[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
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,344)

Search Parameters:
Keywords = retinal degeneration

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 8206 KiB  
Article
Fundus Image Deep Learning Study to Explore the Association of Retinal Morphology with Age-Related Macular Degeneration Polygenic Risk Score
by Adam Sendecki, Daniel Ledwoń, Aleksandra Tuszy, Julia Nycz, Anna Wąsowska, Anna Boguszewska-Chachulska, Andrzej W. Mitas, Edward Wylęgała and Sławomir Teper
Biomedicines 2024, 12(9), 2092; https://doi.org/10.3390/biomedicines12092092 - 13 Sep 2024
Viewed by 231
Abstract
Background: Age-related macular degeneration (AMD) is a complex eye disorder with an environmental and genetic origin, affecting millions worldwide. The study aims to explore the association between retinal morphology and the polygenic risk score (PRS) for AMD using fundus images and deep learning [...] Read more.
Background: Age-related macular degeneration (AMD) is a complex eye disorder with an environmental and genetic origin, affecting millions worldwide. The study aims to explore the association between retinal morphology and the polygenic risk score (PRS) for AMD using fundus images and deep learning techniques. Methods: The study used and pre-processed 23,654 fundus images from 332 subjects (235 patients with AMD and 97 controls), ultimately selecting 558 high-quality images for analysis. The fine-tuned DenseNet121 deep learning model was employed to estimate PRS from single fundus images. After training, deep features were extracted, fused, and used in machine learning regression models to estimate PRS for each subject. The Grad-CAM technique was applied to examine the relationship between areas of increased model activity and the retina’s morphological features specific to AMD. Results: Using the hybrid approach improved the results obtained by DenseNet121 in 5-fold cross-validation. The final evaluation metrics for all predictions from the best model from each fold are MAE = 0.74, MSE = 0.85, RMSE = 0.92, R2 = 0.18, MAPE = 2.41. Grad-CAM heatmap evaluation showed that the model decisions rely on lesion area, focusing mostly on the presence of drusen. The proposed approach was also shown to be sensitive to artifacts present in the image. Conclusions: The findings indicate an association between fundus images and AMD PRS, suggesting that deep learning models may effectively estimate genetic risk for AMD from retinal images, potentially aiding in early detection and personalized treatment strategies. Full article
(This article belongs to the Special Issue Emerging Issues in Retinal Degeneration)
Show Figures

Figure 1

Figure 1
<p>Flow chart of fundus images quality assessment and selection.</p>
Full article ">Figure 2
<p>Flow chart of training and validation procedures in the hybrid model for PRS estimation based on fundus images.</p>
Full article ">Figure 3
<p>Scatter plots for the results of all folds test sets comparing true and estimated PRS values and the distributions in the control and AMD groups (<b>a</b>) and a linear regression model fit (<b>b</b>).</p>
Full article ">
24 pages, 7717 KiB  
Article
Novel Therapeutic Effects of Euphorbia heterophylla L. Methanol Extracts in Macular Degeneration Caused by Blue Light in A2E-Laden ARPE-19 Cells and Retina of BALB/c Mice
by Ayun Seol, Ji-Eun Kim, You-Jeong Jin, Hee-Jin Song, Yu-Jeong Roh, Tae-Ryeol Kim, Eun-Seo Park, Ki-Ho Park, So-Hae Park, Muhammad Salah Uddin, Sang-Woo Lee, Young-Woo Choi and Dae-Youn Hwang
Pharmaceuticals 2024, 17(9), 1193; https://doi.org/10.3390/ph17091193 - 10 Sep 2024
Viewed by 332
Abstract
Natural products with high antioxidant activity are considered as innovative prevention strategies to effectively prevent age-related macular degeneration (AMD) in the early stage because the generation of reactive oxygen species (ROS) leading to the development of drusen is reported as an important cause [...] Read more.
Natural products with high antioxidant activity are considered as innovative prevention strategies to effectively prevent age-related macular degeneration (AMD) in the early stage because the generation of reactive oxygen species (ROS) leading to the development of drusen is reported as an important cause of this disease. To investigate the prevention effects of the methanol extracts of Euphorbia heterophylla L. (MEE) on AMD, its effects on the antioxidant activity, inflammatory response, apoptosis pathway, neovascularization, and retinal tissue degeneration were analyzed in N-retinylidene-N-retinylethanolamine (A2E)-landed spontaneously arising retinal pigment epithelia (ARPE)-19 cells and BALB/c mice after exposure to blue light (BL). The MEE contained 10 active components and showed high free radical scavenging activity against 2,2-diphenyl-1-picrylhydrazyl (DPPH), 2,2′-azino-bis (3-ethylbenzothiazoline-6-sulfonic acid) (ABTS), and nitric oxide (NO) radicals. The pretreatments of high-dose MEE remarkably suppressed the production of intracellular ROS (88.2%) and NO (25.2%) and enhanced (SOD) activity (84%) and the phosphorylation of nuclear factor erythroid 2–related factor 2 (Nrf2) in A2E + BL-treated ARPE-19 cells compared to Vehicle-treated group. The activation of the inducible nitric oxide synthase (iNOS)-induced cyclooxygenase-2 (COX-2) mediated pathway, inflammasome activation, and expression of inflammatory cytokines was significantly inhibited in A2E + BL-treated ARPE-19 cells after the MEE pretreatment. The activation of the apoptosis pathway and increased expression of neovascular proteins (36% for matrix metalloproteinase (MMP)-9) were inhibited in the MEE pretreated groups compared to the Vehicle-treated group. Furthermore, the thickness of the whole retina (31%), outer nuclear layer (ONL), inner nuclear layer (INL), and photoreceptor layer (PL) were significantly increased by the MEE pretreatment of BALB/c mice with BL-induced retinal degeneration. Therefore, these results suggest that the MEE, with its high antioxidative activity, protects against BL-induced retinal degeneration through the regulation of the antioxidative system, inflammatory response, apoptosis, and neovascularization in the AMD mouse model. Full article
(This article belongs to the Section Natural Products)
Show Figures

