[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
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
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
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
remove_circle_outline
remove_circle_outline

Search Results (4,355)

Search Parameters:
Keywords = ocular

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 8796 KiB  
Article
Scattering Improves Temporal Resolution of Vision: A Pilot Study on Brain Activity
by Francisco J. Ávila
Photonics 2025, 12(1), 23; https://doi.org/10.3390/photonics12010023 (registering DOI) - 30 Dec 2024
Abstract
Temporal vision is a vital aspect of human perception, encompassing the ability to detect changes in light and motion over time. Optical scattering, or straylight, influences temporal visual acuity and the critical flicker fusion (CFF) threshold, with potential implications for cognitive visual processing. [...] Read more.
Temporal vision is a vital aspect of human perception, encompassing the ability to detect changes in light and motion over time. Optical scattering, or straylight, influences temporal visual acuity and the critical flicker fusion (CFF) threshold, with potential implications for cognitive visual processing. This study investigates how scattering affects CFF using an Arduino-based psychophysical device and electroencephalogram (EEG) recordings to analyze brain activity during CFF tasks under scattering-induced effects. A cohort of 30 participants was tested under conditions of induced scattering to determine its effect on temporal vision. Findings indicate a significant enhancement in temporal resolution under scattering conditions, suggesting that scattering may modulate the temporal aspects of visual perception, potentially by altering neural activity at the temporal and frontal brain lobes. A compensation mechanism is proposed to explain neural adaptations to scattering based on reduced electrical activity in the visual cortex and increased wave oscillations in the temporal lobe. Finally, the combination of the Arduino-based flicker visual stimulator and EEG revealed the excitatory/inhibitory stimulation capabilities of the high-frequency beta oscillation based on the alternation of an achromatic and a chromatic stimulus displayed in the CFF. Full article
(This article belongs to the Special Issue New Technologies for Human Visual Function Assessment)
Show Figures

Figure 1

Figure 1
<p>Electrode position configured for the study. ch#1: O1; ch#2: O2; ch#3: T5; ch#4: T6; ch#5: F7; ch#6: F8; ch#7: Fp1 and ch#8: Fp2. GND and REF (gray and green circles) correspond to ground signal and reference electrodes, respectively.</p>
Full article ">Figure 2
<p>CFF values as a function of chromatic red (<b>a</b>), green (<b>b</b>), blue (<b>c</b>), and achromatic (<b>d</b>) stimuli for normal vision and after inducing ocular scattering. An asterisk indicates a statistical difference with <span class="html-italic">p</span> value lower than 0.05.</p>
Full article ">Figure 3
<p>Raw EEG signals for natural viewing conditions (Control) and each electrode location for a total recording of 1 s acquired from one volunteer. The upper and lower dashed lines represent ± 2 mV amplitude levels.</p>
Full article ">Figure 4
<p>EEGs filtered for delta, theta, alpha, beta, and gamma for a steady achromatic visual stimulus (control) and for flicker visual stimuli at the CFF for red, green, blue, and achromatic (white) colors. The recording time corresponds to 1 s. The amplitude signal is given in millivolts.</p>
Full article ">Figure 5
<p>EEGs filtered for delta, theta, alpha, beta, and gamma for a steady achromatic visual stimulus affected by scattering (Scattering) and for scattering-induced flicker visual stimuli at the CFF for red, green, blue, and white (achromatic) colors. The recording time corresponds to 1 s. The amplitude signal is given in millivolts.</p>
Full article ">Figure 6
<p>Average band power (ABP) for the 8-channel activity for control (blue bars), chromatic CFF flicker stimuli ((<b>a</b>–<b>c</b>) red, green, and blue bars), and achromatic ((<b>d</b>) gray bars) as a function of the filtered wave band. Asterisks indicate those wave bands for which significant differences were found between control and CFF flicker stimuli with significance level <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Topographic brain activity maps filtered by wave band corresponding to the chromatic flicker stimuli (green and blue CFF), for which a statistical change in ABP with respect the control measurements were found. The scale color bar is in millivolts.</p>
Full article ">Figure 8
<p>Average band power (ABP) for the 8-channel activity for control (<b>a</b>), chromatic CFF flicker stimuli ((<b>c</b>–<b>e</b>) red, green, and blue bars), and achromatic ((<b>b</b>) gray bars) as a function of the filtered wave band. Asterisks indicate those wave bands for which significant differences were found between control and CFF flicker.</p>
Full article ">Figure 9
<p>Topographic maps of electrical brain activity as a function of the visual flicker condition. The scale bar is expressed in millivolts.</p>
Full article ">Figure 10
<p>Energy density changes (%) for each electrode channel as a function of the visual stimuli. Blue bars correspond to scattering-free retinal flickering, and red bars the results under scattering effects.</p>
Full article ">Figure 11
<p>Morlet Wavelet for control (<b>a</b>), pure scattering (<b>c</b>), blue CFF stimulus (<b>b</b>), and scattering-induced blue CFF stimulus (<b>d</b>) and transforms and their corresponding power spectral density (PSD) values [(<b>e</b>,<b>f</b>)] for channel O2.</p>
Full article ">Figure 12
<p>Morlet Wavelet for control (<b>a</b>), pure scattering (<b>c</b>), blue CFF stimulus (<b>b</b>), and scattering-induced blue CFF stimulus (<b>d</b>) and transforms and their corresponding power spectral density (PSD) values [(<b>e</b>,<b>f</b>)] for channel T6.</p>
Full article ">Figure 13
<p>Morlet Wavelet for control (<b>a</b>), pure scattering (<b>c</b>), blue CFF stimulus (<b>b</b>), and scattering-induced blue CFF stimulus (<b>d</b>) and transforms and their corresponding power spectral density (PSD) values [(<b>e</b>,<b>f</b>)] for channel F8.</p>
Full article ">Figure 14
<p>Morlet Wavelet for control (<b>a</b>), pure scattering (<b>c</b>), blue CFF stimulus (<b>b</b>), and scattering-induced blue CFF stimulus (<b>d</b>) and transforms and their corresponding power spectral density (PSD) values [(<b>e</b>,<b>f</b>)] for channel Fp2.</p>
Full article ">
13 pages, 1868 KiB  
Case Report
Postpartum Exudation of Idiopathic Quiescent Macular Neovascularization: A Narrative Review with a Related Case Report
by Livio Vitiello, Maddalena De Bernardo, Ilaria De Pascale, Giulio Salerno, Alfonso Pellegrino and Nicola Rosa
Life 2025, 15(1), 31; https://doi.org/10.3390/life15010031 (registering DOI) - 30 Dec 2024
Abstract
The abnormal growth of irregular new blood vessels into the subretinal or intraretinal space is known as macular neovascularization (MNV). People over 50 are often affected by this disorder, which is typically brought on by age-related macular degeneration. In addition, MNV can be [...] Read more.
The abnormal growth of irregular new blood vessels into the subretinal or intraretinal space is known as macular neovascularization (MNV). People over 50 are often affected by this disorder, which is typically brought on by age-related macular degeneration. In addition, MNV can be found in people under 50 years of age, who may present primary ophthalmic diseases such as pathological myopia, angioid streaks, traumatic choroidal rupture, or suspected ocular histoplasmosis syndrome. However, it is important to consider a specific set of young individuals who may develop MNV even in the absence of pathological myopia or other identifiable inflammatory, peripapillary, post-traumatic, or degenerative fundus abnormalities. This latter condition is classified as idiopathic MNV. After a literature review focused on young patients affected by one of these two clinical entities, we report the case of a Caucasian young woman suffering for four years from an idiopathic and quiescent MNV that started exuding after childbirth, probably due to the induction with oxytocin, and was treated with intravitreal Aflibercept 2 mg injections. Full article
(This article belongs to the Special Issue Eye Diseases: Diagnosis and Treatment, 3rd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) OCT color fundus image of the quiescent MNV, showing the presence of slightly hyperpigmented flat lesion, with indefinite edges, in the macular region. (<b>b</b>) B-scan OCT image of the same retinal lesion, characterized by the visualization of two hyperreflective layers (double-layer sign): retinal pigment epithelium (red arrows) and the Bruch membrane (white arrows) with a major axis on the horizontal plane, with the presence of reflective material inside (yellow asterisk), in the absence of subretinal fluid.</p>
Full article ">Figure 2
<p>Top image: B-scan OCT of the quiescent MNV, showing the “double-layer sign” and the presence of hyperreflective material under the retinal pigment epithelium, in the absence of subretinal fluid. Bottom images: 3 × 3 mm OCT angiogram of the superficial (<b>a</b>) and deep (<b>b</b>) capillary plexus. The OCT angiogram of the choriocapillaris (<b>c</b>) displays a well-defined neovascular network under the retinal pigment epithelium (red arrows).</p>
Full article ">Figure 3
<p>B-scan OCT reveals an enlargement of the known retinal lesion, with a maximum retinal thickness of 433 microns, fuzzy borders, and subretinal hyperreflective material, which is a sign of progression of type 1 to type 2 MNV (red arrows), in the presence of intraretinal and subretinal fluid (yellow asterisks). All are imaging signs of MNV activation.</p>
Full article ">Figure 4
<p>OCT B-scan one month after the sixth intravitreal injection. The maximum retinal thickness decreased to 337 microns, and it is possible to detect the presence of some outer retinal tubulations and small intraretinal cysts (blue square), with more regular borders of the retinal lesion and a decrease in the fibrous scar area.</p>
Full article ">Figure 5
<p>OCT B-scan three months after the last intravitreal injection, showing a retinal thickness increase (368 microns) and a recurrent worsening of intraretinal fluid, associated with a concomitant worsening of visual acuity and metamorphopsia.</p>
Full article ">
22 pages, 541 KiB  
Review
Retinopathy in Metabolic Dysfunction-Associated Steatotic Liver Disease
by Myrsini Orfanidou and Stergios A. Polyzos
Medicina 2025, 61(1), 38; https://doi.org/10.3390/medicina61010038 (registering DOI) - 30 Dec 2024
Viewed by 157
Abstract
Metabolic dysfunction-associated steatotic liver disease (MASLD) is a multisystemic disease, i.e., influencing various organ systems beyond the liver and, thus, contributing to comorbidities. Characterized by excessive fat accumulation in the hepatocytes, MASLD is frequently linked to metabolic syndrome components, such as obesity, insulin [...] Read more.
Metabolic dysfunction-associated steatotic liver disease (MASLD) is a multisystemic disease, i.e., influencing various organ systems beyond the liver and, thus, contributing to comorbidities. Characterized by excessive fat accumulation in the hepatocytes, MASLD is frequently linked to metabolic syndrome components, such as obesity, insulin resistance, dyslipidemia, and hypertension. Therefore, exploring the intricate connection between MASLD and other organ systems, including the eyes, seems to be essential. In this context, retinopathy has been investigated for its potential association with MASLD, since both conditions share common pathogenetic pathways. Chronic low-grade inflammation, oxidative stress, insulin resistance, and endothelial dysfunction are only some of those mechanisms contributing to disease progression and, possibly, determining the bidirectional interplay between the liver and retinal pathology. This narrative review aims to summarize data concerning the multisystemicity of MASLD, primarily focusing on its potential association with the eyes and, particularly, retinopathy. Identifying this possible association may emphasize the need for early screening and integrated management approaches that address the liver and eyes as interconnected components within the framework of a systemic disease. Further research is necessary to delineate the precise mechanisms and develop targeted interventions to mitigate the bidirectional impact between the liver and eyes, aiming to reduce the overall burden of disease and improve patient outcomes. Full article
(This article belongs to the Special Issue Retinal and Choroidal Vascular Disease)
Show Figures

