes encompass assessments of best-corrected visual acuity, fundoscopy, optical coherence tomography, fundus autofluorescence, fluorescein angiography, electrophysiological tests, and genetic testing. This study focuses on the application of adaptive optics (AO), a non-invasive retinal examination, for the assessment of patients with IRDs. AO facilitates the high-quality, detailed observation of retinal photoreceptor structures (cones and rods) and enables the quantitative analysis of parameters such as cone density (DM), cone spacing (SM), cone regularity (REG), and Voronoi analysis (N%6). AO examinations were conducted on eyes diagnosed with Stargardt disease (STGD, N=36), cone dystrophy (CD, N=9), and cone-rod dystrophy (CRD, N=8), and on healthy eyes (N=14). There were significant differences in the DM, SM, REG, and N%6 parameters between the healthy and IRD-affected eyes (p<0.001 for DM, SM, and REG; p=0.008 for N%6). The mean DM in the CD, CRD, and STGD groups was 8900.39/mm2, 9296.32/mm2, and 16,209.66/mm2, respectively, with a significant inter-group difference (p=0.006). The mean SM in the CD, CRD, and STGD groups was 12.37 μm, 14.82 μm, and 9.65 μm, respectively, with a significant difference observed between groups (p=0.002). However, no significant difference was found in REG and N%6 among the CD, CRD, and STGD groups. Significant differences were found in SM and DM between CD and STGD (p=0.014 for SM; p=0.003 for DM) and between CRD and STGD (p=0.027 for SM; p=0.003 for DM). Our findings suggest that AO holds significant potential as an impactful diagnostic tool for IRDs. Full article
Show Figures

Figure 1

Figure 1
<p>Eye fundus image of a patient with Stargardt disease (DRI OCT Triton; Topcon). Observe the ‘bull’s eye’ maculopathy (indicated by arrows), pigment deposits (indicated by asterisks), and the presence of yellow-white flecks (highlighted with dots) in the perifoveal area.</p>
Full article ">Figure 2
<p>Adaptive optics image showcasing the photoreceptors of a healthy eye (Rtx1™; Imagine Eyes, France). Individual cones are distinctly visualized (visible as white and grayish dots), and the cone mosaic image appears undisrupted.</p>
Full article ">Figure 3
<p>Adaptive optics image demonstrating the photoreceptors of an eye with Stargardt disease (Rtx1™; Imagine Eyes, France). Observe the disruption in the cone mosaic (examples indicated by light blue arrows) and the presence of ‘dark spaces’ (examples highlighted with red asterisks). The area with inadequate visualization of the cone mosaic is marked with green X symbols.</p>
Full article ">Figure 4
<p>Eye fundus image of a macula with cone dystrophy (DRI OCT Triton; Topcon). Observe the ‘bull’s eye’ maculopathy (indicated by arrows).</p>
Full article ">Figure 5
<p>Eye fundus image of a macula with cone-rod dystrophy (DRI OCT Triton; Topcon). Note the ‘bull’s eye’ maculopathy (indicated by arrows), pigment deposits in the perifoveal area (marked with an asterisk), and the pallor of the optic nerve disc (marked with dark blue X symbol).</p>
Full article ">Figure 6
<p>Adaptive optics image of the photoreceptors in an eye with cone dystrophy (Rtx1™; Imagine Eyes, France). Observe the disruption of the cone mosaic (examples indicated by light blue arrows) and the presence of ‘dark spaces’ (examples indicated by red asterisks). The areas with poor visualization of the cone mosaic are marked with green X symbols.</p>
Full article ">Figure 7
<p>Adaptive optics image of the photoreceptors in an eye with cone-rod dystrophy (Rtx1™; Imagine Eyes, France). Observe the disruption of the cone mosaic (examples indicated by light blue arrows) and the presence of ‘dark spaces’ (examples indicated by red asterisks). Areas with poor visualization of the cone mosaic are marked with green X symbols.</p>
Full article ">Figure 8
<p>Comparison of DM among CD, CRD, and STGD groups. DM: cone density [1/mm<math display="inline"><semantics><msup><mrow/><mn>2</mn></msup></semantics></math>]; CD: cone dystrophy; CRD: cone-rod dystrophy; STGD: Stargardt disease.</p>
Full article ">Figure 9
<p>Comparison of SM among the CD, CRD, and STGD groups. SM: cone spacing [<math display="inline"><semantics><mi mathvariant="sans-serif">μ</mi></semantics></math>m]; CD: cone dystrophy; CRD: cone-rod dystrophy; STGD: Stargardt disease.</p>
Full article ">
Back to TopTop