Figure 1

Figure 1
<p>Determination of the main components in the MEE, using LC-MS analysis. Ten compounds including oleamide, linoleamide, palmitoleamide, quercetin, quercetrin, kaempferol, ellagic acid, β-sitosterol, octanoic acid, and linoleic acid were detected as distinct peaks in the chromatogram. Abbreviations: LC-MS, liquid chromatography–mass spectrometry; MEE, methanol extracts of <span class="html-italic">Euphorbia heterophylla</span> L.</p>
Full article ">Figure 2
<p>Free radical scavenging activity of the MEE. (<b>A</b>) DPPH radical scavenging activity. (<b>B</b>) ABTS radical scavenging activity. (<b>C</b>) NO radical scavenging activity. The activity of each radical was determined at 0.1 mM radicals and varying concentrations of the MEE (1–1000 μg/mL). The dotted line represented the trend pattern. The free radical scavenging analyses were performed on three MEE samples, and the optical density was measured twice for each well. Data are reported as the mean ± SD. Abbreviations: DPPH, 2,2-diphenyl-1-picrylhydrazyl; ABTS, 2,2′-azino-bis (3-ethylbenzothiazoline-6-sulfonic acid); NO, nitric oxide; IC<sub>50</sub>, half-maximal inhibitory concentration; MEE, methanol extracts of <span class="html-italic">Euphorbia heterophylla</span> L.</p>
Full article ">Figure 3
<p>Detection of intracellular ROS and NO in MEE + A2E + BL-treated ARPE-19 cells. (<b>A</b>) Fluorescence image of DCF stained ARPE-19 cells. The fluorescence of the cells was observed at 200× magnification. (<b>B</b>) Number of DCF stained cells. The preparation of DCFH-DA-stained cells was performed on two to three wells per group, and the positive cells were counted in two fields of view (67,500 mm<sup>2</sup>) in each well. (<b>C</b>) NO concentration. NO concentration in the culture supernatants was measured using the Griess reagent. The NO concentration analyses were performed using three wells per each group, and the assay for each sample was analyzed twice. Data are reported as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05 vs. No group. # <span class="html-italic">p</span> &lt; 0.05 vs. Vehicle + A2E + BL-treated group. Abbreviations: NO, nitric oxide; DCFHDA, 2,7-Dichlorofluorescin diacetate; MEE, methanol extracts of <span class="html-italic">Euphorbia heterophylla</span> L.; A2E, N-retinylidene-N-retinylethanolamine; BL, blue light.</p>
Full article ">Figure 4
<p>Determination of SOD, Nrf2, and p-Nrf2 proteins levels and SOD activity in MEE + A2E + BL-treated ARPE-19 cells. (<b>A</b>) Expression level of SOD, Nrf2, and p-Nrf2 proteins. After preparation of the total cell lysates from MEE + A2E + BL-treated ARPE-19 cells, SOD, Nrf2, and p-Nrf2 proteins were analyzed using specific antibodies. The expression level of each protein was normalized to β-actin. The cell lysates were prepared from two to three dishes per group and the Western blots were analyzed twice for each sample. (<b>B</b>) Level of SOD activity. After preparation of total cell lysates from the MEE + A2E + BL-treated ARPE-19 cells, SOD activity was detected using a specific assay kit. One unit of SOD was defined as the amount of the enzyme in the MEE solution (20 µL) that inhibits the reduction reaction of WST-1 with the superoxide anion by 50%. Data are reported as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05 vs. No group. # <span class="html-italic">p</span> &lt; 0.05 vs. Vehicle + A2E + BL-treated group. Abbreviations: SOD, superoxide dismutase; Nrf2, nuclear factors erythroid 2-related factors; MEE, methanol extracts of <span class="html-italic">Euphorbia heterophylla</span> L.; A2E, N-retinylidene-N-retinylethanolamine; BL, blue light; WST-1, water-soluble tetrazolium salt-1; ARPE, arising retinal pigment epithelia.</p>
Full article ">Figure 5
<p>Expression of key regulators on the iNOS-induced COX-2 mediated pathway, inflammasome pathway and transcription of inflammatory cytokines in MEE + A2E + BL-treated ARPE-19 cells. (<b>A</b>) Expression level of COX-2 and iNOS. (<b>B</b>) Expression level of NLRP3, ASC, Cas-1, and Cleaved Cas-1. After preparation of the total cell lysates from MEE + A2E + BL-treated ARPE-19 cells, the expression level of each protein was analyzed using specific antibodies and normalized to β-actin. The cell lysates were prepared from two to three dishes per group and the Western blots were analyzed twice for each sample. (<b>C</b>) mRNA levels of TNF-α, IL-1β, IL-6, and NF-κB. The mRNA levels of each gene were analyzed using specific primers and normalized to β-actin. The total RNAs were purified from cells of two to three dishes per group and RT-qPCR analysis was conducted twice for each sample. Data are reported as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05 vs. No group. # <span class="html-italic">p</span> &lt; 0.05 vs. Vehicle + A2E + BL-treated group. Abbreviations: COX-2, cyclooxygenase-2; iNOS, inducible nitric oxide synthase; NLRP3, NLR family pyrin domain containing 3; ASC, apoptosis-associated speck-like protein; Cas-1, Caspase-1; TNF-α, tumor necrosis factor α; IL, Interleukin; NF-κB, nuclear factor kappa-light-chain-enhancer of activated B; MEE, methanol extracts of <span class="html-italic">Euphorbia heterophylla</span> L.; A2E, N-retinylidene-N-retinylethanolamine; BL, blue light; ARPE, arising retinal pigment epithelia.</p>
Full article ">Figure 5 Cont.
<p>Expression of key regulators on the iNOS-induced COX-2 mediated pathway, inflammasome pathway and transcription of inflammatory cytokines in MEE + A2E + BL-treated ARPE-19 cells. (<b>A</b>) Expression level of COX-2 and iNOS. (<b>B</b>) Expression level of NLRP3, ASC, Cas-1, and Cleaved Cas-1. After preparation of the total cell lysates from MEE + A2E + BL-treated ARPE-19 cells, the expression level of each protein was analyzed using specific antibodies and normalized to β-actin. The cell lysates were prepared from two to three dishes per group and the Western blots were analyzed twice for each sample. (<b>C</b>) mRNA levels of TNF-α, IL-1β, IL-6, and NF-κB. The mRNA levels of each gene were analyzed using specific primers and normalized to β-actin. The total RNAs were purified from cells of two to three dishes per group and RT-qPCR analysis was conducted twice for each sample. Data are reported as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05 vs. No group. # <span class="html-italic">p</span> &lt; 0.05 vs. Vehicle + A2E + BL-treated group. Abbreviations: COX-2, cyclooxygenase-2; iNOS, inducible nitric oxide synthase; NLRP3, NLR family pyrin domain containing 3; ASC, apoptosis-associated speck-like protein; Cas-1, Caspase-1; TNF-α, tumor necrosis factor α; IL, Interleukin; NF-κB, nuclear factor kappa-light-chain-enhancer of activated B; MEE, methanol extracts of <span class="html-italic">Euphorbia heterophylla</span> L.; A2E, N-retinylidene-N-retinylethanolamine; BL, blue light; ARPE, arising retinal pigment epithelia.</p>
Full article ">Figure 6
<p>Level of cell deaths parameters in the MEE + A2E + BL-treated ARPE-19 cells. (<b>A</b>) Relative level of cell viability and (<b>B</b>) morphology of the MEE + A2E + BL-treated ARPE-19 cells. After pretreatment with three different dosages of the MEE for 24 h, their viability was analyzed using the MTT assay. The MTT assays were performed on samples from two to three wells per group, and the optical density was measured twice for each well. (<b>C</b>) The number of specific cells stained with Annexin V and 7-AAD. After staining, the number of stained cells was analyzed using a Muse Cell Analyzer. Annexin V and 7-AAD staining were performed from two to three wells per group, and the number of specific cells was counted twice for each well. The red lines represented four different stages of cell death, while individual cell was marked with a red dot. (<b>D</b>) Protein expression level of Bax, Bcl-2, Cas-3, and Cleaved Cas-3. The expression level of each protein was normalized to β-actin. The cell lysates were prepared from two to three dishes per group and the Western blots were analyzed twice for each sample. Data are reported as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05 vs. No group. # <span class="html-italic">p</span> &lt; 0.05 vs. Vehicle + A2E + BL-treated group. Abbreviations: MTT, 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide; 7-AAD, 7-aminoactinomycin D; Bax, Bcl-2-associated X protein; Bcl-2, B-cell lymphoma 2; Cas-3, Caspase-3; MEE, methanol extracts of <span class="html-italic">Euphorbia heterophylla</span> L.; A2E, N-retinylidene-N-retinylethanolamine; BL, blue light; ARPE, arising retinal pigment epithelia.</p>
Full article ">Figure 7
<p>Expression levels of angiogenic factors in MEE + A2E + BL-treated ARPE-19 cells. (<b>A</b>) Expression of angiogenic proteins. After collections of the total cell lysates from the MEE + A2E + BL-treated ARPE-19 cells, the MMP-2, MMP-9, and VEGF proteins were analyzed using specific antibodies. The expression level of each protein was normalized to β-actin. The cell lysates were prepared from two to three dishes per group and the Western blots were analyzed twice for each sample. (<b>B</b>) Transcription level of VEGF gene. The mRNA levels of this gene were analyzed using specific primers and normalized to β-actin. The total RNAs were purified from the cells of two to three dishes per group and RT-qPCR analysis was conducted twice for each sample. Data are reported as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05 vs. No group. # <span class="html-italic">p</span> &lt; 0.05 vs. Vehicle + A2E + BL-treated group. Abbreviations: MMP, matrix metalloproteinase; VEGF, vascular endothelial growth factor; MEE, methanol extracts of <span class="html-italic">Euphorbia heterophylla</span> L.; A2E, N-retinylidene-N-retinylethanolamine; BL, blue light; ARPE, arising retinal pigment epithelia.</p>
Full article ">Figure 8
<p>Histological structures in the retina and expression of iNOS and COX-2 proteins of MEE + BL-treated BALB/c mice. (<b>A</b>) Histology of the retina and thickness of whole retina. Histopathological alterations such as the thickness of the whole retina, OS, ONL, INL, and IPL were observed in the retina section stained with H&amp;E solution at 200× magnification. The thickness of whole retina represented bold arrow. The thicknesses of retina were measured using the Image J program 1.52a. Three different points including thickest point, thinnest point and medium thickness point were measured and presented as their mean values. The preparation of the H&amp;E-stained section was performed on three to five mice per group and the thickness was measured twice for each stained tissue. Expression of (<b>B</b>) iNOS and (<b>C</b>) COX-2 proteins in the retina. After staining three different antibodies, the stable DAB developed retinal section was observed at 400× magnification using light microscopy. The preparation of the IHC stained section was performed on three to five mice per group, and the distribution of each protein was analyzed twice for each stained tissue. Data are reported as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05 vs. No group. # <span class="html-italic">p</span> &lt; 0.05 vs. Vehicle + A2E + BL-treated group. Abbreviations: OS, outer segment, ONL, outer nuclear layer; INL, inner nuclear layer; IPL, inner plexiform layer; iNOS, inducible nitric oxide synthase; COX-2, cyclooxygenase-2; MEE, methanol extracts of <span class="html-italic">Euphorbia heterophylla</span> L.; A2E, N-retinylidene-N-retinylethanolamine; BL, blue light; H&amp;E, hematoxylin, and eosin; DAB, 3,3′-diaminobenzidine; IHC, immunohistochemistry.</p>
Full article ">
28 pages, 1393 KiB  
Review
Review: Neuroprotective Nanocarriers in Glaucoma
by Kun Pei, Maria Georgi, Daniel Hill, Chun Fung Jeffrey Lam, Wei Wei and Maria Francesca Cordeiro
Pharmaceuticals 2024, 17(9), 1190; https://doi.org/10.3390/ph17091190 - 10 Sep 2024
Viewed by 303
Abstract
Glaucoma stands as a primary cause of irreversible blindness globally, characterized by the progressive dysfunction and loss of retinal ganglion cells (RGCs). While current treatments primarily focus on controlling intraocular pressure (IOP), many patients continue to experience vision loss. Therefore, the research focus [...] Read more.
Glaucoma stands as a primary cause of irreversible blindness globally, characterized by the progressive dysfunction and loss of retinal ganglion cells (RGCs). While current treatments primarily focus on controlling intraocular pressure (IOP), many patients continue to experience vision loss. Therefore, the research focus has shifted to therapeutic targets aimed at preventing or delaying RGC death and optic nerve degeneration to slow or halt disease progression. Traditional ocular drug administration, such as eye drops or oral medications, face significant challenges due to the eye’s unique structural and physiological barriers, which limit effective drug delivery. Invasive methods like intravitreal injections can cause side effects such as bleeding, inflammation, and infection, making non-invasive delivery methods with high bioavailability very desirable. Nanotechnology presents a promising approach to addressing these limitations in glaucoma treatment. This review summarizes current approaches involving neuroprotective drugs combined with nanocarriers, and their impact for future use. Full article
Show Figures

Figure 1

Figure 1
<p>Anatomy of the retina and ganglion cells.</p>
Full article ">Figure 2
<p>Ocular drug delivery systems.</p>
Full article ">Figure 3
<p>Nanocarrier formulations for ocular drug delivery.</p>
Full article ">
24 pages, 1646 KiB  
Article
The Impact of ARMS2 (rs10490924), VEGFA (rs3024997), TNFRSF1B (rs1061622), TNFRSF1A (rs4149576), and IL1B1 (rs1143623) Polymorphisms and Serum Levels on Age-Related Macular Degeneration Development and Therapeutic Responses
by Dzastina Cebatoriene, Alvita Vilkeviciute, Greta Gedvilaite-Vaicechauskiene, Monika Duseikaite, Akvile Bruzaite, Loresa Kriauciuniene, Dalia Zaliuniene and Rasa Liutkeviciene
Int. J. Mol. Sci. 2024, 25(17), 9750; https://doi.org/10.3390/ijms25179750 - 9 Sep 2024
Viewed by 307
Abstract
Age-related macular degeneration (AMD) is a major global health problem as it is the leading cause of irreversible loss of central vision in the aging population. Anti-vascular endothelial growth factor (anti-VEGF) therapies are effective but do not respond optimally in all patients. This [...] Read more.
Age-related macular degeneration (AMD) is a major global health problem as it is the leading cause of irreversible loss of central vision in the aging population. Anti-vascular endothelial growth factor (anti-VEGF) therapies are effective but do not respond optimally in all patients. This study investigates the genetic factors associated with susceptibility to AMD and response to treatment, focusing on key polymorphisms in the ARMS2 (rs10490924), IL1B1 (rs1143623), TNFRSF1B (rs1061622), TNFRSF1A (rs4149576), VEGFA (rs3024997), ARMS2, IL1B1, TNFRSF1B, TNFRSF1A, and VEGFA serum levels in AMD development and treatment efficacy. This study examined the associations of specific genetic polymorphisms and serum protein levels with exudative and early AMD and the response to anti-VEGF treatment. The AA genotype of VEGFA (rs3024997) was significantly associated with a 20-fold reduction in the odds of exudative AMD compared to the GG + GA genotypes. Conversely, the TT genotype of ARMS2 (rs10490924) was linked to a 4.2-fold increase in the odds of exudative AMD compared to GG + GT genotypes. In females, each T allele of ARMS2 increased the odds by 2.3-fold, while in males, the TT genotype was associated with a 5-fold increase. Lower serum IL1B levels were observed in the exudative AMD group compared to the controls. Early AMD patients had higher serum TNFRSF1B levels than controls, particularly those with the GG genotype of TNFRSF1B rs1061622. Exudative AMD patients with the CC genotype of TNFRSF1A rs4149576 had lower serum TNFRSF1A levels compared to the controls. Visual acuity (VA) analysis showed that non-responders had better baseline VA than responders but experienced decreased VA after treatment, whereas responders showed improvement. Central retinal thickness (CRT) reduced significantly in responders after treatment and was lower in responders compared to non-responders after treatment. The T allele of TNFRSF1B rs1061622 was associated with a better response to anti-VEGF treatment under both dominant and additive genetic models. These findings highlight significant genetic and biochemical markers associated with AMD and treatment response. This study found that the VEGFA rs3024997 AA genotype reduces the odds of exudative AMD, while the ARMS2 rs10490924 TT genotype increases it. Lower serum IL1B levels and variations in TNFRSF1B and TNFRSF1A levels were linked to AMD. The TNFRSF1B rs1061622 T allele was associated with better anti-VEGF treatment response. These markers could potentially guide risk assessment and personalized treatment for AMD. Full article
Show Figures