Figure 1

Figure 1
<p>The potential association between MASLD and retinopathy. Insulin resistance and hyperglycemia, which are common in MASLD, promote oxidative stress and glycation end-product formation, thus breaking down the blood–retinal barrier and increasing the risk of retinopathy. MASLD also contributes to endothelial dysfunction, by reducing nitric oxide availability and increasing retinal vascular stiffness, thus leading to features of retinopathy, like capillary leakage and neovascularization. Additionally, chronic inflammation and lipid dysmetabolism exacerbate microvascular damage through pro-inflammatory cytokines, dyslipidemia, and lipid peroxidation, further linking MASLD to retinal disease. Abbreviations: CRP, C-reactive protein; DR, diabetic retinopathy; HDL, high-density lipoprotein; IL-6, interleukin-6; LDL, low-density lipoprotein; MASLD, metabolic dysfunction-associated steatotic liver disease; TG, triglycerides; TNF-a, tumor necrosis factor-a.</p>
Full article ">
10 pages, 469 KiB  
Article
Evaluation of Ocular Irritation Sensitivity: Implications of Clinical Parameters, Pain Sensitivity, and Tear Neuromediator Profiles
by Hyeon-Jeong Yoon, Ja Young Moon, Hyun Jee Kim, Sodam Park, Ji Suk Choi, Hoon-In Choi, Seoyoung Kim and Kyung Chul Yoon
J. Clin. Med. 2025, 14(1), 138; https://doi.org/10.3390/jcm14010138 (registering DOI) - 29 Dec 2024
Viewed by 210
Abstract
Background/Objectives: Sensitivity to ocular irritation varies among individuals, being influenced by clinical, subjective, and biochemical factors. This study aimed to evaluate individual variability in ocular irritation sensitivity, focusing on clinical parameters, pain perception, and tear neuromediator profiles. Methods: Sixty female participants aged 20–40 [...] Read more.
Background/Objectives: Sensitivity to ocular irritation varies among individuals, being influenced by clinical, subjective, and biochemical factors. This study aimed to evaluate individual variability in ocular irritation sensitivity, focusing on clinical parameters, pain perception, and tear neuromediator profiles. Methods: Sixty female participants aged 20–40 were classified into high-sensitivity and low-sensitivity groups based on their response to an irritant (Tween20). Clinical assessments included the ocular surface disease index (OSDI), tear break-up time (TBUT), Schirmer test, and corneal touch threshold measured with the Cochet–Bonnet esthesiometer. Pain sensitivity was assessed using the pain sensitivity questionnaire (PSQ), and tear neuromediators were quantified in tear samples before and after stimulation. The concentrations of calcitonin gene-related peptide (CGRP), nerve growth factor, neuropeptide Y, vasoactive intestinal peptide (VIP), and substance P were measured using an enzyme-linked immune sorbent assay (ELISA). Results: The high-sensitivity group exhibited significantly higher OSDI scores (p = 0.038). No significant differences were observed in TBUT, corneal staining scores, or Schirmer’s test results. The PSQ results revealed that the high-sensitivity group had lower total and moderate pain scores (p = 0.037 and p = 0.040, respectively). An analysis of the tear neuromediator showed elevated baseline CGRP levels (p = 0.017) and a significant post-stimulation increase in substance P (p = 0.021) in the high-sensitivity group. Conclusions: These findings emphasize the value of combining clinical, subjective, and biochemical measures to understand sensitivity to ocular irritation. This comprehensive approach may guide the development of safer cosmetic formulations and improve safety assessment protocols. Full article
(This article belongs to the Section Ophthalmology)
Show Figures

Figure 1

Figure 1
<p>Comparison of tear neuromediators between the sensitivity and no-sensitivity groups of subjects, (<b>A</b>) calcitonin gene-related peptide (CGRP), (<b>B</b>) nerve growth factor (NGF), (<b>C</b>) neuropeptide Y (NPY), (<b>D</b>) vasoactive intestinal peptide (VIP), and (<b>E</b>) substance P (SP; * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
11 pages, 3000 KiB  
Article
Variabilities in Retinal Hemodynamics Across the Menstrual Cycle in Healthy Women Identified Using Optical Coherence Tomography Angiography
by Vlad Constantin Donica, Alexandra Lori Donica, Irina Andreea Pavel, Ciprian Danielescu, Anisia Iuliana Alexa and Camelia Margareta Bogdănici
Life 2025, 15(1), 22; https://doi.org/10.3390/life15010022 (registering DOI) - 28 Dec 2024
Viewed by 210
Abstract
Background: Numerous conditions, both physiological and pathological, can influence changes in the retinal vascular architecture. In order to be able to highlight pathological aspects of systemic diseases with ocular activity, it is necessary to understand how physiological fluctuations can influence circulation at the [...] Read more.
Background: Numerous conditions, both physiological and pathological, can influence changes in the retinal vascular architecture. In order to be able to highlight pathological aspects of systemic diseases with ocular activity, it is necessary to understand how physiological fluctuations can influence circulation at the retinal level. The present study attempts to evaluate retinal and choroidal vascular and structural changes in healthy female subjects over the course of a menstrual cycle using OCT-A. Methods: We analyzed 22 eyes from healthy reproductive women with a regular menstrual cycle. We performed five OCT-A scans of the subjects every 7–8 days over the course of a month starting from the first day of the menstrual cycle and ending with the first day of the next cycle, measuring perfusion density in the superficial and deep vascular plexuses, choroidal thickness, and FAZ perimeter. Results: There are physiological variations in retinal hemodynamics that can be identified using OCT-A, choroidal thickness having statistically significant increased values in the parafoveal nasal sector during the ovulatory phase (289.18 µm) compared to the early follicular phase (281.9 µm), and the subfoveal sector during the ovulatory phase (319.04 µm) compared to the early follicular phase (308.27 µm). Conclusions: These findings along with abnormally small FAZ perimeters indicate that the menstrual cycle phase should be considered whenever interpreting OCT-A results. Further studies that include larger cohorts, control groups, and hormone serum levels are necessary to confirm and correlate retinal vascular alterations and the phase of the menstrual cycle using OCT-A. Full article
Show Figures