Figure 1

Figure 1
<p>Serum IL1B levels were measured in patients with early AMD vs. control group (<b>A</b>) and exudative AMD vs. control groups (<b>B</b>). <span class="html-italic">p</span>-values marked with bold indicate statistically significant <span class="html-italic">p</span>-values, significant when <span class="html-italic">p</span> &lt; 0.05; Mann–Whitney U test was used.</p>
Full article ">Figure 2
<p>Serum TNFRSF1B levels were measured in patients with early AMD vs. control group (<b>A</b>) and exudative AMD vs. control group (<b>B</b>). <span class="html-italic">p</span>-values marked with bold indicate statistically significant <span class="html-italic">p</span>-values, significant when <span class="html-italic">p</span> &lt; 0.05; Mann–Whitney U test was used.</p>
Full article ">Figure 3
<p>Serum TNFRSF1B levels were measured in patients with exudative AMD vs. control group and compared between <span class="html-italic">TNFRSF1B</span> rs1061622 genotypes (<b>A</b>) and between early AMD vs. control group (<b>B</b>). <span class="html-italic">p</span>-values marked with bold indicate statistically significant <span class="html-italic">p</span>-values, significant when <span class="html-italic">p</span> &lt; 0.05; Mann–Whitney U test was used.</p>
Full article ">Figure 4
<p>Serum TNFRSF1A levels were measured in patients with exudative AMD vs. control group and compared between <span class="html-italic">TNFRSF1A</span> rs4149576 genotypes (<b>A</b>) and between early AMD vs. control group (<b>B</b>). <span class="html-italic">p</span>-values marked with bold indicate statistically significant <span class="html-italic">p</span>-values, significant when <span class="html-italic">p</span> &lt; 0.05; Mann–Whitney U test was used.</p>
Full article ">
15 pages, 941 KiB  
Article
Impact of Intra-Retinal Fluids on Changes in Retinal Ganglion Cell and Nerve Fiber Layers in Neovascular AMD under Anti-VEGF Therapy
by Yaser Abu Dail, Berthold Seitz, Haris Sideroudi and Alaa Din Abdin
J. Clin. Med. 2024, 13(17), 5318; https://doi.org/10.3390/jcm13175318 - 8 Sep 2024
Viewed by 358
Abstract
Purpose: To investigate the influence of intraretinal fluid (IRF) on change in retinal nerve fiber layer (RNFL) and retinal ganglion cell layer (RGCL) and thickness in patients with naive neovascular AMD under anti-VEGF treatment. Design: post hoc analysis. Methods: 97 [...] Read more.
Purpose: To investigate the influence of intraretinal fluid (IRF) on change in retinal nerve fiber layer (RNFL) and retinal ganglion cell layer (RGCL) and thickness in patients with naive neovascular AMD under anti-VEGF treatment. Design: post hoc analysis. Methods: 97 eyes of 83 patients on continuous therapy with intravitreal anti-vascular endothelial growth factors (anti-VEGF) and a follow-up of 24 months were included. RGCL and RNFL thickness in the perifoveal (-O), parafoveal (PF), and nasal areas and number of injections (IVI) were recorded before the first IVI as well as 1 and 2 years after initiating treatment and compared longitudinally and between groups with and without IRF. Results: The group with IRF at baseline had a higher RNFL thickness at baseline and showed a significant reduction in RNFL-PF between baseline and first and second follow-ups (p < 0.001) but not between first and second follow-ups. The group without IRF showed no significant reduction in RNFL over time. The presence of IRF was not associated with a reduction in RNFL-O or RNFL-nasal. RGCL thickness decreased significantly in both groups with and without IRF after 2 years. Number of IVIs showed no significant correlation to RNFL or RGCL after stratification for the presence of IRF. Conclusions: The presence of IRF has a significant influence on RNFL thickness at baseline as well as on its changes over time during anti-VEGF therapy. The preoperative presence of IRF should be considered when comparing changes in RNFL thickness after IVI. Full article
(This article belongs to the Section Ophthalmology)
Show Figures

Figure 1

Figure 1
<p>Development of the thickness of the retinal ganglion cell layer (RGCL) parafoveal (PF) and in the outer ring of the ETDRS grid (O) in the overall group and in the groups with and without subretinal fluid (SRF). Statistically significant changes (<span class="html-italic">p</span> &lt; 0.05) are marked with double arrows.</p>
Full article ">Figure 2
<p>Development of the thickness of the retinal nerve fiber layer (RNFL), both parafoveal (PF) and in the outer ring of the ETDRS grid (O), in the overall group, and in the groups with and without intraretinal fluid (IRF). Statistically significant changes (<span class="html-italic">p</span> &lt; 0.05) are marked with double arrows.</p>
Full article ">
18 pages, 29602 KiB  
Article
Slc4a7 Regulates Retina Development in Zebrafish
by Youyuan Zhuang, Dandan Li, Cheng Tang, Xinyi Zhao, Ruting Wang, Di Tao, Xiufeng Huang and Xinting Liu
Int. J. Mol. Sci. 2024, 25(17), 9613; https://doi.org/10.3390/ijms25179613 - 5 Sep 2024
Viewed by 345
Abstract
Inherited retinal degenerations (IRDs) are a group of genetic disorders characterized by the progressive degeneration of retinal cells, leading to irreversible vision loss. SLC4A7 has emerged as a candidate gene associated with IRDs, yet its mechanisms remain largely unknown. This study aims to [...] Read more.
Inherited retinal degenerations (IRDs) are a group of genetic disorders characterized by the progressive degeneration of retinal cells, leading to irreversible vision loss. SLC4A7 has emerged as a candidate gene associated with IRDs, yet its mechanisms remain largely unknown. This study aims to investigate the role of slc4a7 in retinal development and its associated molecular pathogenesis in zebrafish. Morpholino oligonucleotide knockdown, CRISPR/Cas9 genome editing, quantitative RT-PCR, eye morphometric measurements, immunofluorescent staining, TUNEL assays, visual motor responses, optokinetic responses, rescue experiments, and bulk RNA sequencing were used to assess the impact of slc4a7 deficiency on retinal development. Our results demonstrated that the knockdown of slc4a7 resulted in a dose-dependent reduction in eye axial length, ocular area, and eye-to-body-length ratio. The fluorescence observations showed a significant decrease in immunofluorescence signals from photoreceptors and in mCherry fluorescence from RPE in slc4a7-silenced morphants. TUNEL staining uncovered the extensive apoptosis of retinal cells induced by slc4a7 knockdown. Visual behaviors were significantly impaired in the slc4a7-deficient larvae. GO and KEGG pathway analyses reveal that differentially expressed genes are predominantly linked to aspects of vision, ion channels, and phototransduction. This study demonstrates that the loss of slc4a7 in larvae led to profound visual impairments, providing additional insights into the genetic mechanisms predisposing individuals to IRDs caused by SLC4A7 deficiency. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>The expression profile of <span class="html-italic">SLC4A7 (slc4a7</span>) in both humans and zebrafish. (<b>A</b>) Profiling of human single-cell RNA sequencing unveils <span class="html-italic">SLC4A7</span> expression in various ocular tissues, including the retina, cornea, iris, sclera, and choroid. (<b>B</b>) Single-cell transcriptome profiling and <span class="html-italic">SLC4A7</span> gene signatures of the human retina. The upper tSNE plot showing major cell subsets in human retina. The lower tSNE plot of all cells colored by enrichment of <span class="html-italic">SLC4A7</span> gene signatures. (<b>C</b>) Detailed expression pattern of <span class="html-italic">SLC4A7</span> across various cell types in the human retina. (<b>D</b>) Human subcellular localization of <span class="html-italic">SLC4A7</span> provided by COMPARTMENTS. (<b>E</b>,<b>F</b>) qRT-PCR display the time series (<b>E</b>) and the tissue-specific (<b>F</b>) expression pattern of <span class="html-italic">slc4a7</span> in zebrafish larvae. Bar plots are shown as the mean ± std. Label of human and zebrafish indicating the species source of the data. The normalized expression values refer to the standardized measure of gene expression levels across different cells or cell types in a single-cell RNA sequencing dataset, as calculated and provided by the online analysis tool, Single Cell Portal. The relative expression values were calculated relative to the expression of the reference gene β-actin. The dashed lines indicate detailed information about the corresponding tissues or figures. Human or zebrafish icons indicate the species origin of the data.</p>
Full article ">Figure 2
<p><span class="html-italic">Slc4a7</span>-deficient zebrafish morphants exhibited marked ocular changes. (<b>A</b>) Whole-body view (scale bar = 500 μm), lateral view (scale bar = 100 μm), and vertical view (scale bar = 100 μm) of zebrafish larvae at 3 dpf. The ocular area is indicated by white circles. The ocular axis is indicated by a red dashed line. The equatorial axis is indicated by a blue dashed line. (<b>B</b>–<b>E</b>) The measurement of ocular axis length, equatorial length, ocular area, and the ratio of ocular length to body length. <span class="html-italic">n</span> = 25 in each group. Bar plots are shown as the mean ± s.e.m. Data were analyzed using one-way ANOVA followed by Tukey’s post hoc tests, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 indicate significant differences from the control 1.00 ng group. The scattered circles, squares, and triangles of various colors in the bar chart denote individual values of zebrafish eye-related parameters for different groups.</p>
Full article ">Figure 2 Cont.
<p><span class="html-italic">Slc4a7</span>-deficient zebrafish morphants exhibited marked ocular changes. (<b>A</b>) Whole-body view (scale bar = 500 μm), lateral view (scale bar = 100 μm), and vertical view (scale bar = 100 μm) of zebrafish larvae at 3 dpf. The ocular area is indicated by white circles. The ocular axis is indicated by a red dashed line. The equatorial axis is indicated by a blue dashed line. (<b>B</b>–<b>E</b>) The measurement of ocular axis length, equatorial length, ocular area, and the ratio of ocular length to body length. <span class="html-italic">n</span> = 25 in each group. Bar plots are shown as the mean ± s.e.m. Data were analyzed using one-way ANOVA followed by Tukey’s post hoc tests, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 indicate significant differences from the control 1.00 ng group. The scattered circles, squares, and triangles of various colors in the bar chart denote individual values of zebrafish eye-related parameters for different groups.</p>
Full article ">Figure 3
<p>Loss of <span class="html-italic">slc4a7</span> leads to retinal abnormalities in zebrafish larvae. (<b>A</b>) Immunostaining for RCVRN in <span class="html-italic">slc4a7</span>-deficient and control AB zebrafish strains at 5 dpf. (<b>B</b>) The fluorescence imaging depicts retinal pigment epithelium cells and amacrine cells in <span class="html-italic">slc4a7</span>-deficient and control Tg (gad1b:mCherry) strain at 5 dpf. (<b>C</b>) The fluorescence imaging focusing on blood vessel endothelial cells in <span class="html-italic">slc4a7</span>-deficient and control Tg(kdrl:mCherry) strains at 5 dpf. (<b>D</b>) The fluorescence imaging illustrates Müller cells in <span class="html-italic">slc4a7</span>-deficient and control Tg (gfap:egfp) strains at 5 dpf. (<b>E</b>) Statistical results for normalized fluorescence intensity of RCVRN. (<b>F</b>,<b>G</b>) Statistical results for normalized fluorescence intensity of mCherry signal in amacrine cells (ACs), (<b>F</b>), and the RPE layer (<b>G</b>). (<b>H</b>) Statistical results for vascular network thickness surrounding photoreceptors. (<b>I</b>) Statistical results for the normalized fluorescence intensity of the eGFP signal in Müller cell layer. Scale bar = 50 μm. Bar plots are shown as the mean ± s.e.m. The <span class="html-italic">t</span>-test was performed between the two groups. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001. GCL: ganglion cell layer; IPL: inner plexiform layer; INL: inner nuclear layer; OPL: outer plexiform layer; ONL: outer nuclear layer.</p>
Full article ">Figure 4
<p>Knockdown of <span class="html-italic">slc4a7</span> resulted in a marked elevation of apoptosis in the zebrafish retina. TUNEL assay was used to detect apoptosis in larval retinas at 5 dpf. Knockdown of <span class="html-italic">slc4a7</span> led to a marked elevation in the number of TUNEL<sup>+</sup> cells in the retina at 5 days dpf. Scale bars = 20 μm. Blue indicates cell nucleus stained by DAPI. Red indicates cells with a positive TUNEL reaction.</p>
Full article ">Figure 5
<p>Zebrafish morphants lacking <span class="html-italic">slc4a7</span> exhibited impaired visual behaviors. (<b>A</b>,<b>B</b>) VMR testing results are depicted in the line charts for <span class="html-italic">slc4a7</span>-deficient zebrafish morphants. (<b>C</b>,<b>D</b>) Scatter plots with bar show the quantification of ON responses and OFF responses. Each point represents the average activity of 12 larvae within the group at the moment of each light environment change. Error bars represent standard deviation (STD). (<b>E</b>,<b>F</b>) Frequency distributions of eye movements (times/minute) in zebrafish larvae are illustrated in the stacked boxes. Data were analyzed using one-way ANOVA with Tukey’s post hoc tests, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001. The scattered circles of various colors in the bar chart denote individual values of zebrafish eye-related parameters for different groups.</p>
Full article ">Figure 6
<p>Transcriptional profiling of bulbus oculi from <span class="html-italic">slc4a7</span>-deficient zebrafish morphants. (<b>A</b>) Principal component analysis (PCA) of the expressed genes showing <span class="html-italic">slc4a7</span> and control sample separation. (<b>B</b>) Heatmap of sample-to-sample distances. (<b>C</b>) Volcano plot showing highlights of DEGs from <span class="html-italic">slc4a7</span> KD eyes compared with control eyes. (<b>D</b>) GO analysis identified the top 30 most significant GO terms in the <span class="html-italic">slc4a7</span> 1.00 ng group compared to the control 1.00 ng group. (<b>E</b>) Significantly enriched KEGG pathways (top 20) in the <span class="html-italic">slc4a7</span> 1.00 ng group compared to the control 1.00 ng group.</p>
Full article ">Figure 7
<p><span class="html-italic">Slc4a7</span> mRNA compensation rescues phenotypes in slc4a7-deficient morphants. (<b>A</b>) Enlarged vertical and lateral views of larval eyeballs. Larvae injected with slc4a7 MO and full-length slc4a7 mRNA exhibit normal-sized eyeballs at 3 dpf. Scale bar = 100 mm. (<b>B</b>–<b>D</b>) Statistical analysis of axial length and ocular area. (<b>E</b>,<b>F</b>) VMR testing demonstrates significant recovery of both ON and OFF responses in slc4a7-deficient zebrafish compensated with mRNA compared to those without compensation. <span class="html-italic">n</span> = 20 in each group. Rescue experiments were repeated three times. Bar plots represent the mean ± s.e.m. Data were analyzed using one-way ANOVA followed by Tukey’s post hoc test, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, indicate significant differences from the <span class="html-italic">slc4a7</span> 1.00 ng group. ns means no significance. The scattered circles of various colors in the bar chart denote individual values of zebrafish eye-related parameters for different groups.</p>
Full article ">
12 pages, 2669 KiB  
Article
Inflammation and Vasculitis Related to Brolucizumab
by António Campos, Carolina Mota, Francisco Caramelo, Nuno Oliveira, Sara Silva and João Sousa
J. Clin. Med. 2024, 13(17), 5208; https://doi.org/10.3390/jcm13175208 - 2 Sep 2024
Viewed by 451
Abstract
Background/objectives: To compare the prevalence of intra-ocular inflammation (IOI) between brolucizumab and aflibercept in neovascular age-related macular degeneration (nAMD) after intra-vitreal injections (IVI) and to compare the IOI odds ratios (ORs) of both therapies with the prevalence of septic endophthalmitis after IVI [...] Read more.
Background/objectives: To compare the prevalence of intra-ocular inflammation (IOI) between brolucizumab and aflibercept in neovascular age-related macular degeneration (nAMD) after intra-vitreal injections (IVI) and to compare the IOI odds ratios (ORs) of both therapies with the prevalence of septic endophthalmitis after IVI that was previously reported in the literature. Methods: A total of 468 IVI of brolucizumab (117 eyes) were compared with 2884 IVI of aflibercept (305 eyes) regarding IOI and occlusive retinal vasculitis (RV) from December 2021 to June 2023 in this retrospective study. The OR was calculated for both anti-VEGF agents and was compared with the relative risk of septic endophthalmitis after IVI. Results: There were four eyes with unilateral IOI related to brolucizumab (3.42%), one presenting uveitis (0.85%), two vitritis (1.71%) and the last one presenting occlusive RV (0.85%), compared with two eyes presenting unilateral IOI (anterior uveitis, 0.66%) and none with RV from the aflibercept cohort. The incidence of IOI per injection with brolucizumab (0.855%) was significantly higher compared with aflibercept (0.069%, p = 0.004). The OR of IOI related to brolucizumab IVI compared with septic endophthalmitis was 20 times greater (1.49 for aflibercept, p = 0.646, versus 20.15 for brolucizumab, p < 0.001). The OR of RV with brolucizumab compared with septic endophthalmitis was 4.6. Conclusion: Data from our department suggest a much higher risk of IOI and occlusive retinal vasculitis after brolucizumab when compared with aflibercept. The risk of IOI and severe sight-threatening complications related to brolucizumab is greater than the risk of septic endophthalmitis after any IVI. Full article
(This article belongs to the Section Ophthalmology)
Show Figures

Figure 1

Figure 1
<p>Optical coherence tomography (OCT) findings in the event with retinal artery vasculitis and artery occlusion. (<b>A</b>). On q8 aflibercept, there were subretinal fluid and intra-retinal cysts, despite a BCVA of 20/50 +2 L. (<b>B</b>). After the second injection of brolucizumab, there was no fluid in the retina, but there was edema of the inner retinal layers and diffuse atrophy of the retina. There was submacular scarring and extensive atrophy of the ellipsoid zone. BCVA dropped to counting fingers. (<b>C</b>). Before discharge, there was extensive atrophy and some retinal cysts before the patient resumed aflibercept. BCVA was &lt; 20/200.</p>
Full article ">Figure 2
<p>Color fundus photography (CFP). There were diffuse vascular sheathing involving arteries in the four quadrants, narrowing of the arterial and vein diameters, and cotton wool spots, but notably, retinal hemorrhages were absent.</p>
Full article ">Figure 3
<p>Fluorescein angiography. (<b>A</b>) (40″) and (<b>B</b>) (50″). Early choroidal multifocal hypofluorescence related to hypoperfusion (asterisks) and retinal arterial filing defects in the form of amputation of flux in the superior nasal and inferior temporal arteries (arrows), along with superior temporal and ciliary arteries’ delay. (<b>C</b>) (1′09″) and (<b>D</b>) (2′09″). Wide extensions of non-perfusion were present in the nasal and temporal areas, with venous narrowing and late filling as well as macular ischemia (arrowheads).</p>
Full article ">
28 pages, 5832 KiB  
Article
Bioactive Glial-Derived Neurotrophic Factor from a Safe Injectable Collagen–Alginate Composite Gel Rescues Retinal Photoreceptors from Retinal Degeneration in Rabbits
by Tingyu Hu, Ting Zhou, Rajesh Kumar Goit, Ka Cheung Tam, Yau Kei Chan, Wai-Ching Lam and Amy Cheuk Yin Lo
Mar. Drugs 2024, 22(9), 394; https://doi.org/10.3390/md22090394 - 30 Aug 2024
Viewed by 897
Abstract
The management of vision-threatening retinal diseases remains challenging due to the lack of an effective drug delivery system. Encapsulated cell therapy (ECT) offers a promising approach for the continuous delivery of therapeutic agents without the need for immunosuppressants. In this context, an injectable [...] Read more.
The management of vision-threatening retinal diseases remains challenging due to the lack of an effective drug delivery system. Encapsulated cell therapy (ECT) offers a promising approach for the continuous delivery of therapeutic agents without the need for immunosuppressants. In this context, an injectable and terminable collagen–alginate composite (CAC) ECT gel, designed with a Tet-on pro-caspase-8 system, was developed as a safe intraocular drug delivery platform for the sustained release of glial-cell-line-derived neurotrophic factor (GDNF) to treat retinal degenerative diseases. This study examined the potential clinical application of the CAC ECT gel, focusing on its safety, performance, and termination through doxycycline (Dox) administration in the eyes of healthy New Zealand White rabbits, as well as its therapeutic efficacy in rabbits with sodium-iodate (SI)-induced retinal degeneration. The findings indicated that the CAC ECT gel can be safely implanted without harming the retina or lens, displaying resistance to degradation, facilitating cell attachment, and secreting bioactive GDNF. Furthermore, the GDNF levels could be modulated by the number of implants. Moreover, Dox administration was effective in terminating gel function without causing retinal damage. Notably, rabbits with retinal degeneration treated with the gels exhibited significant functional recovery in both a-wave and b-wave amplitudes and showed remarkable efficacy in reducing photoreceptor apoptosis. Given its biocompatibility, mechanical stability, controlled drug release, terminability, and therapeutic effectiveness, our CAC ECT gel presents a promising therapeutic strategy for various retinal diseases in a clinical setting, eliminating the need for immunosuppressants. Full article
(This article belongs to the Special Issue Marine-Derived Biomaterials for Tissue Regeneration)
Show Figures