Figure 1

Figure 1
<p>OCT-A plexus segmentation (<b>a</b>), perfusion density (<b>b</b>), and FAZ measurement (<b>c</b>).</p>
Full article ">Figure 2
<p>SVP–PD fluctuations in each quadrant over the course of a menstrual cycle.</p>
Full article ">Figure 3
<p>DVP–PD fluctuations in each quadrant over the course of a menstrual cycle.</p>
Full article ">Figure 4
<p>SVP and DVP–PD fluctuations in the fovea over the course of a menstrual cycle.</p>
Full article ">Figure 5
<p>PP–PD fluctuations in each quadrant over the course of a menstrual cycle.</p>
Full article ">Figure 6
<p>Choroidal thickness values in each quadrant over the course of a menstrual cycle.</p>
Full article ">Figure 7
<p>FAZ values in each quadrant over the course of a menstrual cycle.</p>
Full article ">Figure 8
<p>Small FAZ perimeters in healthy patients.</p>
Full article ">
57 pages, 2020 KiB  
Review
Therapeutic Potential of Herbal Medicines in Combating Particulate Matter (PM)-Induced Health Effects: Insights from Recent Studies
by Aekkhaluck Intharuksa, Warunya Arunotayanun, Mingkwan Na Takuathung, Yaowatat Boongla, Siripat Chaichit, Suthiwat Khamnuan and Anchalee Prasansuklab
Antioxidants 2025, 14(1), 23; https://doi.org/10.3390/antiox14010023 - 27 Dec 2024
Viewed by 392
Abstract
Particulate matter (PM), particularly fine (PM2.5) and ultrafine (PM0.1) particles, originates from both natural and anthropogenic sources, such as biomass burning and vehicle emissions. These particles contain harmful compounds that pose significant health risks. Upon inhalation, ingestion, or dermal [...] Read more.
Particulate matter (PM), particularly fine (PM2.5) and ultrafine (PM0.1) particles, originates from both natural and anthropogenic sources, such as biomass burning and vehicle emissions. These particles contain harmful compounds that pose significant health risks. Upon inhalation, ingestion, or dermal contact, PM can penetrate biological systems, inducing oxidative stress, inflammation, and DNA damage, which contribute to a range of health complications. This review comprehensively examines the protective potential of natural products against PM-induced health issues across various physiological systems, including the respiratory, cardiovascular, skin, neurological, gastrointestinal, and ocular systems. It provides valuable insights into the health risks associated with PM exposure and highlights the therapeutic promise of herbal medicines by focusing on the natural products that have demonstrated protective properties in both in vitro and in vivo PM2.5-induced models. Numerous herbal medicines and phytochemicals have shown efficacy in mitigating PM-induced cellular damage through their ability to counteract oxidative stress, suppress pro-inflammatory responses, and enhance cellular defense mechanisms. These combined actions collectively protect tissues from PM-related damage and dysfunction. This review establishes a foundation for future research and the development of effective interventions to combat PM-related health issues. However, further studies, including in vivo and clinical trials, are essential to evaluate the safety, optimal dosages, and long-term effectiveness of herbal treatments for patients under chronic PM exposure. Full article
(This article belongs to the Special Issue Environmental Pollution and Oxidative Stress)
Show Figures

Figure 1

Figure 1
<p>Sources of hazardous chemicals and particulate matter, the size distribution of PM, and the impact of PM on human health. Created using BioRender (Intharuksa, A., 2024). Available at <a href="https://BioRender.com/b97y489" target="_blank">https://BioRender.com/b97y489</a> (accessed on 20 November 2024).</p>
Full article ">Figure 2
<p>Key processes and molecular mechanisms underlying PM-mediated toxicity in human cells. Typically, PM can enter the human body through three major routes: inhalation, ingestion, and dermal absorption. Penetrated PM triggers inflammatory responses with the release of multiple proinflammatory cytokines as well as induction of intracellular ROS generation. Inside the cell, an elevated level of ROS further leads to activation of key signaling pathways related to inflamma-tion, particularly MAPKs and NF-kB, where solid arrow-headed black lines denote activation, dashed arrow-headed black lines indicate nuclear translocation, and solid right-angled blue ar-row represents gene transcription. Created using BioRender (Prasansuklab, A., 2025). Available at <a href="https://BioRender.com/b88q782" target="_blank">https://BioRender.com/b88q782</a> (accessed on 20 November 2024).</p>
Full article ">Figure 3
<p>A schematic illustration summarizing the health impacts of PM exposure on human body systems and some potential natural products against PM toxicities. Rounded rectangles surrounding each body system represent natural products possessing beneficial biological properties to counteract the harmful effects of PM exposure, which green boxes indicate phytochemicals, pink boxes indicate plant extracts, and yellow boxes indicate herbal formulations. Created using BioRender (Prasansuklab, A., 2025). Available at <a href="https://BioRender.com/j55x263" target="_blank">https://BioRender.com/j55x263</a> (accessed on 20 November 2024).</p>
Full article ">
20 pages, 2143 KiB  
Article
Thermosensitive In Situ Ophthalmic Gel for Effective Local Delivery and Antifungal Activity of Ketoconazole Nanoparticles
by Chutima Chaiwut, Sarin Tadtong, Puriputt Akachaipaibul, Jutamas Jiaranaikulwanitch, Sudarshan Singh, Siriporn Okonogi, Dwi Marlina Syukri and Chuda Chittasupho
Gels 2025, 11(1), 13; https://doi.org/10.3390/gels11010013 - 27 Dec 2024
Viewed by 282
Abstract
Fungal keratitis is a severe ocular infection caused by pathogenic fungi, leading to potential vision loss if untreated. Current antifungal treatments face limitations such as low solubility, poor corneal penetration, and limited therapeutic options. This study aimed to develop a thermosensitive in situ [...] Read more.
Fungal keratitis is a severe ocular infection caused by pathogenic fungi, leading to potential vision loss if untreated. Current antifungal treatments face limitations such as low solubility, poor corneal penetration, and limited therapeutic options. This study aimed to develop a thermosensitive in situ gel incorporating ketoconazole nanoparticles (NPs) to enhance drug solubility, stability, and antifungal activity. Ketoconazole NPs were prepared using the solvent displacement method, achieving a particle size of 198.25 ± 27.51 nm, encapsulation efficiency of 94.08 ± 0.51%, polydispersity index of 0.42 ± 0.08, and a positive zeta potential value of +10.08 ± 0.19 mV. The NPs exhibited sustained zero-order release kinetics. The optimized NPs were incorporated into a poloxamer-based in situ gel, demonstrating a gelation temperature of 34.67 ± 0.58 °C and the shortest gelation time. The formulation provided a 5-fold increase in solubility and a 10-fold improvement in drug release compared to pure ketoconazole. Stability studies confirmed the gel retained its physicochemical and rheological properties for three months under various storage conditions. The in situ gel showed sustained release, effective antifungal activity against Malassezia furfur, and good tolerability, suggesting it as a promising alternative for treating fungal keratitis with improved bioavailability and patient compliance. Full article
(This article belongs to the Special Issue Recent Advances in Gels Engineering for Drug Delivery (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Physical stability of ketoconazole-loaded NPs (NPs) over a 3-month storage period at 30 °C. (<b>A</b>) Particle size; (<b>B</b>) PDI; (<b>C</b>) zeta potential (*** indicates <span class="html-italic">p</span> &lt; 0.001 and **** indicates <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 2
<p>Cumulative release profile of ketoconazole from NPs over an 8 h incubation period in phosphate-buffered saline (pH 7.4) at 33 ± 0.5 °C.</p>
Full article ">Figure 3
<p>Chemical stability of ketoconazole-loaded NPs over a 3-month storage period at three different temperatures (4 °C, 30 °C, and 45 °C).</p>
Full article ">Figure 4
<p>Gelation time of in situ gel loaded with ketoconazole NPs during a 3-month incubation period at different storage temperatures (4 °C, 30 °C, and 45 °C). Statistical significance compared to initial gelation time is indicated as follows: * indicates <span class="html-italic">p</span> &lt; 0.05 and **** indicates <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Gelation temperature of in situ gel loaded with ketoconazole NPs over a 3-month incubation period at different storage temperatures (4 °C, 30 °C, and 45 °C). Statistical significance compared to initial gelation temperature is indicated as follows: * indicates <span class="html-italic">p</span> &lt; 0.05, ** indicates <span class="html-italic">p</span> &lt; 0.01, and **** indicates <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>pH stability of in situ gel loaded with ketoconazole NPs over a 3-month incubation period at different storage temperatures (4 °C, 30 °C, and 45 °C). Statistical significance compared to the initial pH value is indicated as follows: * indicates <span class="html-italic">p</span> &lt; 0.05, **** indicates <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p>Viscosity-temperature profile of in situ gel formulations with and without ketoconazole-loaded NPs.</p>
Full article ">Figure 8
<p>The viscosity of ketoconazole-loaded NPs stored at three different temperatures (4 °C, 30 °C, and 45 °C) over a 3-month incubation period. * indicates <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
14 pages, 4205 KiB  
Article
Ocular Aberrations and Retinal Thickness Variations After Moderate-Term Reading on Electronic Devices by Age
by María Arcas-Carbonell, Elvira Orduna-Hospital, María Mechó-García, María Munarriz-Escribano and Ana Sanchez-Cano
Photonics 2025, 12(1), 16; https://doi.org/10.3390/photonics12010016 - 27 Dec 2024
Viewed by 251
Abstract
Background: This study aims to evaluate subjective visual fatigue and objective optical and morphological changes in ocular structures after intermediate-duration reading on an iPad and an Ebook across different age groups. Methods: The sample included 108 right eyes from healthy subjects aged 18 [...] Read more.
Background: This study aims to evaluate subjective visual fatigue and objective optical and morphological changes in ocular structures after intermediate-duration reading on an iPad and an Ebook across different age groups. Methods: The sample included 108 right eyes from healthy subjects aged 18 to 66 years. The participants read for 20 min on an Ebook and another 20 min on an iPad under controlled illumination conditions. Aberrometry and retinal optical coherence tomography (OCT) measurements were taken before and after each reading session. Parameters such as total aberration, high-order aberration (HOA), low-order aberration (LOA), and retinal thickness in the nine Early Treatment Diabetic Retinopathy Study (ETDRS) areas were measured. The sample was analyzed as a whole and divided into five age groups by decade. Results: This study included 66 women (61.11%) and 42 men (38.89%), with an average age of 36.58 years (±14.83). The aberrometry results revealed significant differences in the total root mean square (RMSTOTAL) after reading on both devices (p = 0.001). Low-order aberrations (RMSLOA) also changed significantly (p = 0.001 for Ebook, p = 0.002 for the iPad), but high-order aberrations (RMSHOA) did not. Central retinal thickness increased significantly after reading on the Ebook (p < 0.001) but not on the iPad. The peripheral retinal thickness did not change significantly. Conclusion: Moderate-duration reading increases LOA and central retinal thickness, with variations by age group and more pronounced effects from the Ebook, whereas HOA remains unaffected. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Elements for the proper conduct of the study. (1) Controlled lighting box. (2) Chinrest. (3) Stand for placing the electronic device. (4) Dispositive of lecture (in this case, the iPad). (<b>B</b>) Normalized spectral irradiance (W/m<sup>2</sup>) of the ambient light reaching the corneal plane while reading.</p>
Full article ">Figure 2
<p>(<b>A</b>) Tomographic section of the central retina. (<b>B</b>) Fundus image displaying central retinal thickness in the Early Treatment Diabetic Retinopathy Study (ETDRS) grid. (<b>C</b>) Tomographic section of the peripheral retina. (<b>D</b>) Fundus image displaying peripheral retinal thickness in the ETDRS grid. All the images correspond to the right eye of the same subject.</p>
Full article ">Figure 3
<p>Comparative graphical representation of the root mean square (RMS) in µm (<b>A</b>) and the spherical equivalent (M) in D (<b>B</b>) before and after reading by devices considering all the subjects. Comparative graphical representation of the RMSs in µm (<b>C</b>) and M in D (<b>D</b>) before and after reading by device stratified by age group from G1 to G5. Low-order aberrations (LOAs), high-order aberrations (HOAs), 3rd-order aberrations (RMSs), 4th-order aberrations (RMSs), and 5th-order aberrations (RMSs). The Wilcoxon signed-rank test for two related samples was assessed, and the <span class="html-italic">p</span>-value is depicted above the graphical bars, marked with an asterisk (*), indicating statistically significant differences when <span class="html-italic">p</span> &lt; 0.05. The potential options are illustrated with a solid line when the baseline measurement is compared with the Ebook reading, and a dashed line is generated when the baseline measurement is compared with the iPad.</p>
Full article ">Figure 4
<p>Comparative graphical representation of the Zernike polynomial values of 2nd-order (<b>A</b>), 3rd, 4th, 5th, and 6th-order (<b>B</b>) polynomials before and after reading, using the two devices employed for this purpose, considering differentiated by age groups from G1 to G5. The Wilcoxon signed-rank test for two related samples was assessed, and the <span class="html-italic">p</span>-value is depicted above the graphical bars, marked with an asterisk (*), indicating statistically significant differences when <span class="html-italic">p</span> &lt; 0.05. The potential options are illustrated with a solid line when the baseline measurement is compared with the Ebook reading. The dashed line compares the baseline with the iPad, and the dotted line compares the Ebook with the iPad.</p>
Full article ">Figure 5
<p>Comparison between baseline measurements and measurements taken after continuous reading on the iPad and Ebook (<b>A</b>) in each of the 9 Early Treatment Diabetic Retinopathy Study (ETDRS) areas, the total volume (<b>B</b>) and the average total thickness (<b>C</b>) of the central retina of all subjects. Comparison between baseline measurements and measurements taken after continuous reading on the iPad and Ebook (<b>D</b>) in each of the 9 ETDRS areas, the average total thickness (<b>E</b>) and the total volume (<b>F</b>) of the central retina when differentiated by age groups from G1 to G5. The Wilcoxon signed-rank test for two related samples was assessed, and the <span class="html-italic">p</span>-value is depicted above the graphical bars, marked with an asterisk (*), indicating statistically significant differences when <span class="html-italic">p</span> &lt; 0.05. The potential options are illustrated with a solid line when the baseline measurement is compared with the Ebook reading. The dashed line compares the baseline with the iPad, and the dotted line compares the Ebook with the iPad.</p>
Full article ">Figure 6
<p>(<b>A</b>) The comparison of image quality between the two reading situations, iPad and Ebook, depending on the subjectively perceived effort stated by participants, where 0 represents “none”, and 5 represents “a lot”. The orange bars indicate the number of subjects for each answer choice after reading onthe Ebook, along with the linear trend line for these data in the same color. The blue bars represent the corresponding results for the iPad device. (<b>B</b>) The comparison of visual quality between the two reading conditions, iPad and Ebook, based on the clarity perceived by the participants while reading, where 0 represents “clear”, 1 represents “slightly blurry”, and 2 represents “blurry”. The orange pie portions indicate the number of subjects selecting each response option after reading onthe Ebook, whereas the blue pie segments show the corresponding results for the iPad, as indicated in the legend.</p>
Full article ">
18 pages, 1024 KiB  
Review
Effects of Selected Antioxidants on Electroretinography in Rodent Diabetic Retinopathy
by Radosław Dutczak and Marita Pietrucha-Dutczak
Antioxidants 2025, 14(1), 21; https://doi.org/10.3390/antiox14010021 - 27 Dec 2024
Viewed by 259
Abstract
Electroretinography (ERG) is a non-invasive technique for evaluating the retinal function in various ocular diseases. Its results are useful for diagnosing ocular disorders and assessing disease progression or treatment effectiveness. Since numerous studies are based on animal models, validating the ERG results from [...] Read more.
Electroretinography (ERG) is a non-invasive technique for evaluating the retinal function in various ocular diseases. Its results are useful for diagnosing ocular disorders and assessing disease progression or treatment effectiveness. Since numerous studies are based on animal models, validating the ERG results from animals is pivotal. The first part of this paper presents basic information on the types of ERG tests used on rodents, and the second part describes the recorded functional changes in rodents’ retinas when various antioxidant treatments for diabetic retinopathy were used. Our study showed that among the tests for diabetic retinopathy diagnosis in rodents, full-field ERG is accurate and the most commonly used, and pattern ERG and the photopic negative response of the flash ERG tests are rarely chosen. Furthermore, antioxidants generally protect retinas from functional losses. Their beneficial influence is expressed in the preserved amplitudes of the a- and b-waves and the oscillatory potentials. However, prolonging the drug exposure showed that the antioxidants could delay the onset of adverse changes but did not stop them. Future studies should concentrate on how long-term antioxidant supplementation affects the retinal function. Full article
Show Figures