Figure 1

Figure 1
<p>Schematic diagrams illustrating a CAC ECT gel designed for drug delivery and equipped with a controllable biosafety switch. (<b>A</b>), Gel safety and performance were assessed in healthy NZW rabbits. (<b>B</b>), Gel termination can be achieved by Dox administration when necessary and by its biosafety termination mechanism. CAC ECT, collagen–alginate composite encapsulated cell therapy. CAC ECT, collagen–alginate composite encapsulated cell therapy. NZW, New Zealand White. HEK, human embryonic kidney. GDNF, glial-cell-derived neurotrophic factor. CMV, cytomegalovirus.</p>
Full article ">Figure 2
<p>Intravitreal injection of CAC ECT gels induced no changes in retinal function and retinal cytoarchitecture (<b>A</b>). The photopic and scotopic electroretinographic responses indicated retinal function of different groups, including the unoperated control, three-gel, and six-gel groups at 2 weeks. Photopic 5 a-wave and b-wave (mean ± SEM, n = 11, 4, and 3 animals for unoperated control, three-gel, and six-gel groups, respectively), scotopic 0.01 a-wave and b-wave (mean ± SEM, n = 11, 4, and 6 animals for unoperated control, three-gel, and six-gel groups, respectively), and scotopic 10 a-wave and b-wave (mean ± SEM, n = 12, 6, and 4 animals for unoperated control, three-gel, and six-gel groups, respectively). (<b>B</b>) Representative images of H&amp;E staining in rabbit retina: unoperated control, operated control, three-gel, and six-gel. Scale bar, 50 μm. (<b>C</b>) The thickness of ONL, OPL, INL, and IPL in rabbits sacrificed at 2 weeks post operation. No significant changes in retinal thickness were obtained after one-way ANOVA followed by Bonferroni’s post hoc comparisons tests. (mean ± SEM, n = 5, 3, 5, and 5 animals for unoperated control, operated control, three-gels, and six-gel, respectively). CAC ECT, collagen–alginate composite encapsulated cell therapy. ONL, outer nuclear layer. OPL, outer plexiform layer. INL, inner nuclear layer. IPL, inner plexiform layer. GCL, ganglion cell layer.</p>
Full article ">Figure 3
<p>CAC ECT gel may cause stress in the retina, as assessed by gliosis in Müller cells. (<b>A</b>,<b>B</b>) CAC ECT gels were injected into the vitreous cavity, and rabbits were euthanized 2 weeks later. Immunostaining of retinal paraffin was performed to detect microglial cells (Iba-1) (<b>A</b>) and astrocytes (GFAP) (<b>B</b>), with nuclear staining by DAPI (blue). Scale bar, 50 µm. CAC ECT, collagen–alginate composite encapsulated cell therapy.</p>
Full article ">Figure 4
<p>In vivo performance of CAC ECT gel after 2-week implantation. (<b>A</b>) Morphology of gels on fabrication day and retrieved gels after 2-week implantation. Scale bar, 100 μm. (<b>B</b>) Diameter of gel before and after implantation (mean ± SEM, n = 6, 6, and 6 for pre-implantation level, three-gel, and six-gel groups, respectively); one-way ANOVA with Dunnett’s post hoc test demonstrated no significant difference. (<b>C</b>) Diameter of acellular outer region (mean ± SEM, n = 6, 6, and 6 for pre-implantation level, three-gel, and six-gel groups, respectively). One-way ANOVA with Dunnett’s post hoc test demonstrated no significant difference. (<b>D</b>) SEM images showing the microstructure of the CAC ECT gel, focusing on gel surface and core region after 2-week implantation. Top row: A porous IPN network consisting of collagen and alginate assembled on the surface of the CAC gel and a magnified view of the porous network. Bottom row: cells were surrounded by collagen fibrils and CAC matrix and a magnified view of the internal structure (*: alginate, +: collagen, #: cell). CAC, collagen–alginate composite. SEM, scanning electron microscopy. IPN, interpenetrating network. ns, not significant.</p>
Full article ">Figure 5
<p>Retrieved gels were mostly viable at 2 weeks and secreted GDNF in a dose-dependent manner. (<b>A</b>) Morphology (left column) and Live/Dead image (right column) of retrieved CAC ECT gels from the three-gel and six-gel groups after 2 weeks of implantation. Scale bar, 100 μm. (<b>B</b>) Viability of retrieved gels from the three-gel and six-gel groups at 2 weeks assessed by the Live/Dead assay (mean ± SD, n = 6, and 6 for the three-gel and six-gel groups, respectively). Unpaired <span class="html-italic">t</span>-test demonstrated no significant differences. ns, not significant. (<b>C</b>) Vitreous GDNF level measured by ELISA assay (mean ± SEM, n = 5, 3, 5, and 5 for unoperated control, operated control, three-gel group, and six-gel group, respectively). **** <span class="html-italic">p</span> &lt; 0.0001 by one-way ANOVA followed by Bonferroni’s post hoc comparisons tests. GDNF, glial-cell-derived neurotrophic factor. ANOVA, analysis of variance. ELISA, enzyme-linked immunosorbent assay.</p>
Full article ">Figure 6
<p>One week of Dox administration can decrease encapsulated cell viability and terminate gel functionality, without affecting retinal morphology and causing changes in retinal function. (<b>A</b>) Phase-contrast microscopic images of non-Dox-treated gels and Dox-treated gel. Scale bar, 100 μm. (<b>B</b>) Live/Dead image of retrieved gel from the Dox-treated group showing the distribution of viable cells and dying cells. Scale bar, 100 μm. (<b>C</b>) Relative cell viability of non-Dox-treated group and Dox-treated group assessed by MTS assay. (mean ± SD, n = 11, and 6 for Non-Dox and + Dox group). **** <span class="html-italic">p</span> &lt; 0.0001 by unpaired <span class="html-italic">t</span>-test. (<b>D</b>) Accumulated GDNF levels in rabbit vitreous by GDNF ELISA after 1 week of oral Dox administration. (mean ± SEM, n = 5, and 4 for Non-Dox and +Dox group, respectively). *** <span class="html-italic">p</span> &lt; 0.001 unpaired <span class="html-italic">t</span>-test. (<b>E</b>) H&amp;E-stained retinal sections demonstrated no major change in retinal cytoarchitecture. Scale bar, 50 μm. (<b>F</b>) The thickness of various retinal layers including ONL, OPL, INL, IPL, and GCL was measured. (mean ± SEM, n = 5, 4 and 4 for Unoperated control, Non-Dox, and +Dox, respectively). A one-way ANOVA with Bonferroni post hoc test demonstrated non-significant differences. (<b>G</b>) The ERG response indicating the retinal functions of different groups, including Unoperated control, Non-Dox and +Dox. Amplitudes of major wave components quantified included photopic 5 a- and b-wave (mean ± SEM, n = 6, 6 and 4 animals for Unoperated control, non-Dox, and +Dox, respectively), scotopic 0.01 a-wave and b-wave (mean ± SEM, n = 6, 6 and 4 animals for Unoperated control, non-Dox, and +Dox, respectively) and scotopic 10 a-wave and b-wave (mean ± SEM, n = 6, 6 and 4 animals for Unoperated control, non-Dox, and +Dox, respectively). Statistical analysis using one-way ANOVA with Bonferroni post hoc test demonstrated no significant differences. Dox, doxycycline. GDNF, glial-cell-derived neurotrophic factor. SEM, standard deviation of the mean. MTS, dimethylthiazol-carboxymethoxyphenyl-sulfophenyl-tetrazolium. ONL, outer nuclear layer. OPL, outer plexiform layer. INL, inner nuclear layer. IPL, inner plexiform layer. GCL, ganglion cell layer.</p>
Full article ">Figure 7
<p>The therapeutic efficacy of CAC ECT gels in rabbits with retinal degeneration induced by SI. (<b>A</b>) Scotopic a-wave and b-wave electroretinographic responses were measured at 10 cd.s/m<sup>2</sup>, indicating retinal functions in normal rabbit eyes (baseline) and SI-induced rabbit eyes, including no-gel-injected eyes, and eyes injected with 3 and six gels. (mean ± SEM, n = 11, 8, 7, and 6 independent animals for baseline, no gel, three-gel, and six-gel groups). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001 by one-way ANOVA followed by Dunnett’s post-test. (<b>B</b>) Representative image of H&amp;E-stained retinal sections from healthy rabbits (baseline), rabbits receiving SI only, and rabbits with three-gel and six-gel treatment. Scale bar, 50 µm. (<b>C</b>) The thickness of ONL containing the photoreceptor nuclei was significantly higher in rabbits treated with CAC ECT gels when compared with no-gel-treated rabbits after 2-week implantation. (mean ± SEM, n = 6, 8, 7, and 6 for baseline, no gel, three-gel, and six-gel groups). * <span class="html-italic">p</span> &lt; 0.05 by one-way ANOVA followed by Dunnett post <span class="html-italic">t</span>-test. (<b>D</b>) The number of ONL nuclei per mm was significantly higher in rabbits with three-gel and six-gel injection compared with no-gel-treated rabbits after 2-week implantation. (mean ± SEM, n = 6, 8, 7, and 6 for baseline, no gel, three-gel, and six-gel group). **** <span class="html-italic">p</span> &lt; 0.0001 by one-way ANOVA followed by Dunnett post <span class="html-italic">t</span>-test. (<b>E</b>) Representative images showing the distribution of apoptotic cells in the retina detected by TUNEL assay and nuclei were stained by DAPI. Scale bar, 50 µm. (<b>F</b>) The number of TUNEL-positive apoptotic cells in ONL significantly decreased in SI-induced rabbits treated with three-gel and six-gel compared with no-gel-treated rabbits after 2-week implantation. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 by one-way ANOVA followed by Dunnett’s post-test. (<b>G</b>) Live/Dead assay showed that the six-gel-treated and six-gel-treated animals contained mostly living cells (green). Scale bar, 100 μm. (<b>H</b>) Increased vitreous GDNF levels only detected in rabbits with gel treatment. ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001 by one-way ANOVA with Bonferroni post hoc. CAC ECT, collagen–alginate composite encapsulated cell therapy. SI, sodium iodate. ANOVA, analysis of variance. ONL, outer nuclear layer. GDNF, glial-cell-derived neurotrophic factor. TUNEL, terminal deoxynucleotidyl transferase dUTP nick end labeling. DAPI, 4′,6-diamidino-2-phenylindole.</p>
Full article ">Figure 8
<p>Schematic diagram showing the detailed preparation of CAC ECT gels. (<b>A</b>) HEK/293/GDNF/Tet-on pro-Casp8 cells were trypsinized. (<b>B</b>) The HEK/293/GDNF/Tet-on pro-Casp8 cell pellet was mixed with collagen and alginate. (<b>C</b>) An aliquot of 2 μL of the gel mixture was prepared. (<b>D</b>) The 2 μL gel mixture was gently transferred to the glass molds. (<b>E</b>) The glass molds containing the 2 μL gel were collected into a 50 mL Falcon tube. (<b>F</b>) The glass molds were then placed in an incubator at 37 °C for 75 min to initiate collagen gelation. (<b>G</b>) The glass molds were then introduced to a 100 mM CaCl<sub>2</sub> solution followed by a CaCl<sub>2</sub> bath to gelate the alginate. (<b>H</b>) The gels were collected from the molds and cultured in a DMEM-HG medium before intravitreal gel injection.</p>
Full article ">Figure 9
<p>Intravitreal injection of CAC ECT gels on healthy NZW rabbits. (<b>A</b>) The surrounding ocular tissue was first sterilized by diluted betadine. Then, eyelids were spread apart using a speculum. (<b>B</b>) One incision (2 mm from the limbus) into the vitreous was made using a 20-gauge blade followed by an insertion of a pipette tip containing a 2-μL CAC ECT gel into the vitreous.</p>
Full article ">Figure 10
<p>Experimental design illustrating the (<b>A</b>) in vivo safety assessment, (<b>B</b>) in vivo performance evaluation, and (<b>C</b>) in vivo termination for the CAC ECT gel in healthy NZW rabbits and (<b>D</b>) therapeutic efficacy in a sodium-iodate-induced retinal degeneration model in rabbits. CAC ECT, collagen–alginate composite encapsulated cell therapy.</p>
Full article ">
9 pages, 2383 KiB  
Article
An Association between HTRA1 and TGF-β2 in the Vitreous Humor of Patients with Chorioretinal Vascular Diseases
by Yoko Fukushima, Shizuka Takahashi, Machiko Nakamura, Tatsuya Inoue, Yusuke Fujieda, Toshiyuki Sato, Shingo Noguchi, Motokazu Tsujikawa, Hirokazu Sakaguchi and Kohji Nishida
J. Clin. Med. 2024, 13(17), 5073; https://doi.org/10.3390/jcm13175073 (registering DOI) - 27 Aug 2024
Viewed by 465
Abstract
Background: The aim of this paper was to investigate the protein concentrations of high-temperature requirement A 1 (HTRA1) and transforming growth factor-β (TGF-β) in the vitreous humor of patients with chorioretinal vascular diseases. Methods: This study measured protein [...] Read more.
Background: The aim of this paper was to investigate the protein concentrations of high-temperature requirement A 1 (HTRA1) and transforming growth factor-β (TGF-β) in the vitreous humor of patients with chorioretinal vascular diseases. Methods: This study measured protein concentrations of HTRA1, TGF-β13, and vascular endothelial growth factor A (hereinafter called VEGF) in the vitreous humor from seven eyes of patients with chorioretinal vascular diseases (age-related macular degeneration, diabetic macular edema, and retinal vein occlusion) and six control eyes (idiopathic epiretinal membrane and macular hole). We analyzed the mutual relationship among the protein levels. Results: The protein levels of HTRA1 and VEGF were significantly increased in the chorioretinal vascular disease group compared with the control group (1.57 ± 0.79 ×109 mol/mL vs. 0.68 ± 0.79 ×109 mol/mL, p = 0.039; 3447.00 ± 3423.47 pg/mL vs. 35.33 ± 79.01 pg/mL, p = 0.046, respectively). TGF-β2 levels were not significantly different between groups (2222.71 ± 1151.25 pg/mL for the chorioretinal vascular disease group vs. 1918.83 ± 744.01 pg/mL for the control group, p = 0.62). The concentration of HTRA1 was strongly associated with TGF-β2 levels in the vitreous humor, independent of VEGF (r = 0.80, p = 0.0010). Conclusions: We revealed that vitreous HTRA1 was increased in patients with chorioretinal vascular diseases and strongly correlated with TGF-β2. Full article
(This article belongs to the Special Issue An Update on Retinal Diseases: From Diagnosis to Treatment)
Show Figures

Figure 1

Figure 1
<p>The distribution of the individual vitreous concentrations of (<b>A</b>) HTRA1, (<b>B</b>) TGF-<math display="inline"><semantics> <msub> <mi mathvariant="sans-serif">β</mi> <mn>2</mn> </msub> </semantics></math>, (<b>C</b>) TGF-<math display="inline"><semantics> <msub> <mi mathvariant="sans-serif">β</mi> <mn>3</mn> </msub> </semantics></math>, and (<b>D</b>) VEGF by group. Note that HTRA1 levels were measured as <math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>9</mn> </mrow> </msup> </semantics></math> mol/L, and TGF-<math display="inline"><semantics> <msub> <mi mathvariant="sans-serif">β</mi> <mn>2</mn> </msub> </semantics></math>, TGF-<math display="inline"><semantics> <msub> <mi mathvariant="sans-serif">β</mi> <mn>3</mn> </msub> </semantics></math>, and VEGF were measured as pg/mL. <span class="html-italic">p</span>-values are indicated in the plots, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>Scatter plots and linear regression lines of the concentrations of (<b>A</b>) TGF-<math display="inline"><semantics> <msub> <mi mathvariant="sans-serif">β</mi> <mn>2</mn> </msub> </semantics></math> and HTRA1, (<b>B</b>) HTRA1 and VEGF, and (<b>C</b>) VEGF and TGF-<math display="inline"><semantics> <msub> <mi mathvariant="sans-serif">β</mi> <mn>2</mn> </msub> </semantics></math>. The equation, correlation coefficient, and statistical significance are provided in the plots, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Representative images obtained from patients with AMD and DME. The levels of HTRA1 and TGF-<math display="inline"><semantics> <msub> <mi mathvariant="sans-serif">β</mi> <mn>2</mn> </msub> </semantics></math> are elevated in the eyes of patients with more severe hyperpermeability. The HTRA1, TGF-<math display="inline"><semantics> <msub> <mi mathvariant="sans-serif">β</mi> <mn>2</mn> </msub> </semantics></math>, and VEGF concentrations in the eyes are given below the images. FA in AMD cases (<b>A</b>,<b>B</b>) was obtained by SPECTRALIS<sup>®</sup> HRA (Heidelberg Engineering, Heidelberg, Germany), and in DME cases (<b>C</b>,<b>D</b>) by Optos<sup>®</sup> (Optos, Marlborough, MA, USA). OCT was obtained by SPECTRALIS for all disease types.</p>
Full article ">Figure 3 Cont.
<p>Representative images obtained from patients with AMD and DME. The levels of HTRA1 and TGF-<math display="inline"><semantics> <msub> <mi mathvariant="sans-serif">β</mi> <mn>2</mn> </msub> </semantics></math> are elevated in the eyes of patients with more severe hyperpermeability. The HTRA1, TGF-<math display="inline"><semantics> <msub> <mi mathvariant="sans-serif">β</mi> <mn>2</mn> </msub> </semantics></math>, and VEGF concentrations in the eyes are given below the images. FA in AMD cases (<b>A</b>,<b>B</b>) was obtained by SPECTRALIS<sup>®</sup> HRA (Heidelberg Engineering, Heidelberg, Germany), and in DME cases (<b>C</b>,<b>D</b>) by Optos<sup>®</sup> (Optos, Marlborough, MA, USA). OCT was obtained by SPECTRALIS for all disease types.</p>
Full article ">
15 pages, 2437 KiB  
Article
Metformin Alleviates Inflammation and Induces Mitophagy in Human Retinal Pigment Epithelium Cells Suffering from Mitochondrial Damage
by Maija Toppila, Sofia Ranta-aho, Kai Kaarniranta, Maria Hytti and Anu Kauppinen
Cells 2024, 13(17), 1433; https://doi.org/10.3390/cells13171433 - 27 Aug 2024
Viewed by 437
Abstract
Mitochondrial malfunction, excessive production of reactive oxygen species (ROS), deficient autophagy/mitophagy, and chronic inflammation are hallmarks of age-related macular degeneration (AMD). Metformin has been shown to activate mitophagy, alleviate inflammation, and lower the odds of developing AMD. Here, we explored the ability of [...] Read more.
Mitochondrial malfunction, excessive production of reactive oxygen species (ROS), deficient autophagy/mitophagy, and chronic inflammation are hallmarks of age-related macular degeneration (AMD). Metformin has been shown to activate mitophagy, alleviate inflammation, and lower the odds of developing AMD. Here, we explored the ability of metformin to activate mitophagy and alleviate inflammation in retinal pigment epithelium (RPE) cells. Human ARPE-19 cells were pre-treated with metformin for 1 h prior to exposure to antimycin A (10 µM), which induced mitochondrial damage. Cell viability, ROS production, and inflammatory cytokine production were measured, while autophagy/mitophagy proteins were studied using Western blotting and immunocytochemistry. Metformin pre-treatment reduced the levels of proinflammatory cytokines IL-6 and IL-8 to 42% and 65% compared to ARPE-19 cells exposed to antimycin A alone. Metformin reduced the accumulation of the autophagy substrate SQSTM1/p62 (43.9%) and the levels of LC3 I and II (51.6% and 48.6%, respectively) after antimycin A exposure. Metformin also increased the colocalization of LC3 with TOM20 1.5-fold, suggesting active mitophagy. Antimycin A exposure increased the production of mitochondrial ROS (226%), which was reduced by the metformin pre-treatment (84.5%). Collectively, metformin showed anti-inflammatory and antioxidative potential with mitophagy induction in human RPE cells suffering from mitochondrial damage. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Metabolic activity (<b>A</b>) and LDH leakage (<b>B</b>) were measured following a 24 h metformin exposure. Cells were exposed to different metformin concentrations (10, 15, 20, 25, 30, 35, or 40 mM). Absorbance values of the exposed cells were compared to those of untreated control cells, the level of which was set to be 100% (<b>A</b>) or 1 (<b>B</b>). Data were combined from three experiments containing four parallel samples in each group per experiment (n = 12). Results are presented as mean ± standard error of the mean (SEM). *** <span class="html-italic">p</span> ≤ 0.0001, ** <span class="html-italic">p</span> ≤ 0.001, * <span class="html-italic">p</span> ≤ 0.05; ns = not statistically significant (Mann–Whitney U test).</p>
Full article ">Figure 2
<p>The effect of metformin on the levels of IL-6 (<b>A</b>), IL-8 (<b>B</b>), and LDH (<b>C</b>) after RPE cells were exposed to metformin (1 h, Met) and antimycin A (24 h, Aa). Aa was dissolved and diluted in DMSO, and an identical amount of DMSO served as a vehicle control. Data were combined from three experiments containing four parallel samples in each group per experiment (n = 12). Results are presented as mean ± SEM. **** <span class="html-italic">p</span> ≤ 0.00001, *** <span class="html-italic">p</span> ≤ 0.0001, ** <span class="html-italic">p</span> ≤ 0.001; ns = not statistically significant (Mann–Whitney U test).</p>
Full article ">Figure 3
<p>Potential mechanism behind the anti-inflammatory effects of metformin in cultured ARPE-19 cells. The levels of NF-κB (<b>A</b>) or phosphorylated ERK1/2 (<b>B</b>), p38 (<b>C</b>), and SAPK-JNK (<b>D</b>) after 1 h metformin (Met, 15 mM) and 2 h antimycin A (Aa, 10 µM) treatment. Aa was dissolved and diluted in DMSO, and an identical amount of DMSO served as a vehicle control. Data were combined from three experiments containing 2 (<b>A</b>) or 3 (<b>B</b>–<b>D</b>) parallel samples in each group per experiment (<b>A</b>, n = 6; <b>B</b>–<b>D</b>, n = 9). Results are presented as mean ± SEM. ** <span class="html-italic">p</span> ≤ 0.001; ns = not statistically significant (Mann–Whitney U test).</p>
Full article ">Figure 4
<p>Mitochondrial membrane potential and mitochondrial quality were studied using the MitoTracker Red staining under a confocal microscope at 65× magnification after 1 h metformin (Met, 15 mM) and 24 h antimycin A (Aa, 10 µM) treatment (<b>A</b>). The relative brightness of mitochondria was calculated using the ImageJ software (<b>B</b>). Aa was dissolved and diluted in DMSO, and an identical amount of DMSO served as a vehicle control. Data were combined from three experiments containing 1 or 3 parallel samples in each group per experiment. Three cells were calculated per frame (n = 21). The scale bar equals to 10 µm. Results are presented as mean ± SEM. **** <span class="html-italic">p</span> ≤ 0.00001, *** <span class="html-italic">p</span> ≤ 0.0001, * <span class="html-italic">p</span> ≤ 0.05 (Mann–Whitney U test).</p>
Full article ">Figure 5
<p>The levels of autophagy markers afterARPE-19 cells were exposed to metformin for 1 h (Met, 15 mM) and to antimycin A for 6 h (Aa, 10 µM). p62/SQSTM1 (<b>A</b>), LC3-I (<b>B</b>), and LC3-II (<b>C</b>) were measured using the Western blot technique, with a representative image being shown (<b>D</b>). Aa was dissolved and diluted in DMSO, and an identical amount of DMSO served as a vehicle control. Data were combined from three experiments containing 2 parallel samples in each group per experiment (n = 6). Results are presented as mean ± SEM. ** <span class="html-italic">p</span> ≤ 0.001, * <span class="html-italic">p</span> ≤ 0.05; ns = not statistically significant (Mann–Whitney U test).</p>
Full article ">Figure 6
<p>The messenger RNA levels of SQSTM1 (<b>A</b>), MAP1LC3B (<b>B</b>), and MTOR (<b>C</b>) were measured after 1 h metformin (Met, 15 mM) and 3 h antimycin A (Aa, 10 µM) treatment. Aa was dissolved and diluted in DMSO, with an identical amount of DMSO serving as a vehicle control. Data were combined from three experiments containing 3 parallel samples in each group per experiment (n = 9). Results are presented as mean ± SEM. *** <span class="html-italic">p</span> ≤ 0.0001, * <span class="html-italic">p</span> ≤ 0.05; ns = not statistically significant (Mann–Whitney U test).</p>
Full article ">Figure 7
<p>The colocalization of LC3 (red) and TOM20 (green) using Manders colocalization coefficient (<b>A</b>) was studied from images taken under a confocal microscope at 65× magnification (<b>B</b>) after ARPE-19 cells were exposed to metformin (Met, 15 mM) for 1 h and to antimycin A (Aa, 10 µM) for 24 h. Aa was dissolved and diluted in DMSO, with an identical amount of DMSO serving as a vehicle control. Data were combined from 4 experiments containing 3 parallel samples in each group per experiment. Three cells were calculated per frame (n = 21). The scale bar equals to 10 µm. Results are presented as mean ± SEM. * <span class="html-italic">p</span> ≤ 0.05; ns = not statistically significant (Mann–Whitney U test).</p>
Full article ">Figure 8
<p>The levels of total cellular ROS (<b>A</b>) and mitochondrial ROS (<b>B</b>) after ARPE-19 cells were exposed to metformin (Met, 15 mM) for 1 h and to antimycin A (Aa, 10 µM) for 1 h (<b>A</b>), 2 h (<b>B</b>), or 3 h (<b>C</b>). Aa was dissolved and diluted in DMSO, with an identical amount of DMSO serving as a vehicle control. Data were combined from three experiments containing 6 parallel samples in each group per experiment (n = 18, (<b>A</b>)) or 4 parallel samples in each group per experiment (n = 12, (<b>B</b>)). Results are presented as mean ± SEM. **** <span class="html-italic">p</span> ≤ 0.00001, *** <span class="html-italic">p</span> ≤ 0.0001, ** <span class="html-italic">p</span> ≤ 0.001; ns = not statistically significant (Mann–Whitney U test).</p>
Full article ">
11 pages, 1464 KiB  
Review
Transorbital Alternating Current Stimulation in Glaucoma: State of the Art from Neurophysiological Bases to Clinical Practice
by Giuseppe Granata, Sharon Delicati and Benedetto Falsini
Optics 2024, 5(3), 353-363; https://doi.org/10.3390/opt5030026 - 27 Aug 2024
Viewed by 485
Abstract
Recovery after visual loss is a key goal of neuroscience and treatments able to improve visual function are still largely lacking. Glaucoma, one of the leading causes of visual disability in the world, is usually associated with elevated intraocular pressure (IOP), but a [...] Read more.
Recovery after visual loss is a key goal of neuroscience and treatments able to improve visual function are still largely lacking. Glaucoma, one of the leading causes of visual disability in the world, is usually associated with elevated intraocular pressure (IOP), but a subset of “normal tension glaucoma” patients develop damage without ever manifesting high IOP. Sometimes, even in patients with good control of IOP, retinal ganglion cell degeneration can progress to forward blindness. Moreover, usually the damage already caused by the disease remains. These considerations underline the need to find new, effective treatments and solutions to add to the standard ones. In this paper, we expose the most important data supporting the use of alternating current stimulation, including the theoretical bases of this approach, in glaucoma. Full article
Show Figures