Figure 1

Figure 1
<p>Types of waves recorded in various electroretinography tests. (ffERG—full-field ERG; mfERG—multifocal ERG; PERG—pattern ERG; PhNR—the photopic negative response; RGC—retinal ganglion cells; OPs—oscillatory potentials; AD—Alzheimer’s disease; PD—Parkinson’s disease).</p>
Full article ">Figure 2
<p>Results processing methods in rodent diabetes. (PhNR—the photopic negative response; OPs—oscillatory potentials).</p>
Full article ">
27 pages, 12325 KiB  
Article
Optimized Prime Editing of Human Induced Pluripotent Stem Cells to Efficiently Generate Isogenic Models of Mendelian Diseases
by Rodrigo Cerna-Chavez, Alba Ortega-Gasco, Hafiz Muhammad Azhar Baig, Nathan Ehrenreich, Thibaud Metais, Michael J. Scandura, Kinga Bujakowska, Eric A. Pierce and Marcela Garita-Hernandez
Int. J. Mol. Sci. 2025, 26(1), 114; https://doi.org/10.3390/ijms26010114 - 26 Dec 2024
Viewed by 355
Abstract
Prime editing (PE) is a CRISPR-based tool for genome engineering that can be applied to generate human induced pluripotent stem cell (hiPSC)-based disease models. PE technology safely introduces point mutations, small insertions, and deletions (indels) into the genome. It uses a Cas9-nickase (nCas9) [...] Read more.
Prime editing (PE) is a CRISPR-based tool for genome engineering that can be applied to generate human induced pluripotent stem cell (hiPSC)-based disease models. PE technology safely introduces point mutations, small insertions, and deletions (indels) into the genome. It uses a Cas9-nickase (nCas9) fused to a reverse transcriptase (RT) as an editor and a PE guide RNA (pegRNA), which introduces the desired edit with great precision without creating double-strand breaks (DSBs). PE leads to minimal off-targets or indels when introducing single-strand breaks (SSB) in the DNA. Low efficiency can be an obstacle to its use in hiPSCs, especially when the genetic context precludes the screening of multiple pegRNAs, and other strategies must be employed to achieve the desired edit. We developed a PE platform to efficiently generate isogenic models of Mendelian disorders. We introduced the c.25G>A (p.V9M) mutation in the NMNAT1 gene with over 25% efficiency by optimizing the PE workflow. Using our optimized system, we generated other isogenic models of inherited retinal diseases (IRDs), including the c.1481C>T (p.T494M) mutation in PRPF3 and the c.6926A>C (p.H2309P) mutation in PRPF8. We modified several determinants of the hiPSC PE procedure, such as plasmid concentrations, PE component ratios, and delivery method settings, showing that our improved workflow increased the hiPSC editing efficiency. Full article
(This article belongs to the Special Issue Molecular Research in Retinal Degeneration)
Show Figures