Figure 1

Figure 1
<p>The picture summarizes the model proposed for glaucoma progression. The Figure, modified from Porciatti et al. 2012 [<a href="#B4-optics-05-00026" class="html-bibr">4</a>], shows the mismatch (asterisk) between structural and functional impairment. After a prolonged stressful condition (the increase of IOP in the case of glaucoma), the autoregulatory mechanisms to sustain normal RGC function fail. The dysfunctional RGCs (straight face) can come back to a normal condition (smiley face) under less stressful conditions (or during rehabilitating treatments, as for electrical stimulation) or otherwise proceed to death (frowny face) if the stress continues to act. X-axis: time course of the disease. Y-axis: percentage of residual function and structure. Continuous line: residual structure. Dashed line: residual function. Asterisk: mismatch.</p>
Full article ">Figure 2
<p>The Figure shows the visual field improvement of patients with retinitis pigmentosa before (<b>left</b>) and after (<b>right</b>) 10 consecutive days of trACS.</p>
Full article ">Figure 3
<p>The Figure summarizes the different effects of electrical stimulation on the visual pathway (taking together transorbital, transcorneal, and transpalpebral montage). On the left are reported the effects of stimulation at the eye and optic nerve level. On the right are reported the effects at the brain level. The result of the stimulation is probably due to a combination of these mechanisms in different proportions according to the pathology, the severity of the clinical picture, and the personal predisposition. The clinical effect can be the improvement of visual function and, for chronic disease, the slowing of disease progression. The mouse symbol refers to data obtained only from animal experiments.</p>
Full article ">
13 pages, 5853 KiB  
Article
HSPB4/CRYAA Protect Photoreceptors during Retinal Detachment in Part through FAIM2 Regulation
by Cagri G. Besirli, Madhu Nath, Jingyu Yao, Mercy Pawar, Angela M. Myers, David Zacks and Patrice E. Fort
Neurol. Int. 2024, 16(5), 905-917; https://doi.org/10.3390/neurolint16050068 - 26 Aug 2024
Viewed by 264
Abstract
Our previous study discussed crystallin family induction in an experimental rat model of retinal detachment. Therefore, we attempted to evaluate the role of α-crystallin in photoreceptor survival in an experimental model of retinal detachment, as well as its association with the intrinsically neuroprotective [...] Read more.
Our previous study discussed crystallin family induction in an experimental rat model of retinal detachment. Therefore, we attempted to evaluate the role of α-crystallin in photoreceptor survival in an experimental model of retinal detachment, as well as its association with the intrinsically neuroprotective protein Fas-apoptotic inhibitory molecule 2 (FAIM2). Separation of retina and RPE was induced in rat and mouse eyes by subretinal injection of hyaluronic acid. Retinas were subsequently analyzed for the presence αA-crystallin (HSPB4) and αB-crystallin (HSPB5) proteins using immunohistochemistry and immunoblotting. Photoreceptor death was analyzed using terminal deoxynucleotidyl transferase-mediated deoxyuridine triphosphate nick end labeling (TUNEL) staining and cell counts. The 661W cells subjected to FasL were used as a cell model of photoreceptor degeneration to assess the mechanisms of the protective effect of αA-crystallin and its dependence on its phosphorylation on T148. We further evaluated the interaction between FAIM2 and αA-crystallin using a co-immunoprecipitation assay. Our results showed that α-crystallin protein levels were rapidly induced in response to retinal detachment, with αA-crystallin playing a particularly important role in protecting photoreceptors during retinal detachment. Our data also show that the photoreceptor intrinsically neuroprotective protein FAIM2 is induced and interacts with α-crystallins following retinal detachment. Mechanistically, our work also demonstrated that the phosphorylation of αA-crystallin is important for the interaction of αA-crystallin with FAIM2 and their neuroprotective effect. Thus, αA-crystallin is involved in the regulation of photoreceptor survival during retinal detachment, playing a key role in the stabilization of FAIM2, serving as an important modulator of photoreceptor cell survival under chronic stress conditions. Full article
(This article belongs to the Collection Advances in Neurodegenerative Diseases)
Show Figures

Figure 1

Figure 1
<p>Retinal detachment is associated with differential expression of αA- and αB-crystallin. αA- (<b>A</b>) and αB-crystallin (<b>B</b>) transcripts levels were analyzed by quantitative PCR in detached and contralateral attached retinas at 1, 3, and 7 days post-detachment (<span class="html-italic">n</span> = 6/gp/time point). αA- and αB-crystallin protein levels (<b>C</b>) were then analyzed by immunoblotting in similar tissues, and a representative image is shown, along with a graphic representation of the corresponding quantification (the arrow denotes the specific band of interest for CryAB). While both αA- and αB-crystallin were regulated post-transcriptionally from within 1 day of retinal detachment, only αA-crystallin transcriptional expression was increased by retinal detachment transiently at the 3-day time point. * significantly different from attached (* <span class="html-italic">p</span> ≤ 0.05 or ** <span class="html-italic">p</span> ≤ 0.01). CryAA: αA- crystallin; CryAB: αB-crystallin; GAPDH: glyceraldehyde 3-phosphate dehydrogenase; A: attached; D1: day 1; D2: day 2; D7: day 7.</p>
Full article ">Figure 2
<p>Loss of αA-crystallin affects αB-crystallin levels in response to retinal detachment, but the opposite is not true. αA- and αB-crystallin protein levels were analyzed by immunoblotting in retinas from wild-type (WT), αA-crystallin KO, and αB-crystallin KO mice at 1, 3, and 7 days post-detachment (<span class="html-italic">n</span> = 6/gp/time point). A representative image is shown for the WT (<b>A</b>), αA-crystallin knockout (<b>B</b>), and αB-crystallin knockout (<b>C</b>) mice, along with a graphic representation of the corresponding quantification. Both αA- and αB-crystallin protein levels were increased in WT mice. αA-crystallin knockout displayed no significant increase in αB-crystallin expression, while increased basal expression of αA-crystallin, as well as a detachment-induced increase, was observed in αB-crystallin knockout mice. * significantly different from attached (* <span class="html-italic">p</span> ≤ 0.05 or ** <span class="html-italic">p</span> ≤ 0.01). CryAA: αA- crystallin; CryAB: αB-crystallin; WT: wild type; GAPDH: glyceraldehyde 3-phosphate dehydrogenase; A: attached; D: detached; D1: day 1; D2: day 2; D7: day 7.</p>
Full article ">Figure 3
<p>Loss of αA-crystallin, but not αB-crystallin, leads to enhanced detachment-induced photoreceptor cell death and retinal thinning. Retinal cell death (<b>A</b>) and retinal thickness (<b>B</b>) were measured on retinal cross-sections of WT mice, αA-crystallin KO mice (Ko-CryAA), αB-crystallin KO mice (Ko-CryAB), and αA/αB-crystallin double-knockout mice (Ko-CryAA/AB) at 3 days and 2 months post-retinal detachment (<span class="html-italic">n</span> = 9/gp/time point). The number of TUNEL-positive cells was significantly increased in the detached retinas of the Ko-CryAA and Ko-CryAA/AB mice but not in those of the Ko-CryAB mice. Representative images of TUNEL staining used for cell counts are shown in (<b>C</b>). Consistent with the observed levels of cell death, retinal thickness, especially that of ONL, was reduced in the Ko-CryAA and Ko-CryAA/AB mice but not in the Ko-CryAB mice. * significantly different from WT mice (* <span class="html-italic">p</span> ≤ 0.05 or ** <span class="html-italic">p</span> ≤ 0.01). CryAA: αA-crystallin; CryAB: αB-crystallin; WT: wild type). OS: outer segments; ONL: outer nuclear layer; INL: inner nuclear layer; GCL: ganglion cell layer.</p>
Full article ">Figure 4
<p>αA-crystallin and αB-crystallin are expressed in photoreceptor inner and outer segments following retinal detachment. Immunohistochemistry for αA- (<b>A</b>,<b>B</b>) and αB-crystallins (<b>E</b>,<b>F</b>) was performed on cross-sections from attached (<b>A</b>,<b>E</b>) and detached (<b>B</b>,<b>F</b>) retinas of WT mice, respectively (<span class="html-italic">n</span> = 3). Consistent with the protein expression data, the immunostaining results show increased immunoreactivity of both α-crystallins, specifically in photoreceptors, in response to detachment ((<b>C</b>) attached vs. (<b>D</b>) detached insert for αA-crystallin corresponding to white rectangles in A and B respectively). OS: outer segments; IS: inner segments; ONL: outer nuclear layer; INL: inner nuclear layer.</p>
Full article ">Figure 5
<p>FAIM2 expression in stressed photoreceptors requires the presence of α-crystallins. Immunohistochemistry for FAIM2 was performed on cross-sections from attached (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>) and detached (<b>B</b>,<b>D</b>,<b>F</b>,<b>H</b>) retinas of WT (<b>A</b>,<b>B</b>), αA-crystallin KO (<b>C</b>,<b>D</b>), αB-crystallin KO (<b>E</b>,<b>F</b>), and αA/αB-crystallin double-knockout (<b>G</b>,<b>H</b>) mice (<span class="html-italic">n</span> = 3/gp). As shown in the inserts (<b>I</b>–<b>L</b> corresponding to white rectangles in <b>B</b>, <b>D</b>, <b>F</b> and <b>H</b>), FAIM2 immunoreactivity was significantly increased in photoreceptor segments following detachment only in the WT mice. OS: outer segments; IS: inner segments; ONL: outer nuclear layer; INL: inner nuclear layer.</p>
Full article ">Figure 6
<p>FAIM2 interacts strongly with αA-crystallin. Representative immunoblots for FAIM2 and αA- (<b>left panel</b>) and αB-crystallin (<b>right panel</b>) following co-immunoprecipitation using tissue extracts from attached and detached rat retinas (<span class="html-italic">n</span> = 6). Increased levels of αA-, but not αB-crystallin, were pulled-down with FAIM2 in detached retinas compared to attached retinas.</p>
Full article ">Figure 7
<p>FAIM2’s interaction with αA-crystallin is T148 phosphorylation-dependent. (<b>A</b>) Representative images of 661W cells before and after 8 h of Fas pathway activation (<span class="html-italic">n</span> = 3 independent experiments with three replicates each). Higher-magnification (insert corresponding to the white rectangles on 8 h full fields) images are provided to better appreciate the larger number of pyknotic cells (white arrows denote some examples) in the empty vector- and T148A-CryAA-overexpressing cells compared to the wild-type and T148D-CryAA-overexpressing 661W cells. (<b>B</b>) Representative immunoblots (membranes were cut prior to hybridization for improved signal/noise) for FAIM2 and αA-crystallin following co-immunoprecipitation using cell lysates from Fas-activated photoreceptors overexpressing either the wild-type, phosphomimetic (T148D) or non-phosphorylatable (T148A) form of αA-crystallin. Increased levels of the phosphomimetic form (T148D) were pulled-down with FAIM2 in comparison to the non-phosphorylatable (T148A) form of αA-crystallin. (IP: immunoprecipitation; IB: immunoblot; EV: empty vector; WT: wild type; scale bars are 25 μm (large field) and 10 μm (insert).).</p>
Full article ">
10 pages, 2979 KiB  
Article
Rotenone-Induced Optic Nerve Damage and Retinal Ganglion Cell Loss in Rats
by Yasuko Yamamoto, Takazumi Taniguchi and Atsushi Shimazaki
Biomolecules 2024, 14(9), 1047; https://doi.org/10.3390/biom14091047 - 23 Aug 2024
Viewed by 419
Abstract
Rotenone is a mitochondrial complex I inhibitor that causes retinal degeneration. A study of a rat model of rotenone-induced retinal degeneration suggested that this model is caused by indirect postsynaptic N-methyl-D-aspartate (NMDA) stimulation triggered by oxidative stress-mediated presynaptic intracellular calcium signaling. To elucidate [...] Read more.
Rotenone is a mitochondrial complex I inhibitor that causes retinal degeneration. A study of a rat model of rotenone-induced retinal degeneration suggested that this model is caused by indirect postsynaptic N-methyl-D-aspartate (NMDA) stimulation triggered by oxidative stress-mediated presynaptic intracellular calcium signaling. To elucidate the mechanisms by which rotenone causes axonal degeneration, we investigated morphological changes in optic nerves and the change in retinal ganglion cell (RGC) number in rats. Optic nerves and retinas were collected 3 and 7 days after the intravitreal injection of rotenone. The cross-sections of the optic nerves were subjected to a morphological analysis with axon quantification. The axons and somas of RGCs were analyzed immunohistochemically in retinal flatmounts. In the optic nerve, rotenone induced axonal swelling and degeneration with the incidence of reactive gliosis. Rotenone also significantly reduced axon numbers in the optic nerve. Furthermore, rotenone caused axonal thinning, fragmentation, and beading in RGCs on flatmounts and decreased the number of RGC soma. In conclusion, the intravitreal injection of rotenone in rats induced morphological abnormities with a reduced number of optic nerve axons and RGC axons when the RGC somas were degenerated. These findings help elucidate the pathogenesis of optic neuropathy induced by mitochondrial dysfunction. Full article
(This article belongs to the Special Issue Retinal Diseases: Molecular Mechanisms and Therapies)
Show Figures