Figure 1

Figure 1
<p>PE3 prime editing strategy to generate the isogenic model for an <span class="html-italic">NMNAT1</span> c.25G&gt;A (p.V9M) mutation. (<b>A</b>). WGS analysis of the genome of the hiPSC line IMR90-clone 4 depicts no variants in the region of interest for <span class="html-italic">NMNAT1</span> editing compared to the reference human genome hg38. (<b>B</b>). Two gRNAs out of ten were selected for molecular cloning based on the distance to the desired edit position (<b>C</b>). pegRNA1 and (<b>D</b>). pegRNA2 anneal to the complementary DNA strand. Cas9 Nickase nicks the opposite strand, allowing for PBS annealing and reverse transcription along the RT template that incorporates the desired single nucleotide edit (red letter).</p>
Full article ">Figure 2
<p>Validation of PE in HEK-293T cells. (<b>A</b>). Each pegRNA has 3 key components: a gRNA or spacer, a scaffold, and an extension with the PBS and the RTT sequence, which introduces the desired edit. (<b>B</b>). Confluent HEK cells were co-transfected using Lipofectamine 3000 with the pegRNA, the PEmax editor, and a nicking guide plasmid harboring a reporter EGFP cassette. Then, 48 h post-transfection, approximately 80–90% of the HEK cells expressed GFP. Scale bar = 50 μm. (<b>C</b>). Three days after plasmid delivery, genomic DNA was extracted, and PE was analyzed by NGS. pegRNA1 was designed to introduce a G&gt;A edit at +3 position from the nicking site in <span class="html-italic">NMNAT1,</span> and pegRNA2 was designed to introduce the same edit at +14 position from the nicking site. (<b>D</b>). Editing efficiencies were determined by NGS and expressed as the percentage of alleles with G•C target converted to T•A. NGS showed editing only occurred with PE3 with low efficiency: 3.40% for pegRNA1 and 0.42% for peg RNA2.</p>
Full article ">Figure 3
<p>Optimization of PE of hiPSCs for <span class="html-italic">NMNAT1</span> c.25G&gt;A. (<b>A</b>). Optimization of electroporation conditions for the efficient delivery of PE components in the cells using 1 μg of PEmax editor, 90 ng of nicking guide RNA, and 240 ng of pegRNA1. Three different electroporation parameters were tested. The best condition was 1100 V, 20 ms, and 2 pulses, resulting in the highest cell survival and the highest percentage of edited alleles detected by next-generation sequencing (NGS). (<b>B</b>). Combinatorial screening of different concentrations of PE components. Using the optimal electroporation parameters, a total of 9 different combinations of PE components were assessed to optimize PE efficiency. PEmax was set at 900 ng, and three doses of pegRNA (low, 120 ng; medium (mid), 180 ng; and high, 240 ng) and three doses of nicking guide RNA (low, 90 ng; medium (mid), 120 ng; and high 180 ng) were included. (<b>C</b>). NGS analysis of the bulk electroporated hiPSCs showed the maximum editing efficiency (9.60%) was achieved with high pegRNA1 and low nicking gRNA. (<b>D</b>). PE efficiency under optimized conditions demonstrated a substantial increase in the percentage of GFP-positive cells (66.80%) post-electroporation that translated into a notable increase in editing efficiency. Scale bar = 50 μm.</p>
Full article ">Figure 4
<p>Generation of <span class="html-italic">NMNAT1</span><sup>V9M/V9M</sup> isogenic clones. (<b>A</b>). Genotype confirmation and (<b>B</b>). determination of the PE efficiency for <span class="html-italic">NMNAT1</span> c.25 G&gt;A using Sanger sequencing. PE3 efficiency accounted for over 25.00% of edited clones. Data in (<b>A</b>) are representative of <span class="html-italic">n</span> ≥ 2 independent replicates.</p>
Full article ">Figure 5
<p>Prime editing strategy to generate the isogenic model for <span class="html-italic">PRPF3</span> c.1482C&gt;T (p.T494M) mutation and for <span class="html-italic">PRPF8</span> c.6926A&gt;C (p.H2309P) mutation. (<b>A</b>). The gRNA1 anneals with the complementary DNA strand. Cas9 Nickase nicks the PAM-containing strand of the target DNA. The PBS anneals with the PAM-containing strand, and RTase extends the 3′ end using the RT template. (<b>B</b>). pegRNA was selected for molecular cloning based on the distance to the mutation and sequence length. (<b>C</b>). The gRNA1 anneals with the complementary DNA strand. Cas9 Nickase nicks the PAM-containing strand of the target DNA. The PBS anneals with the PAM-containing strand, and RTase extends the 3′ end using the RT template. (<b>D</b>). pegRNA was selected for molecular cloning based on the distance to the mutation and sequence length.</p>
Full article ">Figure 6
<p>Generation of <span class="html-italic">PRPF3</span> isogenic clones. (<b>A</b>). Genotype confirmation and (<b>B</b>). determination of the PE efficiency with high pegRNA for <span class="html-italic">PRPF3</span> c.1482 C&gt;T (p.T494M) using NGS and Sanger sequencing. Data in (<b>A</b>) are representative of <span class="html-italic">n</span> ≥ 2 independent replicates.</p>
Full article ">Figure 7
<p>Generation of <span class="html-italic">PRPF8</span> isogenic clones. (<b>A</b>). Genotype confirmation and (<b>B</b>). determination of the PE efficiency with high and maximal pegRNAs for <span class="html-italic">PRPF8</span> c.6926 A&gt;C (p.H2309P) using NGS and Sanger sequencing. Data in (<b>A</b>) are representative of <span class="html-italic">n</span> ≥ 2 independent replicates.</p>
Full article ">Figure 8
<p>Characterization of hiPSC wild-type <span class="html-italic">NMNAT1<sup>+/+</sup></span>, and homozygous <span class="html-italic">NMNAT1</span><sup>−/−</sup> clones showing immunofluorescence (IF) images of SOX2<sup>+</sup>, SSEA4<sup>+</sup>, OCT<sup>+</sup>, and NANOG<sup>+</sup> cells for pluripotency markers; embryoid bodies exhibiting NESTIN<sup>+</sup> and GFAP<sup>+</sup> (ectoderm), SMA<sup>+</sup> (mesoderm), and SOX17<sup>+</sup> (endoderm) for germ layer makers; and chromosomal copy number variation analysis. Representative images from <span class="html-italic">n</span> &gt; 5. Scale bar = 100 μm.</p>
Full article ">Figure 9
<p>Prime editing (PE) pipeline. Our PE pipeline includes the in silico design of the pegRNA components and molecular cloning to ensemble the pegRNAs, the nicking, and the PEmax editor plasmids. We also performed a full QC analysis of the hiPSC lines prior to editing, including evaluating pluripotency and chromosomal abnormalities and assessing the differentiation capacity. Electroporation with different combinations of PE components was performed on highly confluent cultures to generate isogenic models for the <span class="html-italic">NMNAT1</span>, <span class="html-italic">PRPF3</span>, and <span class="html-italic">PRPF8</span> mutations. After 2–5 days, electroporated hiPSC lines were dissociated into single cells and passaged at low-density seeding (LDS). In parallel, NGS was performed to confirm editing has occurred in the bulk cultures and estimate the number of clones that need to be recovered from the LDS cultures. Single colonies from LDS cultures were manually dissected and passaged by mechanical disruption into 96-well plates, where the desired edit was confirmed through Sanger sequencing analysis. Confirmed edited clones were expanded first to 12-well plates and then to 6-well plates for banking. Cryopreserved seed and master banks were obtained in parallel to the QC analysis of the selected clones.</p>
Full article ">
14 pages, 898 KiB  
Article
The Prevalence of Accommodative and Binocular Dysfunctions in Children with Reading Difficulties
by Ilze Ceple, Aiga Svede, Evita Serpa, Evita Kassaliete, Liva Volberga, Rita Mikelsone, Asnate Berzina, Angelina Ganebnaya, Linda Krauze and Gunta Krumina
Life 2025, 15(1), 7; https://doi.org/10.3390/life15010007 - 25 Dec 2024
Viewed by 258
Abstract
Uncorrected refractive error and unsatisfactory performance on several clinical accommodation and binocular vision tests are more common in children who struggle with reading. The aim of the current study is to explore the prevalence of accommodative and binocular dysfunctions in children with and [...] Read more.
Uncorrected refractive error and unsatisfactory performance on several clinical accommodation and binocular vision tests are more common in children who struggle with reading. The aim of the current study is to explore the prevalence of accommodative and binocular dysfunctions in children with and without reading difficulties. Reading performance was assessed with the Acadience Reading (formerly DIBELS Next) test adjusted and validated for the Latvian language. Children with (N = 39) and without (N = 43) reading difficulties underwent thorough assessment of their subjective refraction, as well as binocular and accommodation functions. The results demonstrate no difference in the prevalence of complaints between children with and without reading difficulties (26% and 23%, respectively). However, children with reading difficulties more frequently present with significant uncorrected refractive errors and/or accommodative and binocular dysfunctions than children without reading difficulties (69% and 47%, respectively). According to the findings, even in cases where a child does not exhibit any ocular or visual complaints, a comprehensive visual function evaluation should be required for any child who struggles with reading. Full article
(This article belongs to the Special Issue Vision Science and Optometry)
Show Figures

Figure 1

Figure 1
<p>Frequency of visual complaints in children with reading difficulties and without reading difficulties (the control group).</p>
Full article ">Figure 2
<p>Frequency of significant refractive errors and/or accommodative and binocular dysfunctions in children with reading difficulties and without reading difficulties (the control group). * Statistically significantly difference (Chi-square test of independence <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Relative frequency of significant refractive errors and accommodative and binocular dysfunctions in children with reading difficulties and without reading difficulties (the control group). * Statistically significantly difference (Chi-square test of independence <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
26 pages, 7691 KiB  
Article
Neuroprotective Effect of the Combination of Citicoline and CoQ10 in a Mouse Model of Ocular Hypertension
by José A. Matamoros, Sara Rubio-Casado, José A. Fernández-Albarral, Miguel A. Martínez-López, Elena Salobrar-García, Eva M. Marco, Victor Paleo-García, Rosa de Hoz, Inés López-Cuenca, Lorena Elvira-Hurtado, Lidia Sánchez-Puebla, José M. Ramírez, Juan J. Salazar, Meritxell López-Gallardo and Ana I. Ramírez
Antioxidants 2025, 14(1), 4; https://doi.org/10.3390/antiox14010004 - 24 Dec 2024
Viewed by 444
Abstract
Glaucoma is a neurodegenerative disease characterized by the loss of retinal ganglion cells (RGCs), with intraocular pressure (IOP) being its primary risk factor. Despite controlling IOP, the neurodegenerative process often continues. Therefore, substances with neuroprotective, antioxidant, and anti-inflammatory properties could protect against RGC [...] Read more.
Glaucoma is a neurodegenerative disease characterized by the loss of retinal ganglion cells (RGCs), with intraocular pressure (IOP) being its primary risk factor. Despite controlling IOP, the neurodegenerative process often continues. Therefore, substances with neuroprotective, antioxidant, and anti-inflammatory properties could protect against RGC death. This study investigated the neuroprotective effects on RGCs and visual pathway neurons of a compound consisting of citicoline and coenzyme Q10 (CoQ10) in a mouse model of unilateral, laser-induced ocular hypertension (OHT). Four groups of mice were used: vehicle group (n = 6), citicoline + CoQ10 group (n = 6), laser–vehicle group (n = 6), and laser–citicoline + CoQ10 group (n = 6). The citicoline + CoQ10 was administered orally once a day starting 15 days before laser treatment, continuing until sacrifice (7 days post-laser). Retinas, the dorsolateral geniculate nucleus (dLGN), the superior colliculus (SC), and the visual cortex (V1) were analyzed. The citicoline + CoQ10 compound used in the laser–citicoline + CoQ10 group demonstrated (1) an ocular hypotensive effect only at 24 h post-laser; (2) prevention of Brn3a+ RGC death in OHT eyes; and (3) no changes in NeuN+ neurons in the dLGN. This study demonstrates that the oral administration of the citicoline + CoQ10 combination may exert a neuroprotective effect against RGC death in an established rodent model of OHT. Full article
(This article belongs to the Special Issue Antioxidants and Retinal Diseases—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Scheme of the experimental groups of this study.</p>
Full article ">Figure 2
<p>Intraocular pressure (IOP) graphics. (<b>A</b>) IOP in different study groups after induction of hypertension. (<b>B</b>) IOP data in the different study groups 24 h after induction. Data are mean ± s.e.m.; each data point denotes an individual measure of the IOP. Abbreviations: ocular hypertension (OHT); coenzyme Q10 (CoQ10). Statistical significance indicators: **** <span class="html-italic">p</span> &lt; 0.0001 vehicle vs. OHT; ++++ <span class="html-italic">p</span> &lt; 0.0001 citicoline + CoQ10 vs. OHT–citicoline + CoQ10; ### <span class="html-italic">p</span> &lt; 0.001, #### <span class="html-italic">p</span> &lt; 0.0001 OHT vs. contralateral; ^^^ <span class="html-italic">p</span> &lt; 0.001, ^^^^ <span class="html-italic">p</span> &lt; 0.0001 OHT–citicoline + CoQ10 vs. contralateral–citicoline + CoQ10; aaaa <span class="html-italic">p</span> &lt; 0.0001 for OHT vs. OHT–citicoline + CoQ10.</p>
Full article ">Figure 3
<p>Comparison of total number of Brn3a+ retinal ganglion cells (RGCs) in the different study groups treated with the combination of citicoline and CoQ10 or with vehicle 7 days after OHT induction. Data are mean ± s.e.m.; each data point denotes an individual measure of the Brn3a + RGCs per area of 0.1502 mm<sup>2</sup>. Abbreviations: ocular hypertension (OHT); coenzyme Q10 (CoQ10). Statistical significance indicators: **** <span class="html-italic">p</span> &lt; 0.0001 vehicle vs. OHT; #### <span class="html-italic">p</span> &lt; 0.0001 OHT vs. contralateral; aaaa <span class="html-italic">p</span> &lt; 0.0001 for OHT vs. OHT–citicoline + CoQ10.</p>
Full article ">Figure 4
<p>Immunohistochemical images of anti-Brn3a-stained retinal whole mounts 7 days after OHT induction. (<b>A</b>) Vehicle, (<b>B</b>) citicoline + CoQ10, (<b>C</b>) OHT, (<b>D</b>) OHT–citicoline + CoQ10, (<b>E</b>) contralateral, (<b>F</b>) contralateral–citicoline + CoQ10. Abbreviations: ocular hypertension (OHT); coenzyme Q10 (CoQ10).</p>
Full article ">Figure 5
<p>Comparison of number of Brn3a+ retinal ganglion cells (RGCs) in the four retinal sectors, superior, inferior, nasal, and temporal, in the different study groups 7 days after OHT induction. Data are mean ± s.e.m.; each data point denotes an individual measure of the Brn3a + RGCs per area of 0.1502 mm<sup>2</sup>. Abbreviations: ocular hypertension group (OHT); coenzyme Q10 (CoQ10). Statistical significance indicators: ** <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 vehicle vs. OHT; ## <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 OHT vs. contralateral; aaaa <span class="html-italic">p</span> &lt; 0.0001 OHT vs. OHT–citicoline + CoQ10.</p>
Full article ">Figure 6
<p>Comparison of the number of Brn3a+ retinal ganglion cells (RGCs) in untreated and citicoline and CoQ10-treated hypertensive eyes in the four retinal sectors, superior, temporal, nasal and inferior, 7 days after OHT induction. Data are mean ± s.e.m.; each data point denotes an individual measure of the Brn3a + RGCs per area of 0.1502 mm<sup>2</sup>. Abbreviations: ocular hypertension (OHT); coenzyme Q10 (CoQ10). Statistical significance indicators: In OHT group: * <span class="html-italic">p</span> &lt; 0.05 for superior vs. temporal, ** <span class="html-italic">p</span> &lt; 0.01 superior vs. nasal, **** <span class="html-italic">p</span> &lt; 0.0001 superior vs. inferior. In OHT–citicoline + CoQ10 group: + <span class="html-italic">p</span> &lt; 0.05 superior vs. temporal.</p>
Full article ">Figure 7
<p>Comparison of number of Brn3a+ retinal ganglion cells (RGCs) in the three retinal areas, peripapillary, intermediate, and peripheral, in the different study groups 7 days after OHT induction. Data are mean ± s.e.m.; each data point denotes an individual measure of the Brn3a + RGCs per area of 0.1502 mm<sup>2</sup>. Abbreviations: ocular hypertension group (OHT); coenzyme Q10 (CoQ10). Statistical significance indicators: **** <span class="html-italic">p</span> &lt; 0.0001 vehicle vs. OHT; ## <span class="html-italic">p</span> &lt; 0.01, #### <span class="html-italic">p</span> &lt; 0.0001 OHT vs. contralateral; aaaa <span class="html-italic">p</span> &lt; 0.0001 OHT vs. OHT–citicoline + CoQ10.</p>
Full article ">Figure 8
<p>Comparison of the number of Brn3a+ retinal ganglion cells (RGCs) in untreated and citicoline and CoQ10-treated hypertensive eyes in the three retinal areas: peripapillary, intermediate, and peripheral, 7 days after OHT induction. Data are mean ± s.e.m.; each data point denotes an individual measure of the Brn3a + RGCs per area of 0.1502 mm<sup>2</sup>. Abbreviations: ocular hypertension (OHT); coenzyme Q10 (CoQ10). Statistical significance indicators: In OHT group: ** <span class="html-italic">p</span> &lt; 0.01 peripheral vs. peripapillary; **** <span class="html-italic">p</span> &lt; 0.0001 peripheral vs. intermediate. In OHT–citicoline + CoQ10 group: ++++ <span class="html-italic">p</span> &lt; 0.0001 peripheral vs. peripapillary; ++++ <span class="html-italic">p</span> &lt; 0.0001 peripheral vs. intermediate.</p>
Full article ">Figure 9
<p>Comparison of total number of melanopsin+ intrinsically photosensitive retinal ganglion cells (ipRGCs) in the different study groups treated with the combination of citicoline and CoQ10 or with vehicle 7 days after OHT induction. Data are mean ± s.e.m.; each data point denotes an individual measure of the melanopsin+ ipRGCs per area of 0.1502 mm<sup>2</sup>. Abbreviations: ocular hypertension (OHT); coenzyme Q10 (CoQ10).</p>
Full article ">Figure 10
<p>Immunohistochemical images of anti-melanopsin-stained retinal whole mounts 7 days after ocular hypertension induction. (<b>A</b>) Vehicle, (<b>B</b>) citicoline + CoQ10, (<b>C</b>) OHT, (<b>D</b>) OHT–citicoline + CoQ10, (<b>E</b>) contralateral, (<b>F</b>) contralateral–citicoline + CoQ10. Abbreviations: ocular hypertension (OHT); coenzyme Q10 (CoQ10).</p>
Full article ">Figure 11
<p>Comparison of number of melanopsin+ intrinsically photosensitive retinal ganglion cells (ipRGCs) in the four retinal sectors, superior, inferior, nasal, and temporal, in the different study groups 7 days after OHT induction. Data are mean ± s.e.m.; each data point denotes an individual measure of the melanopsin+ ipRGCs per area of 0.1502 mm<sup>2</sup>. Abbreviations: ocular hypertension (OHT); coenzyme Q10 (CoQ10).</p>
Full article ">Figure 12
<p>Comparison of total number of melanopsin+ intrinsically photosensitive retinal ganglion cells (ipRGCs) in the three retinal areas, peripapillary, intermediate, and peripheral, in the different study groups 7 days after OHT induction. Data are mean ± s.e.m.; each data point denotes an individual measure of the melanopsin+ ipRGCs per area of 0.1502 mm<sup>2</sup>. Abbreviations: ocular hypertension (OHT); coenzyme Q10 (CoQ10).</p>
Full article ">Figure 13
<p>Pearson’s correlation coefficient (r) analysis between the number of Brn3a+ RGCs and maximum IOP in the untreated eyes (vehicle and OHT–vehicle groups) and the treated eyes (citicoline + CoQ10 and OHT–citicoline + CoQ10).</p>
Full article ">Figure 14
<p>Pearson’s correlation coefficient (r) analysis between the number of melanopsin+ ipRGC and maximum IOP in the untreated eyes (vehicle and OHT–vehicle groups) and the treated eyes (citicoline + CoQ10 and OHT–citicoline + CoQ10).</p>
Full article ">Figure 15
<p>Analysis of NeuN+ cells (OD) of the dorsolateral geniculate nucleus in the total (<b>A</b>), central zone (<b>B</b>), and peripheral zone (<b>C</b>) in the different study groups 7 days after ocular hypertension induction. Data are mean ± s.e.m.; each data point denotes an individual measure of the NeuN+ cells (OD). Abbreviations: ocular hypertension group (OHT); coenzyme Q10 (CoQ10). Statistical significance indicators: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 for vehicle vs. OHT dLGN right.</p>
Full article ">Figure 16
<p>Immunohistochemical images of brain sections of the dorsolateral geniculate nucleus stained with anti-NeuN 7 days after induction of ocular hypertension. (<b>A</b>) Vehicle, (<b>B</b>) citicoline + CoQ10, (<b>C</b>) OHT dLGN right, (<b>D</b>) OHT dLGN right–citicoline + CoQ10, (<b>E</b>) contralateral dLGN left, (<b>F</b>) contralateral dLGN left–citicoline + CoQ10. Abbreviations: ocular hypertension (OHT); coenzyme Q10 (CoQ10).</p>
Full article ">Figure 17
<p>Analysis of NeuN+ cells (OD) of the superior colliculus (SC) in the different study groups treated with the combination of citicoline and CoQ10 or with vehicle 7 days after OHT induction. Data are mean ± s.e.m.; each data point denotes an individual measurement of the NeuN+ cells (OD). Abbreviations: ocular hypertension (OHT); coenzyme Q10 (CoQ10).</p>
Full article ">Figure 18
<p>Analysis of NeuN+ cells (OD) of the visual cortex (V1) in the different study groups treated with the combination of citicoline and CoQ10 or with vehicle 7 days after ocular hypertension induction. Data are mean ± s.e.m.; each data point denotes an individual measure of the NeuN+ cells (OD). Abbreviations: ocular hypertension (OHT); coenzyme Q10 (CoQ10).</p>