Figure 1

Figure 1
<p>Axon swelling, axon degeneration, and reactive gliosis in the optic nerves after rotenone injection. After the intravitreal injection of rotenone, the axons in the optic nerves were morphologically evaluated on day 3 and day 7. The optic nerves were counter-stained with toluidine blue. The entire cross-section was imaged using a 100× oil immersion objective lens. The yellow arrows indicate axon swelling. The white arrows indicate degenerative axons. The red asterisk indicates reactive gliosis. Dimethyl sulfoxide was used as the vehicle control. The scale bar corresponds to 50 µm.</p>
Full article ">Figure 2
<p>Quantification of axon numbers in the optic nerve after the rotenone injection. The axon numbers in the optic nerves were assessed using the AxoNet plugin from ImageJ on day 3 and day 7 after the rotenone injection. Each value represents the mean ± standard error of the mean (SEM) of three eyes. Three rats were used. ** <span class="html-italic">p</span> &lt; 0.01, compared to the vehicle (Welch’s <span class="html-italic">t</span>-test). NS: not significant.</p>
Full article ">Figure 3
<p>Rotenone-induced loss of axonal density and the axonal beading of RGC. The TUJ1 (axonal maker) staining of retinal flatmounts was conducted for the assessment of axon density and axonal beading following the rotenone injection at day 3 and day 7 (20× magnification) (<b>a</b>). The blue rectangle delineates the area’s higher magnification in panel a. The white arrows indicate axonal beading (<b>b</b>). The scale bar corresponds to 100 µm.</p>
Full article ">Figure 4
<p>Immunohistochemistry of RGCs by RNA-binding protein with multiple splicing (RBPMS) and TUJ1 antibodies on retinal flatmounts. The double staining of RBPMS (RGC maker) and TUJ1 of RGCs was performed on retinal flatmounts following the rotenone injection on day 3 and day 7. Red represents RBPMS; green represents TUJ1 (20× magnification). The scale bar corresponds to 100 µm.</p>
Full article ">Figure 5
<p>Quantification of RGCs after the rotenone injection. The RBPMS-positive cells were identified using an upright epifluorescence microscope. The CellProfiler software was used for the quantification of RGC numbers. Each value represents the mean ± SEM of 4–5 eyes. Four rats were used. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01, compared to the vehicle (Welch’s <span class="html-italic">t</span>-test).</p>
Full article ">
11 pages, 453 KiB  
Review
NAD+ and Niacin Supplementation as Possible Treatments for Glaucoma and Age-Related Macular Degeneration: A Narrative Review
by Mohamed R. Gemae, Mario D. Bassi, Patrick Wang, Eric K. Chin and David R.P. Almeida
Nutrients 2024, 16(16), 2795; https://doi.org/10.3390/nu16162795 - 21 Aug 2024
Viewed by 938
Abstract
Glaucoma and age-related macular degeneration (AMD) are progressive retinal diseases characterized by increased oxidative stress, inflammation, and mitochondrial dysfunction. This review investigates the potential therapeutic benefits of NAD+ and niacin supplementation in managing glaucoma and AMD. A literature search was conducted encompassing keywords [...] Read more.
Glaucoma and age-related macular degeneration (AMD) are progressive retinal diseases characterized by increased oxidative stress, inflammation, and mitochondrial dysfunction. This review investigates the potential therapeutic benefits of NAD+ and niacin supplementation in managing glaucoma and AMD. A literature search was conducted encompassing keywords such as “niacin”, “NAD”, “glaucoma”, “AMD”, and “therapeutics”. NAD+ depletion is associated with increased oxidative stress and mitochondrial dysfunction in glaucoma and AMD. Niacin, a precursor to NAD+, has shown promise in replenishing NAD+ levels, improving choroidal blood flow, and reducing oxidative damage. Animal studies in glaucoma models indicate that nicotinamide (NAM) supplementation preserves RGC density and function. Large-scale population-based studies indicate an inverse correlation between niacin intake and glaucoma prevalence, suggesting a preventative role. Randomized controlled trials assessing niacin supplementation showed significant improvements in visual field sensitivity and inner retinal function, with a dose-dependent relationship. In AMD, nicotinamide supplementation may improve rod cell function and protect against oxidative stress-induced damage. Cross-sectional studies reveal that individuals with AMD have a lower dietary intake of niacin. Further studies suggest niacin’s role in improving choroidal blood flow and dilating retinal arterioles, potentially mitigating ischemic damage and oxidative stress in AMD. Beyond current management strategies, NAD+ and niacin supplementation may offer novel therapeutic avenues for glaucoma and AMD. Further research is warranted to elucidate their efficacy and safety in clinical settings. Full article
(This article belongs to the Special Issue Diet and Supplements in the Prevention and Treatment of Eye Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The mitochondrial states of elevated superoxide production. The reduced coenzyme Q pool and reduced NAD+ lead to increased O<sub>2−</sub> production.</p>
Full article ">
13 pages, 845 KiB  
Article
Measuring Geographic Atrophy Area Using Column-Based Machine Learning Software on Spectral-Domain Optical Coherence Tomography versus Fundus Auto Fluorescence
by Or Shmueli, Adi Szeskin, Ilan Benhamou, Leo Joskowicz, Yahel Shwartz and Jaime Levy
Bioengineering 2024, 11(8), 849; https://doi.org/10.3390/bioengineering11080849 - 19 Aug 2024
Viewed by 483
Abstract
Background: The purpose of this study was to compare geographic atrophy (GA) area semi-automatic measurement using fundus autofluorescence (FAF) versus optical coherence tomography (OCT) annotation with the cRORA (complete retinal pigment epithelium and outer retinal atrophy) criteria. Methods: GA findings on FAF and [...] Read more.
Background: The purpose of this study was to compare geographic atrophy (GA) area semi-automatic measurement using fundus autofluorescence (FAF) versus optical coherence tomography (OCT) annotation with the cRORA (complete retinal pigment epithelium and outer retinal atrophy) criteria. Methods: GA findings on FAF and OCT were semi-automatically annotated at a single time point in 36 pairs of FAF and OCT scans obtained from 36 eyes in 24 patients with dry age-related macular degeneration (AMD). The GA area, focality, perimeter, circularity, minimum and maximum Feret diameter, and minimum distance from the center were compared between FAF and OCT annotations. Results: The total GA area measured on OCT was 4.74 ± 3.80 mm2. In contrast, the total GA measured on FAF was 13.47 ± 8.64 mm2 (p < 0.0001), with a mean difference of 8.72 ± 6.35 mm2. Multivariate regression analysis revealed a significant correlation between the difference in area between OCT and FAF and the total baseline lesion perimeter and maximal lesion diameter measured on OCT (adjusted r2: 0.52; p < 0.0001) and the total baseline lesion area measured on FAF (adjusted r2: 0.83; p < 0.0001). Conclusions: We report that the GA area measured on FAF differs significantly from the GA area measured on OCT. Further research is warranted in order to determine the clinical relevance of these findings. Full article
(This article belongs to the Special Issue Artificial Intelligence Applications in Ophthalmology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Distribution of GA area annotated using OCT (<b>a</b>) and FAF (<b>b</b>). (<b>a</b>) Distribution of the number of eyes plotted against GA area (in mm<sup>2</sup>) measured using OCT. The solid line represents a Gaussian fit of the data. (<b>b</b>) Distribution of the number of eyes plotted against GA area (in mm<sup>2</sup>) measured using FAF. The solid line represents a Gaussian fit of the data. GA = Geographic atrophy; OCT = Optical coherence tomography; FAF = Fundus auto-fluorescence.</p>
Full article ">Figure 2
<p>An index case illustrating the difference in GA area between OCT and FAF. (<b>a</b>) The GA annotation on an OCT B-scan using the cRORA criteria (in cyan) was projected on the corresponding IR image, showing a measured GA area of 11 mm<sup>2</sup>. (<b>b</b>) The GA annotation on FAF (in red), with a measured GA area of 26 mm<sup>2</sup>. (<b>c</b>) Both the OCT annotation from (<b>a</b>) (cyan) and the FAF annotation from (<b>b</b>) (red outline) were projected on the FAF image. (<b>d</b>) The B-scan on the right corresponds to the hyper-reflective area on the green line in the IR image on the left (the areas between the two yellow arrows in the IR image and B-scan match). Right: The cRORA evident in the OCT B-scan (indicated by the blue columns) corresponds only partially to the hypo-autofluorescent areas seen in the level of the B-scan section on FAF in (<b>c</b>). GA = Geographic atrophy; OCT = Optical coherence tomography; FAF = Fundus auto-fluorescence; cRORA = Complete retinal pigment epithelium and outer retinal atrophy.</p>
Full article ">
Back to TopTop