Full article ">
17 pages, 11444 KiB  
Article
Oxidative Stress, Inflammation and Altered Glucose Metabolism Contribute to the Retinal Phenotype in the Choroideremia Zebrafish
by Cécile Méjécase, Neelima Nair, Hajrah Sarkar, Pablo Soro-Barrio, Maria Toms, Sophia Halliday, Katy Linkens, Natalia Jaroszynska, Constance Maurer, Nicholas Owen and Mariya Moosajee
Antioxidants 2024, 13(12), 1587; https://doi.org/10.3390/antiox13121587 - 23 Dec 2024
Viewed by 358
Abstract
Reactive oxygen species (ROS) within the retina play a key role in maintaining function and cell survival. However, excessive ROS can lead to oxidative stress, inducing dysregulation of metabolic and inflammatory pathways. The chmru848 zebrafish models choroideremia (CHM), an X-linked chorioretinal dystrophy, [...] Read more.
Reactive oxygen species (ROS) within the retina play a key role in maintaining function and cell survival. However, excessive ROS can lead to oxidative stress, inducing dysregulation of metabolic and inflammatory pathways. The chmru848 zebrafish models choroideremia (CHM), an X-linked chorioretinal dystrophy, which predominantly affects the photoreceptors, retinal pigment epithelium (RPE), and choroid. In this study, we examined the transcriptomic signature of the chmru848 zebrafish retina to reveal the upregulation of cytokine pathways and glia migration, upregulation of oxidative, ER stress and apoptosis markers, and the dysregulation of glucose metabolism with the downregulation of glycolysis and the upregulation of the oxidative phase of the pentose phosphate pathway. Glucose uptake was impaired in the chmru848 retina using the 2-NBDG glucose uptake assay. Following the overexpression of human PFKM, partial rescue was seen with the preservation of photoreceptors and RPE and increased glucose uptake, but without modifying glycolysis and oxidative stress markers. Therapies targeting glucose metabolism in CHM may represent a potential remedial approach. Full article
(This article belongs to the Special Issue Antioxidants and Retinal Diseases—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Cytokine pathway is upregulated in <span class="html-italic">chm<sup>ru848</sup></span> zebrafish retina. <span class="html-italic">chm<sup>ru848</sup></span> zebrafish samples are indicated in magenta and wt samples are indicated in cyan.</p>
Full article ">Figure 2
<p>Oxidative (<b>a</b>), ER stress (<b>b</b>) and apoptosis (<b>c</b>) markers are upregulated in the <span class="html-italic">chm<sup>ru848</sup></span> zebrafish retina. <span class="html-italic">chm<sup>ru848</sup></span> zebrafish are indicated in magenta and wt samples are indicated in cyan.</p>
Full article ">Figure 3
<p>Glycolysis is downregulated in favour of the pentose phosphate pathway in <span class="html-italic">chm<sup>ru848</sup></span> zebrafish retina. (<b>a</b>,<b>b</b>) Glycolysis pathways are affected in <span class="html-italic">chm<sup>ru848</sup></span> retina. (<b>a</b>) <span class="html-italic">chm<sup>ru848</sup></span> zebrafish are indicated in magenta and wt samples are indicated in cyan. (<b>b</b>) Scheme adapted from <a href="https://www.wikipathways.org/instance/WP1356" target="_blank">https://www.wikipathways.org/instance/WP1356</a> (accessed on 18 December 2024) [<a href="#B27-antioxidants-13-01587" class="html-bibr">27</a>,<a href="#B28-antioxidants-13-01587" class="html-bibr">28</a>]. Significantly downregulated genes are indicated in blue, while upregulated genes are in red. (<b>c</b>) Pentose phosphate pathway is upregulated in the <span class="html-italic">chm<sup>ru848</sup></span> retina compared to wt.</p>
Full article ">Figure 4
<p><span class="html-italic">PFKM</span> overexpression improves retinal phenotype but not photoreceptor cell death at 5 dpf. (<b>a</b>) Human <span class="html-italic">PFKM</span> expression was detected 5 days post-injection in <span class="html-italic">chm<sup>ru848</sup></span> and wt zebrafish. (<b>b</b>) mRNA-injected <span class="html-italic">chm<sup>ru848</sup></span> showed a more preserved photoreceptor layer (white asterisk) with better retinal lamination and a thicker RPE layer compared to the uninjected mutants (<span class="html-italic">n</span> = 9 zebrafish per group). Scale bar: 50 μm; scale bar zoom in: 20 μm. (<b>c</b>) Cell viability is unchanged after <span class="html-italic">PFKM</span> overexpression in wt and <span class="html-italic">chm<sup>ru848</sup></span> zebrafish (n = 9 zebrafish per group). Scale bar = 50 μm. ONL: outer nuclear layer; INL: inner nuclear layer. (<b>d</b>,<b>e</b>) Expression of anti-apoptotic (<b>d</b>) and pro-apoptotic (<b>e</b>) markers were analysed using RT-qPCR at 5 dpf in zebrafish eyes, from uninjected and injected wt and <span class="html-italic">chm<sup>ru848</sup></span>. Data are expressed as mean ± SEM from n = 3 (with 20 eyes per group). Statistical significance was determined by one-way ANOVA. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.005; **** <span class="html-italic">p</span> &lt; 0.0001. “●”—wt; “▲”—wt + <span class="html-italic">PFKM</span>; “▼”—<span class="html-italic">chm<sup>ru848</sup></span>; “◆”—<span class="html-italic">chm<sup>ru848</sup></span> + <span class="html-italic">PFKM</span>.</p>
Full article ">Figure 5
<p><span class="html-italic">PFKM</span> overexpression rescues 2-NBDG uptake in photoreceptors in <span class="html-italic">chm<sup>ru848</sup></span> retina at 5 dpf. Scale bar = 50 μm. 2-NBDG is in green; nucleus stained with DAPI (blue). * <span class="html-italic">p</span> &lt; 0.05; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p><span class="html-italic">PFKM</span> overexpression does not improve activity of the glycolysis pathway or oxidative stress 5 dpf in <span class="html-italic">chm<sup>ru848</sup></span> zebrafish. Expression of genes involved in the glycolysis pathway (<b>a</b>), in the pentose phosphate oxidative branch (<b>b</b>) and in oxidative stress (<b>c</b>) were analysed using RT-qPCR at 5dpf in zebrafish eyes, from uninjected and injected wt and <span class="html-italic">chm<sup>ru848</sup></span>. Data are expressed as mean ± SEM from n = 3 (with 20 eyes per group). Statistical significance was determined by One-way ANOVA. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.005; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001. “●”—wt; “▲”—wt + <span class="html-italic">PFKM</span>; “▼”—<span class="html-italic">chm<sup>ru848</sup></span>; “◆”—<span class="html-italic">chm<sup>ru848</sup></span> + <span class="html-italic">PFKM</span>.</p>
Full article ">
8 pages, 209 KiB  
Brief Report
Efficacy of Cefiderocol Against Endophthalmitis Isolates
by Brennan Schilling, Michael Hii, Hazel Q. Shanks, Eric G. Romanowski, Jonathan B. Mandell, Robert M. Q. Shanks and Michael Zegans
Antibiotics 2024, 13(12), 1236; https://doi.org/10.3390/antibiotics13121236 - 23 Dec 2024
Viewed by 411
Abstract
Background/Objectives: Endophthalmitis is an intraocular microbial infection that can lead to permanent blindness, even with prompt anti-microbial therapy. Multi-drug-resistant organisms are on the rise, potentially limiting the efficacy of current empiric antibiotic therapies of intravitreal ceftazidime and vancomycin. Cefiderocol is a recent FDA- [...] Read more.
Background/Objectives: Endophthalmitis is an intraocular microbial infection that can lead to permanent blindness, even with prompt anti-microbial therapy. Multi-drug-resistant organisms are on the rise, potentially limiting the efficacy of current empiric antibiotic therapies of intravitreal ceftazidime and vancomycin. Cefiderocol is a recent FDA- and EMA-approved antibiotic for multi-drug-resistant Gram-negative bacteria. Methods: To better understand its potential utility in the treatment of ocular infections, the MIC of cefiderocol was compared to ceftazidime and amikacin in endophthalmitis bacterial isolates using Epsilometer testing. Because vancomycin is commonly given concomitantly as part of empiric endophthalmitis treatment, possible synergistic and antagonistic effects of concomitant vancomycin and cefiderocol were also evaluated. Results: Cefiderocol was found to have lower MIC values compared to ceftazidime for Pseudomonadales or Enterobacterales species. When comparing the MICs of cefiderocol and vancomycin, there appeared to be no antagonism between the two antibiotics. Conclusions: This is the first report exploring the use of cefiderocol in endophthalmitis strains. The results of this study show this is a promising antibiotic for multi-drug-resistant Gram-negative organisms but further research is needed to investigate its intraocular safety profile. Full article
(This article belongs to the Collection Antibiotics in Ophthalmology Practice)
12 pages, 1673 KiB  
Article
Effects on Posture of a Two-Diopter Horizontal Prism Base Out on the Non-Dominant Eye
by Davide Marini, Giovanni Rubegni, Lorenzo Sarti, Alessandra Rufa, Marco Mandalà, Fabio Ferretti, Gian Marco Tosi and Mario Fruschelli
J. Clin. Med. 2024, 13(24), 7847; https://doi.org/10.3390/jcm13247847 - 23 Dec 2024
Viewed by 329
Abstract
Background/Objectives: Ocular proprioception is implicated in balance control and heterophoria is associated with abnormal posture, though previous research focused mainly on the role of vertical phoria and the use of vertical prisms. This study aims to evaluate whether ocular misalignment and prismatic [...] Read more.
Background/Objectives: Ocular proprioception is implicated in balance control and heterophoria is associated with abnormal posture, though previous research focused mainly on the role of vertical phoria and the use of vertical prisms. This study aims to evaluate whether ocular misalignment and prismatic correction of horizontal phoria affect posture. Methods: Sixty-nine (N = 69) young healthy subjects were included and equally divided by horizontal distance phoria: orthophoria (n = 23), esophoria (n = 23) and exophoria (n = 23). A prism of low power (two-diopter) was placed base out on the non-dominant eye, reducing misalignment in esophorics and increasing it in exophorics more than in orthophorics. Dynamic computerized posturography was performed with the sensory organization test protocol (SOT) of the EquiTest® NeuroCom® version 8 platform both without and with prism, always maintaining subjects unaware of prism use. A mixed model for repeated measures analysis of variance was run to evaluate the main effect of prism and the interaction effect of prism with baseline phoria. Results: Composite movement strategy score without prism was 88.1 ± 2.8% (ankle-dominant strategy) and slightly increased to 89.0 ± 3.1% with prism insertion (p = 0.004), further shifting toward ankle strategy. Composite equilibrium score without prism was 80.3 ± 6.5% and remained stable with prism insertion (81.3 ± 8.2%, p = 0.117), medio-lateral and antero-posterior projection of center of gravity did not displace significantly under prism insertion (p = 0.652 and p = 0.270, respectively). At baseline, posturographic parameters were statistically independent of individual phoria, and no significant interaction between prism insertion and individual phoria was documented for any parameters (p > 0.05 for all). Secondary analysis and pairwise comparisons confirmed that the effect of prism was strongly selective on condition SOT 5 (eyes-closed, platform sway-referenced) with improvement of equilibrium (70.4 ± 9.7% with prism vs. 65.7 ± 11.6% without) and more use of ankle strategy (81.6 ± 5.3% with prism vs. 78.2 ± 6.0% without), without any interaction of phoria and ocular dominance, while the other conditions were comparable with and without prism. Conclusions: A two-diopter prism base out on the non-dominant eye induces the body to use the ankle joint more independently of individual phoria, suggesting a small improvement in postural control, while maintaining oscillations of the center of gravity unaltered. Prism seems to enhance the function of vestibular system selectively. Phoria adjustments with prismatic correction enable intervention in postural behavior. Extraocular muscles could act as proprioceptors influencing postural stability. Full article
(This article belongs to the Section Ophthalmology)
Show Figures

Figure 1

Figure 1
<p>Distributions of horizontal distance phoria (Δ, prismatic diopter) and stereoacuity (log arcsec) <span class="html-italic">without</span> and <span class="html-italic">with prism</span>. Column and bar respectively represent median and interquartile range of phoria, and mean and standard error of stereoacuity. Wilcoxon signed-rank test statistical significance (<span class="html-italic">P</span>) and effect size (<span class="html-italic">r</span>) showed a significant exophoric shift in all groups after prism insertion. Two-way ANOVA statistical significance (<span class="html-italic">P</span>) and partial effect size (η<sup>2</sup>) of main effect (<span class="html-italic">Prism</span>) and interaction (<span class="html-italic">Prism × Phoria</span>) showed a significant small increase in logarithm of stereoacuity seconds of arc for all groups.</p>
Full article ">Figure 2
<p>Composite equilibrium (CES) and composite movement strategy scores (CMS) <span class="html-italic">without</span> (at baseline) and <span class="html-italic">with prism</span>, sorted by <span class="html-italic">baseline phoria</span>. A higher CSS denotes a better performance (less sway); a higher CMS indicates a more predominant ankle strategy (versus hip). Dot and bar represent mean and standard error, respectively. Two-way ANOVA statistical significance (<span class="html-italic">P</span>) and partial effect size (η<sup>2</sup>) of main effect (<span class="html-italic">Prism</span>) and interaction (<span class="html-italic">Prism × Phoria</span>) showed only a significant shift of CMS toward ankle strategy for all groups.</p>
Full article ">Figure 3
<p>Medio-lateral (ML/Px) and antero-posterior (AP/Py) projection of center of gravity (COG) <span class="html-italic">without</span> (at baseline) and <span class="html-italic">with prism</span>, sorted by <span class="html-italic">baseline phoria</span>. Right and forward displacements are positive, left and backward negative. Dot and bar represent mean and standard error, respectively. Two-way ANOVA statistical significance (<span class="html-italic">P</span>) and partial effect size (η<sup>2</sup>) of main effect (<span class="html-italic">Prism</span>) and interaction (<span class="html-italic">Prism × Phoria</span>) showed no significant displacement of COG, though COG-Px displacement of exophoric subjects was divergent (rightward) from esophoric and orthophoric subjects (leftward).</p>
Full article ">
Back to TopTop